Cancer Drug Discovery and Development Series Editor: Beverly A. Teicher
For further volumes: http://www.springer.com/series/7625
Rebecca G. Bagley Editor
The Tumor Microenvironment
Editor Rebecca G. Bagley Genzyme Corporation Framingham MA, USA
[email protected]
ISBN 978-1-4419-6614-8 e-ISBN 978-1-4419-6615-5 DOI 10.1007/978-1-4419-6615-5 Springer New York Dordrecht Heidelberg London Library of Congress Control Number: 2010934382 © Springer Science+Business Media, LLC 2010 All rights reserved. This work may not be translated or copied in whole or in part without the written permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in connection with any form of information storage and retrieval, electronic adaptation, computer soft-ware, or by similar or dissimilar methodology now known or hereafter developed is forbidden. The use in this publication of trade names, trademarks, service marks, and similar terms, even if they are not identified as such, is not to be taken as an expression of opinion as to whether or not they are subject to proprietary rights. Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)
Preface
The fact that tumors are composed of both tumor cells and host cells has long been known. These tumor-associated cells include vascular endothelial cells and pericytes, as well as inflammatory cells such as neutrophils, monocytes, macrophages, mast cells and eosinophils, and lymphocytes. The tumor cells also interact with stromal cells and with elements of the tissue extracellular matrix. What has been less appreciated is the role that these cells could have in modulating the growth, invasion, and metastasis of the tumor. Early on, the elements of what we now call the tumor microenvironment were considered to be more or less innocent bystanders to the role of the tumor cells as they grew and invaded local sites. Today, there is an increased understanding of the critical role of the tumor microenvironment as dramatically influencing the course of tumor development and dissemination. This volume represents a superb compilation of the latest thoughts and data regarding the role of each essential component of the tumor microenvironment in cancer development and progression. Perhaps, the earliest recognition of the role of nonmalignant cells as cancer regulators was the recognition that lymphocytes can participate in what was termed “immune surveillance” in the 1960s. Our understanding of tumor immunity has improved markedly since then, and there are now successful clinical studies showing the potential use of immune-based therapies in cancer treatment. The role of tumor infiltrating lymphocytes is nicely discussed here in the chapter by Fu and Yu, with a special emphasis on the potential prognostic implications of the presence of these cells at a tumor site. Natural killer cells, first shown to be involved in tumor surveillance by Herlyn in the 1980s, also have multiple effects on tumors as discussed in the chapter by Arai. Additional immune and inflammatory cells are also discussed. These include macrophages, which have effects on both tumor cells and blood vessels, and mast cells, which have a potent and somewhat underappreciated effect on tumor growth that is clarified in the chapter by Angelidou and Zhang. A picture emerges of the tumor as a city filled with diverse cellular inhabitants who operate in the role of a “zoning board” in telling the tumor how fast it can expand and into what spaces it can grow. An extremely important regulator of tumor growth and metastasis is the tumor vasculature. Again, tumors were known to be filled with functional blood vessels for many decades before Judah Folkman explained how they might influence tumor v
vi
Preface
growth. His theory of tumor angiogenesis predicted that anti-angiogenic agents might be developed to combat cancer and this prediction has been borne out. In many cases, the predominant signal for angiogenesis is tissue hypoxia and a very strong section of this book deals with the effects of hypoxia on the tumor and on the vascular components of the tumor. The potential translational aspects of this field is just emerging and a compelling discussion of the novel therapies that target the hypoxia induced transcription factor, HIF-1-alpha can be found in the book’s opening chapter by Rapisarda and Melillo. Along with the immune cells, inflammatory cells, and vascular cells, the tumor surrounds stromal cells including fibroblasts and other mesenchymal cells. The interplay between stromal cells and tumor cells is an extremely vibrant area of study and is nicely discussed in the chapter by Orimo. Of course, a principal function of these cells is to produce and remodel the elements of the extracellular matrix as described in the contribution by DeClerck. These proteins can present a barrier to tumor invasion and metastasis. Excellent and comprehensive discussions of the role of collagens and of fibronectin as well as their binding partners the integrins and tetraspanins can be found in the section on the extracellular matrix. The modulation of the matrix by cancer cells is discussed in detail in the chapter on matrix metalloproteases and cancer cell invasion by Zucker and Cao. The interplay of tumor cells and the microenvironment is mirrored in the complex interactions of tumor cells, fibroblasts, matrix elements, and the various cells that produce either matrix metalloproteases or their natural inhibitors. No contemporary discussion of the role of the tumor microenvironment could be complete without an in-depth examination of the existence and potential role of stem cells in the establishment of human tumors and in their metastatic dissemination. This is a relatively new field and comes with the attendant controversies that often characterize new areas of investigation. This can make for very exciting reading, as it does in the chapters on cancer stem cells by Giordano, on endothelial cell precursors by Shaked and on mesenchymal stem cells by Banerjee. It is safe to say that the exact contribution of all of these progenitor, precursor and stem cell populations to cancer is an emerging and truly important field of inquiry. These chapters beautifully capture the scientific foment that characterizes the study of these cells and their role in cancer. Perhaps, the most exciting aspect of the increased attention paid to the tumor microenvironment is the promise that new approaches to cancer therapy can be derived from our paying attention to these non-tumor elements and to their interactions with the tumor cells. This is beautifully captured throughout this book. One senses an emerging optimism that the tumor microenvironment is not only important, but that it represents a treasure trove of potential therapeutic opportunities such as the ones described in the chapter on secreted growth factors as therapeutic targets by Teicher. Whether one starts out as a skeptic, or as a devotee, this excellent and comprehensive compilation will educate the reader and stimulate new ideas about tumors as an integrated combination of cellular, secreted, stromal, and matrix elements that each represents a potential therapeutic target. Boston, MA
Bruce Zetter
Contents
Part I: Physiological Parameters 1 Combination Strategies Targeting Hypoxia Inducible Factor 1 (HIF-1) for Cancer Therapy.................................................... Annamaria Rapisarda and Giovanni Melillo
3
2 The Tumor Microenvironment: New Insights into Regulation of Tumor pH by Carbonic Anhydrases.............................. Pawel Swietach, Adrian L. Harris, and Richard D. Vaughan-Jones
23
3 Hypoxia, Gene Expression, and Metastasis........................................... Olga V. Razorenova and Amato J. Giaccia 4 Molecular Mechanisms Regulating Expression and Function of Cancer-Associated Carbonic Anhydrase IX..................... Jaromir Pastorek and Silvia Pastorekova 5 Glycolytic Pathway as a Target for Tumor Inhibition.......................... Weiqin Lu and Peng Huang
43
59 91
Part II: Malignant Cells 6 Aberrant DNA Methylation in Cancer Cells......................................... Toshikazu Ushijima
121
7 DNA Repair and Redox Signaling.......................................................... Mark R. Kelley, Millie M. Georgiadis, and Melissa L. Fishel
133
8 Cancer Stem Cells and Microenvironment............................................ Mario Federico and Antonio Giordano
169
9 Epithelial–Mesenchymal Transition in Development and Diseases...................................................................... Yadi Wu and Binhua P. Zhou
187 vii
viii
Contents
10 Invasion and Metastasis......................................................................... Douglas M. Noonan, Giuseppina Pennesi, and Adriana Albini 11 Dormancy of Disseminated Tumor Cells: Reciprocal Crosstalk with the Microenvironment.................................................. Paloma Bragado, Aparna C. Ranganathan, and Julio A. Aguirre-Ghiso
213
229
Part III: Vasculature and Stroma 12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome of Antiangiogenic Therapy: New Perspectives on an Ongoing Controversy.................................... Robert S. Kerbel, Francesco Bertolini, and Yuval Shaked 13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth....................................................................... Pravin J. Mishra and Debabrata Banerjee 14 Integrin Signaling in Lymphangiogenesis............................................ Barbara Garmy-Susini 15 Role of Pericytes in Resistance to Antiangiogenic Therapy................................................................................................... Koji Matsuo, Chunhua Lu, Mian M.K. Shazad, Robert L. Coleman, and Anil K. Sood 16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas........................................................................... Urszula M. Polanska, Kieran T. Mellody, and Akira Orimo
257
275 289
311
325
Part IV: Immune-Mediated Cells 17 Mast Cells and Tumor Microenvironment.......................................... Theoharis C. Theoharides, Konstantinos-Dionysios Alysandratos, Asimenia Angelidou, and Bodi Zhang
353
18 Macrophages in the Tumor Microenvironment.................................. Monica Escorcio-Correia and Thorsten Hagemann
371
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes........................................................................................... Ping Yu and Yang-Xin Fu
385
Contents
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation................................................................................ Adam Yagui-Beltrán, Qizhi Tang, and David M. Jablons 21 Natural Killer Cells for Adoptive Immunotherapy............................ Jonathan E. Benjamin and Sally Arai
ix
409 431
Part V: Extracellular Matrix 22 Fibronectin.............................................................................................. Andreas Menrad
457
23 Collagen in Cancer................................................................................. Janelle L. Lauer and Gregg B. Fields
477
24 Integrins and Cancer............................................................................. Laurie G. Hudson and M. Sharon Stack
509
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis................................................................................ Stanley Zucker and Jian Cao 26 Tetraspanins and Cancer Metastasis.................................................... Margot Zöller
531 555
Part VI: Secreted Proteins 27 Chemokines and Metastasis.................................................................. Kalyan C. Nannuru, Seema Singh, and Rakesh K. Singh
601
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis............................................................ Sonia B. Jakowlew
633
29 Cooperative Interactions Between Integrins and Growth Factor Signaling in Pathological Angiogenesis.................................... Jennifer Roth, Eric Tweedie, and Peter C. Brooks
673
30 The Extracellular Matrix and the Growth and Survival of Tumors......................................................................... Yves A. DeClerck
695
x
Contents
31 Secreted Growth Factors as Therapeutic Targets............................... Beverly A. Teicher
711
32 Adrenomedullin...................................................................................... Rebecca G. Bagley
733
Index................................................................................................................
749
Contributors
Julio A. Aguirre-Ghiso Division of Hematology and Oncology, Departments of Medicine and Department of Otolaryngology, Tisch Cancer Institute, Mount Sinai School of Medicine, One Gustave L Levy Place, Box 1079, New York, NY 10029, USA Adriana Albini IRCCS Multimedica, Via Fantoli 16/15, 20138 Milano, Italy Konstantinos-Dionysios Alysandratos, MD Tufts University School of Medicine, 136 Harrison Ave., Boston, MA 02111, USA Asimenia Angelidou, MD Tufts University School of Medicine, 136 Harrison Ave., Boston, MA 02111, USA Sally Arai, MD Division of Blood and Marrow Transplantation, Stanford School of Medicine, 300 Pasteur Drive, H3249 MC 5623, Stanford, CA 94305, USA, Rebecca G. Bagley Genzyme Corporation, 49 New York Ave, Framingham, MA 01701, USA Debabrata Banerjee Department of Medicine and Pharmacology, Robert Wood Johnson Medical School, The Cancer Institute of New Jersey, University of Medicine and Dentistry of New Jersey, 195 Little Albany Street, New Brunswick, NJ 08901, USA Jonathan E. Benjamin Division of Blood and Marrow Transplantation, Stanford University Medical Center, Stanford, CA, USA
xi
xii
Contributors
Francesco Bertolini, MD/PhD Department of Hematology-Oncology, European Institute of Oncology, Milan, Italy Paloma Bragado, PhD Division of Hematology and Oncology, Department of Medicine and Department of Otolaryngology, Mount Sinai School of Medicine – NYU, New York, NY 10029, USA Peter C. Brooks Maine Medical Center Research Institute, Center for Molecular Medicine, 81 Research Drive, Scarborough, ME, USA Jian Cao, PhD Stony Brook University, Room 004, Life Sciences Building, Stony Brook, NY 11794-5200, USA Robert L. Coleman, MD Department of Gynecologic Oncology, M. D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA; Center for RNA Interference and Non-Coding RNA, University of Texas, Houston, TX, USA Yves A. DeClerck Departments of Pediatrics and Biochemistry and Molecular Biology, University of Southern California and The Saban Research Institute of Childrens Hospital Los Angeles, Los Angeles, CA 20027, USA Monica Escorcio-Correia, PhD Barts and The London School of Medicine and Dentistry, Institute of Cancer, Charterhouse Square, London, UK EC1M 6BQ Mario Federico, MD Sbarro Health Research Organization, Temple University, BioLife Science Building, Suite 333, 1900 N. 12th Street, Philadelphia, PA 19122, USA; Section of Clinical Oncology, University of Palermo, Palermo, Italy Gregg B. Fields Department of Biochemistry, University of Texas Health Science Center, 7703 Floyd Curl Drive, San Antonio, TX 78229, USA Melissa L. Fishel, PhD Herman B. Wells Center for Pediatric Research, Indiana University, 980 W. Walnut Street, Indianapolis, IN 46202, USA
Contributors
Yang-Xin Fu, MD/PhD Department of Pathology and Section of Dermatology/ Department of Medicine, University of Chicago, 5841 S. Maryland Ave, MC 3083, Chicago, IL 60637, USA Barbara Garmy-Susini Unité mixte Inserm U858, Institut de Médecine Moléculaire de Rangueil, IFR 150, 1, Avenue Jean Poulhès, BP 84225 31432, Toulouse Cedex 4, France Millie M. Georgiadis, PhD Herman B. Wells Center for Pediatric Research, Indiana University, 980 W. Walnut Street, Indianapolis, IN 46202, USA Amato J. Giaccia Division of Radiation and Cancer Biology, Stanford University, Stanford, CA 94305, USA Antonio Giordano Biology Department, Sbarro Health Research Organization, Temple University, Philadelphia, PA, USA Thorsten Hagemann Barts and The London School of Medicine and Dentistry, Institute of Cancer, Charterhouse Square, London, EC1M 6BQ, UK Adrian L. Harris Weatherall Institute of Molecular Medicine, Oxford University, Oxford OX3 9DS, UK Peng Huang Department of Molecular Pathology, The University of Texas MD, Anderson Cancer Center, 1515 Holcombe Boulevard, Houston, TX 77030, USA Laurie G. Hudson, PhD University of New Mexico College of Pharmacy, Albuquerque, NM, USA David Jablons, MD Department of Surgery, University of California San Francisco, San Francisco, CA 94143, USA Sonia B. Jakowlew National Cancer Institute, Center for Cancer Training, Cancer Training Branch, Bethesda, MD 20892, USA
xiii
xiv
Contributors
Mark R. Kelley Department of Pediatrics, Section of Hematology/Oncology, Herman B Wells Center for Pediatric Research, Indiana University, Indianapolis, IN 46202, USA Robert S. Kerbel, PhD Department of Cellular and Molecular Biology, Sunnybrook Health Sciences Centre, University of Toronto, Toronto, ON, Canada Janelle L. Lauer, PhD The University of Texas Health Science Center, 7703 Floyd Curl Drive, San Antonio, TX 78229-3900, USA Chunhua Lu, MD/PhD Department of Gynecologic Oncology, M. D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA Weiqin Lu, PhD MD Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA Koji Matsuo, MD Department of Gynecologic Oncology, M. D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA Giovanni Melillo SAIC-Frederick, Inc., DTP-Tumor Hypoxia Laboratory, Developmental Therapeutics Program, National Cancer Institute – Frederick, Building 432, Room 218, Frederick, MD 21702, USA Kieran T. Mellody, BSc./MPhil Paterson Institute for Cancer Research, The University of Manchester, Wilmslow Road, Manchester, England M20 4BX Andreas Menrad Antibody Therapeutics Genzyme Europe Research, 310 Cambridge Science Park, Milton Road, Cambridge, CB4 OWG, UK Pravin J. Mishra, MS Department of Medicine, Robert Wood Johnson Medical School, The Cancer Institute of New Jersey, University of Medicine and Dentistry of New Jersey, 195 Little Albany St, New Brunswick, NJ 08901, USA Kalyan C. Nannuru, PhD Department of Pathology and Microbiology, University of Nebraska Medical Center, 985900 Nebraska Medical Center, Omaha, NE 68198-5900, USA
Contributors
xv
Douglas M. Noonan, PhD Universita degli Studi dell’Insubria, Varese, Italy Akira Orimo CR-UK Stromal-Tumour Interaction Group, Paterson Institute for Cancer Research, The University of Manchester, Wilmslow Road, Manchester, M20 4BX, UK Jaromir Pastorek Centre of Molecular Medicine, Slovak Academy of Sciences, Dubravska cesta 9, 845 05 Bratislava, Slovak Republic Silvia Pastorekova Centre of Molecular Medicine, Slovak Academy of Sciences, Dubravask cesta 9, Bratislava 845 05, Slovak Republic Giuseppina Pennesi, MD IRCCS Multimedica, Via Fantoli 16/15, 20138 Milano, Italy Urszula M. Polanska, Msc/PhD Paterson Institute for Cancer Research, The University of Manchester, Wilmslow Road, Manchester, England M20 4BX Aparna C. Ranganathan, PhD Division of Hematology and Oncology, Department of Medicine and Department of Otolaryngology, Mount Sinai School of Medicine – NYU, New York, NY 10029, USA Annamaria Rapisarda, PhD SAIC-Frederick, Inc., DTP-Tumor Hypoxia Laboratory, Developmental Therapeutics Program, National Cancer Institute – Frederick, Building 432, Room 218, Frederick, MD 21702, USA Olga V. Razorenova, PhD Division of Radiation and Cancer Biology, Stanford University School of Medicine, 269 Campus Drive, Stanford, CA 94305, USA Jennifer M. Roth, MS Center for Molecular Medicine, Maine Medical Center Research Institute, 81 Research Drive, Scarborough, ME 04074, USA Yuval Shaked Department of Molecular Pharmacology, Rappaport Faculty of Medicine, Technion – Israel Institute of Technology, Haifa, Israel
xvi
Contributors
M. Sharon Stack University of Missouri School of Medicine, M214C Medical Sciences Building, 1 Hospital Drive, Columbia, MO 65212, USA Mian M. K. Shazad, MS/MD Department of Gynecologic Oncology, M.D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA; Department of Obstetrics and Gynecology, Baylor College of Medicine, Houston, TX, USA Rakesh K. Singh Department of Pathology and Microbiology, University of Nebraska Medical Center, 985900 Nebraska Medical Center, Omaha, NE 68198-5900, USA Seema Singh, PhD Department of Pathology and Microbiology, University of Nebraska Medical Center, 985900 Nebraska Medical Center, Omaha, NE 68198-5900, USA Anil K. Sood Department of Gynecologic Oncology, Cancer Biology, M. D. Anderson Cancer Center, 1515 Holcombe Blvd, Houston, TX 77030, USA Center for RNA Interference and Non-Coding RNA, University of Texas, Houston, TX, USA Pawel Swietach, DPhil Deparment of Physiology, Anatomy & Genetics, Oxford University, Oxford OX1 3PT, UK Qizi Tang, PhD Department of Surgery, University of California San Francisco, San Francisco, CA 94143, USA Beverly A. Teicher Genzyme Corporation, 49 New York Avenue, Framingham, MA 01701-9322, USA
[email protected] Theoharis C. Theoharides Tufts University School of Medicine, 136 Harrison Ave., Boston, MA 02111, USA Eric Tweedie, BS Center for Molecular Medicine, Maine Medical Center Research Institute, 81 Research Drive, Scarborough, ME 04074, USA
Contributors
xvii
Toshikazu Ushijima National Cancer Center Research Institute, 5-1-1 Tsukiji, Chuo-ku, Tokyo, Japan Richard D. Vaughan-Jones, PhD Deparment of Physiology, Anatomy & Genetics, Oxford University, Oxford OX1 3PT, UK Yadi Wu, PhD Department of Molecular and Biomedical Pharmacology, Markey Cancer Center and University of Kentucky School of Medicine, BBSRB room B336, 741 South Limestone, Lexington, KY 40506-0509, USA Adam Yagui-Beltrán Department of Surgery, Division of Adult Thoracic Surgery, The Helen Diller Family Comprehensive Cancer Center Thoracic Surgery, University of California San Francisco, 1600 Divisadero, Room A-743, San Francisco, CA 94143-1724, USA Ping Yu The Committee on Immunology and Department of Pathology and Section of Dermatology/Department of Medicine, University of Chicago, 5841 S. Maryland Ave, Chicago, IL 60637, USA Bruce Zettler, PhD Childrens’s Hospital, Boston Harvarel Medical School, 300 Lougwood Ave., Boston, MA 02115, USA Bodi Zhang, MD/MPH Tufts University School of Medicine, 136 Harrison Ave., Boston, MA 02111, USA Binhua P. Zhou Markey Cancer Center, University of Kentucky School of Medicine, Lexington, KY, USA and Department of Molecular and Cellular Biochemistry, University of Kentucky School of Medicine, Lexington, KY, USA
xviii
Margot Zöller Department of Tumor Cell Biology, University Hospital of Surgery, Im Neuenheimer Feld 365, D-69120, Heidelberg, Germany and Department of Tumor Progression and Immune Defense, German Cancer Research Center, Heidelberg, Germany and Department of Applied Genetics, University of Karlsruhe, Karlsruhe, Germany Stanley Zucker Veterans Affairs Medical Center, Northport, NY 11768, USA and Stony Brook University, Stony Brook, NY 11794, USA
Contributors
Part I
Physiological Parameters
Chapter 1
Combination Strategies Targeting Hypoxia Inducible Factor 1 (HIF-1) for Cancer Therapy Annamaria Rapisarda and Giovanni Melillo
Abstract Solid tumors often present regions of decreased oxygen levels (hypoxia) due to an imbalance between increased oxygen consumption and insufficient oxygen delivery from the aberrant tumor vasculature. Intratumor hypoxia is associated with altered cellular metabolism, an invasive and metastatic phenotype, as well as resistance to radiation and chemotherapy. The discovery of Hypoxia Inducible Factor-1 (HIF-1), a transcription factor critically involved in cellular responses to hypoxia and tumor progression, has provided evidence of a potential molecular target of intratumor hypoxia that could be exploited for the development of novel cancer therapeutics. A growing number of small molecule inhibitors of HIF-1, which act by distinct molecular mechanisms, have been described so far. However, HIF-1 expression in human cancers is focal and heterogeneous, consistent with the possibility that single agent HIF-1 inhibitors may have limited clinical activity. It is then plausible that combination strategies aimed at maximizing the clinical potential of HIF-1 inhibition may be more effective. We will discuss current approaches used for targeting HIF-1, emphasizing in particular opportunities for rationally designed combination strategies aimed at exploiting vulnerable features of the tumor microenvironment.
Introduction Whereas each tumor type reflects a heterogeneous collection of genetic and epigenetic alterations, hypoxia represents a unifying characteristic of human tumors, being a widely represented feature of the tumor microenvironment. Intratumor hypoxia not only affects cells that are located far from a functional blood vessel but also originates from an imbalance between increased oxygen consumption of rapidly G. Melillo (*) DTP-Tumor Hypoxia Laboratory, Bldg. 432, Room 218, SAIC – Frederick, Inc., NCI at Frederick, Frederick, MD 21702, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_1, © Springer Science+Business Media, LLC 2010
3
4
A. Rapisarda and G. Melillo
growing tumor cells and inadequate oxygen supply from a structurally and functionally abnormal tumor vasculature. The fluctuation of oxygen levels in the tumor microenvironment results in the induction of an array of responses in tumor cells, such as inactivation of apoptotic pathways and activation of pro-survival pathways, induction of a more invasive and metastatic phenotype, switch to a glycolytic metabolism and induction of angiogenesis, responses aimed at providing adaptation to, or escape from, the hypoxic microenvironment. Clinically, hypoxia is associated with resistance to standard treatment, including chemotherapy and radiation therapy, and is predictive of metastasis and poor outcome in a variety of tumor types (Hockel and Vaupel 2001; Brown and Giaccia 1998; Brizel et al. 1996; Semenza 2007). Hypoxia controls the expression of hundreds of genes, many of which are regulated by Hypoxia Inducible Factor-1 (HIF-1), a master regulator of the transcriptional response to oxygen deprivation (Semenza 2007; Manalo et al. 2005). HIF-1 is a heterodimeric protein comprised of a constitutively expressed HIF-1b (also known as aryl hydrocarbon receptor nuclear translocator, ARNT) subunit and a HIF-a subunit, which is regulated by oxygen concentrations (Wang et al. 1995; Wang and Semenza 1995). Under normoxic onditions, HIF-a is rapidly hydroxylated on two proline residues by oxygen-dependent prolyl hydroxylases (PHDs) and targeted for ubiquitylation and proteasomal degradation in a VHL-dependent fashion. In addition, HIF-1a can be hydroxylated on an asparagine residue (Asn803) by Factor Inhibiting HIF-1 (FIH), hence inhibiting the recruitment of the co-activator p300/CBP and transcriptional activity. In contrast, under hypoxic conditions the HIF-a subunit is stabilized and translocates to the nucleus where it dimerizes with HIF-1b and, by binding to hypoxia responsive elements (HREs), activates transcription of hundreds of target genes involved in key steps of tumorigenesis, including angiogenesis, metabolism, proliferation, metastasis, and differentiation (Semenza 2008) (Fig. 1.1). Overwhelming evidence indicates that HIF-a (HIF-1a and/or HIF2a) is indeed over-expressed in the majority of human cancers (Zhong et al. 1999; Talks et al. 2000; Bos et al. 2001), where it is associated with patient mortality and poor response to treatment (Table 1.1) (Aebersold et al. 2001; Birner et al. 2000, 2001b; Bos et al. 2003; Koukourakis et al. 2006). Notably, increased HIF-1a expression is associated with a variety of genetic alterations frequently detected in human cancers, including loss of function of tumor suppressor genes, such as VHL, PTEN, and CDKN2A (which encodes ARF) or oncogenic gain of function, such as RAS, SRC, and BCR-ABL (Semenza 2003). In addition, the involvement of receptor tyrosine kinase-dependent signaling pathways, including EGFR, HER2/Neu, and the PI3K/AKT/mTOR and MAPK pathways, in the induction of HIF-1a expression also suggests that many of these pathways converge on or implicate HIF-1 in mediating cell survival and growth (Pouyssegur et al. 2006; Majumder et al. 2004; Pore et al. 2006; Hudson et al. 2002). Therefore, HIF-1 has become an attractive target for the development of novel cancer therapeutics and despite the challenges associated with the discovery and development of small molecule inhibitors of transcription factors, many HIF-1 inhibitors which may potentially be useful for cancer therapy have been described so far (Onnis et al. 2009; Melillo 2007; Welsh et al. 2006; Melillo 2006).
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
5
Angiogenic factors
(VEGF, FGF2, PDGF, PIGF, Tie2)
VHL binding Proteosomal degradation
Inhibition of transcriptional activity
P402 OH
Drug resistance
(Telomerase, Oct4)
(MDR1/ABCB-1, ABCG-2)
Glycolitic metabolism, pH regulation
HIF- α P564 OH HIF- α N803OH PHDs
RTKs/mTOR
(GLUT1, PDK1, PGK1, HK2, LDH-A, CA9, MCT4)
FIH
Survival, Proliferation
HIF- α
HIF-1 translation
mRNA transcription
O2 concentration
Immortalization and stem cells
β
α
Dimerization Cofactors recruitment DNA binding
(Cyclin D1, VEGF, IGF2, TGF-α)
Migration (CXCR4, SDF-1 α)
Invasion Metastasis (UPAR, Met, LOX)
Fig. 1.1 Regulation of Hypoxia Inducible factor-a (HIF-a) and HIF-1 dependent gene expression. Under normal oxygen conditions, HIF-a is continuously translated and rapidly degraded through the VHL-proteasome pathway following hydroxylation of proline residues by PHDs. Factor inhibiting HIF-1 mediates the hydroxylation of an asparagine residue, thus inhibiting the recruitment of the co-activator p300/CBP and HIF-1 transcriptional activity. All these processes are inhibited when oxygen concentration decreases, hence the HIF-a subunit is stabilized, translocates to the nucleus where it dimerizes with HIF-1b and induces the transcription of a large number of genes involved in crucial aspects of tumorigenesis. In addition, receptor tyrosine kinase-dependent signaling pathways can increase the rate of HIF-1a translation, at least in part, through the activation of mTOR
Small Molecule Inhibitors of HIF-1 The approach most widely used so far to identify HIF-1 inhibitors has been cell-based high throughput screens (HTS) using reporter gene assays (Rapisarda et al. 2002). The main advantage of cell-based HTS is the potential for the identification of unrecognized pathways associated with the activation of the target being investigated. However, determining the mechanism of action of active “hits” may be time consuming and chances to identify highly specific inhibitors is very low. Indeed, the majority of HIF-1 inhibitors described so far lack specificity, which complicates the pharmacological validation of HIF-1 inhibitors for cancer therapy, and either target signaling pathways that are involved in HIF-1 regulation or inhibit HIF-1a protein accumulation and/or transcriptional activity. Examples of agents that target signaling pathways implicated in the HIF-1a expression include
6
A. Rapisarda and G. Melillo
Table 1.1 Human cancers in which HIF-1a over-expression has been associated with patient mortality Cancer type
Reference
Astrocytoma (diffuse) Bladder (superficial urothelial)a Bladder (transitional cells) Breast Breast (c-Erb2 positive) Breast (positive linfonode) Breast (negative linfonode) Cervix (early stage) Cervix (radiation therapy) Cervix (grade IB-IIIB, radiation therapy) Endometrial Gastric Gastrointestinal stromal tumor (stomach) Lung (nonsmall cell lung carcinoma) Melanoma (malignant)b Oligodendrioglioma Oropharynx-squamous cell carcinoma Ovariana Pancreas
Korkolopoulou et al. (2004) Theodoropoulos et al. (2004) Theodoropoulos et al. (2005) Vleugel et al. (2005) and Dales et al. (2005) Giatromanolaki et al. (2007) Schindl et al. (2002) Bos et al. (2003) Birner et al. (2000) Burri et al. (2003) Bachtiary et al. (2003) Sivridis et al. (2002) Griffiths et al. (2007) Takahashi et al. (2003) Swinson et al. (2004) Giatromanolaki et al. (2003) Birner et al. (2001a) Aebersold et al. (2001) Birner et al. (2001b) Sun et al. (2007)
Combination of HIF-1a over-expression and mutant p53 HIF-2a over-expression
a
b
inhibitors of mTOR (Majumder et al. 2004; Hudson et al. 2002), AKT (Tan et al. 2005), Her2/Neu (Laughner et al. 2001), EGFR (Pore et al. 2006; Koukourakis et al. 2003; Zhong et al. 2000), and Bcr-Abl (Mayerhofer et al. 2002) (Fig. 1.2). Since HIF-1a accumulation is controlled primarily at the level of protein translation or protein degradation, many of the HIF-1 inhibitors identified thus far do indeed decrease HIF-1a protein levels by a number of distinct molecular mechanisms. HIF-1 inhibitors that affect protein translation, although by mechanisms not always clearly elucidated, include topoisomerase I (topotecan and EZN-2208, a pegylated form of SN38) (Rapisarda et al. 2004a, b; Sapra et al. 2008) and topoisomerase II inhibitors (NSC644221) (Creighton-Gutteridge et al. 2007), microtubule stabilizing (taxanes and epothilones) and destabilizing agents (Mabjeesh et al. 2003; Escuin et al. 2005), cardiac glycosides (Zhang et al. 2008), PX-478, which also increases HIF-1a degradation (Koh et al. 2008) and farnesyltransferase inhibitors (Han et al. 2005). Agents that appear to affect primarily HIF-1a degradation include inhibitors of the chaperon protein Hsp90 (17-AAG, 17-DMAG) (Isaacs et al. 2002; Mabjeesh et al. 2002), and HDAC inhibitors (vorinostat, LAQ824, FK228) (Ellis et al. 2009), which may also block HIF-1a transcriptional activity (Kong et al. 2006), and YC-1, which activates soluble guanylate cyclase (sGC) (Kim et al. 2006a). Another mechanism by which HIF-1 may be inhibited is at the level of transactivation. In addition to chetomin (Kung et al. 2004), which
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
HIF-1 β
α
Co-factors Recruitment Transcriptional activity DNA binding
NSC50352
7 Chetomin Bortezomib Vorinostat Amphotericin
HIF-1-dependent gene expression
Echinomycin Polyamides Doxorubicin
Dimerization
HIF-1 β 17AAG,17DMAG YC-1, SCH66336 NS398, ibuprofen celecoxib
HIF-1 α
RTKs/mTOR
Translation Degradation
Temsirolimus Everolimus Gefitinib Erlotinib Cetuximab Herceptin
TPT,SN38 NSC644221, Digoxin Candidaspongiolide PX478, 2ME2
Proteasome HIF-1 α mRNA
HIF-1 α antisense RNA Aminoflavone
Fig. 1.2 Signaling pathways targeted by small molecule inhibitors of HIF-1
was originally identified as a small molecule that interferes with the interaction between HIF-1a C-TAD and p300/CBP, inhibition of the proteasome (bortezomib, Velcade®) has also been associated with blockade of HIF-1 transcriptional activity, despite the paradoxical increase of HIF-1a protein, although the exact mechanism of action remains to be fully elucidated (Shin et al. 2008; Kaluz et al. 2008). Finally, modulation of HIF-1a mRNA levels using antisense oligonucleotides (EZN-2968) (Greenberger et al. 2008) or the AhR ligand aminoflavone (Terzuoli et al. submitted) also leads to inhibition of HIF-1 expression. However, in order to identify HIF-1 small molecule inhibitors with a higher degree of selectivity, efforts are ongoing to target protein–protein interaction, a challenging yet potentially highly rewarding strategy, in an attempt to interfere with HIF-a/HIF-1b dimerization (Yang et al. 2005; Park et al. 2006; Scheuermann et al. 2009). In addition, examples of HIF-1 inhibitors that affect HIF-1 DNA binding to the HRE have been provided, including synthetic polyamides (Olenyuk et al. 2004), echinomycin (Kong et al. 2005), and anthracyclines (Lee et al. 2009), which represent a conceptually attractive strategy to inhibit HIF-1. Many of the HIF-1 inhibitors described above are either FDA approved or are in early clinical development (Table 1.2). Results of ongoing clinical trials of HIF-1 inhibitors in patients with cancer will provide information regarding both inhibition of HIF-1 and potential clinical activity of this emerging therapeutic approach.
8
A. Rapisarda and G. Melillo
Table 1.2 HIF-1 inhibitors FDA approved or in clinical development with potential mechanism of action Agent
Target
Mechanism
Reference
EZN-2968 Aminoflavone
HIF-1a mRNA AhR
Greenberger et al. (2008) Terzuoli, submitted
Topotecan EZN-2208 PX-478 Erlotinib, gefitinib, cetuximab Trastuzumab Temsirolimus, everolimus 17AAG, 17DMAG
Topoisomerase 1 Topoisomerase 1 unknown EGFR
Degradation HIF-1a mRNA processing Translation Translation Translation Signaling
Her2/Neu mTOR
Translation Translation
HSP90
Degradation
Doxorubicin Bortezomib
DNA Proteasome
Vorinostat, LAQ824, FK228
HDAC
DNA binding Transcriptional activity Degradation, transcriptional activity
Rapisarda et al. (2004b) Sapra et al. (2008) Welsh et al. (2004) Pore et al. (2006) and Luwor et al. (2005) Koukourakis et al. (2003) Del et al. (2006) and Wan et al. (2006) Isaacs et al. (2002) and Mabjeesh et al. (2002) Lee et al. (2009) Kaluz et al. (2006) Kong et al. (2006), Mie et al. (2003), and Qian et al. (2006)
Targeting HIF-1: Single Agent or Combination? The potential therapeutic benefit of HIF-1 inhibition has been extensively documented in preclinical models using both genetic and pharmacological tools, and consistent inhibition of tumor growth and angiogenesis has been demonstrated in different tumor types (Semenza 2003; Melillo 2006). However, the lack of selective HIF-1 inhibitors and reliable biomarkers associated with HIF-1 inhibition in tumor tissue, associated with the hardly predictive nature of results obtained in xenograft models, have significantly hampered the validation of HIF-1 as a target in human cancers. For instance, it is still largely unclear whether tumor regression should be an expected, although desirable, outcome of single agent HIF-1 inhibition. Evidence has been provided that HIF-1 inhibition might be most effective when used in early, rather than late, stages of tumor progression (Li et al. 2005), a situation rarely encountered in clinical oncology practice, where the majority of patients present with metastatic disease. In addition, even conceptually, inhibition of HIF-1 expressing cells, as a single agent strategy, may not be particularly effective for a number of reasons, including the focal and heterogeneous expression of HIF-1a in solid tumors, the reliance of cancer cells on HIF-1-independent pathways in oxygenated areas of the tumor, and the overall well-established redundancy of oncogenic signaling pathways that may be hardly affected by single agent strategies. In addition, the broad involvement of HIF-1 in a number of biological pathways that are
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
9
relevant for tumorigenesis provides unique opportunities for combination therapies that may take advantage of the different approaches available in the current therapeutic armamentarium. A better understanding of the role that HIF-1 plays in the biology of human cancers and in the context of available therapeutic strategies will provide valuable information for the design of rational combinations that may target vulnerable aspects of the tumor microenvironment.
Molecularly Targeted Agents and HIF-1 Inhibition Combination strategies aimed at targeting multiple and redundant signaling pathways that contribute to oncogenesis are a mainstay of current therapeutic efforts in oncology for the potential to both prevent the emergence of resistant clones and translate in a better therapeutic outcome. Indeed, combinations of molecularly targeted agents, with or without chemotherapy, are currently being tested in virtually any tumor type with the goal of assessing safety of administration and potential therapeutic efficacy. As mentioned above, several novel targeted agents currently in clinical development or FDA approved also inhibit HIF-1, raising the possibility that combination strategies incorporating these agents may already provide examples in which inhibition of HIF-1 may potentially contribute to the therapeutic efficacy. Experimental evidence in support of this conclusion has been provided by combination of the mTOR inhibitor rapamycin with the HDAC inhibitor LBH589, which showed increased antitumor activity in prostate and renal cell cancer xenografts models (Verheul et al. 2008). Interestingly, the concomitant inhibition of HDACs and mTOR resulted in a greater inhibition of HIF-1a protein levels relative to what was achieved by each agent alone, possibly due to the activity of these inhibitors at different levels of HIF-1a regulation including translation, protein stability, and transcriptional activity. In addition, experiments conducted in colon cancer xenograft models using low doses of rapamycin in combination with the camptothecin analog irinotecan (a topoisomerase I inhibitor) showed a marked increase in antitumor activity associated with a profound inhibition of HIF-1a protein accumulation that was not observed with either agent alone (Pencreach et al. 2009). Interestingly, previous studies in a human glioma xenograft model had suggested that daily, but not intermittent, administration of single agent topotecan, a topoisomerase I HIF-1 inhibitor, downregulated HIF-1a expression in tumor tissue and caused inhibition of tumor growth and angiogenesis (Rapisarda et al. 2004c). These examples emphasize how HIF-1 inhibition may be an underappreciated therapeutic consequence of molecularly targeted agents currently in clinical development. A better understanding of the potential contribution of HIF-1 inhibition to therapeutic activity of these combinations may help in selecting patients more likely to respond and in designing optimal schedules of administration to maximize inhibition of HIF-1-dependent responses.
10
A. Rapisarda and G. Melillo
Hypoxic Cells Are More Resistant to Chemotherapy and Radiation Therapy It has long been recognized that hypoxic cells are more resistant to conventional therapeutic approaches, including chemotherapy and radiation therapy. Likewise, it is well appreciated that resistant hypoxic cells may contribute to treatment failure and poor prognosis of patients whose cancers are more hypoxic. Resistance to radiotherapy and chemotherapy in hypoxic tumors has been attributed to a multitude of mechanisms, including direct effects (decreased generation of oxygendependent radical formation or decreased drug accessibility to the more distant hypoxic cells) and indirect effects (clonal selection of cells adapted to hypoxia) (Brown and Giaccia 1998; Vaupel et al. 2001). Hypoxia exerts selective pressure on cells for loss of p53, regulates the transcription of several genes pivotal for the selection of a resistant phenotype, such as multidrug resistance gene 1 (MDR-1) and ABCG-2 and ABCB-1 (ABC transporters), induces expression of VEGF and other pro-angiogenic factors (important for the survival of endothelial cells following radiation), downregulates genes involved in DNA mismatch repair (MSH2, MSH6, NSB1, and BRCA1) leading to increased genomic instability (Huang 2008), and finally induces genes important for survival of cancer cells (Bertout et al. 2008; Huang 2008). Many of the effects described above are mediated by HIF-1; therefore, it is conceivable that inhibition of HIF-1 may contribute in a number of ways to combination therapies by counteracting mechanisms that are activated in the tumor microenvironment to generate resistance and by sensitizing cancer cells to available therapeutic strategies.
Combination of HIF-1 Inhibitors with Chemotherapy The potential contribution of HIF-1a to cellular resistance to chemotherapy has been suggested by evidence in both experimental models and the clinical setting. Initial clinical observations indicated that high levels of HIF-1a expression were associated with incomplete responses to radiation and chemotherapy in patients with head and neck squamous cancers (Aebersold et al. 2001; Koukourakis et al. 2003). Formal evidence that HIF-1a may be implicated in mediating resistance to conventional chemotherapeutic agents was provided by studies showing that MDR-1, one of the most widely expressed efflux pumps that mediate resistance to chemotherapy, is a direct HIF-1 target gene (Comerford et al. 2002). In addition, it was shown that HIF-1a deficient mouse embryonic fibroblasts were more sensitive to carboplatin, etoposide (a topoisomerase II inhibitor), and irradiation (agents that cause double-strand DNA breaks) but not to the topoisomerase 1 inhibitor SN38 (that generates single-strand DNA breaks) (Unruh et al. 2003), suggesting that HIF-1-mediated resistance to chemotherapy cannot simply be explained by upregulation of MDR-1. Resistance to etoposide, a
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
11
topoisomerase II inhibitor, also appears to correlate with HIF-1-dependent induction of anti-apoptotic genes and downregulation of topoisomerase IIa (Sermeus et al. 2008; Sullivan and Graham 2009). The potential synergistic activity of HIF-1 inhibition in combination with chemotherapy has been formally tested in a glioma xenograft model in which HIF-1a expression was downregulated by an inducible shRNA system (Li et al. 2006). Earlier evidence had indicated that HIF-1a downregulation by shRNA effectively inhibited tumor growth, although the effect was more pronounced in early rather than late stages of tumor progression (Li et al. 2005). The combination of the cytotoxic drug temozolomide with HIF-1a knockdown exhibited a super-additive therapeutic effect compared with either approach alone, demonstrating that HIF-1 inhibition impacted on sensitivity to chemotherapy (Li et al. 2006). However, the same authors showed no benefit of a combination between HIF-1 inhibition and the angiogenesis inhibitor ABT-869 (a multitargeted receptor tyrosine kinase inhibitor). It is interesting to note that following treatment with ABT-869, large tumors that are less likely to be responsive to HIF-1a inhibition became sensitive to HIF-1a knockdown, suggesting that anti-angiogenic treatment might partially alleviate the “resistance” of large tumors to HIF-1a knockdown and sensitize tumors to HIF-1 inhibition. Further studies in preclinical models, ultimately with validation in clinical trials, are required to fully explore the potential of HIF-1 inhibition to modulate cellular sensitivity to chemotherapeutic agents and to block stress pathways induced by chemotherapy in the tumor microenvironment. Combination of HIF-1 Inhibitors with Radiation Therapy Hypoxia is an important factor contributing to tumor radioresistance, a phenomenon initially explained by the reduced generation of oxygen radicals in hypoxic tissues and consequent decrease in radiation-induced DNA damage (Moeller et al. 2007). In reality, more complex biological pathways are implicated in the effects of radiation therapy on hypoxic tumors. Clinical studies have suggested that high levels of HIF-1a expression correlate with poor local control in patients with oropharyngeal cancers (Aebersold et al. 2001) and evidence from preclinical studies has indicated that HIF-1 deficient tumors are more radiosensitive than their wild-type counterpart (Zhang et al. 2004; Moeller et al. 2005; Williams et al. 2005). Moreover, it has been shown that radiation increases HIF-1a levels and induces the expression of HIF-1-dependent genes in tumors (Moeller et al. 2004), consistent with the possibility that HIF-1 activation may be part of compensatory pathways that might mediate resistance to radiation therapy. However, further studies have shown that HIF-1 may have divergent effects on radiosensitivity. On the one hand, HIF-1 might be involved in the induction of apoptosis by increasing p53 phosphorylation following radiation treatment, thus sensitizing cancer cells to radiation (Moeller et al. 2005). On the other side, it is well appreciated that the tumor vasculature is an important target of radiation therapy and a major determinant of response
12
A. Rapisarda and G. Melillo
(Garcia-Barros et al. 2003), thus the induction of angiogenic factors, including VEGF, by HIF-1 may actually protect endothelial cells from radiation-induced apoptosis (Gupta et al. 2002). Indeed, combination of the HIF-1 inhibitor YC-1 with radiation blocked radiation induced HIF-1 upregulation and resulted in significant vascular destruction, suggesting that combining HIF-1 blockade with radiation may increase antitumor efficacy (Moeller et al. 2004). In addition, the HIF-1 inhibitor PX-478 increased therapeutic efficacy of radiation by blocking HIF-1-dependent effects on the tumor microenvironment in addition to a direct effect on tumor cell survival (Schwartz et al. 2009). Finally, recent evidence also suggests that HIF-2a inhibition sensitizes renal cancer cells to radiation therapy by increasing p53 activity and inducing apoptosis (Bertout et al. 2009). Given the complexity of the role that HIF signaling may play in the response of tumor tissue to radiation therapy, careful consideration should be given to the identification of tumor types and features of the tumor microenvironment that may predict for or be associated with a positive interaction between radiation therapy and HIF-1 inhibition.
Intratumor Hypoxia as a Potential Mechanism of Resistance to Anti-angiogenic Therapies The excitement associated with the introduction of anti-angiogenic agents for cancer therapy has been partially blunted by the limited single agent activity of this therapeutic strategy and by the development of resistance, a phenomenon originally unanticipated for an approach targeted to “normal,” genetically stable endothelial cells. While anti-angiogenic therapies have shown clinical benefit in combination with chemotherapy in metastatic colorectal cancer, advanced nonsmall cell lung cancer, and metastatic breast cancer (Ferrara 2005; Shojaei and Ferrara 2007; Ellis and Hicklin 2008), the single agent activity has been well below expectations and has been demonstrated only in few tumor types, including in renal cell carcinoma and more recently gliomas. More importantly, clinical benefit has only been transient and most tumors eventually become resistant and relapse. In addition to “intrinsic” resistance in patients who do not respond to anti-angiogenic agents, several mechanisms of “acquired” resistance have been proposed, including upregulation of alternative pro-angiogenic signals, production of pro-angiogenic factors by stromal cells, recruitment of bone marrow derived pro-angiogenic cells, increased pericyte coverage of the vasculature, and activation of invasive phenotypes (Rapisarda and Mellilo 2009). Interestingly, many of the responses implicated in resistance to anti-angiogenic therapy may be mediated by the products of genes induced by HIF-1. The consequences of effective anti-angiogenic therapy on the tumor microenvironment are still poorly understood and partly controversial, which limits our ability to develop rational combination therapies. On the one hand, “normalization” of tumor vasculature has been proposed, with consequent decrease in interstitial pressure and
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
13
better delivery of chemotherapy (Jain 2005); on the other hand, vascular “regression” has been demonstrated in several experimental models and appears to be a logical consequence of effective inhibition of tumor vasculature (Kerbel and Folkman 2002). The latter scenario is associated with an increase in intratumor hypoxia, which not only might be mechanistically involved in mediating resistance to anti-angiogenic therapies but also might provide unique opportunities for combination strategies (Paez-Ribes et al. 2009; Loges et al. 2009). Indeed, while expression of HIF-1a in solid tumors is focal and heterogeneous, detected predominantly in perinecrotic areas, “therapy-induced” hypoxia, which may be associated with the administration of antiangiogenic therapies, will lead to a more persistent and protracted expression of HIF-1a. Higher levels of HIF-1 in the tumor microenvironment may not only contribute to drug resistance in general, but also provide a mechanism of “escape” for tumor cells overcoming the potential therapeutic effects of anti-angiogenic agents. Thus, it is conceivable that combination of anti-angiogenic therapy with HIF-1 inhibition may have synergistic therapeutic activity by targeting HIF-1 in the context of a “hypoxic stressed” tumor microenvironment. Combination of Anti-angiogenic Therapies and HIF-1 Inhibitors To test the hypothesis that anti-angiogenic therapies may be more efficacious in combination with HIF-1 inhibition, we combined topotecan, a topoisomerase I inhibitor that downregulates HIF-1a protein in vitro and in vivo (Rapisarda et al. 2004b, c), with the anti-VEGF antibody bevacizumab in a glioma xenograft model. Indeed, we found that inhibition of HIF-1a by topotecan in a hypoxic stressed tumor microenvironment resulted in a more pronounced antitumor effect, relative to either agent alone. The effects on tumor growth were associated with significantly decreased HIF-1 transcriptional activity and reduced tumor cell proliferation (Rapisarda et al. 2009), consistent with the hypothesis that targeting HIF-1a activity may abrogate compensatory pathways required for cancer cell survival. Moreover, recent studies showed that genetic disruption of both HIF-1a and HIF-2a expression in colon cancer xenografts improves tumor response to sunitinib (a multitargeted receptor tyrosine kinase inhibitor) (Burkitt et al. 2009). This effect was mediated by a marked decrease in tumor angiogenesis and perfusion, through inhibition of multiple pro-angiogenic factors, as well as a marked decrease in tumor cells proliferation. To further support the role that HIF-1 inhibition might play in combination with anti-angiogenic therapies, it is interesting to point out that the combination of bevacizumab and irinotecan (another topoisomerase I inhibitor that also inhibits HIF-1) has shown clinical benefit in glioblastoma patients with a six-month overall survival of 62–77% (Vredenburgh et al. 2007; Chen et al. 2007). Further studies are required to better define the patient population that may benefit from these combination approaches, in particular by better understanding when and to what extent anti-angiogenic therapy is indeed associated with increased intratumor hypoxia and HIF-1a expression.
14
A. Rapisarda and G. Melillo
Cancer Cell Metabolism and the Hypoxic Tumor Microenvironment More than 80 years ago, Otto Warburg observed that cancer cells avidly take up glucose and produce lactic acid under aerobic conditions, a process subsequently referred to as the Warburg effect or aerobic glycolysis. The exact molecular mechanism(s) regulating cancer cells dependence on glycolysis remain unclear, but it is likely that a combination of effects due to oncogenic alterations and the hypoxic tumor microenvironment play a role. Many genetic alterations associated with the cancer phenotype promote glycolysis and, at the same time, suppress oxidative phosphorylation. For instance, p53 plays an important role in regulating both “aerobic” and “anaerobic” respiration, while c-Myc expression activates the transcription of several glycolytic enzymes (Kroemer and Pouyssegur 2008). The largest functional group of genes consistently upregulated by HIF-1 in a number of cell types is associated with glucose metabolism. HIF-1 increases the expression of the transporters necessary for the entry of glucose into the cell, the genes involved in the enzymatic breakdown of glucose to pyruvate and the enzymes involved in the metabolism of pyruvate. In hypoxic cells, pyruvate is converted by lactate dehydrogenase (LDH) to lactate, which is then released into the extracellular space. HIF-1 also controls cellular oxygen consumption by (a) inducing the expression of PDH kinase 1 (PDK1), which inhibits pyruvate dehydrogenase (PDH) and the conversion of pyruvate to acetyl-CoA, ultimately decreasing mitochondrial respiration (Papandreou et al. 2006; Kim et al. 2006b) and (b) reducing mitochondrial mass, by counteracting the stimulatory action of c-Myc on mitochondrial biogenesis (Zhang et al. 2007). It is conceivable then that these metabolic changes utilized by cancer cells to survive in a hypoxic, nutrient deprived tumor microenvironment might also offer therapeutic opportunities to selectively target hypoxic cells and spare normal tissues.
HIF-1 Inhibitors in Combination Strategies Targeting Tumor Metabolism Cancer cells, but not normal cells, are thought to be “addicted” to glycolysis and attempts are being pursued to explore the possibility to target glycolysis for cancer therapy. However, the activity of 2-deoxyglucose (2-DG, a competitive inhibitor of glucose), as single agent, has been limited by evidence of toxicities in clinical trials (Denko 2008). Interestingly, genetic knockdown of HIF-1a, as well as treatment with CCI-779 (an inhibitor of mTOR that decreases HIF-1 activity), sensitizes cancer cells in vitro to 2-DG (Maher et al. 2007; Wangpaichitr et al. 2008), suggesting that HIF-1 inhibitors might be used to increase the therapeutic efficacy of 2-DG. An original approach has been proposed to increase oxygen consumption in cancer cells by reversing the inhibition of mitochondrial function using either the HIF-1 inhibitor echinomycin or the PDK-1 inhibitor dichloroacetate (DCA) (Cairns et al. 2007). The pharmacological induced increase in oxygen consumption, associated
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
15
with the limited oxygen supply due to the dysfunctional tumor vasculature, made the tumors more hypoxic and sensitized cancer cells to the hypoxic cytotoxin tirapazamine. These data suggest that HIF-1 inhibitors might have a potential role in combination strategies aimed at targeting tumor metabolism or oxygen consumption.
Conclusion In the complex biology of human cancers, HIF-1 is yet another attractive target for the development of cancer therapeutics. The role of HIF-1 in human cancers has been extensively investigated and accumulating evidence implicates HIF-1 in a number of biological processes associated with resistance to therapy and tumor progression. Hence, HIF-1 inhibition appears as a logical therapeutic strategy and many small molecule inhibitors of HIF-1 have been described, several of which are validated in preclinical models. However, we have still limited understanding of when and to what extent inhibition of HIF-1 in cancer patients may be effective. Single agent studies are required to have a better appreciation of the biological consequence associated with HIF-1 inhibition in human cancers, yet the lack of selective pharmacological inhibitors of HIF-1 significantly hinders this task. On the other hand, we have learned over the last few years that targeting multiple signaling pathways deregulated in cancer cells may be a more effective therapeutic strategy. The broad involvement of HIF-1 in many biological processes associated with and required for tumor progression provides unique opportunities for the development of combination therapies. A better understanding of the role played by HIF-1 in pathways associated with the administration of chemotherapy, radiation and anti-angiogenic therapies may unveil vulnerable aspects of the tumor microenvironment that could be exploited for the rational design of combination strategies. Acknowledgments The authors would like to thank members of the Tumor Hypoxia Laboratory and Dr. R. H. Shoemaker for helpful discussion. This project has been funded in whole or in part with Federal funds from the National Cancer Institute, National Institutes of Health, under Contract No. N01-CO-12400. The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. government. This research was supported in part by the Developmental Therapeutics Program, DCTD, of the National Cancer Institute, NIH.
References Aebersold DM, Burri P, Beer KT et al (2001) Expression of hypoxia-inducible factor-1alpha: a novel predictive and prognostic parameter in the radiotherapy of oropharyngeal cancer. Cancer Res 61:2911–2916.
16
A. Rapisarda and G. Melillo
Bachtiary B, Schindl M, Potter R et al (2003) Overexpression of hypoxia-inducible factor 1alpha indicates diminished response to radiotherapy and unfavorable prognosis in patients receiving radical radiotherapy for cervical cancer. Clin Cancer Res 9:2234–2240. Bertout JA, Patel SA, Simon MC (2008) The impact of O2 availability on human cancer. Nat Rev Cancer 8:967–975. Bertout JA, Majmundar AJ, Gordan JD et al (2009) HIF2{alpha} inhibition promotes p53 pathway activity, tumor cell death, and radiation responses. Proc Natl Acad Sci USA 106:14391–14396. Birner P, Schindl M, Obermair A et al (2000) Overexpression of hypoxia-inducible factor 1alpha is a marker for an unfavorable prognosis in early-stage invasive cervical cancer. Cancer Res 60:4693–4696. Birner P, Gatterbauer B, Oberhuber G et al (2001a) Expression of hypoxia-inducible factor-1 alpha in oligodendrogliomas: its impact on prognosis and on neoangiogenesis. Cancer 92:165–171. Birner P, Schindl M, Obermair A et al (2001b) Expression of hypoxia-inducible factor 1alpha in epithelial ovarian tumors: its impact on prognosis and on response to chemotherapy. Clin Cancer Res 7:1661–1668. Bos R, Zhong H, Hanrahan CF et al (2001) Levels of hypoxia-inducible factor-1 alpha during breast carcinogenesis. J Natl Cancer Inst 93:309–314. Bos R, van der GP, Greijer AE et al (2003) Levels of hypoxia-inducible factor-1alpha independently predict prognosis in patients with lymph node negative breast carcinoma. Cancer 97:1573–1581. Brizel DM, Scully SP, Harrelson JM et al (1996) Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer Res 56:941–943. Brown JM, Giaccia AJ (1998) The unique physiology of solid tumors: opportunities (and problems) for cancer therapy. Cancer Res 58:1408–1416. Burkitt K, Chun SY, Dang DT et al (2009) Targeting both HIF-1 and HIF-2 in human colon cancer cells improves tumor response to sunitinib treatment. Mol Cancer Ther 8(5):1148–1156. Burri P, Djonov V, Aebersold DM et al (2003) Significant correlation of hypoxia-inducible factor1alpha with treatment outcome in cervical cancer treated with radical radiotherapy. Int J Radiat Oncol Biol Phys 56:494–501. Cairns RA, Papandreou I, Sutphin PD et al (2007) Metabolic targeting of hypoxia and HIF1 in solid tumors can enhance cytotoxic chemotherapy. Proc Natl Acad Sci USA 104:9445–9450. Chen W, Delaloye S, Silverman DH et al (2007) Predicting treatment response of malignant gliomas to bevacizumab and irinotecan by imaging proliferation with [18F] fluorothymidine positron emission tomography: a pilot study. J Clin Oncol 25:4714–4721. Comerford KM, Wallace TJ, Karhausen J et al (2002) Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (MDR1) gene. Cancer Res 62:3387–3394. Creighton-Gutteridge M, Cardellina JH, Stephen AG et al (2007) Cell type-specific, topoisomerase II-dependent inhibition of hypoxia-inducible factor-1alpha protein accumulation by NSC 644221. Clin Cancer Res 13:1010–1018. Dales JP, Garcia S, Meunier-Carpentier S et al (2005) Overexpression of hypoxia-inducible factor HIF-1alpha predicts early relapse in breast cancer: retrospective study in a series of 745 patients. Int J Cancer 116:734–739. Del BD, Ciuffreda L, Trisciuoglio D et al (2006) Antiangiogenic potential of the mammalian target of rapamycin inhibitor temsirolimus. Cancer Res 66:5549–5554. Denko NC (2008) Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat Rev Cancer 8:705–713. Ellis LM, Hicklin DJ (2008) Pathways mediating resistance to vascular endothelial growth factortargeted therapy. Clin Cancer Res 14:6371–6375. Ellis L, Hammers H, Pili R (2009) Targeting tumor angiogenesis with histone deacetylase inhibitors. Cancer Lett 280(2):145–153. Escuin D, Kline ER, Giannakakou P (2005) Both microtubule-stabilizing and microtubuledestabilizing drugs inhibit hypoxia-inducible factor-1alpha accumulation and activity by disrupting microtubule function. Cancer Res 65:9021–9028.
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
17
Ferrara N (2005) VEGF as a therapeutic target in cancer. Oncology 69(Suppl 3):11–16. Garcia-Barros M, Paris F, Cordon-Cardo C et al (2003) Tumor response to radiotherapy regulated by endothelial cell apoptosis. Science 300:1155–1159. Giatromanolaki A, Sivridis E, Kouskoukis C et al (2003) Hypoxia-inducible factors 1alpha and 2alpha are related to vascular endothelial growth factor expression and a poorer prognosis in nodular malignant melanomas of the skin. Melanoma Res 13:493–501. Giatromanolaki A, Koukourakis MI, Simopoulos C et al (2007) Metastatic cancer cells from c-erbB-2 negative primary breast cancer maintain the original c-erbB-2/HIF1alpha phenotype. Cancer Biol Ther 6:153–155. Greenberger LM, Horak ID, Filpula D et al (2008) A RNA antagonist of hypoxia-inducible factor1alpha, EZN-2968, inhibits tumor cell growth. Mol Cancer Ther 7:3598–3608. Griffiths EA, Pritchard SA, Valentine HR et al (2007) Hypoxia-inducible factor-1alpha expression in the gastric carcinogenesis sequence and its prognostic role in gastric and gastro-oesophageal adenocarcinomas. Br J Cancer 96:95–103. Gupta VK, Jaskowiak NT, Beckett MA et al (2002) Vascular endothelial growth factor enhances endothelial cell survival and tumor radioresistance. Cancer J 8:47–54. Han JY, Oh SH, Morgillo F et al (2005) Hypoxia-inducible factor 1alpha and antiangiogenic activity of farnesyltransferase inhibitor SCH66336 in human aerodigestive tract cancer. J Natl Cancer Inst 97:1272–1286. Hockel M, Vaupel P (2001) Tumor hypoxia: definitions and current clinical, biologic, and molecular aspects. J Natl Cancer Inst 93:266–276. Huang LE (2008) Carrot and stick: HIF-alpha engages c-Myc in hypoxic adaptation. Cell Death Differ 15:672–677. Hudson CC, Liu M, Chiang GG et al (2002) Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Mol Cell Biol 22:7004–7014. Isaacs JS, Jung YJ, Mimnaugh EG et al (2002) Hsp90 regulates a von Hippel Lindau-independent hypoxia-inducible factor-1 alpha-degradative pathway. J Biol Chem 277:29936–29944. Jain RK (2005) Antiangiogenic therapy for cancer: current and emerging concepts. Oncology (Williston Park) 19:7–16. Kaluz S, Kaluzova M, Stanbridge EJ (2006) Proteasomal inhibition attenuates transcriptional activity of hypoxia-inducible factor 1 (HIF-1) via specific effect on the HIF-1alpha C-terminal activation domain. Mol Cell Biol 26:5895–5907. Kaluz S, Kaluzova M, Stanbridge EJ (2008) Comment on the role of FIH in the inhibitory effect of bortezomib on hypoxia-inducible factor-1. Blood 111:5258–5259. Kerbel R and Folkman J (2002) Clinical translation of angiogenesis inhibitors. Nat Rev Cancer 2:727–739. Kim HL, Yeo EJ, Chun YS et al (2006a) A domain responsible for HIF-1alpha degradation by YC-1, a novel anticancer agent. Int J Oncol 29:255–260. Kim JW, Tchernyshyov I, Semenza GL et al (2006b) HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 3:177–185. Koh MY, Spivak-Kroizman T, Venturini S et al (2008) Molecular mechanisms for the activity of PX-478, an antitumor inhibitor of the hypoxia-inducible factor-1alpha. Mol Cancer Ther 7:90–100. Kong D, Park EJ, Stephen AG et al (2005) Echinomycin, a small-molecule inhibitor of hypoxiainducible factor-1 DNA-binding activity. Cancer Res 65:9047–9055. Kong X, Lin Z, Liang D et al (2006) Histone deacetylase inhibitors induce VHL and ubiquitinindependent proteasomal degradation of hypoxia-inducible factor 1alpha. Mol Cell Biol 26:2019–2028. Koukourakis MI, Simopoulos C, Polychronidis A et al (2003) The effect of trastuzumab/docatexel combination on breast cancer angiogenesis: dichotomus effect predictable by the HIFI alpha/ VEGF pre-treatment status? Anticancer Res 23:1673–1680.
18
A. Rapisarda and G. Melillo
Korkolopoulou P, Patsouris E, Konstantinidou AE et al (2004) Hypoxia-inducible factor 1alpha/ vascular endothelial growth factor axis in astrocytomas. Associations with microvessel morphometry, proliferation and prognosis. Neuropathol Appl Neurobiol 30:267–278. Koukourakis MI, Bentzen SM, Giatromanolaki A et al (2006) Endogenous markers of two separate hypoxia response pathways (hypoxia inducible factor 2 alpha and carbonic anhydrase 9) are associated with radiotherapy failure in head and neck cancer patients recruited in the CHART randomized trial. J Clin Oncol 24:727–735. Kroemer G and Pouyssegur J (2008) Tumor cell metabolism: cancer’s Achilles’ heel. Cancer Cell 13:472–482. Kung AL, Zabludoff SD, France DS et al (2004) Small molecule blockade of transcriptional coactivation of the hypoxia-inducible factor pathway. Cancer Cell 6:33–43. Laughner E, Taghavi P, Chiles K et al (2001) HER2 (neu) signaling increases the rate of hypoxiainducible factor 1alpha (HIF-1alpha) synthesis: novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Mol Cell Biol 21:3995–4004. Lee K, Qian DZ, Rey S et al (2009) Anthracycline chemotherapy inhibits HIF-1 transcriptional activity and tumor-induced mobilization of circulating angiogenic cells. Proc Natl Acad Sci USA 106:2353–2358. Li L, Lin X, Staver M et al (2005) Evaluating hypoxia-inducible factor-1alpha as a cancer therapeutic target via inducible RNA interference in vivo. Cancer Res 65:7249–7258. Li L, Lin X, Shoemaker AR et al (2006) Hypoxia-inducible factor-1 inhibition in combination with temozolomide treatment exhibits robust antitumor efficacy in vivo. Clin Cancer Res 12:4747–4754. Loges S, Mazzone M, Hohensinner P et al (2009) Silencing or fueling metastasis with VEGF inhibitors: antiangiogenesis revisited. Cancer Cell 15:167–170. Luwor RB, Lu Y, Li X et al (2005) The antiepidermal growth factor receptor monoclonal antibody cetuximab/C225 reduces hypoxia-inducible factor-1 alpha, leading to transcriptional inhibition of vascular endothelial growth factor expression. Oncogene 24:4433–4441. Mabjeesh NJ, Post DE, Willard MT et al (2002) Geldanamycin induces degradation of hypoxiainducible factor 1alpha protein via the proteosome pathway in prostate cancer cells. Cancer Res 62:2478–2482. Mabjeesh NJ, Escuin D, LaVallee TM et al (2003) 2ME2 inhibits tumor growth and angiogenesis by disrupting microtubules and dysregulating HIF. Cancer Cell 3:363–375. Maher JC, Wangpaichitr M, Savaraj N et al (2007) Hypoxia-inducible factor-1 confers resistance to the glycolytic inhibitor 2-deoxy-D-glucose. Mol Cancer Ther 6:732–741. Majumder PK, Febbo PG, Bikoff R et al (2004) mTOR inhibition reverses Akt-dependent prostate intraepithelial neoplasia through regulation of apoptotic and HIF-1-dependent pathways. Nat Med 10:594–601. Manalo DJ, Rowan A, Lavoie T et al (2005) Transcriptional regulation of vascular endothelial cell responses to hypoxia by HIF-1. Blood 105:659–669. Mayerhofer M, Valent P, Sperr WR et al (2002) BCR/ABL induces expression of vascular endothelial growth factor and its transcriptional activator, hypoxia inducible factor-1alpha, through a pathway involving phosphoinositide 3-kinase and the mammalian target of rapamycin. Blood 100:3767–3775. Melillo G (2006) Inhibiting hypoxia-inducible factor 1 for cancer therapy. Mol Cancer Res 4:601–605. Melillo G (2007) Targeting hypoxia cell signaling for cancer therapy. Cancer Metastasis Rev 26:341–352. Mie LY, Kim SH, Kim HS et al (2003) Inhibition of hypoxia-induced angiogenesis by FK228, a specific histone deacetylase inhibitor, via suppression of HIF-1alpha activity. Biochem Biophys Res Commun 300:241–246. Moeller BJ, Cao Y, Li CY et al (2004) Radiation activates HIF-1 to regulate vascular radiosensitivity in tumors: role of reoxygenation, free radicals, and stress granules. Cancer Cell 5:429–441. Moeller BJ, Dreher MR, Rabbani ZN et al (2005) Pleiotropic effects of HIF-1 blockade on tumor radiosensitivity. Cancer Cell 8:99–110.
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
19
Moeller BJ, Richardson RA, Dewhirst MW (2007) Hypoxia and radiotherapy: opportunities for improved outcomes in cancer treatment. Cancer Metastasis Rev 26:241–248. Olenyuk BZ, Zhang GJ, Klco JM et al (2004) Inhibition of vascular endothelial growth factor with a sequence-specific hypoxia response element antagonist. Proc Natl Acad Sci USA 101: 16768–16773. Onnis B, Rapisarda A, Melillo G (2009) Development of HIF-1 inhibitors for cancer therapy. J Cell Mol Med in press.. Paez-Ribes M, Allen E, Hudock J et al (2009) Antiangiogenic therapy elicits malignant progression of tumors to increased local invasion and distant metastasis. Cancer Cell 15:220–231. Papandreou I, Cairns RA, Fontana L et al (2006) HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab 3:187–197. Park EJ, Kong D, Fisher R et al (2006) Targeting the PAS-A domain of HIF-1alpha for development of small molecule inhibitors of HIF-1. Cell Cycle 5:1847–1853. Pencreach E, Guerin E, Nicolet C et al (2009) Marked activity of irinotecan and rapamycin combination toward colon cancer cells in vivo and in vitro is mediated through cooperative modulation of the mammalian target of rapamycin/hypoxia-inducible factor-1alpha axis. Clin Cancer Res 15:1297–1307. Pore N, Jiang Z, Gupta A et al (2006) EGFR tyrosine kinase inhibitors decrease VEGF expression by both hypoxia-inducible factor (HIF)-1-independent and HIF-1-dependent mechanisms. Cancer Res 66:3197–3204. Pouyssegur J, Dayan F, Mazure NM (2006) Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature 441:437–443. Qian DZ, Kachhap SK, Collis SJ et al (2006) Class II histone deacetylases are associated with VHL-independent regulation of hypoxia-inducible factor 1 alpha. Cancer Res 66:8814–8821. Rapisarda A, Melillo G (2009) Role of the hypoxic tumor microenvironment in the resistance to anti-angiogenic therapies. Drug Resist Updat 12:74–80. Rapisarda A, Uranchimeg B, Scudiero DA et al (2002) Identification of small molecule inhibitors of hypoxia-inducible factor 1 transcriptional activation pathway. Cancer Res 62:4316–4324. Rapisarda A, Shoemaker RH, Melillo G (2004a) Targeting topoisomerase I to inhibit hypoxia inducible factor 1. Cell Cycle 3:172–175. Rapisarda A, Uranchimeg B, Sordet O et al (2004b) Topoisomerase I-mediated inhibition of hypoxia-inducible factor 1: mechanism and therapeutic implications. Cancer Res 64:1475–1482. Rapisarda A, Zalek J, Hollingshead M et al (2004c) Schedule-dependent inhibition of hypoxiainducible factor-1alpha protein accumulation, angiogenesis, and tumor growth by topotecan in U251-HRE glioblastoma xenografts. Cancer Res 64:6845–6848. Rapisarda A, Hollingshead M, Uranchimeg B et al (2009) Increased antitumor activity of bevacizumab in combination with hypoxia inducible factor-1 inhibition. Mol Cancer Ther 8:1867–1877. Sapra P, Zhao H, Mehlig M et al (2008) Novel delivery of SN38 markedly inhibits tumor growth in xenografts, including a camptothecin-11-refractory model. Clin Cancer Res 14:1888–1896. Scheuermann TH, Tomchick DR, Machius M et al (2009) Artificial ligand binding within the HIF2alpha PAS-B domain of the HIF2 transcription factor. Proc Natl Acad Sci USA 106:450–455. Schindl M, Schoppmann SF, Samonigg H et al (2002) Overexpression of hypoxia-inducible factor 1alpha is associated with an unfavorable prognosis in lymph node-positive breast cancer. Clin Cancer Res 8:1831–1837. Schwartz DL, Powis G, Thitai-Kumar A et al (2009) The selective hypoxia inducible factor-1 inhibitor PX-478 provides in vivo radiosensitization through tumor stromal effects. Mol Cancer Ther 8:947–958. Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3:721–732. Semenza GL (2007) Hypoxia and cancer. Cancer Metastasis Rev 26:223–224. Semenza GL (2008) Hypoxia-inducible factor 1 and cancer pathogenesis. IUBMB Life 60:591–597.
20
A. Rapisarda and G. Melillo
Sermeus A, Cosse JP, Crespin M et al (2008) Hypoxia induces protection against etoposide-induced apoptosis: molecular profiling of changes in gene expression and transcription factor activity. Mol Cancer 7:27. Shin DH, Chun YS, Lee DS et al (2008) Bortezomib inhibits tumor adaptation to hypoxia by stimulating the FIH-mediated repression of hypoxia-inducible factor-1. Blood 111:3131–3136. Shojaei F, Ferrara N (2007) Antiangiogenic therapy for cancer: an update. Cancer J 13:345–348. Sivridis E, Giatromanolaki A, Gatter KC et al (2002) Association of hypoxia-inducible factors 1alpha and 2alpha with activated angiogenic pathways and prognosis in patients with endometrial carcinoma. Cancer 95:1055–1063. Sullivan R, Graham CH (2009) Hypoxia prevents etoposide-induced DNA damage in cancer cells through a mechanism involving hypoxia-inducible factor 1. Mol Cancer Ther 8:1702–1713. Sun HC, Qiu ZJ, Liu J et al (2007) Expression of hypoxia-inducible factor-1 alpha and associated proteins in pancreatic ductal adenocarcinoma and their impact on prognosis. Int J Oncol 30: 1359–1367. Swinson DE, Jones JL, Cox G et al (2004) Hypoxia-inducible factor-1 alpha in non small cell lung cancer: relation to growth factor, protease and apoptosis pathways. Int J Cancer 111:43–50. Takahashi R, Tanaka S, Hiyama T et al (2003) Hypoxia-inducible factor-1alpha expression and angiogenesis in gastrointestinal stromal tumor of the stomach. Oncol Rep 10:797–802. Talks KL, Turley H, Gatter KC et al (2000) The expression and distribution of the hypoxiainducible factors HIF-1alpha and HIF-2alpha in normal human tissues, cancers, and tumorassociated macrophages. Am J Pathol 157:411–421. Tan C, de Noronha RG, Roecker AJ et al (2005) Identification of a novel small-molecule inhibitor of the hypoxia-inducible factor 1 pathway. Cancer Res 65:605–612. Theodoropoulos VE, Lazaris AC, Sofras F et al (2004) Hypoxia-inducible factor 1 alpha expression correlates with angiogenesis and unfavorable prognosis in bladder cancer. Eur Urol 46:200–208. Theodoropoulos VE, Lazaris AC, Kastriotis I et al (2005) Evaluation of hypoxia-inducible factor 1alpha overexpression as a predictor of tumour recurrence and progression in superficial urothelial bladder carcinoma. BJU Int 95:425–431. Unruh A, Ressel A, Mohamed HG et al (2003) The hypoxia-inducible factor-1 alpha is a negative factor for tumor therapy. Oncogene 22:3213–3220. Vaupel P, Kelleher DK, Hockel M (2001) Oxygen status of malignant tumors: pathogenesis of hypoxia and significance for tumor therapy. Semin Oncol 28:29–35. Verheul HM, Salumbides B, van EK et al (2008) Combination strategy targeting the hypoxia inducible factor-1 alpha with mammalian target of rapamycin and histone deacetylase inhibitors. Clin Cancer Res 14:3589–3597. Vleugel MM, Greijer AE, Shvarts A et al (2005) Differential prognostic impact of hypoxia induced and diffuse HIF-1alpha expression in invasive breast cancer. J Clin Pathol 58:172–177. Vredenburgh JJ, Desjardins A, Herndon JE et al (2007) Bevacizumab plus irinotecan in recurrent glioblastoma multiforme. J Clin Oncol 25:4722–4729. Wan X, Shen N, Mendoza A et al (2006) CCI-779 inhibits rhabdomyosarcoma xenograft growth by an antiangiogenic mechanism linked to the targeting of mTOR/Hif-1alpha/VEGF signaling. Neoplasia 8:394–401. Wang GL and Semenza GL (1995) Purification and characterization of hypoxia-inducible factor 1. J Biol Chem 270:1230–1237. Wang GL, Jiang BH, Rue EA et al (1995) Hypoxia-inducible factor 1 is a basic-helix-loop-helixPAS heterodimer regulated by cellular O2 tension. Proc Natl Acad Sci USA 92:5510–5514. Wangpaichitr M, Savaraj N, Maher J et al (2008) Intrinsically lower AKT, mammalian target of rapamycin, and hypoxia-inducible factor activity correlates with increased sensitivity to 2-deoxy-D-glucose under hypoxia in lung cancer cell lines. Mol Cancer Ther 7:1506–1513. Welsh S, Williams R, Kirkpatrick L et al (2004) Antitumor activity and pharmacodynamic properties of PX-478, an inhibitor of hypoxia-inducible factor-1alpha. Mol Cancer Ther 3:233–244.
1 Combination Strategies Targeting Hypoxia Inducible Factor 1
21
Welsh SJ, Koh MY, Powis G (2006) The hypoxic inducible stress response as a target for cancer drug discovery. Semin Oncol 33:486–497. Williams KJ, Telfer BA, Xenaki D et al (2005) Enhanced response to radiotherapy in tumors deficient in the function of hypoxia-inducible factor-1. Radiother Oncol 75:89–98. Yang J, Zhang L, Erbel PJ et al (2005) Functions of the Per/ARNT/Sim domains of the hypoxiainducible factor. J Biol Chem 280:36047–36054. Zhang X, Kon T, Wang H et al (2004) Enhancement of hypoxia-induced tumor cell death in vitro and radiation therapy in vivo by use of small interfering RNA targeted to hypoxia-inducible factor-1alpha. Cancer Res 64:8139–8142. Zhang H, Gao P, Fukuda R et al (2007) HIF-1 inhibits mitochondrial biogenesis and cellular respiration in VHL-deficient renal cell carcinoma by repression of C-MYC activity. Cancer Cell 11:407–420. Zhang H, Qian DZ, Tan YS et al (2008) Digoxin and other cardiac glycosides inhibit HIF-1alpha synthesis and block tumor growth. Proc Natl Acad Sci USA 105:19579–19586. Zhong H, De Marzo AM, Laughner E et al (1999) Overexpression of hypoxia-inducible factor 1alpha in common human cancers and their metastases. Cancer Res 59:5830–5835. Zhong H, Chiles K, Feldser D et al (2000) Modulation of hypoxia-inducible factor 1alpha expression by the epidermal growth factor/phosphatidylinositol 3-kinase/PTEN/AKT/FRAP pathway in human prostate cancer cells: implications for tumor angiogenesis and therapeutics. Cancer Res 60:1541–1545.
Chapter 2
The Tumor Microenvironment: New Insights into Regulation of Tumor pH by Carbonic Anhydrases Pawel Swietach, Adrian L. Harris, and Richard D. Vaughan-Jones
Abstract Intracellular pH (pHi) is a powerful modulator of cell function. There is a narrow range of pHi, usually between 7.0 and 7.5, over which cell growth and development is optimal. Metabolism, particularly cellular respiration, tends to disturb this favorable steady state. The survival of cells undergoing such an acid challenge relies on efficient mechanisms for regulating pHi. In solid tumors, pHi regulation is challenged by their heavy demand for energy and poor blood supply. Inadequately perfused cancer cells deprive the milieu of oxygen, forcing a switchover to the less energy efficient, anaerobic mode of respiration. To meet energy demands, cells upregulate respiration and excrete more acid which lowers extracellular pH (pHe). The tumor extracellular milieu is believed to exert Darwinian selection in favor of cancer cells and against normal cells, based on the cell’s capacity to protect pHi under hypoxia and low pHe. In this chapter, we describe the pathways operating in tumors for removing acid and regulating pHi, with particular emphasis on membrane transporters and carbonic anhydrase (CA) enzymes. We discuss these pathways in terms of survival adaptations and possible targets for anticancer therapy.
Biological Importance of pH The hydrogen atom is very common in the human body, and despite being the lightest of elements, it comprises about a tenth of total body mass. Virtually all hydrogen atoms are bonded by means of electron transfer or sharing with other elements. In many such compounds, the hydrogen nucleus can dissociate reversibly from the remainder of the molecule. In the case of the most common isotope 1 H, the hydrogen nucleus is solely a proton, and therefore the H+ dissociation and
A.L. Harris (*) Weatherall Institute of Molecular Medicine, Oxford University, Oxford OX3 9DS, UK e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_2, © Springer Science+Business Media, LLC 2010
23
24
P. Swietach et al.
association reactions are dubbed deprotonation and protonation, respectively. Water, the biological solvent that contributes towards 60% of the body’s mass, also undergoes reversible deprotonation, and in pure form equilibrates at a “neutral” H+ ion concentration ([H+]) of 1027 M. Acids and bases are defined as molecules that increase and decrease [H+], respectively. To avoid the use of exponential notation, the acidity/basicity of a solution is commonly expressed as the negative logarithm of molar [H+], and given the symbol pH (Boron 2004). Many biologically relevant molecules are weak acids and weak bases. Proteins, for instance, are chains of weak acids/bases. Weak acids/bases exist in dynamic equilibrium between their protonated and unprotonated forms, the concentration ratio of which is very sensitive to pH. Since the H+ ion has very high charge density, pH will have a major impact on the physico-chemical properties of weak acids/ bases. This is particularly relevant to proteins, where charge distribution is linked to structure and function (Boron 2004; Swietach et al. 2007). Indeed, intracellular pH (pHi) is a potent and universal regulator of cellular physiology (Boron 2004; Swietach et al. 2007; Roos and Boron 1981; Vaughan-Jones et al. 2008; Isfort et al. 1993; McConkey and Orrenius 1996). Displacements of pHi by only a fraction of a unit can lead to patho-physiological responses and apoptotic death (Park et al. 1999). Cells must defend their optimal pHi by corrective (homeostatic) mechanisms. A major challenge to these housekeeping mechanisms is metabolic acid production, a process that takes place in all respiring cells throughout their life. To counteract metabolic acid loading, excess intracellular H+ ions must be removed across the cell membrane into the extracellular space and out of the body at the lungs and kidneys. Despite its tiny radius, the highly-polarizing H+ ion is poorly permeant across the lipid bilayer of cell membranes. To allow transmembrane H+ - traffic, cells have evolved transport proteins for H+ ions (e.g. Na+/H+ exchange), or other means of facilitated H+ ion diffusion (e.g. its reaction with HCO3– to form membrane-permeant CO2). The performance of these mechanisms can be compromised by the accumulation of extracellular acid, either as a result of allosteric H+ - inhibition of transport proteins (Vaughan-Jones and Wu 1990; Aronson 1985) or by the collapse diffusion gradients that drive acid-efflux (Swietach et al. 2008, 2009). Through these mechanisms, extracellular pH (pHe) can feedback onto pHi homeostasis. In turn, pHe depends on the diffusive coupling across the extracellular space and with blood. Efficient pHi regulation is central to the survival of metabolically-active tissues, particularly those with relatively poor blood perfusion. Solid tumors are characterized by a high acid production rate (Gatenby and Gillies 2004; Gillies et al. 2008) and inadequate blood supply (Swietach et al. 2007; Vaupel et al. 1989; Kallinowski et al. 1989; Stubbs et al. 2000). Although cancer cells, by definition, harbor a number of mutated gene products, the bulk of cellular protein is no different from normal cells and bears a similar pHi sensitivity, carving a narrow pHi range for survival. The growth of tumors may depend critically on the ability of cancer cells to remove intracellular acid more efficiently than normal cells. The conditions that are characteristic of tumors, notably hypoxia and low pHe (Griffiths et al. 2001; Gatenby et al. 2006), may be tolerated by cancer cells, but not by normal cells, thereby selecting against the latter (Fang et al. 2008). In this chapter, we describe
2 The Tumor Microenvironment
25
the mechanisms for controlling pHi in solid tumors, with particular emphasis on the role of CO2/HCO3– buffer and its enzyme catalyst, carbonic anhydrase (CA).
Sources of Cellular Acid Cellular Respiration A large number of biochemical reactions produce or consume H+ ions, but production of lactic acid by anaerobic respiration and carbon dioxide (CO2) by aerobic respiration are quantitatively the most important acid fluxes that can alter pHi. Glucose, the major respiratory substrate in tumors, produces 2 × H+ plus 2 × lactate (i.e. ionized lactic acid) by the cytosolic O2-independent glycolytic pathway, or 6 × CO2 plus 6 × H2O through O2-dependent mitochondrial pathways. In addition, during periods of nucleic acid synthesis, glucose can yield 1 × CO2 through an additional cytosolic pathway called the pentose phosphate shunt (Helmlinger et al. 2002). Upon hydration, CO2 yields H+ and HCO3−. Assuming equilibrium at pH 7, the molar H+ – yield per aerobically and anaerobically respired glucose is 5.3 and 2.0, respectively (aciddissociation constants for CO2 and lactic acid are 1026.2 and 1024 M, respectively). Tumors demand a particularly high respiratory rate to energize their substantial growth and development. Glucose yields almost 20 times more energy (ATP molecules) when it is respired to CO2 rather than lactic acid. The carbon atoms of lactic acid are less oxidized than those in CO2, and still have considerable energy trapped in their bonds. Indeed, lactic acid can be respired further to CO2. Based on energy efficiency alone, it would be sensible for tumors to prefer aerobic respiration. But to benefit from this higher energy yield, tumors need functional mitochondria and adequate O2 supply. As solid tumors develop, they grow away from their basement membrane and vasculature. This increases the diffusion distance for O2 delivery. Furthermore, the elevated O2 demand in cancer cells that are closer to blood vessels greatly reduces O2 availability in more distal cells, and shortens the distance over which cells are adequately oxygenated. This distance, known as the Krogh radius (Krogh 1919), can be as little as a few cell layers thick in tumors (Dewhirst et al. 1994). Beyond the Krogh radius, tissue becomes hypoxic. In addition to inadequate and highly disorganized vasculature (Thomlinson and Gray 1955), the extent of hypoxia within tumors is increased by variable and discontinuous blood flow (Vaupel et al. 1989; Helmlinger et al. 1997). The size of the hypoxic region can be visualized by pimonidazole staining (Fig. 2.1a). Low O2 tension (<1%) blocks the degradation of hypoxia-inducible factor (HIF), a transcription factor that otherwise has a very short lifetime under normoxia. HIF was discovered as an inducer of erythropoietin expression (Semenza and Wang 1992), and is now an established regulator of genes involved in glucose metabolism (e.g. Glut-1 for glucose uptake, Fig. 2.1b) and pH regulation (e.g. CA9, a carbonic anhydrase isoform, Fig. 2.1c) (Semenza 2003). Hypoxia forces glucose to be respired anaerobically by blocking O2-dependent mitochondrial reactions and
26
P. Swietach et al.
Fig. 2.1 Tumor hypoxia and its correlation with hypoxia-induced proteins, modified from Airley et al. (2003). Serial sections of cervix carcinoma biopsies, immuno-stained for (a) pimonidazole, (b) Glut-1, and (c) CAIX, showing similarities in distribution and pattern. Magnification 200×. Shaded arrows point in the direction of decreasing O2, as assessed by increasing pimonidazole staining
by HIF-induced expression of pyruvate dehydrogenase kinase (PDK) and lactate dehydrogenase (LDH), which divert pyruvate away from mitochondrial reactions (PDK) and towards lactic acid production (LDH) (Semenza et al. 1994).
The Warburg Effect Under uncompromised O2 supply, it is in the interest of cellular energetics to respire aerobically. This phenomenon, called the Pasteur effect, is a coordinated effort by adenosine nucleotides (Ramaiah et al. 1964), redox potentials (Cerdan et al. 2006), and transcription factors (Hardie 2003) to favor the production of CO2 over lactic acid. In an apparent violation of the Pasteur effect, cancer cells typically show constitutive upregulation of lactic acid production even under normoxic conditions. This observation, called the Warburg effect (Gatenby and Gillies 2004; Warburg 1930), is central to cancer biology and is proposed to arise as a result of the stabilization of HIF by lactic acid (Lu et al. 2002), aberrant HIF degradation by mutated von Hippel–Lindau protein (Ivanov et al. 1998), or dysfunctional transcription factors, such as c-myc, which regulate carbohydrate turnover (Liao and Dickson 2000). The persistence of the Warburg effect in cancer cells must provide a benefit for survival that outweighs the lower ATP yield of anaerobic respiration. One advantage of the Warburg effect is its lack of dependence on O2 supply, which is highly unreliable in tumors (Gatenby and Gillies 2004). Fluctuating blood flow has been proposed to promote cancer development by selecting for cells that survive repeated hypoxia-reoxygenation cycles (Gatenby and Gillies 2004; Kimura et al. 1996), i.e. those supporting the Warburg phenomenon. A second plausible benefit of the Warburg effect is the greater acid-yield per ATP molecule generated, seven-fold higher than aerobic respiration. It has been suggested that lactic acid overproduction sets up an acidic microenvironment that is critical for the transition from pre-invasive to malignant carcinoma (Gatenby and Gillies 2004). A positive correlation has been established between glycolytic flux (and hence acid production) and tumor aggressiveness and invasiveness (Gatenby
2 The Tumor Microenvironment
27
and Gillies 2004). Since glucose supply is usually plentiful, it is feasible to upregulate lactic acid production (viz., the Warburg effect) and deposit a greater amount of acid than possible with aerobic respiration, while still catering for cellular energy demands. The near-universal enhancement of glucose utilization by tumors, shown by 18fluorodeoxyglucose positron emission tomography (Gambhir 2002), supports this model of enhanced acid production. According to modern hypotheses of cancer progression, an acidic extracellular milieu (a) encourages cancer cell development and (b) is lethal to normal cells. There is growing evidence for an important role for H+-sensitive G-protein coupled receptors (OGR1 (Ludwig et al. 2003) and G2A (Murakami et al. 2004)) in transducing extracellular acidity into a proliferation signal (Huang et al. 2008). In a neuroblastoma cell line, low pHe activates the inositol-1,4,5-trisphosphate (IP3) signaling pathway, which triggers Ca2+ release from intracellular stores and the activation of the MEK/ERK pathway. In addition, acidic pHe has been shown to favor tumor invasiveness by increasing cell migration, decreasing cell adhesion, and disrupting the extracellular matrix (Gatenby and Gillies 2004; Gatenby and Gawlinski 1996; Kato et al. 1992; Martinez-Zaguilan et al. 1996; Schlappack et al. 1991; Stock and Schwab 2009). Cancer cells are believed to survive better in an acidic extracellular microenvironment by acquiring a more potent acid-extruding phenotype that protects the intracellular compartment from acid buildup (Swietach et al. 2007; Gatenby et al. 2007). Inadequate acid extrusion in normal cells may explain the lethality of low pHe therein (Fig. 2.2).
Fig. 2.2 Cartoon of acid production from a normal cell (smooth outline) and a cancer cell (rigged outline) in an hypoxic region of a tumor, beyond the Krogh radius from a blood vessel (wavy cylinder on right). The cancer cell (bottom cell with jagged outline) produces more acid (gray-filled H+ ions) than the normal cell (top cell with smooth outline, emitting white-filled H+ ions). Extracellular acid (produced mainly by the overactive cancer cell) favors cancer cell growth and development, but has detrimental effects on normal cells due to their lack of sufficient acid-extruding mechanisms to protect the intracellular space from acidification
28
P. Swietach et al.
Transport of Acid Across the Surface Membrane Intracellular and Extracellular pH in Tumors For many decades after Otto Warburg’s seminal finding that tumors overproduce lactic acid, it was assumed that both pHi and pHe are acidic in cancers. The low pH detected using microelectrodes (Kallinowski et al. 1989; Wike-Hooley et al. 1984) was deemed to support this hypothesis, until 31P nuclear magnetic resonance (NMR) revealed alkaline pH in solid tumors (Griffiths et al. 1981). These apparently conflicting reports are easily reconciled by considering that 31P NMR reports mainly pHi, whereas electrodes are sensitive mainly to pHe. Subsequent NMR measurements with extracellular probes such as 3-aminopropylphosphonate confirmed low pHe (Gillies et al. 1994). Therefore, the major difference between cancer and normal tissue is that the former has lower pHe, typically around 6.9 (Vaupel et al. 1990), but as low as 6.0 in some reports (Gillies et al. 2002). Tumors, like normal tissue, require modestly alkaline pHi to support growth, as shown by in vitro (Park et al. 1999; Chiche et al. 2009) and in vivo (Chiche et al. 2009) experiments. The maintenance of high pHi at low pHe requires efficient acid extrusion, a phenotype that is now considered characteristic of cancer cells.
Regulation of Tumor pH by Membrane Transport There are four challenges to efficient pH–i regulation in tumors. Firstly, the demand placed on pHi - regulating mechanisms is considerable. Tumors respire glucose at 0.1–1.0 mM/min (Vaupel et al. 1989) and generate acid equivalents at a rate that could promptly disturb resting free [H+] (~0.06 mM at pHi 7.2). Secondly, the transmembrane efflux of CO2 and lactic acid is limited by their membrane permeability and intracellular dissociation into membrane-impermeant ions (H+, HCO3–, lactate). Of the two acid products, lactic acid is less readily excreted than CO2 because (a) lactic acid is a larger and more polar molecule than CO2, and therefore is less able to cross the membrane’s lipid bilayer, (b) H+ and lactate produced anaerobically stay 99.9% dissociated in cytoplasm because of the high acid dissociation constant of lactic acid (10–4 M), whereas CO2 dissociates less strongly, with a ~90% yield at equilibrium (at pH 7; CO2 dissociation constant of 10–6.2 M). For every molecule of weak acid produced, anaerobic respiration leads to more intracellular H+ ion trapping than aerobic respiration, and therefore places a greater demand for H+ extruding mechanisms. However, intracellular conversion of CO2 into H+ and HCO3– can be catalyzed by intracellular isoforms of CA, a ubiquitously expressed enzyme (Swietach et al. 2007; Maren 1967; Supuran 2008). In a diffusion-reaction steady-state, intracellular hydration of CO2 reduces the capacity of its diffusive flux out of cells.
2 The Tumor Microenvironment
29
Thirdly, the share of cellular ATP for active acid extrusion is limited and cannot impinge on ATP reserves for growth and development. This may be particularly rate limiting in tumors that rely on energy inefficient anaerobic respiration. Finally, acid extruding mechanisms must be adequately active at low pHe. Extracellular H+ ions have been shown to have an inhibitory effect on transporters such as Na+/H+ exchange (Vaughan-Jones and Wu 1990; Aronson 1985). This would significantly reduce the capacity for acid extrusion in tumors with low pHe. To compensate for this, cancer cells may express a higher density of acid/base transporters, or express isoforms that are less susceptible to inhibition by low pHe.
Efflux of Metabolic Acid Figure 2.3 illustrates pathways for pHi regulation in tumor cells. Without specific mechanisms for transporting acid equivalents, cells would have to rely on the passive efflux of their acid products. CO2 and, to a lesser extent, lactic acid can cross the lipid bilayer, but efficient efflux requires a facilitated permeation pathway. Such
Fig. 2.3 The role of membrane transporters in facilitating acid efflux. Aerobic (left) and anaerobic (right) respiration of glucose yield the acid products CO2 and lactic acid, respectively. CO2 and, to a lesser degree, lactic acid are membrane-permeant and can exit across the surface membrane. At near-neutral pH, however, a fraction of CO2 and lactic acid will ionize. To control intracellular pH, cells express several transporters to facilitate transmembrane acid removal: AQP, aquaporins; MCT, monocarboxylic acid transport; HE, H+-ion extrusion; BT, HCO3– transport
30
P. Swietach et al.
facilitated transport can occur through aquaporin gas channels for CO2 (Nakhoul et al. 1998) (Fig. 2.3a) and by monocarboxylic acid transporters (MCT) for H++lactate (equivalent to lactic acid) co-transport (Halestrap and Price 1999) (Fig. 2.3b). Isoforms MCT1 and MCT4 have been identified in cancer. Furthermore, hypoxia induces MCT4 expression to match the HIF-induced rise in lactic acid production (Ullah et al. 2006).
Buffering of H+ Ions Like all weak acids, lactic acid and hydrated CO2 undergo deprotonation in aqueous medium. The magnitude of the intracellular [H+] rise can be reduced by means of pH buffering, for example: Buffer + H + ↔ HBuffer + Cells have high intracellular buffering capacity because of the proteins, peptides, amino acids, and other weak acids/bases that they harbor. On their own, buffers offer limited protection from acid loading as they can minimize, but cannot remove, pHi disturbances. Typically, it takes many mmol/L of acid to reduce pHi by one unit (Swietach et al. 2008, 2009). However, chronic acid loading over the lifetime of a cell would eventually saturate buffer sites and drive pHi into the lethal range. Cells can counteract buffer saturation by keeping the concentration of protonated and unprotonated buffer constant. In a so-called open buffer system, the concentration of either form of buffer can be restored by transport to or from the extracellular space. The major “open” buffer system in cells is CO2/HCO3–. Intracellular H+ ions react with HCO3– forming membrane permeant CO2 that can exit cells freely. Intracellular buffering is then restored by HCO32 uptake, e.g. by means of Na+-dependent Cl–/HCO3– exchange (Lee and Tannock 1998) (Fig. 2.3c). The effectiveness of this mechanism at protecting pHi depends critically on the availability of extracellular HCO3– for cellular uptake and on efficient CO2 venting. The extracellular space of most tissues is often assumed to be an infinite compartment with solutes at well-controlled concentrations. However, in tumors with poor blood flow, the extracellular milieu may not provide adequate HCO3– (particularly at low pHe) to perpetuate pH buffering by CO2/HCO3–. Under those circumstances, excess H+ ions may have to be extruded by HCO3–-independent mechanisms.
Extrusion of H+ Ions Excess H+ ions can be removed from cells by means of membrane-bound transporter proteins. High levels of Na+/H+ exchangers (NHE) (Swietach et al. 2008;
2 The Tumor Microenvironment
31
Lee and Tannock 1998; McLean et al. 2000) and vacuolar-type H+ ATPase (V-ATPase) pumps (Martinez-Zaguilan et al. 1993) have been detected in many types of cancer cell (Fig. 2.3d). Fibroblasts transfected with V-ATPase (Gillies et al. 1990) or NHE (Reshkin et al. 2000) become tumorigenic, suggesting an important role for H+ extruders in developing the cancer phenotype. In breast cancer, expression levels of V-ATPase are high and in excess of NHE, and correlate with metastatic potential (Sennoune et al. 2004). Experimental evidence supports the notion that H+ extruders contribute to extracellular acidification and help to keep pHi in a range that is favorable for growth (Sennoune et al. 2004; Fais et al. 2007). The enhanced capacity for secreting acid into the extracellular space may be a means of autocrine/paracrine signaling, in which cell proliferation is triggered via extracellularly activated G-protein coupled receptors sensitive to H+ (Huang et al. 2008). It is noteworthy that H+ extrusion and buffering do not remove the anionic conjugate base of the metabolic acid (lactate and HCO3–); therefore, additional mechanisms are required to remove or metabolize these. HCO3– efflux in exchange for extracellular Cl– can occur by means of transporters belonging to the SLC4 (Alper 2006) and SLC26 (Mount and Romero 2004) families. Recently described pathways for organic anion transport on SLC26 (Mount and Romero 2004) and SLC21 (Hagenbuch and Meier 2004) carriers may offer non-H+-coupled means of transporting lactate or its catabolic derivatives.
Regulation of pH Using Nonrespiratory Sources of H+ Ions In the discussion so far, the substrate for transmembrane acid flux was the acid generated by cellular respiration. In metabolically active cancer cells, this is the most substantial source of acid. However, the finding that solid tumors with genetically attenuated glycolysis are still able to acidify their extracellular space suggests that cancer cells can excrete acid from intracellular H+ donors other than metabolically generated acid (Newell et al. 1993; Yamagata et al. 1998). Intracellular buffers provide a substantial pool of H+ ions that can be used to fine-tune pHi independently of the cell’s metabolic activity. Without this reserve of H+ ions, the only means of fine-tuning pHi would be to adjust the balance between metabolic acid production and its transmembrane removal. If, for instance, the production and excretion of metabolic acid is perfectly matched, but a rise in pHi is desired, then the cell must extrude its reserve H+ ions. NHE is a major candidate for finetuning pHi, because of the plethora of signaling molecules that target its activity (Vaughan-Jones and Swietach 2008). Experiments have shown that NHE-driven intracellular alkalinization triggers early stages of neoplastic transformation (Reshkin et al. 2000), and is concomitant with cell cycle progression (Cardone et al. 2005).
32
P. Swietach et al.
Role of Carbonic Anhydrase in Acid-Equivalent Transport Intracellular and Extracellular Carbonic Anhydrase Isoforms Several steps in the pathways illustrated in Fig. 2.3 depend on the forward and backward rate constants in the equilibrium: CO2 + H 2 O ↔ HCO3− + H + The kinetics of pH buffering by CO2/HCO3– and CO2 hydration (and the reverse reaction) depend critically on CA expression levels. The first description of CA activity was based on the intracellular compartment of red blood cells (Meldrum and Roughton 1933). The high intracellular CA (CAi) activity in these cells is now considered to be an exception rather than rule, as typical CAi levels in most cells accelerate CO2 hydration by several, rather than several thousand, fold (Swietach et al. 2007, 2008, 2009; Leem and Vaughan-Jones 1998). The moderate CA catalysis inside cells is a compromise that (a) supports adequate intracellular CO2/HCO3– buffering kinetics (e.g. for the reaction between intracellular HCO3– and lactic acid, Fig. 2.3c) and (b) prevents excessive hydration of cell-generated CO2 (e.g. from aerobic production, Fig. 2.3a) that would trap H+ and HCO3– inside cells. Known catalytically-active intracellular CA isoforms include CA1, CA2, CA3, CA7, and CA13 (Maren 1967; Supuran 2008). Recently, a number of membrane-bound CA isoforms with extracellular-facing catalytic sites (CAe) have been described, including CA4, CA9, CA12, and CA14 (Maren 1967; Supuran 2008; Geers and Gros 2000; Pastorek et al. 1994). Two of these, CA9 and CA12, are hypoxically induced (Wykoff et al. 2000; Loncaster et al. 2001; Airley et al. 2003) and correlate positively with cancer aggressiveness and poor prognosis (Haapasalo et al. 2006; Driessen et al. 2006; Beasley et al. 2001; Chia et al. 2001; Koukourakis et al. 2001), although CA12 is also induced by estrogen in breast cancer and is associated then with a good outcome. It has been postulated that CAe activity in tumors accelerates acid extrusion (Swietach et al. 2007, 2008, 2009; Chiche et al. 2009; Svastova et al. 2004). Figure 2.4 illustrates the possible mechanisms by which CAe activity achieves such facilitation.
Facilitated CO2 Diffusion Efficient CO2 efflux from aerobically respiring cells (Fig. 2.3a) can be curtailed by CO2 buildup in poorly perfused extracellular spaces. The outward transmembrane [CO2] gradient can be increased by hydrating extracellular CO2 into HCO3– and H+ (Swietach et al. 2009; Svastova et al. 2004) (Fig. 2.4a). Theoretical modeling predicts that optimal CO2 venting out of cells can be achieved by favoring extracellular over intracellular CO2 hydration (Swietach et al. 2009). This can be implemented by expressing CAe at a higher level than CAi. In multicellular spheroid models,
2 The Tumor Microenvironment
33
expression of CA9 activity has been shown to reduce intracellular acidity (Swietach et al. 2008, 2009) and increase extracellular acidity (Swietach et al. 2009) in a manner that is sensitive to CAe inhibitors. There is evidence, at least from some cell lines, that overexpression of CA9 decreases overall CAi activity (Swietach et al. 2008, 2009), thereby shifting the site of CO2 hydration from the intracellular to the extracellular compartment. This CAe catalyzed pathway is an example of facilitated CO2 diffusion, and results in a fall of pHe. An apparent contradiction may be inferred from the role of an hypoxia-inducible enzyme (CA9, CA12) in facilitating the efflux of aerobically generated CO2 (Fig. 2.4a). However, the CA9 gene is among the most hypoxia sensitive (Wykoff et al. 2000). Along a radial gradient of decreasing oxygen, CA9 levels are likely to rise ahead of the downregulation of aerobic respiration. The mechanism illustrated in Fig. 2.4a may operate in the intermediate zone, where hypoxia is sufficient to induce CAe but inadequate to block aerobic respiration. In more proximal, oxygenated regions, CAe activity becomes redundant due to the small cell-to-capillary distance for CO2 diffusion (Swietach et al. 2009). CAe-facilitated CO2 diffusion can also take in anaerobically respiring cells. The substrate for CAe catalysis could be CO2 generated by the pentose phosphate
Fig. 2.4 The role of extracellular carbonic anhydrase (CAe) in facilitating efflux of aerobic (left) and anaerobic (right) acid-products. (a) Aerobically generated CO2 crosses the membrane passively down a diffusion gradient, maintained in the outward direction by CAe-catalyzed hydration of excreted CO2. (b) Anaerobically generated lactic acid (HLac), buffered by intracellular HCO3–, produces CO2. The CO2 thus generated exits the cells down a diffusion gradient maintained in the outward direction by CAe-catalyzed hydration. HCO3– generated extracellularly is transported into the cell by BT (e.g. Na+–HCO3– cotransport) to complete the cycle. (c) The products of intracellular CO2 hydration, H+ and HCO3–, are transported out of cells by H+ extruders (HE, e.g. Na+/H+ exchange) and bicarbonate transporters (BT, e.g. Cl–/HCO3– exchange). CAe-catalyzed buffering of H+ ions with HCO3– minimizes possible inhibition of HE by low pHe. (d) HLac flux on monocarboxylic acid transporters (MCT) is maintained in the outward direction by CAe-catalyzed titration of HLac and HCO3–. Water omitted for clarity. Pathways in the upper half tend to decrease pHe, and those in the lower half tend to increase pHe. All pathways stimulate acid efflux and raise pHi
34
P. Swietach et al.
shunt. A more important source of CO2 is intracellular buffering of lactic acid by HCO3–, a reaction catalyzed by CAi, if present (Figs. 2.3c and 2.4b). In addition to supporting an outward [CO2] gradient, CAe activity also generates extracellular HCO3– that is then taken up by cells to replenish intracellular CO2/HCO3– buffering capacity.
Facilitated H+ Diffusion A fraction of aerobically generated CO2 will undergo intracellular hydration, particularly in the presence of CAi activity. Extrusion of the resulting H+ ions will then reduce pHe, which may feedback negatively on transport activity, e.g. by means of allosteric auto-inhibition, as shown for NHE (Vaughan-Jones and Wu 1990; Aronson 1985). CAe activity can support H+ - transport by catalyzing the buffering of cell-extruded H+ ions with extracellular HCO3– (Fig. 2.4c). As a consequence of such facilitated H+ diffusion, pHe increases. In anaerobically respiring cells, the yield of intracellular lactic acid titration can be compromised, particularly under conditions of limited HCO3– uptake, e.g. inadequate transport activity or low extracellular substrate. Intracellular lactic acid that does not react with HCO3– can be extruded by MCT as H++lactate (Fig. 2.3b). CAe activity can stimulate this pathway as well, again by titrating the extruded acid with extracellular HCO3–. This converts lactic acid into CO2, a more mobile and less ionizing acid. Consequently, pHe increases and maintains a steeper outward gradient for H+-lactate co-transport (Fig. 2.4d). This pathway is another example of facilitated H+ diffusion.
The Transport Metabolon In the aforementioned pathways, CA activity works in functional association with membrane transporters. This functional coupling has been dubbed the “transport metabolon” and some reports have also proposed a structural basis for the CA-transporter interaction (Sterling et al. 2001; Becker et al. 2005). The transport metabolon was originally proposed for the collaboration between CAi isoforms and Cl−/HCO3− exchange (Sterling et al. 2001), but there is now evidence for a similar interaction between CAe isoforms and HCO3− transporters (Alvarez et al. 2003; Morgan et al. 2007). As membrane transport is a diffusion-reaction process, the transport metabolon ensures that, through the provision and consumption of transport substrate, the membrane-translocation step is not rate-limited by the cis supply and trans removal of solute. The functional significance of the transport metabolon has, however, been questioned (Lu et al. 2006), particularly under conditions of freely available and equilibrated extracellular CO2/HCO3− buffer (Swietach et al. 2008).
2 The Tumor Microenvironment
35
The Dominant Role for Carbonic Anhydrase in Tumors The balance between the four CAe-assisted pHi-regulatory pathways illustrated in Fig. 2.4 will depend on (a) the dominant mode of cellular respiration, (b) the expression and activity of transporter proteins, and (c) the availability of substrate, such as extracellular HCO3– (Swietach et al. 2009). In hypoxic regions, and in cancer cells showing a pronounced Warburg effect, lactic acid production dominates. Elsewhere, in regions of tumors with functional mitochondria and adequate oxygenation, aerobic respiration to CO2 will form the bulk of acid production. However, the transmembrane flux of CO2 versus H+ (lactic acid) may not follow the pattern of aerobic versus anaerobic respiration. In several cancer cell lines, HCO3–dependent transport has been shown to dominate pHi regulation over HCO3–independent fluxes, such as NHE (Swietach et al. 2008, 2009; Lee and Tannock 1998). This provides the transport-protein machinery to support intracellular lactic acid titration with HCO3–, and hence transmembrane CO2 efflux in glycolytic cells. In the corresponding tumors, the source of CO2 for transmembrane efflux would vary with distance from blood vessels, from aerobic CO2 production to titration of lactic acid by HCO3–. A high overall CO2 secretion rate is in agreement with in vivo data showing high extracellular CO2 levels in rodent solid tumors (Gullino et al. 1965) and that 70% of acid output in colon cancers is in the form of CO2 (Griffiths et al. 2001; Holm et al. 1995). CAe-assisted CO2 versus H+ diffusion can be distinguished on the basis of the effect of CAe inhibitors on pHe. In spheroids grown from CA9-transfected cells, CAe inhibitors increase pHe (Swietach et al. 2009). This suggests that the major effect of CAe catalysis in this preparation is to facilitate CO2 diffusion (Fig. 2.4a, b). The balance between CAe-assisted CO2 versus H+ diffusion in cancers in vivo awaits such pharmacological investigation. A dominance of the former would offer an elegant mechanistic link between two well-established correlations, that between tumor aggressiveness and low pHe (Vaupel et al. 1989; Gatenby et al. 2006; Gillies et al. 2002) and between tumor aggressiveness and the expression of CA9 and/or CA12 (Chia et al. 2001; Generali et al. 2006; Giatromanolaki et al. 2001; Koukourakis et al. 2006).
Future Directions and Outlook for Therapy Acidification of the extracellular milieu, in combination with powerful mechanisms to protect pHi, are believed to be a central hallmark of the aggressive cancer phenotype. Restoration of alkaline pHe is predicted to remove the survival advantage that cancer cells have over normal cells and to discourage invasive behavior and progression towards malignant forms of the cancer (Fang et al. 2008). In addition, raising pHe reduces ionization of weak basic drugs used in chemotherapy, such as epirubicin (Generali et al. 2006), and thereby increases drug penetrability to intracellular targets (Raghunand et al. 2001).
36
P. Swietach et al.
The use of bicarbonate salts to alkalinize blood, with the anticipated rise in tumor pHe, has been proposed as a means of tackling cancer (Raghunand et al. 2001). A more selective and specific therapy proposes pH-regulating transporters and enzymes as targets. Inhibition of these proteins has been proposed to be cancer’s Achilles heel (Fang et al. 2008). Recently, three targets for inhibition have been postulated as plausible therapeutic approaches to treating cancer: (a) H+ pumps, (b) monocarboxylic acid (H+-lactate) transporters, and (c) extracellular CA isoforms. Inhibition of V-ATPase activity using interference RNA reduces metastasis (Fais et al. 2007), and its pharmacological inhibition (e.g. with omeprazole) has been shown to improve responses to chemotherapy in solid tumors (Luciani et al. 2004) and promote death of B-cell tumors (De Milito et al. 2007). Fibroblasts transfected with the E7 oncogene failed to undergo intracellular alkalinization and neoplastic transformation when incubated with the NHE inhibitor, 5-(N,N-dimethyl) amiloride (Reshkin et al. 2000). Inhibition of MCT by cinnamate derivatives and, putatively, by lonidamine leads to cellular lactic acid accumulation and intracellular acidification, which are shown to be lethal in neuroblastoma cells (Fang et al. 2006). Although the efficacy of inhibiting V-ATPase, NHE, and MCT has been demonstrated in some in vitro cell lines and in vivo animal tumor models, the use of these drugs in anticancer therapy is not without drawbacks. Cancer cells tend to overexpress isoforms of V-ATPase, NHE, and MCT, but many noncancer cells also rely on these transporters for survival. The lack of a basis for tumor specificity would trigger drug toxicity and uncalibrated effects on whole-body physiology. Before experimental evidence for tumor-associated CAe isoforms was produced, the broad spectrum CA inhibitor acetazolamide has been shown to reduce in vivo tumor growth (Teicher et al. 1993). Following the discovery of CA9, the sulfonamide CA9-inhibitor indisulam was demonstrated to have powerful anticancer effects (Owa et al. 1999). The tumor-associated CAe isoforms, CA9 and CA12, offer an inhibitory target that is almost exclusive to cancer (with the exception of normal gastric mucosa). Membrane-impermeant, CA9- and CA12-selective inhibitors would target tumors, with minimal predicted side effects on other tissues. An impressive range of membrane-impermeant CAe inhibitors has been synthesized, and the effectiveness of these as anticancer drugs is currently being researched (Supuran 2008). Prodrugs activated by the extracellular acid environment would be another selective approach. Other therapeutic strategies based on tumor-specific CAe isoforms may include inhibitory antibodies, vaccines to develop T-cell antiCA9 or anti-CA12 responses, and targeted antibody therapy. Regulation of tumor pHi remains an intensively researched and hotly debated topic. Furthering our knowledge about the special adaptations found in tumors should improve our overall understanding of pH homeostasis, and may reveal more efficacious and safer anticancer therapy. Acknowledgments This work has been supported by the Medical Research Council and Royal Society (PS), Cancer Research UK (ALH) and British Heart Foundation (RDV-J).
2 The Tumor Microenvironment
37
References Airley RE, Loncaster J, Raleigh JA, Harris AL, Davidson SE, Hunter RD, West CM, Stratford IJ (2003) GLUT-1 and CAIX as intrinsic markers of hypoxia in carcinoma of the cervix: relationship to pimonidazole binding. Int J Cancer 104:85–91. Alper SL (2006) Molecular physiology of SLC4 anion exchangers. Exp Physiol 91:153–161. Alvarez BV, Loiselle FB, Supuran CT, Schwartz GJ, Casey JR (2003) Direct extracellular interaction between carbonic anhydrase IV and the human NBC1 sodium/bicarbonate co-transporter. Biochemistry 42:12321–12329. Aronson PS (1985) Kinetic properties of the plasma membrane Na+-H+ exchanger. Annu Rev Physiol 47:545–560. Beasley NJ, Wykoff CC, Watson PH, Leek R, Turley H, Gatter K, Pastorek J, Cox GJ, Ratcliffe P, Harris AL (2001) Carbonic anhydrase IX, an endogenous hypoxia marker, expression in head and neck squamous cell carcinoma and its relationship to hypoxia, necrosis, and microvessel density. Cancer Res 61:5262–5267. Becker HM, Hirnet DC, Fecher-Trost, Sultemeyer D, Deitmer JW (2005) Transport activity of MCT1 expressed in Xenopus oocytes is increased by interaction with carbonic anhydrase. J Biol Chem 280:39882–39889. Boron WF (2004) Regulation of intracellular pH. Adv Physiol Educ 28:160–179. Cardone RA, Casavola V, Reshkin SJ (2005) The role of disturbed pH dynamics and the Na+/H+ exchanger in metastasis. Nat Rev Cancer 5:786–795. Cerdan S, Rodrigues TB, Sierra A, Benito M, Fonseca LL, Fonseca CP, Garcia-Martin ML (2006) The redox switch/redox coupling hypothesis. Neurochem Int 48:523–530. Chia SK, Wykoff CC, Watson PH, Han C, Leek RD, Pastorek J, Gatter KC, Ratcliffe P, Harris AL (2001) Prognostic significance of a novel hypoxia-regulated marker, carbonic anhydrase ix, in invasive breast carcinoma. J Clin Oncol 19:3660–3668. Chiche J, Ilc K, Laferriere J, Trottier E, Dayan F, Mazure NM, Brahimi-Horn MC, Pouyssegur J (2009) Hypoxia-inducible carbonic anhydrase IX and XII promote tumor cell growth by counteracting acidosis through the regulation of the intracellular pH. Cancer Res 69: 358–368. De Milito A, Iessi E, Logozzi M, Lozupone F, Spada M, Marino ML, Federici C, Perdicchio M, Matarrese P, Lugini L, Nilsson A, Fais S (2007) Proton pump inhibitors induce apoptosis of human B-cell tumors through a caspase-independent mechanism involving reactive oxygen species. Cancer Res 67:5408–5417. Dewhirst MW, Secomb TW, Ong ET, Hsu R, Gross JF (1994) Determination of local oxygen consumption rates in tumors. Cancer Res 54:3333–3336. Driessen, A., Landuyt W, Pastorekova S, Moons J, Goethals L, Haustermans K, Nafteux P, Penninckx F, Geboes K, Lerut T, Ectors N (2006) Expression of carbonic anhydrase IX (CA IX), a hypoxia-related protein, rather than vascular-endothelial growth factor (VEGF), a pro-angiogenic factor, correlates with an extremely poor prognosis in esophageal and gastric adenocarcinomas. Ann Surg 243:334–340. Fais S, De Milito A, You H, Qin W (2007) Targeting vacuolar H+-ATPases as a new strategy against cancer. Cancer Res 67:10627–10630. Fang J, Quinones QJ, Holman TL, Morowitz MJ, Wang Q, Zhao H, Sivo F, Maris JM, Wahl ML (2006) The H+-linked monocarboxylate transporter (MCT1/SLC16A1): a potential therapeutic target for high-risk neuroblastoma. Mol Pharmacol 70:2108–2115. Fang JS, Gillies RD, Gatenby RA (2008) Adaptation to hypoxia and acidosis in carcinogenesis and tumor progression. Semin Cancer Biol 18:330–337. Gambhir SS (2002) Molecular imaging of cancer with positron emission tomography. Nat Rev Cancer 2:683–693. Gatenby RA, Gawlinski ET (1996) A reaction-diffusion model of cancer invasion. Cancer Res 56:5745–5753.
38
P. Swietach et al.
Gatenby RA, Gillies RJ (2004) Why do cancers have high aerobic glycolysis? Nat Rev Cancer 4:891–899. Gatenby RA, Gawlinski ET, Gmitro AF, Kaylor B, Gillies RJ (2006) Acid-mediated tumor invasion: a multidisciplinary study. Cancer Res 66:5216–5223. Gatenby RA, Smallbone K, Maini PK, Rose F, Averill J, Nagle RB, Worrall L, Gillies RJ (2007) Cellular adaptations to hypoxia and acidosis during somatic evolution of breast cancer. Br J Cancer 97:646–653. Geers C, Gros G (2000) Carbon dioxide transport and carbonic anhydrase in blood and muscle. Physiol Rev 80:681–715. Generali D, Fox SB, Berruti A, Brizzi MP, Campo L, Bonardi S, Wigfield SM, Bruzzi P, Bersiga A, Allevi G, Milani M, Aguggini S, Dogliotti L, Bottini A, Harris AL (2006) Role of carbonic anhydrase IX expression in prediction of the efficacy and outcome of primary epirubicin/ tamoxifen therapy for breast cancer. Endocr Relat Cancer 13:921–930. Giatromanolaki A, Koukourakis MI, Sivridis E, Pastorek J, Wykoff CC, Gatter KC, Harris AL (2001) Expression of hypoxia-inducible carbonic anhydrase-9 relates to angiogenic pathways and independently to poor outcome in non-small cell lung cancer. Cancer Res 61:7992–7998. Gillies RJ, Liu Z, Bhujwalla Z (1994) 31P-MRS measurements of extracellular pH of tumors using 3-aminopropylphosphonate. Am J Physiol 267:C195–203. Gillies RJ, Martinez-Zaguilan R, Martinez GM, Serrano R, Perona R (1990) Tumorigenic 3T3 cells maintain an alkaline intracellular pH under physiological conditions. Proc Natl Acad Sci USA 87:7414–7418. Gillies RJ, Raghunand N, Karczmar GS, Bhujwalla ZM (2002) MRI of the tumor microenvironment. J Magn Reson Imaging 16:430–450. Gillies RJ, Robey I, Gatenby RA (2008) Causes and consequences of increased glucose metabolism of cancers. J Nucl Med 49(Suppl 2):24S–42S. Griffiths JR, Stevens AN, Iles RA, Gordon RE, Shaw D (1981) 31P-NMR investigation of solid tumours in the living rat. Biosci Rep 1:319–325. Griffiths JR, McIntyre DJ, Howe FA, Stubbs M (2001) Why are cancers acidic? A carrier-mediated diffusion model for H+ transport in the interstitial fluid. Novartis Found Symp 240:46–62. Gullino PM, Grantham FH, Smith SH, Haggerty AC (1965) Modifications of the acid-base status of the internal milieu of tumors. J Natl Cancer Inst 34:857–869. Hagenbuch, B., Meier PJ (2004) Organic anion transporting polypeptides of the OATP/SLC21 family: phylogenetic classification as OATP/SLCO superfamily, new nomenclature and molecular/functional properties. Pflugers Arch 447:653–665. Halestrap AP, Price NT (1999) The proton-linked monocarboxylate transporter (MCT) family: structure, function and regulation. Biochem J 343(Pt 2):281–299. Hardie DG (2003) Minireview: the AMP-activated protein kinase cascade: the key sensor of cellular energy status. Endocrinology 144:5179–5183. Helmlinger G, Yuan F, Dellian M, Jain RK (1997) Interstitial pH and pO2 gradients in solid tumors in vivo: high-resolution measurements reveal a lack of correlation. Nat Med 3:177–182. Helmlinger G, Sckell A, Dellian M, Forbes NS, Jain RK (2002) Acid production in glycolysisimpaired tumors provides new insights into tumor metabolism. Clin Cancer Res 8:1284–1291. Haapasalo JA, Nordfors KM, Hilvo M, Rantala IJ, Soini Y, Parkkila AK, Pastorekova S, Pastorek J, Parkkila SM, Haapasalo HK (2006) Expression of carbonic anhydrase IX in astrocytic tumors predicts poor prognosis. Clin Cancer Res 12:473–477. Holm E, Hagmuller E, Staedt U, Schlickeiser G, Gunther HJ, Leweling H, Tokus M, Kollmar HB (1995) Substrate balances across colonic carcinomas in humans. Cancer Res 55:1373–1378. Huang WC, Swietach P, Vaughan-Jones RD, Ansorge O, Glitsch MD (2008) Extracellular acidification elicits spatially and temporally distinct Ca2+ signals. Curr Biol 18:781–785. Isfort RJ, Cody DB, Asquith TN, Ridder GM, Stuard SB, LeBoeuf RA (1993) Induction of protein phosphorylation, protein synthesis, immediate-early-gene expression and cellular proliferation by intracellular pH modulation. Implications for the role of hydrogen ions in signal transduction. Eur J Biochem 213:349–357.
2 The Tumor Microenvironment
39
Ivanov SV, Kuzmin I, Wei MH, Pack S, Geil L, Johnson BE, Stanbridge EJ, Lerman MI (1998) Down-regulation of transmembrane carbonic anhydrases in renal cell carcinoma cell lines by wild-type von Hippel-Lindau transgenes. Proc Natl Acad Sci USA 95:12596–12601. Kallinowski F, Schlenger KH, Runkel S, Kloes M, Stohrer M, Okunieff P, Vaupel P (1989) Blood flow, metabolism, cellular microenvironment, and growth rate of human tumor xenografts. Cancer Res 49:3759–3764. Kimura H, Braun RD, Ong ET, Hsu R, Secomb TW, Papahadjopoulos D, Hong K, Dewhirst MW (1996) Fluctuations in red cell flux in tumor microvessels can lead to transient hypoxia and reoxygenation in tumor parenchyma. Cancer Res 56:5522–5528. Koukourakis MI, Giatromanolaki A, Sivridis E, Simopoulos K, Pastorek J, Wykoff CC, Gatter KC, Harris AL (2001) Hypoxia-regulated carbonic anhydrase-9 (CA9) relates to poor vascularization and resistance of squamous cell head and neck cancer to chemoradiotherapy. Clin Cancer Res 7:3399–3403. Koukourakis MI, Pitiakoudis M, Giatromanolaki A, Tsarouha A, Polychronidis A, Sivridis E, Simopoulos C (2006) Oxygen and glucose consumption in gastrointestinal adenocarcinomas: correlation with markers of hypoxia, acidity and anaerobic glycolysis. Cancer Sci 97: 1056–1060. Krogh A (1919) The number and distribution of capillaries in muscles with calculations of the oxygen pressure head necessary for supplying the tissue. J Physiol 52:409–415. Lee AH, Tannock IF (1998) Heterogeneity of intracellular pH and of mechanisms that regulate intracellular pH in populations of cultured cells. Cancer Res 58:1901–1908. Leem CH, Vaughan-Jones RD (1998) Out-of-equilibrium pH transients in the guinea-pig ventricular myocyte. J Physiol 509 (Pt 2):471–485. Liao DJ, Dickson RB (2000) c-Myc in breast cancer. Endocr Relat Cancer 7:143–164. Loncaster JA, Harris AL, Davidson SE, Logue JP, Hunter RD, Wycoff CC, Pastorek J, Ratcliffe PJ, Stratford IJ, West CM (2001) Carbonic anhydrase (CA IX) expression, a potential new intrinsic marker of hypoxia: correlations with tumor oxygen measurements and prognosis in locally advanced carcinoma of the cervix. Cancer Res 61:6394–6399. Lu H, Forbes RA, Verma A (2002) Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. J Biol Chem 277:23111–23115. Lu J, Daly CM, Parker MD, Gill HS, Piermarini PM, Pelletier MF, Boron WF (2006) Effect of human carbonic anhydrase II on the activity of the human electrogenic Na/HCO3 cotransporter NBCe1-A in Xenopus oocytes. J Biol Chem 281:19241–19250. Luciani F, Spada M, De Milito A, Molinari A, Rivoltini L, Montinaro A, Marra M, Lugini L, Logozzi M, Lozupone F, Federici C, Iessi E, Parmiani G, Arancia G, Belardelli F, Fais S (2004) Effect of proton pump inhibitor pretreatment on resistance of solid tumors to cytotoxic drugs. J Natl Cancer Inst 96:1702–1713. Ludwig MG, Vanek M, Guerini D, Gasser JA, Jones CE, Junker U, Hofstetter H, Wolf RM, Seuwen K (2003) Proton-sensing G-protein-coupled receptors. Nature 425:93–98. Maren TH (1967) Carbonic anhydrase: chemistry, physiology, and inhibition. Physiol Rev 47:595–781. Martinez-Zaguilan R, Lynch RM, Martinez GM, Gillies RJ (1993) Vacuolar-type H(+)-ATPases are functionally expressed in plasma membranes of human tumor cells. Am J Physiol 265:C1015–1029. Martinez-Zaguilan R, Seftor EA, Seftor RE, Chu YW, Gillies RJ, Hendrix MJ (1996) Acidic pH enhances the invasive behavior of human melanoma cells. Clin Exp Metastasis 14:176–186. McConkey DJ, Orrenius S (1996) Signal transduction pathways in apoptosis. Stem Cells 14:619–631. McLean LA, Roscoe J, Jorgensen NK, Gorin FA, Cala PM (2000) Malignant gliomas display altered pH regulation by NHE1 compared with nontransformed astrocytes. Am J Physiol Cell Physiol 278:C676–C688. Meldrum NU, Roughton FJ (1933) Carbonic anhydrase. Its preparation and properties. J Physiol 80:113–142. Morgan PE, Pastorekova S, Stuart-Tilley AK, Alper SL, Casey JR (2007) Interactions of transmembrane carbonic anhydrase, CAIX, with bicarbonate transporters.. Am J Physiol Cell Physiol. 293:C738–C748.
40
P. Swietach et al.
Mount DB, Romero MF (2004) The SLC26 gene family of multifunctional anion exchangers. Pflugers Arch 447:710–721. Murakami N, Yokomizo T, Okuno T, Shimizu T (2004) G2A is a proton-sensing G-protein-coupled receptor antagonized by lysophosphatidylcholine. J Biol Chem 279:42484–42491. Kato Y, Nakayama Y, Umeda M, Miyazaki K (1992) Induction of 103-kDa gelatinase/type IV collagenase by acidic culture conditions in mouse metastatic melanoma cell lines. J Biol Chem 267:11424–11430. Nakhoul NL, Davis BA, Romero MF, Boron WF (1998) Effect of expressing the water channel aquaporin-1 on the CO2 permeability of Xenopus oocytes. Am J Physiol 274:C543–C548. Newell K, Franchi A, Pouyssegur J, Tannock I (1993) Studies with glycolysis-deficient cells suggest that production of lactic acid is not the only cause of tumor acidity. Proc Natl Acad Sci USA 90:1127–1131. Owa T, Yoshino H, Okauchi T, Yoshimatsu K, Ozawa Y, Sugi NH, Nagasu T, Koyanagi N, Kitoh K (1999) Discovery of novel antitumor sulfonamides targeting G1 phase of the cell cycle. J Med Chem 42:3789–3799. Park HJ, Lyons JC, Ohtsubo T, Song CW (1999) Acidic environment causes apoptosis by increasing caspase activity. Br J Cancer 80:1892–1897. Pastorek J, Pastorekova S, Callebaut I, Mornon JP, Zelnik V, Opavsky R, Zat’ovicova M, Liao S, Portetelle D, Stanbridge EJ et al (1994) Cloning and characterization of MN, a human tumorassociated protein with a domain homologous to carbonic anhydrase and a putative helixloop-helix DNA binding segment. Oncogene 9:2877–2888. Ramaiah A, Hathaway JA, Atkinson DE (1964). Adenylate as a Metabolic Regulator. Effect on Yeast Phosphofructokinase Kinetics. J Biol Chem 239:3619–3622. Raghunand N, Mahoney B, van Sluis R, Baggett B, Gillies RJ (2001) Acute metabolic alkalosis enhances response of C3H mouse mammary tumors to the weak base mitoxantrone. Neoplasia 3:227–235. Reshkin SJ, Bellizzi A, Caldeira S, Albarani V, Malanchi I, Poignee M, Alunni-Fabbroni M, Casavola V, Tommasino M (2000) Na+/H+ exchanger-dependent intracellular alkalinization is an early event in malignant transformation and plays an essential role in the development of subsequent transformation-associated phenotypes. FASEB J 14:2185–2197. Roos A, Boron WF (1981) Intracellular pH. Physiol Rev 61:296–434. Schlappack OK, Zimmermann A, Hill RP (1991) Glucose starvation and acidosis: effect on experimental metastatic potential, DNA content and MTX resistance of murine tumour cells. Br J Cancer 64:663–670. Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3:721–732. Semenza GL, Wang GL (1992) A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol Cell Biol 12:5447–5454. Semenza GL, Roth PH, Fang HM, Wang GL (1994) Transcriptional regulation of genes encoding glycolytic enzymes by hypoxia-inducible factor 1. J Biol Chem 269:23757–23763. Sennoune SR, Bakunts K, Martinez GM, Chua-Tuan JL, Kebir Y, Attaya MN, Martinez-Zaguilan R (2004) Vacuolar H+-ATPase in human breast cancer cells with distinct metastatic potential: distribution and functional activity. Am J Physiol Cell Physiol 286:C1443–C1452. Sterling D, Reithmeier RA, Casey JR (2001) A transport metabolon. Functional interaction of carbonic anhydrase II and chloride/bicarbonate exchangers. J Biol Chem 276:47886–47894. Stock C, Schwab A (2009) Protons make tumor cells move like clockwork. Pflugers Arch 458(2):981–992. Stubbs M, McSheehy PM, Griffiths JR, Bashford CL (2000) Causes and consequences of tumour acidity and implications for treatment. Mol Med Today 6:15–19. Supuran CT (2008) Carbonic anhydrases: novel therapeutic applications for inhibitors and activators. Nature Rev (Drug Discoveries) 7:1–14. Svastova E, Hulikova A, Rafajova M, Zat’ovicova M, Gibadulinova A, Casini A, Cecchi A, Scozzafava A, Supuran CT, Pastorek J, Pastorekova S (2004) Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH. FEBS Lett 577:439–445.
2 The Tumor Microenvironment
41
Swietach P, Vaughan-Jones RD, Harris AL (2007) Regulation of tumor pH and the role of carbonic anhydrase 9. Cancer Metastasis Rev 26:299–310. Swietach P, Wigfield S, Cobden P, Supuran CT, Harris AL, Vaughan-Jones RD (2008) Tumorassociated carbonic anhydrase 9 spatially coordinates intracellular pH in three-dimensional multicellular growths. J Biol Chem 283:20473–20483. Swietach P, Patiar S, Supuran CT, Harris AL, Vaughan-Jones RD (2009) The role of carbonic anhydrase 9 in regulating extracellular and intracellular pH in three-dimensional tumor cell growths. J Biol Chem 284:20299–20310. Teicher BA, Liu SD, Liu JT, Holden SA, Herman TS (1993) A carbonic anhydrase inhibitor as a potential modulator of cancer therapies. Anticancer Res 13:1549–1556. Thomlinson RH, Gray LH (1955) The histological structure of some human lung cancers and the possible implications for radiotherapy. Br J Cancer 9:539–549. Ullah MS, Davies AJ, Halestrap AP (2006) The plasma membrane lactate transporter MCT4, but not MCT1, is up-regulated by hypoxia through a HIF-1alpha-dependent mechanism. J Biol Chem 281:9030–9037. Vaughan-Jones RD, Swietach P (2008) Pushing and pulling the cardiac sodium/hydrogen exchanger. Circ Res 103:773–775. Vaughan-Jones RD, Wu ML (1990) Extracellular H+ inactivation of Na(+)-H+ exchange in the sheep cardiac Purkinje fibre. J Physiol 428:441–466. Vaughan-Jones RD, Spitzer KW, Swietach P (2008) Intracellular pH regulation in heart. J Mol Cell Cardiol 46(3):318–313. Vaupel P, Kallinowski F, Okunieff P (1989) Blood flow, oxygen and nutrient supply, and metabolic microenvironment of human tumors: a review. Cancer Res 49:6449–6465. Vaupel P, Kallinowski F, Okunieff P (1990) Blood flow, oxygen consumption and tissue oxygenation of human tumors. Adv Exp Med Biol 277:895–905. Warburg O (1930) The metabolism of tumours. Arnold Constable, London. Wike-Hooley JL, Haveman J, Reinhold HS (1984) The relevance of tumour pH to the treatment of malignant disease. Radiother Oncol 2:343–366. Wykoff CC, Beasley NJ, Watson PH, Turner KJ, Pastorek J, Sibtain A, Wilson GD, Turley H, Talks KL, Maxwell PH, Pugh CW, Ratcliffe PJ, Harris AL (2000) Hypoxia-inducible expression of tumor-associated carbonic anhydrases. Cancer Res 60:7075–7083. Yamagata M, Hasuda K, Stamato T, Tannock IF (1998) The contribution of lactic acid to acidification of tumours: studies of variant cells lacking lactate dehydrogenase. Br J Cancer 77:1726–1731.
Chapter 3
Hypoxia, Gene Expression, and Metastasis Olga V. Razorenova and Amato J. Giaccia
Abstract Metastasis is the primary cause of death from cancer due to the spread of disease throughout the body. Increasing evidence suggests that the hypoxic microenvironment serves as a driving force for the metastatic process. Fifty to sixty percent of solid tumors contain hypoxic areas, where the gene expression is reprogrammed by low oxygen microenvironment leading to aggressive invasive cancer cell behavior. Hypoxia upregulates multiple genes involved in different steps of metastatic process, including angiogenesis, proliferation, migration, invasion, motility, adhesion, ECM remodeling, and survival. Moreover, hypoxia confers tumor cells with chemo- and radio-resistance. At the end of this chapter, we discuss the facts linking hypoxia and cancer stem cells (CSC) mainly through the ability of hypoxic microenvironment to shift cells toward the undifferentiated phenotype. Abbreviations AKT ARNT CC-RCC CNS CSC EMT FAK FIH GSC HIF HRE
protein kinase B arylhydrocarbon receptor nuclear translocator clear cell renal cell carcinoma central nervous system cancer stem cell epithelial–mesenchymal transition focal adhesion kinase factor inhibiting HIF glioma stem cell hypoxia-inducible factor hypoxia-responsive element
A.J. Giaccia (*) Division of Radiation and Cancer Biology, Stanford University, Stanford, CA 94305, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_3, © Springer Science+Business Media, LLC 2010
43
44
IDH1 iPS a-KG LOH LOX MAPK NSC PHDs 1–3 PI-3K pVHL VHL WT
O.V. Razorenova and A.J. Giaccia
isocitrate dehydrogenase 1 induced pluripotent stem cell a-ketoglutarate loss of heterozygosity lysyl oxidase mitogen-activated protein kinase neural stem cell prolyl-4-hydroxylases 1–3 phosphoinositide 3 kinase von Hippel–Lindau protein von Hippel–Lindau gene wild-type
The Link Between Hypoxia and Metastasis Metastatic disease is the primary cause of death from cancer due to its spread beyond the primary affected organ to multiple organs in the body (secondary sites). The metastatic process is complex and can be divided into several major steps: (1) invasion and migration of tumor cells, leading to intravasation; (2) survival in the circulation (resistance to anoikis); (3) adhesion and migration, leading to extravasation; (4) survival in the new microenvironment; and (5) proliferation in the secondary site (Fig. 3.1) (Chan and Giaccia 2007; Nguyen et al. 2009). These processes require the involvement of a variety of effector molecules. It has been estimated that 50–60% of solid tumors contain areas of hypoxic and/or anoxic regions (Vaupel and Mayer 2007). Moreover, hypoxia correlates with poor patient outcome and the presence of metastases in multiple tumor types (Hockel et al. 1993, 1996; Brizel et al. 1996; Nordsmark et al. 1996; Fyles et al. 1998). Despite these correlations, it is not clear whether hypoxia provides the appropriate microenvironment promoting metastatic spread or metastasis is a simple consequence of chemo- and radio-resistance of cells in hypoxic regions. Hypoxic cells survive chemotherapy because many chemotherapeutic agents require proliferating cells to induce cytotoxicity and hypoxic cells divide slowly (Shannon et al. 2003). The second cause of failure is the need for vasculature to deliver drugs to cancer cells, which is absent or defective in hypoxic regions (Kizaka-Kondoh et al. 2003). An additional contribution to chemotherapy resistance is through the hypoxic induction of multidrug transporters, leading to the transport of drugs out of the cell (Comerford et al. 2002; Wartenberg et al. 2003). Radiation therapy relies on molecular oxygen to form cytotoxic, double-strand breaks in the cancer cell DNA, which is limited under hypoxic conditions (Kallman and Dorie 1986; Hall 1994). The majority of studies now provide convincing evidence that hypoxia is not only capable to confer cells with chemo- and radio-resistance, but also shifts gene expression to promote the metastatic process (see below).
3 Hypoxia, Gene Expression, and Metastasis
hypoxia
45
intravasation angiogenic factor secretion
adh
esio
n
angiogenesis
extravasation
organ2
surv i
val
organ1
survival, proliferation
Fig. 3.1 Major steps of the metastatic process
Causes and Consequences of Tumor Hypoxia The physiologic range of oxygen for most tissues is between 4 and 8% oxygen (normoxia) in contrast to the 20% level of oxygen used for tissue culture experiments (Gardner and Corn 2008). For each tissue, the range of oxygen tensions can vary accordingly to need; thus, under physiologic conditions it is difficult to establish a universal value to define hypoxia. In this regard, the functional definition of hypoxia is the state when oxygen delivery does not meet the demands of the given tissue. Cellular hypoxia (<1–5% oxygen) is a common stress in normal development (Huang et al. 2004) and numerous pathological conditions, including cancer and ischemia (Giaccia et al. 2003; Semenza 2003). Hypoxic areas in tumors develop as a result of an imbalance between oxygen supply and consumption (Papandreou et al. 2005; Vaupel and Mayer 2007). The expanding tumor mass increases the distance of certain tumor areas from local blood vessels, limiting the oxygen supply. Additionally, at later stages of tumor growth, when a tumor establishes an extensive vasculature to increase oxygen supply, the defects in blood vessel structure contribute to hypoxia reoxygenation of certain tumor areas (Endrich et al. 1979; Grunt et al. 1985; Shah-Yukich and Nelson 1988; Dewhirst et al. 1989). The malformed tumor vasculature is twisted, tortuous with blind ends and can possess arterio-venous shunts. Moreover, tumor vessels are devoid of their normal multilayer complex structure and are leaky and predisposed to dilation and constriction, resulting in transient changes in tissue oxygenation (Carmeliet 2005). The main consequence of hypoxic exposure is the stabilization of hypoxiainducible factors (HIFs), which are critical regulators of gene transcription (Giaccia et al. 2003). The hypoxic microenvironment of the tumor leads to HIF stabilization, which induces gene expression changes in order to adapt to hypoxia (Semenza
46
O.V. Razorenova and A.J. Giaccia
2003; Lee et al. 2004); some of these changes, in turn, allow the acquisition of prometastatic properties of cells (invasiveness, migration, adhesion, and survival) (Giaccia et al. 2003; Semenza 2003). Indeed, HIF stabilization in solid tumors does not simply correlate with poor patient prognosis, but HIFs are active players in the metastatic process (Le et al. 2004; Chan and Giaccia 2007; Rankin and Giaccia 2008; Qing and Simon 2009). There are three known HIF-a molecules (Rankin and Giaccia 2008): HIF-1a, HIF-2a, and HIF-3a, which are all hypoxia inducible with a similar mechanism of regulation by oxygen. HIF-1a is ubiquitously expressed, while HIF-2a expression is restricted to specific cell types: endothelial cells, glial cells, type II pneumocytes, cardiomyocytes, kidney fibroblasts, interstitial cells of the pancreas and duodenum, and hepatocytes (Wiesener et al. 2003). Our knowledge of HIF-3a is currently very limited and some studies suggest that it might act as a dominantnegative regulator of HIF-1a and HIF-2a by generation of an alternatively spliced isoform (Makino et al. 2001). Both HIF-1a and HIF-2a are detected in a variety of human tumors, and elevated HIF-1a expression correlates with poor patient outcome in head and neck cancer, nasopharyngeal carcinoma, colorectal, pancreatic, breast, cervical, osteosarcoma, endometrial, ovarian, bladder, glioblastoma, and gastric carcinomas, while elevated HIF-2a expression correlates with poor patient outcome in hepatocellular, colorectal carcinoma, melanoma, ovarian, and nonsmall cell lung cancers (Rankin and Giaccia 2008). Hypoxic conditions are stress conditions that trigger the cell to make a decision to live or to die. In response to hypoxia, as a pro-survival signal, HIF-a stability and activity increases (Giaccia et al. 2003; Semenza 2003). HIF-mediated gene expression allows cells to respond to a low oxygen environment by increasing oxygen delivery and metabolically to adapt to decreased oxygen availability. HIFs regulate expression of multiple target genes that are involved in glycolysis, angiogenesis, proliferation, migration, etc., by binding to hypoxia-responsive elements (HREs) (5¢RCGTG3¢) in their genomic regulatory regions (Semenza 2003; Le et al. 2004; Chan and Giaccia 2007).
HIF Regulation by Oxygen In untransformed cells, HIF-a activity is oxygen dependent and is tightly regulated (Fig. 3.2). Under normoxia, HIF-1a is hydroxylated on two conserved proline residues, prolines 402 and 564, by prolyl-4-hydroxylases 1–3 (PHDs 1–3) (Ivan et al. 2001; Jaakkola et al. 2001). This hydroxylation reaction requires oxygen and a-ketoglutarate (a-KG) as substrates, and iron and ascorbate as cofactors. When hydroxylated, HIF-1a is recognized by the pVHL protein, which allows the formation of a complex with Elongins B and C, cullin 2, and RING-H2 protein ring-box1. This complex functions as E3 ubiquitin ligase, and ubiquitinated HIF-1a gets degraded in the proteasome (Pause et al. 1999). Under hypoxic conditions, hydroxylation of HIF-1a decreases in an oxygen-dependent manner as oxygen serves as a substrate in the hydroxylation reaction. HIF-1a without proline hydroxyl marks is
3 Hypoxia, Gene Expression, and Metastasis
Normoxia Presence of oxygen Cofactors: ascorbate Fe2+
Hypoxia Absence of oxygen pVHL
pVHL PHDs
HIFα
47
OH
O2 α-KG
OH
Ub Ub Ub
HIFα
HIFα
FIH
p300/ CBP
p300/ CBP
OH
26S
I
F
α
HIFα HIFβ
transcription of HIF target genes
5’RCGTG 3’YGCAC
metabolism angiogenesis invasion migration proliferation
H
Fig. 3.2 Oxygen-dependent HIF regulation. R, purine; Y, pyrimidine
not recognized by pVHL, becomes stabilized, and dimerizes with its constitutively expressed partner HIF-1b (also known as arylhydrocarbon receptor nuclear translocator, ARNT). HIF-a/HIF-1b; heterodimer acts in the nucleus and activates transcription by recruiting the transcriptional co-activators such as p300/CBP (Giaccia et al. 2003; Semenza 2003). The interaction between HIF and p300/CBP is also oxygen dependent and is regulated by factor inhibiting HIF (FIH) (Mahon et al. 2001). Under normoxia, FIH hydroxylates asparagine 803 in HIF-1a transactivation domain and prevents p300/CBP binding. Thus, HIF activation under hypoxia consists of HIF-a stabilization and effective p300/CBP binding.
HIF Regulation by Genetic Alterations of Upstream Regulators Many studies have shown that hypoxia is not the only mechanism of HIF activation. HIF-a stability and activity are tightly controlled in primary cells by multiple signaling pathways. Not surprisingly cancer cells acquire mutations in genes, involved
48
O.V. Razorenova and A.J. Giaccia
into HIF regulation, allowing its activation in tumors under normoxic conditions. As described above, pVHL plays a central role in regulating HIF stability (Fig. 3.2). Genetic inactivation of the VHL gene results in HIF-a stabilization and activation, leading to upregulation of target gene expression under normoxia (Iliopoulos et al. 1996; Maxwell et al. 1999; Giaccia et al. 2003; Kaelin 2008). Germ-line mutations in VHL result in VHL disease, also known as familial Von Hippel–Lindau syndrome (Latif et al. 1993). Affected individuals are born with one wild-type (WT) allele and another mutant allele. In these individuals the remaining, WT VHL allele, is eliminated by a loss of heterozygosity (LOH). Von Hippel–Lindau syndrome is characterized by the high frequency of development of several cancer types, including central nervous system (CNS) and retinal hemangioblastomas, clear-cell renal cell carcinomas (CC-RCCs), and pheochromocytomas. VHL is also inactivated in the majority of sporadic RCCs and hemangioblastomas (up to 80% of cases) (Gnarra et al. 1993; Sprenger et al. 2001; Maynard and Ohh 2004). Families at risk for developing pheochromocytoma almost invariably harbor VHL missense mutations, in contrast to families without pheochromocytomas, which frequently harbor VHL deletions or truncation mutations (Chen et al. 1995; Zbar et al. 1996). Most likely genotype–phenotype correlations in VHL disease reflect the degree of impairment of pVHL function by different mutations (Hoffman et al. 2001; Li et al. 2007). Thus, VHL is a bona fide tumor suppressor gene. One of the best studied functions of HIF-1a is the regulation of glucose metabolism by glycolytic enzymes. Importantly, cell metabolic pathways feedback to HIF to allow tight control of energy generating process. Thus, mutations in isocitrate dehydrogenase 1 (IDH1), which occur in certain types of brain tumors, were found to impair a-KG synthesis. Because a-KG is a substrate in HIF-a hydroxylation reaction (Fig. 3.2), impairment of IDH1 function leads to increased HIF-a stability (Zhao et al. 2009). Importantly, IDH1 acts as a homodimer and mutations in IDH1 not only inactivate the gene product of the mutated allele, but confirm the mutated protein with the dominant-negative activity toward the WT allele (Zhao et al. 2009). This fact explains why the LOH for IDH1 is not observed in tumors. Thus, IDH1 appears to be another tumor suppressor gene involved in HIF-a destabilization. Activation of phosphoinositide 3 kinase/Akt (PI-3K/Akt) and mitogen-activated protein kinase (MAPK) signaling pathways is frequently observed in multiple types of cancer and it was shown that they can induce HIF activity (Berra et al. 2000; Blancher et al. 2001; Jiang et al. 2001; Sodhi et al. 2001). The PI-3K pathway is activated in cancer by enhanced growth factor signaling (Zelzer et al. 1998), activation of oncogenes (positive regulators) or inactivation of PTEN (Zundel et al. 2000), or TSC2 (Brugarolas et al. 2003) (negative regulators). To date, the detailed mechanism of PI-3K/Akt-mediated HIF regulation is unknown, but several studies show it to be mTOR dependent, where mTOR contributes to increased HIF protein synthesis (Hudson et al. 2002; Majumder et al. 2004). On the other hand, MAPK can directly phosphorylate HIF-a and increase its stabilization. In summary, HIF activity can be regulated at multiple levels: oxygen concentration, a-KG synthesis (substrates in hydroxylation reaction), and VHL mutation
3 Hypoxia, Gene Expression, and Metastasis
49
(impairment of HIF-a ubiquitination). One can envision additional ways to stabilize HIF, including mutations in prolines 402 and 564, proteasome inhibition, and treatment with iron chelators. Indeed, these are widely used as artificial means of HIF stabilization; and HIF mutations are rarely found in tumors.
HIF Target Genes Involved in the Metastatic Process Hypoxia promotes metastasis through the regulation of key genes involved in different steps of metastatic process, including angiogenesis, proliferation, migration, invasion, motility, adhesion, ECM remodeling, and survival (list of hypoxia-regulated genes can be found elsewhere) (Giaccia et al. 2003; Semenza 2003; Le et al. 2004; Maynard and Ohh 2007). It is important to note that approximately 1–1.5% of the genome is transcriptionally regulated by hypoxia, and many of these genes have been shown to be HIF regulated (Harris 2002; Semenza 2003, 2007). For some solid tumors, angiogenesis is a critical step in the metastatic process, whereas in other cancers (e.g. ovarian cancer and scirrhous adenocarcinoma), the vasculature is not involved in the metastatic spread. In these tumors, cells from the primary tumor shed to the peritoneal cavity, where they establish secondary lesions. In this case, the group of genes involved in angiogenesis is not relevant for the metastatic spread of these cancer types. The role of individual hypoxia-inducible genes in metastasis has just started to be investigated. This lack of direct in vivo evidence about their role in the metastatic process comes from the fact that many of hypoxia-regulated genes are affecting primary tumor growth, which complicates studies on their roles in metastatic spread. It is believed that the larger the primary tumor, the more cells from that tumor can colonize the secondary organs to form lesions. On the other hand, the establishment of metastasis from one cell requires extensive cell growth. Moreover many hypoxia-inducible genes are multifunctional, involved in several processes, like VEGF, which stimulates cancer cell migration (process not affecting tumor growth) and angiogenesis (affects tumor growth). Furthermore, HIFs activate a diverse array of target genes, which sometimes have opposite functions, like proapoptotic (transcriptional induction of BNIP3, BNIP3L, RTP801 and NIX, and stabilization of p53) and antiapoptotic (transcriptional induction of ADM, IGF2, IGF-BP1, 2, 3, TGF-a, VEGF, Bcl-w like, Pim1, 2). Thus, a shift in the balance of opposing signals as well as the mutation repertoire of a given cancer cell will affect tumor cell behavior under hypoxic conditions. An additional complexity comes from the observation that HIFs not only directly upregulate gene expression, by binding to target promoters, but also indirectly upregulate and downregulate gene expression through downstream signaling. The most significant example of indirect downregulation of gene expression is the E-cadherin (see below). The epithelial–mesenchymal transition (EMT) is one of the primary cellular mechanisms known to be involved in the metastatic process (Berx et al. 2007; Guarino et al. 2007; Tse and Kalluri 2007; Gavert and Ben-Ze’ev 2008). Epithelial
50
O.V. Razorenova and A.J. Giaccia
cells form cell–cell contacts through adherens junctions, whereas the mesenchymal phenotype is characterized by reduced cell–cell contacts, lost cell polarity, and increased interaction with extracellular matrix via integrins and focal adhesion kinase (FAK) (Avizienyte and Frame 2005; McLean et al. 2005). Epithelial cells, which undergo EMT, show a dedifferentiated morphology and become more motile and invasive. Epithelial cells can be reprogrammed to mesenchymal state by hypoxia and HIF that involve suppression of E-cadherin and increased cell motility (Cannito et al. 2008; Sahlgren et al. 2008; Yang et al. 2008). E-cadherin is a cellular adhesion molecule that regulates cell–cell adhesion (for review, see Hanahan and Weinberg 2000). E-cadherin inactivation enhances metastatic potential and its overexpression in cancer inhibits metastasis (Hanahan and Weinberg 2000). HIF has recently been described as a critical factor for the regulation of E-cadherin expression in ovarian carcinoma (Imai et al. 2003) and VHL-deficient CC-RCC (Esteban et al. 2006; Evans et al. 2007). The HIF-mediated upregulation of E-cadherin repressors Twist, Snail, Slug, and SIP1 serves as a mechanism of E-cadherin downregulation (Evans et al. 2007; Vesuna et al. 2008; Yoshida et al. 2009). Twist, a basic helix-loop-helix transcription factor, is a master regulator of gastrulation and mesoderm specification during embryonic development. HIF-1a was shown to regulate the expression of Twist through the HRE located in the Twist proximal promoter (Yang and Wu 2008; Yang et al. 2008). Co-expression of HIF1a, Twist, and Snail in primary tumors of head and neck squamous cell carcinoma patients correlates with the development of metastases and a poor prognosis (Yang et al. 2008). Importantly, the key role of Twist in metastasis was demonstrated by manipulation of Twist expression in highly metastatic mammary carcinoma cells. The downregulation of Twist expression in these cells led to the inhibition of their ability to metastasize from the mammary fat pad to the lung (Yang et al. 2004, 2006). In human breast cancers, high levels of Twist expression are correlated with invasive lobular carcinoma, a highly infiltrating tumor type associated with loss of E-cadherin expression (Yang et al. 2004). Lysyl oxidase (LOX), an amine oxidase involved in extracellular matrix formation, was shown to be regulated by hypoxia and HIFs (Erler et al. 2006; Erler and Giaccia 2006). Expression of LOX is associated with hypoxia in breast cancer as well as head and neck tumors, and correlates with poor patient outcome, including lower metastasis-free and overall survival rates (Erler et al. 2006; Le et al. 2009). Inhibition of LOX pharmacologically, genetically, or using a neutralizing antibody, in an orthotopic breast cancer model suppressed the metastatic potential of the tumor cells without affecting primary tumor growth (Erler et al. 2006). In vitro LOX regulated cell invasion and migration through regulating the activity of FAK. Moreover, LOX is critical for “premetastatic niche” formation in organs, where it relocates after being secreted by hypoxic primary tumors, crosslinks collagen IV in the basement membrane, and promotes CD11b+ myeloid cell recruitment (Erler et al. 2009). CD11b+ cells adhere to crosslinked collagen IV and produce matrix metalloproteinase-2, which cleaves collagen, enhancing the invasion and recruitment of bone marrow derived cells and metastasizing tumor cells.
3 Hypoxia, Gene Expression, and Metastasis
51
Hypoxia, Cancer Stem Cells, and Metastasis In the recent years, our view of cancer has been transformed by the “cancer stem cell” (CSC) theory (Reya et al. 2001; Pardal et al. 2003). Multiple studies have identified CSC subpopulations of cells within leukemias (Lapidot et al. 1994; Bonnet and Dick 1997), gliomas (Singh et al. 2003, 2004), breast (Al-Hajj et al. 2003), colon (Dalerba et al. 2007; O’Brien et al. 2007; Ricci-Vitiani et al. 2007), pancreatic (Li et al. 2007), bladder (Chan et al. 2009), melanoma (Boiko et al. 2010) cancers, head and neck squamous cell carcinoma (Prince et al. 2007), which have high self-renewal capacity, as well as more differentiated cancer cells (nonCSCs), which have very limited proliferative potential, making them incapable to drive tumor growth. The presence of CSC in tumors along with non-CSC makes tumor a complex hierarchically organized tissue. Moreover, it appears that many cancer therapies, while killing the bulk of tumor cells, fail to eliminate CSCs, which are resistant to chemo- and radiation treatment (Bao et al. 2006; Hambardzumyan et al. 2006; Li et al. 2008; Liu et al. 2006; Todaro et al. 2007; Wulf et al. 2001), allowing them to regenerate new tumors. Interestingly, many parallels can be drawn between CSCs and cancer cells residing in the hypoxic microenvironment. There is a striking correlation between chemo- and radiation resistance of CSCs and cancer cells in hypoxic regions. Increasing evidence suggests that hypoxia has the potential to inhibit tumor cell differentiation (Jogi et al. 2002; Helczynska et al. 2003), and thus plays a direct role in the maintenance of CSCs (Keith and Simon 2007; Hill et al. 2009). The role of hypoxia in general and HIFs in particular in maintaining the undifferentiated and pluripotent state of normal stem cells (SCs) is well documented (Cipolleschi et al. 1993, 2000; Lee et al. 2001; Ezashi et al. 2005; Parmar et al. 2007). Embryonic stem cells (ESCs) express the transcription factor Oct4 regulated by hypoxia during development (Covello et al. 2006). Moreover, the activity of Oct4, Sox2, and c-Myc used to reprogram fibroblasts to induced pluripotent stem (iPS) cell state (Takahashi and Yamanaka 2006; Takahashi et al. 2007; Wernig et al. 2007) is modulated by hypoxia. Oct4 is a HIF-2 target gene (Covello et al. 2006), Sox2 is hypoxia inducible in CD133+ neural CSCs and human ES cells (Forristal et al. 2009; McCord et al. 2009), and activity of c-Myc is induced by HIF-2 and inhibited by HIF-1. HIF-2 potentiates c-Myc activity by enhancing its physical association with Sp1, Miz1, and Max, although the precise mechanisms regulating these events are not yet fully understood (Gordan et al. 2007). HIF-1 antagonizes c-Myc activity by competing for binding to the transcription factor Sp1 under hypoxic conditions resulting in inhibition of c-Myc-dependent cell-cycle progression (Koshiji et al. 2004). Hypoxic regulation of Oct4, Sox2, and c-Myc simultaneously in one cell type needs to be investigated, opening the possibility that hypoxic exposure of fibroblasts or forced HIF-2 overexpression might lead to the iPS phenotype. In addition, Notch was shown to maintain a dedifferentiated state in multiple cell types under hypoxia (Gustafsson et al. 2005), where the formation of HIF-1a/Notch ICD domain complex enhanced Notch signaling (Sahlgren et al. 2008).
52
O.V. Razorenova and A.J. Giaccia
Interestingly, PTEN appears to be the negative regulator of SC pool (Groszer et al. 2001; Zheng et al. 2008), because deletion of PTEN results in an increase of neural stem cells (NSCs) numbers along with tumor expansion. Thus, the normal NSCs are finding a way to circumvent PTEN tumor suppressor function in order to become malignant. As discussed above, PTEN is a documented negative regulator of HIF, acting through the PI3K/Akt pathway (Zundel et al. 2000). In recent years, direct links between hypoxia, HIFs, and CSCs have emerged. Growing CD133+ neural CSCs at 7% oxygen compared to 20% reduced their doubling time and increased their self-renewal potential as reflected by clonogenicity (McCord et al. 2009). Growth of neural CSCs at 7% oxygen also resulted in an increase in the expression levels of the NSC markers CD133 and nestin as well as the SC markers Oct4 and Sox2. In addition, HIF-1 activity was not changed in CD133+ CSCs grown at 7%, and HIF-2 was expressed at higher levels (McCord et al. 2009). Another study showed that HIF-2a and multiple HIF-regulated genes are preferentially expressed in glioma stem cells (GSCs) in comparison to nonGSCs and normal neural progenitors (Li et al. 2009). Targeting HIFs in GSCs inhibited self-renewal, proliferation, and survival in vitro, and attenuated tumor initiation potential of GSCs in vivo. The third study showed that hypoxia (1% oxygen) promoted the self-renewal capacity and maintained the undifferentiated phenotype of CD133+ human glioma-derived CSCs, accompanied with expansion of cells bearing CXCR4 (CD184), CD44 (low), and A2B5 surface markers (Soeda et al. 2009). Knockdown of HIF-1a abrogated the hypoxia-mediated CD133-positive CSC expansion. On the other hand, mTOR signal and HIF-1 were shown to negatively regulate CD133 expression in cancer cell lines under 0.1% oxygen (which is close to anoxia) (Matsumoto et al. 2009). In support of the hypothesis that cells might be shifted into the iPS state by forced HIF expression, glioblastoma nonCSCs were unexpectedly found to possess plasticity under hypoxic conditions (2% oxygen), which helped to maintain self-renewal of glioblastoma CSCs and promoted reprogramming of glioblastoma non-CSCs toward a CSC phenotype (Heddleston et al. 2009). Thus, it looks like the oxygen concentration in the normal and CSC niche is very important and it has to be determined whether physiologic oxygen conditions or conditions of hypoxia/anoxia favor CSC and normal SC maintenance. In this case, the apparent discrepancy between the two reports about CD133 expression under hypoxia can be explained by differential effects on HIF-1 and 2 and by different oxygen levels. Because CSCs are capable of self-renewal, and are responsible for the long-term maintenance of tumor growth, it has been predicted that they might be also involved in metastases (Dalerba and Clarke 2007; Hermann et al. 2007). Thus, tumor initiating CD133+ pancreatic CSC population can be subdivided into CXCR4- (nonmetastatic) and CXCR4+ (metastatic) subpopulations. Bulk CD133+ cells (containing a mixture of CD133+/CXCR4+ and CD133+/CXCR4- cells) and individual CD133+/ CXCR4+ and CD133+/CXCR4- populations show similar tumor growth abilities. However, CD133+/CXCR4- tumors do not form spontaneous metastases during the time of experiment. A similar effect was obtained by pharmacological inhibition of
3 Hypoxia, Gene Expression, and Metastasis
53
CXCR4. CXCR4 is the receptor for the CXCL12/SDF-1 chemokine, which is involved in the control of leukocyte trafficking under physiological conditions. In cancer, CXCR4 was shown to regulate migration and metastasis (Muller et al. 2001; Balkwill 2004); notably, its ligand SDF-1 is highly expressed at common sites of metastasis, including the lung, bone marrow, and liver, and might direct CXCR4-expressing tumor cells to secondary metastatic sites (Kucia et al. 2005; Arya et al. 2007). An interesting twist in this story is that CXCR4 (Staller et al. 2003; Phillips et al. 2005; Zagzag et al. 2006) as well as its ligand SDF-1 (Ceradini et al. 2004) are hypoxia-inducible genes and HIF is a potent inducer of CXCR4 in VHL-deficient RCCs, nonsmall cell lung cancer, glioblastomas, and SDF-1 in endothelial cells (Staller et al. 2003; Ceradini et al. 2004; Phillips et al. 2005; Zagzag et al. 2006).
Conclusion The metastatic process is a complex multistep process requiring programming of cancer cells toward mesenchymal undifferentiated phenotype, with dominating promigratory and pro-survival signals. This phenotype is largely achieved through the tumor microenvironment, where cancer cells reside. The growing amount of evidence now shows that hypoxia creates a powerful niche, capable of shifting gene expression in cancer cells through stimulation of HIF activity toward pro-metastatic phenotype and protecting cells from chemo- and radiation treatment. Development of drugs, specifically targeting hypoxic cells, grows rapidly.
References Al-Hajj M, Wicha MS et al (2003) Prospective identification of tumorigenic breast cancer cells. Proc Natl Acad Sci USA 100(7):3983–3988. Arya M, Ahmed H et al (2007) Clinical importance and therapeutic implications of the pivotal CXCL12-CXCR4 (chemokine ligand-receptor) interaction in cancer cell migration. Tumour Biol 28(3):123–131. Avizienyte E, Frame MC (2005) Src and FAK signalling controls adhesion fate and the epithelialto-mesenchymal transition. Curr Opin Cell Biol 17(5):542–547. Balkwill F (2004) The significance of cancer cell expression of the chemokine receptor CXCR4. Semin Cancer Biol 14(3):171–179. Bao S, Wu Q et al (2006) Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 444(7120):756–760. Berra E, Pages G et al (2000) MAP kinases and hypoxia in the control of VEGF expression. Cancer Metastasis Rev 19(1–2):139–145. Berx G, Raspe E et al (2007) Pre-EMTing metastasis? Recapitulation of morphogenetic processes in cancer. Clin Exp Metastasis 24(8):587–597. Blancher C, Moore JW et al (2001) Effects of ras and von Hippel–Lindau (VHL) gene mutations on hypoxia-inducible factor (HIF)-1alpha, HIF-2alpha, and vascular endothelial growth factor
54
O.V. Razorenova and A.J. Giaccia
expression and their regulation by the phosphatidylinositol 3’-kinase/Akt signaling pathway. Cancer Res 61(19):7349–7355. Boiko AD, Razorenova OV et al (2010) Human melanoma-initiating cells express neural crest nerve growth factor receptor CD271. Nature 466:133–137. Bonnet D, Dick JE (1997) Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med 3(7):730–737. Brizel DM, Scully SP et al (1996) Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer Res 56(5):941–943. Brugarolas JB, Vazquez F et al (2003) TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell 4(2):147–158. Cannito S, Novo E et al (2008) Redox mechanisms switch on hypoxia-dependent epithelial– mesenchymal transition in cancer cells. Carcinogenesis 29(12):2267–2278. Carmeliet P (2005) VEGF as a key mediator of angiogenesis in cancer. Oncology 69(Suppl 3):4–10. Ceradini DJ, Kulkarni AR et al (2004) Progenitor cell trafficking is regulated by hypoxic gradients through HIF-1 induction of SDF-1. Nat Med 10(8):858–864. Chan DA, Giaccia AJ (2007) Hypoxia, gene expression, and metastasis. Cancer Metastasis Rev 26(2):333–339. Chan KS, Espinosa I et al (2009) Identification, molecular characterization, clinical prognosis, and therapeutic targeting of human bladder tumor-initiating cells. Proc Natl Acad Sci USA 106(33):14016–14021. Chen F, Kishida T et al (1995) Germline mutations in the von Hippel–Lindau disease tumor suppressor gene: correlations with phenotype. Hum Mutat 5(1):66–75. Cipolleschi MG, Dello Sbarba P et al (1993) The role of hypoxia in the maintenance of hematopoietic stem cells. Blood 82(7):2031–2037. Cipolleschi MG, Rovida E et al (2000) The expansion of murine bone marrow cells preincubated in hypoxia as an in vitro indicator of their marrow-repopulating ability. Leukemia 14(4):735–739. Comerford KM, Wallace TJ et al (2002) Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (MDR1) gene. Cancer Res 62(12):3387–3394. Covello KL, Kehler J et al (2006) HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth. Genes Dev 20(5):557–570. Dalerba P, Clarke MF (2007) Cancer stem cells and tumor metastasis: first steps into uncharted territory. Cell Stem Cell 1(3):241–242. Dalerba P, Dylla SJ et al (2007) Phenotypic characterization of human colorectal cancer stem cells. Proc Natl Acad Sci USA 104(24):10158–10163. Dewhirst MW, Tso CY et al (1989) Morphologic and hemodynamic comparison of tumor and healing normal tissue microvasculature. Int J Radiat Oncol Biol Phys 17(1):91–99. Endrich B, Reinhold HS et al (1979) Tissue perfusion inhomogeneity during early tumor growth in rats. J Natl Cancer Inst 62(2):387–395. Erler JT, Giaccia AJ (2006) Lysyl oxidase mediates hypoxic control of metastasis. Cancer Res 66(21):10238–10241. Erler JT, Bennewith KL et al (2006) Lysyl oxidase is essential for hypoxia-induced metastasis. Nature 440(7088):1222–1226. Erler JT, Bennewith KL et al (2009) Hypoxia-induced lysyl oxidase is a critical mediator of bone marrow cell recruitment to form the premetastatic niche. Cancer Cell 15(1):35–44. Esteban MA, Tran MG et al (2006) Regulation of E-cadherin expression by VHL and hypoxiainducible factor. Cancer Res 66(7):3567–3575. Evans AJ, Russell RC et al (2007) VHL promotes E2 box-dependent E-cadherin transcription by HIF-mediated regulation of SIP1 and snail. Mol Cell Biol 27(1):157–169. Ezashi T, Das P et al (2005) Low O2 tensions and the prevention of differentiation of hES cells. Proc Natl Acad Sci USA 102(13):4783–4788. Forristal CE, Wright KL et al (2009) Hypoxia inducible factors regulate pluripotency and proliferation in human embryonic stem cells cultured at reduced oxygen tensions. Reproduction 139(1):85–97.
3 Hypoxia, Gene Expression, and Metastasis
55
Fyles AW, Milosevic M et al (1998) Oxygenation predicts radiation response and survival in patients with cervix cancer. Radiother Oncol 48(2):149–156. Gardner LB, Corn PG (2008) Hypoxic regulation of mRNA expression. Cell Cycle 7(13): 1916–1924. Gavert N, Ben-Ze’ev A (2008) Epithelial–mesenchymal transition and the invasive potential of tumors. Trends Mol Med 14(5):199–209. Giaccia A, Siim BG et al (2003) HIF-1 as a target for drug development. Nat Rev Drug Discov 2(10):803–811. Gnarra JR, Glenn GM et al (1993) Molecular genetic studies of sporadic and familial renal cell carcinoma. Urol Clin North Am 20(2):207–216. Gordan JD, Bertout JA et al (2007) HIF-2alpha promotes hypoxic cell proliferation by enhancing c-myc transcriptional activity. Cancer Cell 11(4):335–347. Groszer M, Erickson R et al (2001) Negative regulation of neural stem/progenitor cell proliferation by the Pten tumor suppressor gene in vivo. Science 294(5549):2186–2189. Grunt TW, Lametschwandtner A et al (1985) The vascular pattern of basal cell tumors: light microscopy and scanning electron microscopic study on vascular corrosion casts. Microvasc Res 29(3):371–386. Guarino M, Rubino B et al (2007) The role of epithelial–mesenchymal transition in cancer pathology. Pathology 39(3):305–318. Gustafsson MV, Zheng X et al (2005) Hypoxia requires notch signaling to maintain the undifferentiated cell state. Dev Cell 9(5):617–628. Hall EJ (1994) Molecular biology in radiation therapy: the potential impact of recombinant technology on clinical practice. Int J Radiat Oncol Biol Phys 30(5):1019–1028. Hambardzumyan D, Squatrito M et al (2006) Radiation resistance and stem-like cells in brain tumors. Cancer Cell 10(6):454–456. Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100(1):57–70. Harris AL (2002) Hypoxia – a key regulatory factor in tumour growth. Nat Rev Cancer 2(1):38–47. Heddleston JM, Li Z et al (2009) The hypoxic microenvironment maintains glioblastoma stem cells and promotes reprogramming towards a cancer stem cell phenotype. Cell Cycle 8(20): 3274–3284, Helczynska K, Kronblad A et al (2003) Hypoxia promotes a dedifferentiated phenotype in ductal breast carcinoma in situ. Cancer Res 63(7):1441–1444. Hermann PC, Huber SL et al (2007) Distinct populations of cancer stem cells determine tumor growth and metastatic activity in human pancreatic cancer. Cell Stem Cell 1(3):313–323. Hill RP, Marie-Egyptienne DT et al (2009) Cancer stem cells, hypoxia and metastasis. Semin Radiat Oncol 19(2):106–111. Hockel M, Knoop C et al (1993) Intratumoral pO2 predicts survival in advanced cancer of the uterine cervix. Radiother Oncol 26(1):45–50. Hockel M, Schlenger K et al (1996) Association between tumor hypoxia and malignant progression in advanced cancer of the uterine cervix. Cancer Res 56(19):4509–4515. Hoffman MA, Ohh M et al (2001) von Hippel–Lindau protein mutants linked to type 2C VHL disease preserve the ability to downregulate HIF. Hum Mol Genet 10(10):1019–1027. Huang ST, Vo KC et al (2004) Developmental response to hypoxia. FASEB J 18(12):1348–1365. Hudson CC, Liu M et al (2002) Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Mol Cell Biol 22(20):7004–7014.Iliopoulos O, Levy AP et al (1996) Negative regulation of hypoxia-inducible genes by the von Hippel– Lindau protein. Proc Natl Acad Sci USA 93(20):10595–10599. Imai T, Horiuchi A et al (2003) Hypoxia attenuates the expression of E-cadherin via up-regulation of SNAIL in ovarian carcinoma cells. Am J Pathol 163(4):1437–1447. Ivan M, Kondo K et al (2001) HIFalpha targeted for VHL-mediated destruction by proline hydroxylation: implications for O2 sensing. Science 292(5516):464–468. Jaakkola P, Mole DR et al (2001) Targeting of HIF-alpha to the von Hippel–Lindau ubiquitylation complex by O2-regulated prolyl hydroxylation. Science 292(5516):468–472.
56
O.V. Razorenova and A.J. Giaccia
Jiang BH, Jiang G et al (2001) Phosphatidylinositol 3-kinase signaling controls levels of hypoxiainducible factor 1. Cell Growth Differ 12(7):363–369. Jogi A, Ora I et al (2002) Hypoxia alters gene expression in human neuroblastoma cells toward an immature and neural crest-like phenotype. Proc Natl Acad Sci USA 99(10):7021–7026. Kaelin WG Jr (2008) The von Hippel–Lindau tumour suppressor protein: O2 sensing and cancer. Nat Rev Cancer 8(11):865–873. Kallman RF, Dorie MJ (1986) Tumor oxygenation and reoxygenation during radiation therapy: their importance in predicting tumor response. Int J Radiat Oncol Biol Phys 12(4):681–685. Keith B, Simon MC (2007) Hypoxia-inducible factors, stem cells, and cancer. Cell 129(3):465–472. Kizaka-Kondoh S, Inoue M et al (2003) Tumor hypoxia: a target for selective cancer therapy. Cancer Sci 94(12):1021–1028. Koshiji M, Kageyama Y et al (2004) HIF-1alpha induces cell cycle arrest by functionally counteracting Myc. EMBO J 23(9):1949–1956. Kucia M, Reca R et al (2005) Trafficking of normal stem cells and metastasis of cancer stem cells involve similar mechanisms: pivotal role of the SDF-1-CXCR4 axis. Stem Cells 23(7):879–894. Lapidot T, Sirard C et al (1994) A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 367(6464):645–648. Latif F, Tory K et al (1993) Identification of the von Hippel–Lindau disease tumor suppressor gene. Science 260(5112):1317–1320. Le QT, Denko NC et al (2004) Hypoxic gene expression and metastasis. Cancer Metastasis Rev 23(3–4):293–310. Le QT, Harris J et al (2009) Validation of lysyl oxidase as a prognostic marker for metastasis and survival in head and neck squamous cell carcinoma: Radiation Therapy Oncology Group trial 90–03. J Clin Oncol 27(26):4281–4286. Lee YM, Jeong CH et al (2001) Determination of hypoxic region by hypoxia marker in developing mouse embryos in vivo: a possible signal for vessel development. Dev Dyn 220(2):175–186. Lee JW, Bae SH et al (2004) Hypoxia-inducible factor (HIF-1)alpha: its protein stability and biological functions. Exp Mol Med 36(1):1–12. Li C, Heidt DG et al (2007) Identification of pancreatic cancer stem cells. Cancer Res 67(3):1030–1037. Li L, Zhang L et al (2007) Hypoxia-inducible factor linked to differential kidney cancer risk seen with type 2A and type 2B VHL mutations. Mol Cell Biol 27(15):5381–5392. Li X, Lewis MT et al (2008) Intrinsic resistance of tumorigenic breast cancer cells to chemotherapy. J Natl Cancer Inst 100(9):672–679. Li Z, Bao S et al (2009) Hypoxia-inducible factors regulate tumorigenic capacity of glioma stem cells. Cancer Cell 15(6):501–513. Liu G, Yuan X et al (2006) Analysis of gene expression and chemoresistance of CD133+ cancer stem cells in glioblastoma. Mol Cancer 5:67. Mahon PC, Hirota K et al (2001) FIH-1: a novel protein that interacts with HIF-1alpha and VHL to mediate repression of HIF-1 transcriptional activity. Genes Dev 15(20):2675–2686. Majumder PK, Febbo PG et al (2004) mTOR inhibition reverses Akt-dependent prostate intraepithelial neoplasia through regulation of apoptotic and HIF-1-dependent pathways. Nat Med 10(6):594–601. Makino Y, Cao R et al (2001) Inhibitory PAS domain protein is a negative regulator of hypoxiainducible gene expression. Nature 414(6863):550–554. Matsumoto K, Arao T et al (2009) mTOR signal and hypoxia-inducible factor-1 alpha regulate CD133 expression in cancer cells. Cancer Res 69(18):7160–7164. Maxwell PH, Wiesener MS et al (1999) The tumour suppressor protein VHL targets hypoxiainducible factors for oxygen-dependent proteolysis. Nature 399(6733):271–275. Maynard MA, Ohh M (2004) Von Hippel–Lindau tumor suppressor protein and hypoxia-inducible factor in kidney cancer. Am J Nephrol 24(1):1–13. Maynard MA, Ohh M (2007) The role of hypoxia-inducible factors in cancer. Cell Mol Life Sci 64(16):2170–2180.
3 Hypoxia, Gene Expression, and Metastasis
57
McCord AM, Jamal M et al (2009) Physiologic oxygen concentration enhances the stem-like properties of CD133+ human glioblastoma cells in vitro. Mol Cancer Res 7(4):489–497. McLean GW, Carragher NO et al (2005) The role of focal-adhesion kinase in cancer – a new therapeutic opportunity. Nat Rev Cancer 5(7):505–515. Muller A, Homey B et al (2001) Involvement of chemokine receptors in breast cancer metastasis. Nature 410(6824):50–56. Nguyen DX, Bos PD et al (2009) Metastasis: from dissemination to organ-specific colonization. Nat Rev Cancer 9(4):274–284. Nordsmark M, Hoyer M et al (1996) The relationship between tumor oxygenation and cell proliferation in human soft tissue sarcomas. Int J Radiat Oncol Biol Phys 35(4):701–708. O’Brien CA, Pollett A et al (2007) A human colon cancer cell capable of initiating tumour growth in immunodeficient mice. Nature 445(7123):106–110. Papandreou I, Powell A et al (2005) Cellular reaction to hypoxia: sensing and responding to an adverse environment. Mutat Res 569(1–2):87–100. Pardal R, Clarke MF et al (2003) Applying the principles of stem-cell biology to cancer. Nat Rev Cancer 3(12):895–902. Parmar K, Mauch P et al (2007) Distribution of hematopoietic stem cells in the bone marrow according to regional hypoxia. Proc Natl Acad Sci USA 104(13):5431–5436. Pause A, Peterson B et al (1999) Studying interactions of four proteins in the yeast two-hybrid system: structural resemblance of the pVHL/elongin BC/hCUL-2 complex with the ubiquitin ligase complex SKP1/cullin/F-box protein. Proc Natl Acad Sci USA 96(17): 9533–9538. Phillips RJ, Mestas J et al (2005) Epidermal growth factor and hypoxia-induced expression of CXC chemokine receptor 4 on non-small cell lung cancer cells is regulated by the phosphatidylinositol 3-kinase/PTEN/AKT/mammalian target of rapamycin signaling pathway and activation of hypoxia inducible factor-1alpha. J Biol Chem 280(23):22473–22481. Prince ME, Sivanandan R et al (2007) Identification of a subpopulation of cells with cancer stem cell properties in head and neck squamous cell carcinoma. Proc Natl Acad Sci USA 104(3): 973–978. Qing G, Simon MC (2009) Hypoxia inducible factor-2alpha: a critical mediator of aggressive tumor phenotypes. Curr Opin Genet Dev 19(1):60–66. Rankin EB, Giaccia AJ (2008) The role of hypoxia-inducible factors in tumorigenesis. Cell Death Differ 15(4):678–685. Reya T, Morrison SJ et al (2001) Stem cells, cancer, and cancer stem cells. Nature 414(6859):105–111. Ricci-Vitiani L, Lombardi DG et al (2007) Identification and expansion of human colon-cancerinitiating cells. Nature 445(7123):111–115. Sahlgren C, Gustafsson MV et al (2008) Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc Natl Acad Sci USA 105(17):6392–6397. Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3(10):721–732. Semenza GL (2007) Hypoxia and cancer. Cancer Metastasis Rev 26(2):223–224. Shah-Yukich AA, Nelson AC (1988) Characterization of solid tumor microvasculature: a threedimensional analysis using the polymer casting technique. Lab Invest 58(2):236–244. Shannon AM, Bouchier-Hayes DJ et al (2003) Tumour hypoxia, chemotherapeutic resistance and hypoxia-related therapies. Cancer Treat Rev 29(4):297–307. Singh SK, Clarke ID et al (2003) Identification of a cancer stem cell in human brain tumors. Cancer Res 63(18):5821–5828. Singh SK, Hawkins C et al (2004) Identification of human brain tumour initiating cells. Nature 432(7015):396–401. Sodhi A, Montaner S et al (2001) MAPK and Akt act cooperatively but independently on hypoxia inducible factor-1alpha in rasV12 upregulation of VEGF. Biochem Biophys Res Commun 287(1):292–300. Soeda A, Park M et al (2009) Hypoxia promotes expansion of the CD133-positive glioma stem cells through activation of HIF-1alpha. Oncogene 28(45):3949–3959.
58
O.V. Razorenova and A.J. Giaccia
Sprenger SH, Gijtenbeek JM et al (2001) Characteristic chromosomal aberrations in sporadic cerebellar hemangioblastomas revealed by comparative genomic hybridization. J Neurooncol 52(3):241–247. Staller P, Sulitkova J et al (2003) Chemokine receptor CXCR4 downregulated by von Hippel– Lindau tumour suppressor pVHL. Nature 425(6955):307–311. Takahashi K, Yamanaka S (2006) Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126(4):663–676. Takahashi K, Tanabe K et al (2007) Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131(5):861–872. Todaro M, Alea MP et al (2007) Colon cancer stem cells dictate tumor growth and resist cell death by production of interleukin-4. Cell Stem Cell 1(4):389–402. Tse JC, Kalluri R (2007) Mechanisms of metastasis: epithelial-to-mesenchymal transition and contribution of tumor microenvironment. J Cell Biochem 101(4):816–829. Vaupel P, Mayer A (2007) Hypoxia in cancer: significance and impact on clinical outcome. Cancer Metastasis Rev 26(2):225–239. Vesuna F, van Diest P et al (2008) Twist is a transcriptional repressor of E-cadherin gene expression in breast cancer. Biochem Biophys Res Commun 367(2):235–241. Wartenberg M, Ling FC et al (2003) Regulation of the multidrug resistance transporter P-glycoprotein in multicellular tumor spheroids by hypoxia-inducible factor (HIF-1) and reactive oxygen species. FASEB J 17(3):503–505. Wernig M, Meissner A et al (2007) In vitro reprogramming of fibroblasts into a pluripotent ES-cell-like state. Nature 448(7151):318–324. Wiesener MS, Jurgensen JS et al (2003) Widespread hypoxia-inducible expression of HIF-2alpha in distinct cell populations of different organs. FASEB J 17(2):271–273. Wulf GG, Wang RY et al (2001) A leukemic stem cell with intrinsic drug efflux capacity in acute myeloid leukemia. Blood 98(4):1166–1173. Yang J, Mani SA et al (2004) Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 117(7):927–939. Yang J, Mani SA et al (2006) Exploring a new twist on tumor metastasis. Cancer Res 66(9):4549–4552. Yang MH, Wu KJ (2008) TWIST activation by hypoxia inducible factor-1 (HIF-1): implications in metastasis and development. Cell Cycle 7(14):2090–2096. Yang MH, Wu MZ et al (2008) Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat Cell Biol 10(3):295–305. Yoshida J, Horiuchi A et al (2009) Changes in the expression of E-cadherin repressors, Snail, Slug, SIP1, and Twist, in the development and progression of ovarian carcinoma: the important role of Snail in ovarian tumorigenesis and progression. Med Mol Morphol 42(2):82–91. Zagzag D, Lukyanov Y et al (2006) Hypoxia-inducible factor 1 and VEGF upregulate CXCR4 in glioblastoma: implications for angiogenesis and glioma cell invasion. Lab Invest 86(12): 1221–1232. Zbar B, Kishida T et al (1996) Germline mutations in the Von Hippel–Lindau disease (VHL) gene in families from North America, Europe, and Japan. Hum Mutat 8(4):348–357. Zelzer E, Levy Y et al (1998) Insulin induces transcription of target genes through the hypoxiainducible factor HIF-1alpha/ARNT. EMBO J 17(17):5085–5094. Zhao S, Lin Y et al (2009) Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science 324(5924):261–265. Zheng H, Ying H et al (2008) p53 and Pten control neural and glioma stem/progenitor cell renewal and differentiation. Nature 455(7216):1129–1133. Zundel W, Schindler C et al (2000) Loss of PTEN facilitates HIF-1-mediated gene expression. Genes Dev 14(4):391–396.
Chapter 4
Molecular Mechanisms Regulating Expression and Function of Cancer-Associated Carbonic Anhydrase IX Jaromir Pastorek and Silvia Pastorekova
Abstract Hypoxia and acidosis are typical physiological factors of tumor microenvironment acting in concert to support invasive and metastatic propensity of tumor cells through rearrangement of their gene expression profile and remodeling of their phenotype. Induction of carbonic anhydrase IX (CA IX) has recently been recognized as one of the important molecular events observed in tumor cells exposed to hypoxia. CA IX can serve as a “tag” of hypoxic areas in many tumor types and thereby provides a diagnostic tool as well as a target for selective delivery of anticancer immunotherapy. Moreover, CA IX actively contributes to pH regulation which protects tumor cells from acidosis and thereby provides a target for anticancer therapy based on reduction of tumor cell survival via inhibition of CA IX enzyme activity. Both aspects attract considerable attention but their full appraisal and further development require deeper understanding of the mechanisms behind CA IX expression and functioning. Here, we summarize and discuss the “state of the art” of CA IX field in context of its relevance for tumor microenvironment.
Carbonic Anhydrases Acid–base balance mediated by equilibrium between carbon dioxide and carbonic acid is essential for proper functioning of metabolically active cells, tissues, and organs in virtually all living organisms. Although reversible conversion of carbon dioxide to carbonic acid followed by dissociation of carbonic acid to bicarbonate ion and a proton can proceed spontaneously, turnover of this reaction is not fast enough to satisfactorily supply physiological processes and therefore, it needs to be catalyzed by zinc-binding enzymes traditionally called carbonic anhydrases (CAs).
J. Pastorek (*) Department of Molecular Medicine, Institute of Virology, Slovak Academy of Sciences, Dubravska cesta 9, 845 05 Bratislava, Slovak Republic e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_4, © Springer Science+Business Media, LLC 2010
59
60
J. Pastorek and S. Pastorekova
CAs are present all over the kingdoms of animals, plants, fungi, and bacteria (including archeabacteria), and are encoded by five genetically unrelated families (a–e), each containing diverse isoforms (Supuran 2008). Human CAs belong to a-CA family, which is characterized by a deep and wide active site pocket that contains three histidine residues coordinating catalytic zinc at the bottom of the active site cleft and the fourth histidine residue functioning as a proton shuttle between the active site and solution shifting between inward and outward orientation (Zheng et al. 2008). Twelve out of fifteen human a-CA isoforms are catalytically active (CA I–IV, VA, VB, VI, VII, IX, XII–XIV). In addition to histidine residues, they contain also other conserved amino acids involved in formation of inner and outer layers of the active site cavity. The CA isoezymes differ by their catalytic activity that ranges from extremely high (in case of CA II and CA IX that belong to most efficient enzymes in general) through high/intermediate (CA IV, V, VI, VII, XII, XIV) to very low (CA I, III), as shown in Fig. 4.1. Three acatalytic isoforms (CA VIII, X, XI) lack one of three critical zinc-coordinating histidines. Moreover, there are two CA-related proteins that function as receptor protein tyrosine phosphatases (RPTPb/g). Their incompletely conserved, catalytically silent CA domain pocket serves as a ligand-binding site for neuronal adhesion molecule contactin (Peles et al. 1995).
Fig. 4.1 Schematic illustration of the human CA isoforms, their domain composition, enzyme activity, and subcellular localization. Cytoplasmic and mitochondrial CAs consist only of the CA domain, secreted CA contains a short C-terminal extension, membrane-associated CAs in addition have a transmembrane anchor and, except for CA IV, also the cytoplasmic tail. CA IX is the only isoform containing the N-terminal PG-like domain. Structure of the CA IX isoforms is depicted in more detail on the left side. It includes signal peptide (SP), proteoglycan-like segment (PG), carbonic anhydrase domain (CA), transmembrane anchor (TM), and intracytoplasmic tail (IC). Conserved zinc-binding histidines are indicated by black circles, cysteines are shown as white squares, and intramolecular disulfidic bond is drawn as dotted line. Numbers designate amino acids encompassed within the individual domains
4 Molecular Mechanisms Regulating Expression
61
Evolutionarily oldest human CA isoenzymes contain an extracellular catalytic domain and are either transmembrane (CA IX, XII, XIV), or membrane-associated via GPI anchor (CA IV) or secreted (CA VI). The transmembrane CAs also contain a hydrophobic region and a relatively conserved intracellular tail. In addition, CA IX possesses an N-terminal proteoglycan-like region. The younger CAs (I, II, III, V, VII) are intracellular isoenzymes containing a sole CA domain. Except CA VA and VB that reside in the mitochondrion, all intracellular CAs are located in the cytoplasm. Finally, the evolutionarily youngest cluster includes inactive cytoplasmic isoforms mentioned above. CAs are spread throughout the human body. Some of them are ubiquitous (CA II) or broadly distributed (CA IV, XII, XIII), others are confined to only few tissues/cell types such as muscle and adipocytes (CA III), liver (CA VA), and saliva and milk (CA VI). Cells/organs of the most abundant CA expression include the red blood cells, brain, lung, gut, liver, kidney, and testis, where various concurrently present CA isoforms cooperatively contribute to breathing, synaptic transmission, CSF secretion, production of ocular fluid, bone resorption, secretion of gastric acid, urine acidification, sperm maturation, etc. In metabolically active tissues such as kidney or stomach epithelia, the active CAs located in various subcellular compartments or on the opposite sites of the plasma membrane can interact with bicarbonate transporters or biosynthetic enzymes to form spatially and functionally coordinated protein complexes called metabolons (McMurtrie et al. 2004). In bicarbonate transport metabolon, CAs can improve the transmembrane movement of bicarbonate ion, which is a membrane-impermeable molecule transported by integral membrane bicarbonate transporters and exchangers (Casey 2006). The role of CAs is to increase the local concentration of bicarbonate in order to accelerate its flux across the membrane and thereby regulate intracellular and extracellular pH (McMurtrie et al. 2004). CAs are essential components of physiological processes and disbalance in their expression is associated with several diseases. These include on the one hand osteopetrosis and mental retardation due to CA II deficiency and, on the other hand, glaucoma, diabetic retinopathy, epilepsy, obstructive pulmonary disease, obesity and various neurological diseases linked with abnormally high expression of CA II, III, XII, and other isoforms. Some of these diseases are successfully treated with inhibitors of CA activity used as primary or supplementary therapy, for the others, new derivatives of CA inhibitors are under development as promising drugs (Pastorekova et al. 2004b). Although importance of CAs in cancer has been demonstrated relatively recently, inspections into their expression in tumor cells and tissues have been initiated a long time ago. Several independent studies revealed reduced activity and/or decreased levels of cytosolic isoforms CA I, II, III, XIII in colorectal, hepatocellular, and lung tumors (when compared to their normal counterparts) indicating that these isoforms are predominantly needed for differentiated cells/tissues (Chiang et al. 2002; Kivela et al. 2001; Kummola et al. 2005; Kuo et al. 2003; Mori et al. 1993; Yoshiura et al. 2005). This conclusion has, however, few exceptions, including CA II expression in endothelium of malignant melanomas and meningiomas and a recently discovered
62
J. Pastorek and S. Pastorekova
extremely high expression of CA II in gastrointestinal stromal tumors (Haapasalo et al. 2007; Yoshiura et al. 2005; Parkkila et al. 2010). Relationship of the mitochondrial CA VA and VB isoforms to cancer remains unexplored, although they might be relevant for biosynthesis of metabolic intermediates in proliferating tumor cells. Finally, out of four CA isoforms localized at the cell surface, only CA IX and XII showed consistent link with tumor tissues and are systematically investigated as promising targets for anticancer therapy (Pastorekova et al. 2008). Especially CA IX, which is almost exclusively expressed in tumors, attracts considerable interest with expectation that unraveling its regulation and functional contribution to tumor phenotype will provide clues for development of new anticancer strategies.
Molecular Features of CA IX CA IX (initially named MN antigen) was identified in HeLa cervical carcinoma cells as a density-induced cell surface protein associated with tumor phenotype of human cell lines and tissue specimens (Pastorekova et al. 1992; Zavada et al. 1993). Primary structure of the MN antigen was derived from cDNA and genomic sequences and was found to contain a well-conserved carbonic anhydrase domain located in the large extracellular part of the molecule (Pastorek et al. 1994) (Fig. 4.1). This CA domain shows a significant identity to secreted isozyme CA VI (40.8%) and cytosolic isozyme CA II (35.8%), respectively, and possesses all important amino acids required for the catalytic activity. It is N-glycosylated by 4 kDa high mannose sugar chain attached to N346 and is catalytically active. At the time of MN protein discovery, this was the ninth mammalian CA identified and thus it was renamed to CA IX (Hewett-Emmett and Tashian 1996). In addition to catalytic domain, CA IX (as the only CA isoform) contains an N-terminal PG-like region homologous to keratan sulfate attachment domain of a large proteoglycan aggreccan (Opavsky et al. 1996). According to mass spectrometric analysis, this part of the molecule appears to contain a keratan sulfate glycosaminoglycan chain of about 3.5 kDa that is presumably attached to T115, in the article numbered T78 (without signal peptide) (Hilvo et al. 2008). Presence of such modification might have important biological implications analogous to keratan sulfate chain of CD44 metastasis-related protein, which mediates interactions with constituents of extracellular matrix. However, experimental proof of this concept has not been given for CA IX so far. The PG-like region has also another important feature – it predominantly contains negatively charged acidic amino acids that are mostly arranged within four identical and three imperfect GEEDLP repeats. This results in remarkably acidic character of the PG-like region that seems to have a significant impact on the catalytic activity of CA IX as an intrinsic buffer optimizing CO2 hydration at acidic pH values characteristic for solid tumors (Innocenti et al. 2009). Thus, activity of the entire extracellular portion of CA IX (including PG and CA domains) expressed in baculovirus system exceeds that of CA II
4 Molecular Mechanisms Regulating Expression
63
(the best known catalyst) and reaches maximum at pH values around 6.5 in contrast to other CA isoforms which have optimum catalytic efficiency around pH 7. C-terminal part of the catalytic domain is extended with a little juxtamembrane fragment, followed by a hydrophobic transmembrane (TM) region and a short intracellular tail (IC). The IC tail is well conserved among the transmembrane CA isoforms (CA IX, XII, XIV) and its removal leads to loss of cell surface location of CA IX despite preserved TM region suggesting that IC mediates protein–protein interactions that maintain CA IX at its proper position (Hulikova et al. 2009). In vitro mutagenesis of the submembrane cluster of basic amino acids revealed that IC tail participates in inside-out signaling that affects functions of the extracellular CA IX portion (Hulikova et al. 2009). In addition, the IC tail contains three potential phosphorylation sites (T443, S448, andY449). While Y449 participates in EGFRinduced signal transduction to Akt kinase (Dorai et al. 2005), phosphorylated T443 is critical for modulation of the pH regulatory function of CA IX (manuscript in preparation). Molecular weight of CA IX analyzed under reducing conditions and visualized on Western blot corresponds to 58/54 kDa. Existence of two forms of the protein remains unexplained and the data attributing the double band to different parts of CA IX molecule are still controversial. Interestingly, in non-reducing conditions, CA IX is capable to form oligomers that were initially considered for trimers (or tetramers). This supposition was based on the molecular weight of about 153 kDa and on other indices including the capability of CA IX to fit into envelope of the vesicular stomatitis virus and form the virus pseudotypes (Pastorekova et al. 1992; Zavada et al. 1993). The latter aspect was shown to require trimerization or tetramerization (Kreis and Lodish 1986). In fact, CA IX molecule contains four cysteine residues, three of them located in the catalytic domain and one in the region just above the plasma membrane. While C156 and C336 are placed at the positions analogous to intra-molecular S–S bond-forming cysteines in other CAs, the additional two residues C174 and C409 might be theoretically available for intermolecular S–S bonds. Biochemical and crystallographic studies confirmed the proposed intramolecular S–S bond, but they also suggested that CA domain of CA IX forms dimer stabilized by homologous S–S bridge between C174 corresponding to C41 according to CA I numbering (Alterio et al. 2009). Although results of the biochemical and crystallographic analyses do not support the concept of CA IX trimerization, it is still possible that the discrepancy results from the absence of the regions flanking the catalytic domain, as it was observed also in other cases, such as with Fas receptor or TRAF6 (Scott et al. 2009; Yin et al. 2009). Crystallographic analysis brought very important structural data on the CA IX catalytic domain that is similar to other a-CAs (Alterio et al. 2009). It has a compact globular appearance with a large conical cavity containing zinc ion at the bottom. The entrance to the active site is occupied by three arginine residues R95, R97, and R167, which are not conserved in other CA isoforms and have been proposed to mediate crosstalk of CA domain with acidic PG-like region presumably implicated in steric control of the substrate accession or the proton-transfer reaction. The active site and the N-terminus of the catalytic domain face the extracellular space, whereas
64
J. Pastorek and S. Pastorekova
its C-terminus is oriented toward the plasma membrane. This arrangement is fully compatible with the role of CA IX in extracellular acidification of tumor microenvironment.
CA IX Tissue Distribution CA IX protein expression has been studied in a broad range of human tissues. Most of these studies were performed using the monoclonal antibody M75 that recognizes a repetitive linear epitope in the PG-like domain of both native and denatured CA IX (Zavada et al. 2000) and is therefore suitable for virtually all immunodetection methods including immunohistochemistry on both paraffinembedded and frozen sections (even those archived for years), immunoblotting, immunoprecipitation, flow cytometry, etc. Additional antibodies produced by our team (Zat’ovicova et al. 2003) are mostly directed to CA domain and bind to conformational epitopes (thus they do not work in immunoblotting) similarly to well-known monoclonal antibody G250 that was generated even before the discovery of CA IX against a renal cell carcinoma antigen later shown to be identical to CA IX (Oosterwijk 2008a, b). New human monoclonal antibodies described recently display excellent CA IX binding properties (Ahlskog et al. 2009b). Some of these antibodies have a great potential as diagnostic and therapeutic tools. Noteworthy, a commercial antibody NB100 was advertised to give the same staining pattern as M75, but later demonstration of its significant non-specific reactivity against tubulin considerably reduces its value for detection of CA IX (Li et al. 2009). Overall expression pattern of CA IX is characterized by limited expression in normal tissues and broad distribution in different tumors. Physiological expression of CA IX is linked with epithelial tissues of the gastrointestinal tract with the highest CA IX level in all principal cell types of the glandular gastric mucosa and in the gallbladder epithelium. CA IX is also present in the small intestine and proximal colon, but its level decreases toward rectum (Pastorekova et al. 1997). Intestinal CA IX is expressed in the cryptal epithelial cells that possess the greatest proliferative activity (Saarnio et al. 1998). CA IX is also abundant in the gallbladder mucosa, while the pancreatic ducts show only weak expression. In all gastrointestinal epithelia, CA IX is present in the basolateral membranes, suggesting its possible involvement in intercellular communication, maintenance of tissue integrity, and regulation of basolateral ion transport (Pastorekova et al. 1997). Based on gastric phenotype of CA IX-deficient mice that includes hyperplasia, numerous cysts, and abnormal lineage development but no change in gastric pH, CA IX plays a role in gastric morphogenesis and control of cell proliferation and/or differentiation (Gut et al. 2002). CA IX is also expressed in some other healthy tissues, albeit at much lower levels compared to alimentary tract. These include linings of the body cavity, reproductive tract, ventricular linings of CNS, and choroid plexus (Ivanov et al. 2001). The other normal human tissues are CA IX negative.
4 Molecular Mechanisms Regulating Expression
65
In contrast, CA IX expression is readily detected at significant levels in diverse tumors, including carcinomas of the uterine cervix, kidney, brain, head and neck, esophagus and lung, colon, breast, ovaria, endometrium, vulva, bladder, etc. (Bartosova et al. 2002; Haapasalo et al. 2006; Ivanov et al. 2001; Jarvela et al. 2008; Kowalewska et al. 2005; Liao et al. 1994, 1997; McKiernan et al. 1997; Niemela et al. 2007; Saarnio et al. 1998; Turner et al. 1997; Vermylen et al. 1999). From an unknown reason, CA IX is only occasionally expressed in the prostate carcinomas (Smyth et al. 2009).
Regulation of CA IX Expression Presence of CA IX in cancerous tissues of different origin suggests that it represents a phenomenon common to all solid tumors. In accord with this supposition, it has been proven that transcription of CA IX is principally driven by hypoxia that occurs in variable extent in virtually all tumor tissues. Hypoxia often develops in growing tumors due to insufficient supply of oxygen by irregular and functionally inadequate vasculature. This creates a selective pressure favoring those tumor cells that are capable to induce adaptative responses including shift to anaerobic glycolysis, angiogenesis, reduced proliferation and cell adhesion, increased pH regulation, etc. Consequently, hypoxic cells gain oncogenic metabolism, resistance to conventional anticancer treatment, and more aggressive phenotype with metastatic potential (Harris 2002). At the molecular level, hypoxia leads to reprogramming of the cell’s transcriptional profile, which is principally governed by a master transcription factor HIF, i.e. hypoxia inducible factor (Coleman and Ratcliffe 2007), illustrated in Fig. 4.2. HIF is a heterodimer consisting of an oxygen-regulated a subunit (existing in three isoforms) and a constitutive b subunit. In normoxia, HIF-a is modified by oxygendependent prolyl hydroxylases (PHDs) and an asparaginyl hydroxylase (factor inhibiting HIF, named FIH), enzymes that belong to the Fe(II) and 2-oxoglutarate dioxygenase superfamily. FIH hydroxylates N803 in the C-terminal activation domain of human HIF-1a and thereby prevents its interaction with transcriptional coactivators. Prolyl hydroxylases hydoxylate P564 and P402 in the oxygen-dependent degradation domain, thus leading to HIF-a recognition by pVHL tumor suppressor protein, followed by its rapid ubiquitilation and proteasome degradation (Hirota and Semenza 2005). In hypoxia and/or following oncogenic activation or loss of VHL function, HIF-a escapes modifications by PHDs and FIH, enters the nucleus, dimerizes with the constitutive b subunit of HIF, interacts with coactivators, and forms an active transcriptional complex (Ruas and Poellinger 2005). This complex binds DNA at sites containing the hypoxia-response element (HRE; 5¢-RCGTG-3¢), which is present in the promoters of a wide spectrum of target genes encoding functionally diverse proteins involved in angiogenesis (VEGF, VEGF receptor, transforming growth factor TGF-b), vascular tonus (nitric oxide synthase), erythropoiesis (erythropoietin-1),
66
J. Pastorek and S. Pastorekova
Fig. 4.2 Hypoxia-induced HIF-1-mediated transcriptional activation of CA9 gene. The left side shows the situation in normoxia. When oxygen supply is sufficient, prolyl hydroxylases (PHD) and factor inhibiting HIF (FIH) modify a-subunit of hypoxia-inducible factor 1 (HIF-a). Hydroxylation of conserved prolines in HIF-a molecule leads to its recognition by von Hippel– Lindau tumor suppressor protein (VHL), which triggers HIF-a ubiquitilation and degradation. Hydroxylaion of N-terminal asparagine blocks binding of co-activators. This principally results in the absence of active HIF-a from normoxic cells. The right part shows the situation in hypoxia. Lack of oxygen keeps hydroxylases inactive, HIF-a is not modified and remains protected from recognition by VHL. This results in stabilization and accumulation of HIF-a, translocation to nucleus, dimerization with HIF-b, interaction with co-activators, and formation of active transcription complex, which induces expression of target genes, including CA9. This induction can be modified by additional pathways not shown in the figure and leads to exposure of CA IX protein on the surface of hypoxic cells
glycolysis (glucose transporters GLUT-1,3, lactate dehydrogenase LDH-A, pyruvate dehydrogenase kinase PDK1), ion transport and pH regulation (natrium-proton exchanger NHE1, monocarboxylate transporter MCT4), remodeling of extracellular matrix (matrix metalloproteinase MMP-2, cathepsin D), adhesion and migration (hepatocyte growth factor receptor c-Met, transcription factors Snail/Slug), cell proliferation and survival (insulin growth factor-binding protein), drug resistance (P glycoprotein), and additional cellular phenomena linked with tolerance of tumor cells to hypoxia. Genomic position of HRE differs among the transcriptional targets of HIF, in some cases representing enhancer rather than promoter element (Camenisch et al. 2001). In case of the CA9 gene, HRE is an important component of the core promoter and its position adjacent to transcription initiation site (–3/–10) makes it a key regulatory element principally determining CA9 gene expression (Wykoff et al. 2000). This HRE element binds HIF-1 (rather than HIF-2) and cooperates with a neighboring consensus sequence binding SP1 transcription factor (Kaluz et al. 2002). The HRE-SP1 module is critical for transcriptional activation of the CA9 gene both in hypoxia and in high cell density (creating pericellular hypoxia) and its mutagenesis leads to reduction or complete loss of the promoter activity (Kopacek et al. 2005). Additional regulatory elements present within the CA9 promoter, including AP-1 binding site and position/orientation-independent
4 Molecular Mechanisms Regulating Expression
67
silencer, play secondary roles as they just modulate the outcome of the transcriptional activation triggered by HIF-1-SP1 (Kaluz et al. 1999). As recently discussed, CA9 transcription profile primarily reflects the transcriptional activity (not the level) of HIF-1 (Kaluz et al. 2009). In accord, different pathways that affect HIF-1 can also increase the expression of CA9 gene, such as MAPK and PI3 kinase pathways, Notch pathway, HBx oncoprotein of hepatitis B virus, and v-Src and c-Src oncoproteins (Holotnakova et al. 2010; Kopacek et al. 2005; Sansone et al. 2007; Takacova et al. 2010). Moreover, hypoxia-induced CA9 gene expression requires an intact unfolded protein response (UPR) pathway, which operates via ATF4 transcription factor that binds to CA9 promoter (van den Beucken et al. 2009a, b). Under normoxic conditions, expression of CA9 gene is down-regulated by pVHL tumor suppressor protein, which negatively regulates HIF-1a. However, CA9 expression is not suppressed by the mutated variants of pVHL lacking the elongin binding domain which is required for the interaction of pVHL with elongins B/C and integration of pVHL within a ubiquitin ligase complex (Ivanov et al. 1998). This explains why CA IX protein is overexpressed in a high percentage of renal cell carcinomas (RCC) that frequently carry an inactivating mutation in the VHL gene (Mandriota et al. 2002). Extracellular acidosis, which frequently occurs in tumor microenvironment, contributes to CA9 transcription at least in some tumor cell types, such as those derived from glioblastoma (Ihnatko et al. 2006). CA9 induction by acidosis can be observed in both absence and presence of hypoxia. Promoter analysis revealed involvement of SP1 and HRE promoter elements and showed that acidosis can modulate CA9 expression only when basic CA9 activation through MAPK and/or PI3K pathways is preserved. However, inconsistent effects of acidosis on CA9 expression were obtained in other cell types possibly due to their differential sensitivity to acidosis-associated stress (Willam et al. 2006). Finally, transcription of CA9 gene can be negatively controlled by methylation of CA9 promoter and by treatment with xenobiotics (Ashida et al. 2002; Cho et al. 2000; Jakubickova et al. 2005; Takacova et al. 2009). CA9 is also a subject of post-transcriptional regulation via alternative splicing that produces two types of transcripts – a full-length transcript encoding a plasma membrane-localized and functionally competent CA IX protein, and a truncated mRNA lacking exons 8/9 and producing a cytoplasmic/secreted form of CA IX protein with decreased enzyme activity (Barathova et al. 2008). The longer transcript is regulated by hypoxia and linked to tumor phenotype, whereas the shorter one is expressed at low level independently of hypoxia and is therefore detectable also under normoxia and in the normal tissues. This is very important for the correct design of primers used in the clinical studies of cancer- and/or hypoxia-related expression of CA9, where the alternatively spliced variant can provide a falsepositive signal. As recently demonstrated in non-small cell lung carcinoma, fulllength CA9 transcript is the most accurate surrogate of hypoxic stress and represents the only variant with a prognostic role (Malentacchi et al. 2009). These data indicate the importance of a separate measurement of the two isoforms in cancer and
68
J. Pastorek and S. Pastorekova
the need of an accurate re-evaluation of most studies on the clinical role of CA9 mRNA in cancer diagnosis. At the posttranslational level, CA IX protein shows high stability, with the halflife of about 38 h (as determined in reoxygenated HeLa cells), and therefore represents a marker of both present and past hypoxia (Rafajova et al. 2004). The final amount of CA IX protein is regulated by shedding of its ectodomain (ECD) to culture medium and to body fluids of tumor patients (Zatovicova et al. 2005; Zavada et al. 2003). Constitutive shedding of CA IX ECD releases approximately 10% of the cell-associated protein and is sensitive to metalloproteinase-inhibitor batimastat. Hypoxia maintains the normal rate of the basal ECD release, thus leading to a concomitant increase in cell-associated and extracellular levels of CA IX. The CA IX ECD shedding can be several fold induced by treatment with phorbol12-myristate-13-acetate and pervanadate. This activated shedding is mediated by TNF-a-converting enzyme (TACE/ADAM17). Thus, cleavage of CA IX ectodomain appears to be a regulated process that responds to signal transduction-related stimuli and may contribute to adaptive changes in the protein composition of tumor cells and of their microenvironment. In this context, it is noteworthy that the extracellular domain of CA IX shed to body fluids is a promising circulating cancer marker for patient selection, and monitoring of treatment outcome (Carney 2007). Finally, posttranslational modifications, such as phosphorylation, regulate the CA IX-mediated signal transduction (Dorai et al. 2005) as well as the functional status of CA IX (the latter aspect has just been experimentally proven). Taken all data together, it can be concluded that control of CA9 gene expression is a complex process, which affects biogenesis of CA IX protein at several levels, including the functional activation (as discussed in the text below). It is also evident that hypoxic pathway is a major regulator of CA IX.
Role of CA IX in Cancer Hypoxic regulation of CA9 gene indicates that it is a part of the adaptive mechanisms of tumor cells exposed to hypoxia (and associated acidosis) and that the cells need the CA IX protein to survive these stresses typical for tumor microenvironment. In accord with this supposition, it has been demonstrated that CA IX protein is functionally involved in pH regulation and in cell adhesion–migration–invasion, which are important phenomena contributing to tumor progression. The need for increased pH regulation is related to metabolic changes triggered by hypoxia. Insufficient oxygen delivery to hypoxic tumor areas restricts energy production by oxidative phosphorylation and induces metabolic shift of hypoxic cells to glycolysis, which can produce ATP in the absence of oxygen. For this purpose, HIF-1 coordinates up-regulation of virtually all glycolytic enzymes and glucose transporters, as well as lactate dehydrogenase isoforms LDH-A and LHD5, which convert pyruvate to lactate, and also pyruvate dehydrogenase kinase 1, which prevents entry of pyruvate into the Krebs cycle and reinforces the glycolytic
4 Molecular Mechanisms Regulating Expression
69
metabolism (Brahimi-Horn and Pouyssegur 2007). Although glycolysis is less efficient in energy production than respiration, its metabolic intermediates can be utilized for biosynthetic reactions and this makes it useful for proliferating tumor cells in both anaerobic and aerobic conditions, in accord with the classical Warburg’s observation (Gatenby and Gillies 2004). For that reason, tumor cells depend on oncogenic metabolism, in which glycolysis predominates over respiration to extent depending on oxygen availability and activation of different oncogenic pathways. Glycolysis generates lactate as an end-product, but the oncogenic metabolism produces also excess of protons and CO2 (Helmlinger et al. 2002). These acidic products have to be eliminated from tumor cells to maintain neutral intracellular pH and preserve proliferation and survival. The acidic waste then accumulates in the extracellular space because of its insufficient removal by inadequate tumor vasculature and causes extracellular acidosis (Stubbs et al. 2000). Acidosis develops especially in tumor regions that suffer from hypoxia. This is supported by correlation between the mean profiles of partial oxygen pressure and intratumoral pH values (Helmlinger et al. 1997), as well as by VHL/HIF pathway controlled expression of several pH regulating molecules (Karumanchi et al. 2001). Importantly, acidic extracellular pH (pHe) has been associated with tumor progression via multiple effects including up-regulation of angiogenic factors and proteases, increased invasion, and impaired immune functions (Fukumura and Jain 2007; Raghunand et al. 2003). In addition, it can influence the uptake of anticancer drugs and modulate the response of tumor cells to conventional therapy (Teicher 2009). Tumor cells utilize two main mechanisms to regulate their pH – lactate/proton extrusion and bicarbonate uptake (Fig. 4.3). Lactic acid and protons are extruded mostly by the H+/monocarboxylate transporters (MCT4 and MCT1) and the Na+/H+ exchanger (NHE1). Bicarbonate ions are imported by anion exchangers (e.g. AE2) or Na+/bicarbonate co-transporters (NBC1) and react with intracellular protons, thereby increasing the intracellular production of CO2, which diffuses across the plasma membrane to extracellular space. However, origin of the extracellular bicarbonate ions in an acidic pericellular milieu was unclear until CA IX was described as a component of hypoxic tumor phenotype (Stubbs et al. 2000). CA IX seems to be “designed” for this purpose, because its high catalytic activity is insensitive to lactate and reaches maximum at pH 6.5 typical for acidic tumor microenvironment (both aspects contrasting with other CA isoforms) (Innocenti et al. 2005, 2009). In addition, the catalytic domain of CA IX is localized on the cell surface and its entrance is exposed to extracellular space in the position suitable for accommodating CO2 substrate (Alterio et al. 2009). Moreover, CA IX activity is inhibited by bicarbonate suggesting that it operates mainly in catalysis of CO2 hydration to produce bicarbonate rather than to utilize it (Innocenti et al. 2005). Finally, CA IX spatially and functionally cooperates with bicarbonate exchangers/ transporters in bicarbonate transport metabolon (Morgan et al. 2007). According to a recent concept, bicarbonate locally produced by CA IX from extracellular CO2 is directly consumed for transport and neutralization of intracellular pH in tumor cells, whereas proton produced in the same reaction remains
70
J. Pastorek and S. Pastorekova
Fig. 4.3 Role of CA IX as a component of pH regulatory machinery in cancer cell. Hypoxia induces expression of many proteins involved in metabolism and pH regulation including GLUT-1 and glycolytic enzymes. This causes shift to glycolysis that produces excess of lactate and protons. To preserve neutral intracellular pH, these acidic products are extruded by HIF-1-regulated monocarboxylate transporter (MCT4) and the Na+/H+-exchanger (NHE1) and acidify the extracellular microenvironment. Tumor cells also produce CO2 that diffuses across the plasma membrane. Extracellular CO2 is hydrated in a reaction catalyzed by CA IX, which produces protons and bicarbonate ions. Protons remain outside of cell and further enhance acidosis. Bicarbonate ions are delivered directly to bicarbonate transporters (BT) for accelerated inward transport to cytoplasm where they buffer protons and contribute to intracellular pH neutralization. This produces CO2 that leaves the cell by diffusion and may enter new round of hydration
outside of cells and increases extracellular acidosis (Fig. 4.3). In accord with this concept, CA IX role in pH regulation was demonstrated using different cellular models either with constitutive overexpression of CA IX or with suppression of CA IX by RNA interference. The first proof was obtained using MDCK cells grown in monolayer, in which CA IX could acidify the culture medium in hypoxia but not in normoxia, suggesting that hypoxia activates the catalytic performance of CA IX (Svastova et al. 2004). Additional evidence was obtained in three-dimensional multicellular spheroids generated from transfected HCT116 cells (Swietach et al. 2008, 2009). There, the ectopic expression of CA IX resulted in less prominent core intracellular acidity and in enhanced core extracellular acidity. The data obtained in spheroids and interpreted in the framework
4 Molecular Mechanisms Regulating Expression
71
of tumor microenvironment indicate that CA IX facilitates CO2 excretion from cells and that this effect depends on the cellular CO2/lactic acid emission ratio set by local oxygenation and bicarbonate uptake (Swietach et al. 2009). Finally, in hypoxic LS174Tr tumor cells, CA IX was shown to contribute to extracellular acidification in response to a CO2 load and to maintenance of a neutral-alkaline intracellular pH (Chiche et al. 2009b). This latest study also demonstrated that CA IX preserves tumor cell survival in a range of acidic extracellular pH (6.0– 6.8). Moreover, CA IX silencing led to a 40% reduction in xenograft tumor volume supporting the view that hypoxia-induced CA IX is a significant tumor pro-survival pH-regulating enzyme. Interestingly, parallel suppression of CA IX and XII caused dramatic 85% reduction in tumor growth, suggesting cooperation/ compensation of these transmembrane CA isoforms (although suppression of CA XII alone had no effect). The second function of CA IX is linked to cell adhesion and related processes, such as motility, migration, and invasion, which facilitate tumor progression. Again, the insight into this role of CA IX was obtained in the study of E-cadherin-mediated cell–cell adhesion in CA IX-transfected MDCK cells (Svastova et al. 2003). Ectopic expression of CA IX resulted in decreased intercellular contacts, reduced aggregation, and increased dissociation. Similarly to some oncoproteins, such as EFGR and Muc-1, CA IX acts via competitive interaction with b-catenin, thereby disconnecting E-cadherin from adhesion contacts with actin cytoskeleton. The capability of CA IX to interfere with the function of E-cadherin in those tumors that express both proteins might contribute to acquisition of more aggressive phenotype. Since decreased cell adhesion is a prerequisite of motile/migratory phenotype, it was tempting to examine possible involvement of CA IX. Recent experiments using various cell models have shown that expression of CA IX leads to increased scattering, better wound healing, and improved transwell migration both with and without HGF stimulation (manuscript in preparation). On the other hand, the suppression of CA IX by RNA interference decreases the wound healing capacity of tumor cells (Chiche et al. 2010). In this context, it is important to note that motility and migration require not only coordinated formation and release of cell adhesion contacts, digestion and reorganization of the extracellular matrix, but also intense local ion and water transport across the plasma membrane. All of these operations depend on intracellular and extracellular pH regulation. Thus, cell migration is significantly affected by mechanisms that protect tumor cells from hypoxia and acidosis (Stock and Schwab 2009). This knowledge is integrated into a model, in which protruding edge of migrating cell is occupied and driven by different components of ion transport and pH regulating machinery (including CA IX). On this basis, it is conceivable that the catalytic activity plays a considerable role in CA IX-induced de-adhesion, motility, and migration and that the PG domain, implicated in these processes, requires a cooperation with the CA domain (and possibly also with intracellular tail) of CA IX. Initial experimental data favor this idea (unpublished data).
72
J. Pastorek and S. Pastorekova
Clinical Value of CA IX Association with different solid tumors and up-regulation by hypoxia predisposes CA IX to serve as an intrinsic marker of hypoxia and prognostic indicator (Potter and Harris 2004). Moreover, expression pattern and functional involvement of CA IX in hypoxic tumor phenotype make it a suitable target for anticancer treatment (Pastorekova et al. 2006). Both aspects already found a considerable support in clinical studies, in vivo models, and even in the clinical trials. When considering CA IX as a marker of hypoxia, it is important to take into account temporal parameters of its induction and stability. Based on cell culture data, CA IX protein appears on the cell surface several hours after exposure of cell culture to hypoxia, reaching maximum level at around 16 h (Rafajova et al. 2004). Due to its high stability, CA IX remains detectable for several days also in reoxygenated cells. Moreover, CA9 gene is responsive to a broad range of oxygen concentrations representing severe to moderate hypoxia, with optimum between 1 and 2% atmospheric O2 in culture conditions. These parameters differ from those of HIF-1a, which is both quickly induced and short-lived protein as well as from metabolic labels, such as pimonidazole, which accumulates only in severely hypoxic cells (as indicated by narrow perinecrotic incorporation pattern). Furthermore, HIF-1a staining only detects protein levels, but not the transactivation status, which actually determines induction of transcriptional targets. All these facts can explain why there is incomplete overlap between HIF-1a staining and CA IX staining in tumor tissue specimens, and why there are areas with presence of CA IX in absence of HIF-1a and vice versa. From the similar reasons, partial discrepancy exists also between CA IX and other HIF-1 targets, such as VEGF (which is secreted and considerably regulated by non-hypoxic pathways), GLUT-1 (which shows the highest response at 0.1% O2), and surely also other HIF-regulated molecules that all differ with respect to time and magnitude of induction, stability, and other characteristics (Lendahl et al. 2009). Such enormous variability suggests that there is no ideal intrinsic hypoxia marker, since each hypoxia-induced molecule or metabolic label represents only certain facet of the hypoxic response. Indeed, this can be taken as an advantage – for example, comparison of the spatial distribution of CA IX and pimonidazole allows for the discrimination between hypoxic and reoxygenated areas, which might have different clinical impact (Shin et al. 2007). Nevertheless, based on numerous studies it appears that CA IX is the best available indicator of chronic hypoxia and as such, it also seems to have prognostic and/or predictive value in certain tumor types, particularly derived from brain, lung, breast, colon, cervix, ovaria, etc. Although some groups reported absence of significant association with cancer prognosis, most of the published papers state that CA IX expression can be utilized as an independent predictor of poor outcome after therapy and indicator of worse cancer development (summarized in Table 4.1). In this respect, renal cell carcinomas (particularly of clear cell type) represent a special category of tumors, in which CA IX expression is frequent and high due to genetic defect in VHL gene, and absent or non-functional pVHL protein.
Independent prognostic factor for both progression-free survival and overall survival in high-risk early-stage cervical cancer Elevated expression associated with more frequent distant metastases in early-stage cervical cancer Strong correlation to metastasis, important indicator for more aggressive systemic therapy Up-regulated in hypoxic human cervical tumors and ssociated with poor prognosis
Associated with aggressive phenotype and chemo-resistance of basal-like breast tumors Independent prognostic factor with respect to relapse-free survival Correlates with poor prognosis in invasive breast carcinoma In both univariate and multivariate analysis, survival is significantly inferior in patients with invasive breast tumors expressing CA IX Independent prognostic marker in premenopausal patients with one to three positive lymph nodes and a putative marker of radiation resistance Negative predictor of treatment efficacy in ER-positive patients on the adjuvant tamoxifen after primary chemo-endocrine therapy Associated with a significantly shorter disease-free survival in patients with HIF-a-negative breast tumors Perinecrotic HIF-a overexpression with strong expression of CA IX and GLUT-1 associated with poor prognosis in invasive breast cancer Associated with worse relapse-free and overall survival in an unselected cohort of patients with invasive breast cancer
Cervical carcinoma
Breast carcinoma
Chia et al. (2001) (continued)
Vleugel et al. (2005)
Generali et al. (2006a)
Generali et al. (2006b)
Brennan et al. (2006)
Tan et al. (2009) Crabb et al. (2008) Trastour et al. (2007) Hussain et al. (2007)
Kim et al. (2006) Loncaster et al. (2001)
Kirkpatrick et al. (2008)
Liao et al. (2010)
Table 4.1 Overview of clinical studies on relationship of CA IX to cancer progression, prognosis, and/or treatment outcome (non-RCC tumors) Tumor type Proposed clinical value of CA IX Reference
4 Molecular Mechanisms Regulating Expression 73
Proposed clinical value of CA IX
Co-expression of CA IX and HIF-a associated with poor prognosis in oral squamous cell carcinoma patients Combined CA IX and HIF-a significantly predictive of overall survival in patients with advanced head-and-neck cancer Better local control in radiotherapy-treated laryngeal carcinoma patients having tumors with non-hypoxic profile (low CA IX, low HIF-a) Significant association with postoperative recurrence and poor overall survival in surgically treated oral squamous cell carcinoma High CA IX and Ki-67 expression associated with poorer overall survival and shorter diseasefree survival in patients with squamous cell carcinoma of the tongue Independent correlation with prognosis when combined with GLUT-1
Poorer outcome in patients with oligodendroglial tumors Significant parameter for the survival of patients with grades II/III astrocytic gliomas Poor prognosis in astrocytic tumor patients
Expression by cancer-associated fibroblasts predicts lower survival rate in patients with lung adenocarcinoma Full-length CA IX transcript associated with lymph node involvement, tumor stage, and poor prognosis Indicates poor prognosis for patients with stages I+II lung adenocarcinoma The most reliable hypoxia marker for predicting tumor aggressiveness Indicator for poor disease-free survival in early-stage non-small-cell lung cancer Associated with a poor prognosis in non-small-cell lung cancer
Intensity predicts disease-free survival and disease-specific survival in patients with rectal cancer
Correlates with extremely poor prognosis in esophageal and gastric adenocarcinomas Has a tripartite role as a diagnostic, prognostic, and therapeutic molecular marker in urothelial carcinoma of the bladder Correlates with clinical outcome in transitional cell carcinoma of the bladder Predicts for poor prognosis in patients with deep, large, and high-grade soft tissue sarcoma
Head-and-neck cancer
Brain tumors
Lung carcinoma
Colorectal carcinoma
Other tumors
Table 4.1 (continued)
Tumor type
Hussain et al. (2004) Maseide et al. (2004)
Driessen et al. (2006) Klatte et al. (2009)
Korkeila et al. (2009)
Kon-no et al. (2006) Kim et al. (2005) Kim et al. (2004) Swinson et al. (2003)
Malentacchi et al. (2009)
Nakao et al. (2009)
Jarvela et al. (2008) Korkolopoulou et al. (2007) Haapasalo et al. (2006)
De Schutter et al. (2005)
Kim et al. (2007)
Choi et al. 2008)
Schrijvers et al. (2008)
Kappler et al. (2008)
Eckert et al. (2010)
Reference
74 J. Pastorek and S. Pastorekova
4 Molecular Mechanisms Regulating Expression
75
Resulting activation of HIF pathway leads to constitutive up-regulation of HIF targets including CA IX independently of physiological hypoxia (Mandriota et al. 2002). In contrast to other tumor types, CA IX expression in RCC decreases with increasing tumor grade and stage. This is apparently because CA IX is a target of HIF-1a that predominates early in VHL disease but is later displaced by tumor growth-promoting isoform HIF-2a (Raval et al. 2005). Correspondingly, decreased CA IX levels are independently associated with poor prognosis in advanced RCC (Bui et al. 2003). However, levels of CA IX can be upregulated by interferons (IFN-a and IFN-g), which can increase the therapeutic responses in 5–25% of patients with metastatic RCC (Brouwers et al. 2003b). Furthermore, high CA IX expression was suggested as an important predictor of better outcome in RCC patients receiving interleukin-2-based therapy (Atkins et al. 2005). In several recent reviews, CA IX has been proposed to be the most significant molecular marker described in kidney cancer (Pantuck et al. 2008; Signoretti et al. 2008). Assessment of CA IX expression in biopsies offers a new means for stratification of cancer patients and rational selection of appropriate treatment strategy – an approach highly desirable in recent clinical oncology. However, in certain phases of cancer management, the patients would benefit from non-invasive monitoring of CA IX shed to blood circulation. Data published so far suggest that serum CA IX levels in RCC patients correlate with tumor size and stage of the disease, and support the use of soluble CA IX as a biomarker of progression and postoperative recurrence (Zhou et al. 2009). Furthermore, in vivo imaging of CA IX is recently coming to the scene as a very promising approach for pre-surgical diagnosis of RCC (Divgi et al. 2007; Zhang 2008). The imaging tool called REDECTANE® developed by WILEX AG company is based on radioactively labeled monoclonal antibody G250 originally generated by Oosterwijk and coworkers (Lam et al. 2005). The radiolabeled antibody WX-G250 targets clear cell renal cell carcinoma and accumulates in the tumor tissue via binding to CA IX followed by receptor-mediated internalization (Durrbach et al. 1999). The G250 accumulation can be visualized by means of positron emission tomography (PET) (Divgi et al. 2007). A proof-of-concept study showed that REDECTANE® could predict clear cell RCC with a specificity of 100% and a sensitivity of 94%. Moreover, Phase II clinical study demonstrated that the combination of REDECTANE® with PET and computer tomography (CT) versus the standard use of CT alone could improve the diagnosis of renal masses (http://www.wilex.com).
CA IX Targeting Strategies CA IX usefulness as therapeutic target stems from the following facts: (a) strong association with different tumors and low or absent expression in normal tissues, (b) induction by hypoxia and correlation with aggressive phenotype, (c) localization
76
J. Pastorek and S. Pastorekova
on the cell surface allowing interaction with antibodies or drugs, and (d) direct functional involvement in hypoxic tumor phenotype. These facts create a basis for two principal strategies of targeting CA IX-expressing tumor cells – immunotherapy with specific antibodies inducing cytotoxicity and functional inhibition of the CA IX catalytic activity leading to disruption of pH control. The first approach has been consistently studied in pre-clinical and clinical experiments in patients with renal cell carcinomas. Although advanced RCC show reduced CA IX expression, the cut-off was set to 85% (i.e. the expression is considered low when less than 85% of cells display CA IX) and this is still a very high proportion of targetable cells. Again, the major tool here is the monoclonal antibody G250 (Brouwers et al. 2003a; Mulders et al. 2004). Chimeric G250 (composed of variable regions of murine G250 and constant regions derived from human IgG) is recently produced by WILEX AG under commercial name RENCAREX®. Its effector mechanism is ADCC (antibody-dependent cell cytotoxicity). In Phase I and II studies, RENCAREX® has shown good safety and tolerability and a promising efficacy profile. The purpose of the ongoing Phase III study (called ARISER) is to assess the effect of adjuvant treatment with RENCAREX® on disease-free survival and overall survival in RCC patients with a high risk of recurrence following surgery (nephrectomy). Additional monoclonal antibodies potentially suitable for immunotherapeutic purposes are under development. Some of them display similar binding properties and capacity to internalize as G250 and exhibit anticancer effects in animal models (Zatovicova et al. in press). The second approach – functional inhibition of cancer-related carbonic anhydrases – is only at early stage of investigation and is based on carbonic anhydrase inhibitors (sulfonamides and their derivatives) that are mostly produced and characterized by Supuran and coworkers, who were the first to introduce the concept of their use as anticancer drugs (Supuran 2007). Sulfonamides and related compounds generally show non-selective inhibitory activity toward different CA isoenzymes including CA IX, which can be inhibited by a broad range of compounds (see Table 4.2). Interestingly, efficient CA IX inhibition was observed also with celecoxib and valdecoxib, sulfonamide inhibitors of cyclooxygense 2 (COX-2, a key enzyme of arachidonic acid metabolism involved in colorectal carcinoma), and with inhibitors of steroid sulfatase (implicated in production of active steroids in breast cancer) (Weber et al. 2004; Winum et al. 2003a). This may suggest that the mode of action of these drugs could involve targeting of CAs. As demonstrated by several recent papers, increased selectivity toward CA IX can be successfully achieved by modulating the physical and chemical properties of the compounds via attachment of different side chains and other modifications. Certain alterations can introduce or improve the membrane impermeability so that the inhibitor can bind only or predominantly to extracellularly exposed CAs. The other changes can modify the size or surface topology that fits better into active site cavity of CA IX than into other isoforms. Some types of modifications increase the efficiency of inhibition so that they work even at subnanomolar concentrations. This approach was used for development of several
Acetazolamide Aromatic and heterocyclic sulfonamides Halogenated sulfonamides Aliphatic sulfamates Sulfamates or bis-sulfamates (+ steroid sulfatase inhibitors) E7070 (indisulam) anticancer drug in clinical development Non-steroidal anti-inflammatory inhibitors of COX-2 Positively-charged pyridinium derivatives of sulfanilamides Polyfluorinated sulfonamides Sulfonamides incorporating 1,2,4-triazine moieties Sulfonamides derived from 4-isothiocyanato-benzolamide Carboxylates Bis-sulfamates Sulfonamides incorporating hydrazino moieties Boron-containing sulfamides, sulfonamides, and sulfamates
Abbate et al. (2004) Weber et al. (2004)
Potent inhibition of CA I, II, IX in the range of 15–31 nM Unexpected nanomolar inhibition of CA I, II, IV, IX by celecoxib and valdecoxib First selective, membrane-impermeant inhibitors targeting CA IX
Effective inhibitors of CA IX suitable for boron neutron capture therapy targeting hypoxic tumors
The first subnanomolar and selective CA IX inhibitor CA IX-selective inhibitors with selectivity ratios for CA IX over CA II inhibition in the range of 166–706 Excellent inhibitory properties against CA IX with inhibition constants in the range of 3.2–23 nM Lactate insensitivity of CA IX Low nanomolar CA IX inhibitors with good selectivity ratio Several low nanomolar CA IX inhibitors
Ilies et al. (2003) Winum et al. (2003b) Winum et al. (2003a)
CA IX inhibition profile different from CA I, II, IV Efficient CA IX inhibition in the range of 9–23 nM Efficient CA IX inhibition in the range of 18–63 nM
(continued)
Winum et al. (2005a)
Innocenti et al. (2005) Winum et al. (2005c) Winum et al. (2005b)
Cecchi et al. (2004)
Vullo et al. (2004) Garaj et al. (2004)
Pastorekova et al. (2004a)
Parkkila et al. (2000) Vullo et al. (2003)
Inhibition of in vitro migration of renal carcinoma cells Several nanomolar CA IX inhibitors
Table 4.2 Summary of sulfonamide inhibitors, their derivatives, and other compounds tested against different CA isoforms including CA IX Inhibitor type Inhibition properties Reference
4 Molecular Mechanisms Regulating Expression 77
Imatinib (Glivec/Gleevec) and nilotinib (Tasigna)
Inhibitor-coated gold nanoparticles Coumarin derivatives
Bioreductive nitro-containing sulfonamides Biphenylsulfonamides
Mercaptans increase the inhibitory power (52.8- to 243-fold) over the corresponding oxidized derivatives Several potent and selective inhibitors of CA IX
Aromatic sulfonamides activatable in hypoxic tumors Glycoconjugate benzene sulfonamides Antiepileptic drug sulthiame Glycosyl-thioureido-sulfonamides Potent inhibitor of CA II, VII, IX, XII (6–56 nM) Low nM binding to CA IX and XII, high selectivity ratios in the range of 107–955 Some CA IX inhibitors with excellent selectivity ratios (in the range of 10–1395) Inhibitors of CA IX with cytotoxic activity against human colon, lung, and breast cancer cell lines Selective inhibition of CA IX (membrane impermeant) Bind at the entrance of the active site cavity, one of them highly selective for CA IX Inhibited all 13 catalytically active CA isoforms in the range of 4.1 nM–20.2 mM
Inhibition properties
Table 4.2 (continued)
Inhibitor type
Parkkila et al. (2009)
Stiti et al. (2008) Maresca et al. (2010)
Morsy et al. (2009)
D’Ambrosio et al. (2008)
Temperini et al. (2007) Smaine et al. (2007)
Wilkinson et al. (2006)
Saczewski et al. (2006)
Reference
78 J. Pastorek and S. Pastorekova
4 Molecular Mechanisms Regulating Expression
79
compounds with a reasonable selectivity ratio favoring inhibition activities against CA IX compared to other isoforms, in particular CA II (Table 4.2). Another strategy was used to generate hypoxia-activated inhibitors. Different 2-mercapto-substituted-benzenesulfonamides and their disulfides/sulfones showed consistently increased inhibitory power (52.8- to 243-fold) over the corresponding oxidized (S–S type) derivatives (Saczewski et al. 2006). The best representatives out of these differentially acting derivatives can serve as lead compounds for further development of CA IX -specific inhibitors with therapeutic potential against cancer. Since the inhibition data were obtained using the purified recombinant catalytic domain of CA IX, experiments with native protein in cellular context are needed to determine their effective concentrations and consequences of their application. Noteworthy, fluorescein-conjugated carbonic anhydrase inhibitor thioureidohomosulfanilamide (FITC-CAI) was found to bind only to hypoxic cells expressing CA IX, but neither to CA IX-positive normoxic cells nor the CA IX-negative controls (Cecchi et al. 2005; Svastova et al. 2004). Moreover, the FITC-CAI binding was perturbed by reoxygenation (Dubois et al. 2007). This interesting finding suggests that CA IX active site is accessible to sulfonamide only under hypoxic conditions. Results of in vivo imaging confirm previous in vitro data demonstrating that CAI binding and retention requires exposure to hypoxia (Dubois et al. 2009). Thus, labeled sulfonamides may provide a powerful tool to visualize hypoxic tumors. This may represent a promising approach to patient selection for CA IX-directed therapies based on the functional inhibition of CA IX. CA inhibitors can be used as anticancer drugs to compromise the survival of tumor cells due to perturbed pH control, or in combination with conventional chemotherapeutics to improve their uptake and efficacy due to modulated pH gradient. The first option is based on significance of the proper pH control in tumor cells, particularly in those that suffer from hypoxia. It is well established that extracellular acidosis supports cancer progression, whereas neutral intracellular pH is required for survival of hypoxic cancer cells (Stubbs et al. 2000). Inhibition or reduction of extracellular acidosis by carbonic anhydrase inhibitors thus offers a possible therapeutic approach to retardation of tumor progression (Pouyssegur et al. 2006). On the other hand, inhibition of CA activity can lead to inability of tumor cells to neutralize their intracellular pH and adapt to metabolic stress. In fact, several sulfonamide derivatives were capable of reducing an extracellular acidification and intracellular neutralization mediated by CA IX, suggesting that these compounds interfere with the activity of CA IX in cell culture model (Svastova et al. 2004; Swietach et al. 2009). In addition, changes in extracellular pH can affect therapeutic outcome of conventional chemotherapy by influencing the uptake of weakly electrolytic anticancer drugs (Kozin et al. 2001; Vukovic and Tannock 1997). In particular, reduction of extracellular acidosis can increase the uptake and cytotoxic effects of weakly basic drugs, including doxorubicin. Interestingly, chronic ingestion of sodium bicarbonate solution was shown to enhance the capacity of doxorubicin to decrease the tumor size (Raghunand et al. 2003).
80
J. Pastorek and S. Pastorekova
In line with these suppositions, older but forward-looking study showed that acetazolamide, a classical CA inhibitor, reduced in vivo growth of tumor when given alone and produced additive tumor growth delays when administered in combination with various chemotherapeutic compounds (Teicher et al. 1993). This was recently supported using more sophisticated approach, based on two acetazolamide derivatives containing either a charged fluorophore or an albumin-binding moiety, which restrict binding to carbonic anhydrase IX and XII present on tumor cells (Ahlskog et al. 2009a). In vivo studies showed the preferential targeting of tumor cells by the fluorescent acetazolamide derivative and the ability of the albuminbinding acetazolamide derivative to cause tumor retardation in a SK-RC-52 (RCC) xenograft model of cancer. Additional in vivo experiments with promising preliminary results are underway.
Conclusion Our present knowledge on CA IX clearly suggests that this hypoxia-induced molecule is an important component of the tumor microenvironment and that many tumors need CA IX to survive microenvironmental stresses. This addiction of tumor cells to CA IX makes it a useful marker of hypoxia, prognostic indicator, and therapeutic target. Although main strategies of diagnostic and therapeutic exploitation of CA IX have been rationally designed, their concepts were experimentally proven, practically evaluated (in part), and several specific tools are available; further preclinical and clinical studies are necessary to translate all proposed applications to the benefit of cancer patients.
References Abbate F, Casini A, Owa T, Scozzafava A, Supuran CT (2004) Carbonic anhydrase inhibitors: E7070, a sulfonamide anticancer agent, potently inhibits cytosolic isozymes I and II, and transmembrane, tumor-associated isozyme IX. Bioorg Med Chem Lett 14:217–223 Ahlskog JK, Dumelin CE, Trussel S, Marlind J, Neri D (2009a) In vivo targeting of tumorassociated carbonic anhydrases using acetazolamide derivatives. Bioorg Med Chem Lett 19:4851–4856 Ahlskog JK, Schliemann C, Marlind J, Qureshi U, Ammar A, Pedley RB, Neri D (2009b) Human monoclonal antibodies targeting carbonic anhydrase IX for the molecular imaging of hypoxic regions in solid tumours. Br J Cancer 101:645–657 Alterio V, Hilvo M, Di Fiore A, Supuran CT, Pan P, Parkkila S, Scaloni A, Pastorek J, Pastorekova S, Pedone C, Scozzafava A, Monti SM, De Simone G (2009) Crystal structure of the catalytic domain of the tumor-associated human carbonic anhydrase IX. Proc Natl Acad Sci USA 106:16233–16238 Ashida S, Nishimori I, Tanimura M, Onishi S, Shuin T (2002) Effects of von Hippel-Lindau gene mutation and methylation status on expression of transmembrane carbonic anhydrases in renal cell carcinoma. J Cancer Res Clin Oncol 128:561–568
4 Molecular Mechanisms Regulating Expression
81
Atkins M, Regan M, McDermott D, Mier J, Stanbridge E, Youmans A, Febbo P, Upton M, Lechpammer M, Signoretti S (2005) Carbonic anhydrase IX expression predicts outcome of interleukin 2 therapy for renal cancer. Clin Cancer Res 11:3714–3721 Barathova M, Takacova M, Holotnakova T, Gibadulinova A, Ohradanova A, Zatovicova M, Hulikova A, Kopacek J, Parkkila S, Supuran CT, Pastorekova S, Pastorek J (2008) Alternative splicing variant of the hypoxia marker carbonic anhydrase IX expressed independently of hypoxia and tumour phenotype. Br J Cancer 98:129–136 Bartosova M, Parkkila S, Pohlodek K, Karttunen TJ, Galbavy S, Mucha V, Harris AL, Pastorek J, Pastorekova S (2002) Expression of carbonic anhydrase IX in breast is associated with malignant tissues and is related to overexpression of c-erbB2. J Pathol 197:314–321 Brahimi-Horn MC, Pouyssegur J (2007) Hypoxia in cancer cell metabolism and pH regulation. Essays Biochem 43:165–178 Brennan DJ, Jirstrom K, Kronblad A, Millikan RC, Landberg G, Duffy MJ, Ryden L, Gallagher WM, O’Brien SL (2006) CA IX is an independent prognostic marker in premenopausal breast cancer patients with one to three positive lymph nodes and a putative marker of radiation resistance. Clin Cancer Res 12:6421–6431 Brouwers AH, Buijs WC, Oosterwijk E, Boerman OC, Mala C, De Mulder PH, Corstens FH, Mulders PF, Oyen WJ (2003a) Targeting of metastatic renal cell carcinoma with the chimeric monoclonal antibody G250 labeled with (131)I or (111)In: an intrapatient comparison. Clin Cancer Res 9:3953S–3960S Brouwers H, Frielink C, Oosterwijk E, Oyen WJ, Corstens FH, Boerman OC (2003b) Interferons can upregulate the expression of the tumor associated antigen G250-MN/CA IX, a potential target for (radio)immunotherapy of renal cell carcinoma. Cancer Biother Radiopharm 18:539–547 Bui MH, Seligson D, Han KR, Pantuck AJ, Dorey FJ, Huang Y, Horvath S, Leibovich BC, Chopra S, Liao SY, Stanbridge E, Lerman MI, Palotie A, Figlin RA, Belldegrun AS (2003) Carbonic anhydrase IX is an independent predictor of survival in advanced renal clear cell carcinoma: implications for prognosis and therapy. Clin Cancer Res 9:802–811 Camenisch G, Stroka DM, Gassmann M, Wenger RH (2001) Attenuation of HIF-1 DNA-binding activity limits hypoxia-inducible endothelin-1 expression. Pflugers Arch 443:240–249 Carney WP (2007) Circulating oncoproteins HER2/neu, EGFR and CAIX (MN) as novel cancer biomarkers. Expert Rev Mol Diagn 7:309–319 Casey JR (2006) Why bicarbonate? Biochem Cell Biol 84:930–939 Cecchi A, Hulikova A, Pastorek J, Pastorekova S, Scozzafava A, Winum JY, Montero JL, Supuran CT (2005) Carbonic anhydrase inhibitors. Design of fluorescent sulfonamides as probes of tumor-associated carbonic anhydrase IX that inhibit isozyme IX-mediated acidification of hypoxic tumors. J Med Chem 48:4834–4841 Cecchi A, Winum JY, Innocenti A, Vullo D, Montero JL, Scozzafava A, Supuran CT (2004) Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/tumor-associated carbonic anhydrase isozymes I, II, and IX with sulfonamides derived from 4-isothiocyanatobenzolamide. Bioorg Med Chem Lett 14:5775–5780 Coleman ML, Ratcliffe PJ (2007) Oxygen sensing and hypoxia-induced responses. Essays Biochem 43:1–15 Crabb SJ, Bajdik CD, Leung S, Speers CH, Kennecke H, Huntsman DG, Gelmon KA (2008) Can clinically relevant prognostic subsets of breast cancer patients with four or more involved axillary lymph nodes be identified through immunohistochemical biomarkers? A tissue microarray feasibility study. Breast Cancer Res 10:R6 D’Ambrosio K, Vitale RM, Dogne JM, Masereel B, Innocenti A, Scozzafava A, De Simone G, Supuran CT (2008) Carbonic anhydrase inhibitors: bioreductive nitro-containing sulfonamides with selectivity for targeting the tumor associated isoforms IX and XII. J Med Chem 51: 3230–3237 De Schutter H, Landuyt W, Verbeken E, Goethals L, Hermans R, Nuyts S (2005) The prognostic value of the hypoxia markers CA IX and GLUT 1 and the cytokines VEGF and IL 6 in head and neck squamous cell carcinoma treated by radiotherapy +/– chemotherapy. BMC Cancer 5:42
82
J. Pastorek and S. Pastorekova
Divgi CR, Pandit-Taskar N, Jungbluth AA, Reuter VE, Gonen M, Ruan S, Pierre C, Nagel A, Pryma DA, Humm J, Larson SM, Old LJ, Russo P (2007) Preoperative characterisation of clear-cell renal carcinoma using iodine-124-labelled antibody chimeric G250 (124I-cG250) and PET in patients with renal masses: a phase I trial. Lancet Oncol 8:304–310 Dorai T, Sawczuk IS, Pastorek J, Wiernik PH, Dutcher JP (2005) The role of carbonic anhydrase IX overexpression in kidney cancer. Eur J Cancer 41:2935–2947 Driessen A, Landuyt W, Pastorekova S, Moons J, Goethals L, Haustermans K, Nafteux P, Penninckx F, Geboes K, Lerut T, Ectors N (2006) Expression of carbonic anhydrase IX (CA IX), a hypoxia-related protein, rather than vascular-endothelial growth factor (VEGF), a pro-angiogenic factor, correlates with an extremely poor prognosis in esophageal and gastric adenocarcinomas. Ann Surg 243:334–340 Dubois L, Douma K, Supuran CT, Chiu RK, van Zandvoort MA, Pastorekova S, Scozzafava A, Wouters BG, Lambin P (2007) Imaging the hypoxia surrogate marker CA IX requires expression and catalytic activity for binding fluorescent sulfonamide inhibitors. Radiother Oncol 83:367–373 Dubois L, Lieuwes NG, Maresca A, Thiry A, Supuran CT, Scozzafava A, Wouters BG, Lambin P (2009) Imaging of CA IX with fluorescent labelled sulfonamides distinguishes hypoxic and (re)-oxygenated cells in a xenograft tumour model. Radiother Oncol 92:423–428 Durrbach A, Angevin E, Poncet P, Rouleau M, Chavanel G, Chapel A, Thierry D, Gorter A, Hirsch R, Charpentier B, Senik A, Hirsch F (1999) Antibody-mediated endocytosis of G250 tumor-associated antigen allows targeted gene transfer to human renal cell carcinoma in vitro. Cancer Gene Ther 6:564–571 Eckert AW, Lautner MH, Schutze A, Bolte K, Bache M, Kappler M, Schubert J, Taubert H, Bilkenroth U (2010) Co-expression of Hif1alpha and CAIX is associated with poor prognosis in oral squamous cell carcinoma patients. J Oral Pathol Med 39:313–317 Fukumura D, Jain RK (2007) Tumor microenvironment abnormalities: causes, consequences, and strategies to normalize. J Cell Biochem 101:937–949 Garaj V, Puccetti L, Fasolis G, Winum JY, Montero JL, Scozzafava A, Vullo D, Innocenti A, Supuran CT (2004) Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/ tumor-associated carbonic anhydrase isozymes I, II, and IX with sulfonamides incorporating 1,2,4-triazine moieties. Bioorg Med Chem Lett 14:5427–5433 Gatenby RA, Gillies RJ (2004) Why do cancers have high aerobic glycolysis? Nat Rev Cancer 4:891–899 Generali D, Berruti A, Brizzi MP, Campo L, Bonardi S, Wigfield S, Bersiga A, Allevi G, Milani M, Aguggini S, Gandolfi V, Dogliotti L, Bottini A, Harris AL, Fox SB (2006a) Hypoxiainducible factor-1alpha expression predicts a poor response to primary chemoendocrine therapy and disease-free survival in primary human breast cancer. Clin Cancer Res 12: 4562–4568 Generali D, Fox SB, Berruti A, Brizzi MP, Campo L, Bonardi S, Wigfield SM, Bruzzi P, Bersiga A, Allevi G, Milani M, Aguggini S, Dogliotti L, Bottini A, Harris AL (2006b) Role of carbonic anhydrase IX expression in prediction of the efficacy and outcome of primary epirubicin/ tamoxifen therapy for breast cancer. Endocr Relat Cancer 13:921–930 Gut MO, Parkkila S, Vernerova Z, Rohde E, Zavada J, Hocker M, Pastorek J, Karttunen T, Gibadulinova A, Zavadova Z, Knobeloch KP, Wiedenmann B, Svoboda J, Horak I, Pastorekova S (2002) Gastric hyperplasia in mice with targeted disruption of the carbonic anhydrase gene Car9. Gastroenterology 123:1889–1903 Haapasalo J, Nordfors K, Jarvela S, Bragge H, Rantala I, Parkkila AK, Haapasalo H, Parkkila S (2007) Carbonic anhydrase II in the endothelium of glial tumors: a potential target for therapy. Neuro Oncol 9:308–313 Haapasalo JA, Nordfors KM, Hilvo M, Rantala IJ, Soini Y, Parkkila AK, Pastorekova S, Pastorek J, Parkkila SM, Haapasalo HK (2006) Expression of carbonic anhydrase IX in astrocytic tumors predicts poor prognosis. Clin Cancer Res 12:473–477 Harris AL (2002) Hypoxia – a key regulatory factor in tumour growth. Nat Rev Cancer 2:38–47
4 Molecular Mechanisms Regulating Expression
83
Helmlinger G, Sckell A, Dellian M, Forbes NS, Jain RK (2002) Acid production in glycolysisimpaired tumors provides new insights into tumor metabolism. Clin Cancer Res 8: 1284–1291 Helmlinger G, Yuan F, Dellian M, Jain RK (1997) Interstitial pH and pO2 gradients in solid tumors in vivo: high-resolution measurements reveal a lack of correlation. Nat Med 3: 177–182 Hewett-Emmett D, Tashian RE (1996) Functional diversity, conservation, and convergence in the evolution of the alpha-, beta-, and gamma-carbonic anhydrase gene families. Mol Phylogenet Evol 5:50–77 Hilvo M, Baranauskiene L, Salzano AM, Scaloni A, Matulis D, Innocenti A, Scozzafava A, Monti SM, Di Fiore A, De Simone G, Lindfors M, Janis J, Valjakka J, Pastorekova S, Pastorek J, Kulomaa MS, Nordlund HR, Supuran CT, Parkkila S (2008) Biochemical characterization of CA IX, one of the most active carbonic anhydrase isozymes. J Biol Chem 283:27799–27809 Hirota K, Semenza GL (2005) Regulation of hypoxia-inducible factor 1 by prolyl and asparaginyl hydroxylases. Biochem Biophys Res Commun 338:610–616 Holotnakova T, Tylkova L, Takacova M, Kopacek J, Petrik J, Pastorekova S, Pastorek, J (2010) Role of the HBx oncoprotein in carbonic anhydrase 9 induction. J Med Virol 82:32–40 Hulikova A, Zatovicova M, Svastova E, Ditte P, Brasseur R, Kettmann R, Supuran CT, Kopacek J, Pastorek J, Pastorekova S (2009) Intact intracellular tail is critical for proper functioning of the tumor-associated, hypoxia-regulated carbonic anhydrase IX. FEBS Lett 583:3563–3568 Hussain SA, Ganesan R, Reynolds G, Gross L, Stevens A, Pastorek J, Murray PG, Perunovic B, Anwar MS, Billingham L, James ND, Spooner D, Poole CJ, Rea DW, Palmer DH (2007) Hypoxia-regulated carbonic anhydrase IX expression is associated with poor survival in patients with invasive breast cancer. Br J Cancer 96:104–109 Hussain SA, Palmer DH, Ganesan R, Hiller L, Gregory J, Murray PG, Pastorek J, Young L, James ND (2004) Carbonic anhydrase IX, a marker of hypoxia: correlation with clinical outcome in transitional cell carcinoma of the bladder. Oncol Rep 11:1005–1010 Chia SK, Wykoff CC, Watson PH, Han C, Leek RD, Pastorek J, Gatter KC, Ratcliffe P, Harris AL (2001) Prognostic significance of a novel hypoxia-regulated marker, carbonic anhydrase IX, in invasive breast carcinoma. J Clin Oncol 19:3660–3668 Chiang WL, Chu SC, Yang SS, Li MC, Lai JC, Yang SF, Chiou HL, Hsieh YS (2002) The aberrant expression of cytosolic carbonic anhydrase and its clinical significance in human non-small cell lung cancer. Cancer Lett 188:199–205 Chiche J, Ilc K, Brahimi-Horn MC, Pouyssegur J (2010) Membrane-bound carbonic anhydrases are key pH regulators controlling tumor growth and cell migration. Adv Enzyme Regul 50:20–33 Chiche J, Ilc K, Laferriere J, Trottier E, Dayan F, Mazure NM, Brahimi-Horn MC, Pouyssegur J (2009b) Hypoxia-inducible carbonic anhydrase IX and XII promote tumor cell growth by counteracting acidosis through the regulation of the intracellular pH. Cancer Res 69:358–368 Cho M, Grabmaier K, Kitahori Y, Hiasa Y, Nakagawa Y, Uemura H, Hirao Y, Ohnishi T, Yoshikawa K, Ooesterwijk E (2000) Activation of the MN/CA9 gene is associated with hypomethylation in human renal cell carcinoma cell lines. Mol Carcinog 27:184–189 Choi SW, Kim JY, Park JY, Cha IH, Kim J, Lee S (2008) Expression of carbonic anhydrase IX is associated with postoperative recurrence and poor prognosis in surgically treated oral squamous cell carcinoma. Hum Pathol 39:1317–1322 Ihnatko R, Kubes M, Takacova M, Sedlakova O, Sedlak J, Pastorek J, Kopacek J, Pastorekova S (2006) Extracellular acidosis elevates carbonic anhydrase IX in human glioblastoma cells via transcriptional modulation that does not depend on hypoxia. Int J Oncol 29:1025–1033 Ilies MA, Vullo D, Pastorek J, Scozzafava A, Ilies M, Caproiu MT, Pastorekova S, Supuran CT (2003) Carbonic anhydrase inhibitors. Inhibition of tumor-associated isozyme IX by halogenosulfanilamide and halogenophenylaminobenzolamide derivatives. J Med Chem 46:2187–2196 Innocenti A, Pastorekova S, Pastorek J, Scozzafava A, De Simone G, Supuran CT (2009) The proteoglycan region of the tumor-associated carbonic anhydrase isoform IX acts as anintrinsic
84
J. Pastorek and S. Pastorekova
buffer optimizing CO2 hydration at acidic pH values characteristic of solid tumors. Bioorg Med Chem Lett 19:5825–5828 Innocenti A, Vullo D, Scozzafava A, Casey JR, Supuran C (2005) Carbonic anhydrase inhibitors. Interaction of isozymes I, II, IV, V, and IX with carboxylates. Bioorg Med Chem Lett 15:573–578 Ivanov S, Liao SY, Ivanova A, Danilkovitch-Miagkova A, Tarasova N, Weirich G, Merrill MJ, Proescholdt MA, Oldfield EH, Lee J, Zavada J, Waheed A, Sly W, Lerman MI, Stanbridge EJ (2001) Expression of hypoxia-inducible cell-surface transmembrane carbonic anhydrases in human cancer. Am J Pathol 158:905–919 Ivanov SV, Kuzmin I, Wei MH, Pack S, Geil L, Johnson BE, Stanbridge EJ, Lerman MI (1998) Down-regulation of transmembrane carbonic anhydrases in renal cell carcinoma cell lines by wild-type von Hippel-Lindau transgenes. Proc Natl Acad Sci USA 95:12596–12601 Jakubickova L, Biesova Z, Pastorekova S, Kettmann R, Pastorek J (2005) Methylation of the CA9 promoter can modulate expression of the tumor-associated carbonic anhydrase IX in dense carcinoma cell lines. Int J Oncol 26:1121–1127 Jarvela S, Parkkila S, Bragge H, Kahkonen M, Parkkila AK, Soini Y, Pastorekova S, Pastorek J, Haapasalo H (2008) Carbonic anhydrase IX in oligodendroglial brain tumors. BMC Cancer 8:1 Kaluz S, Kaluzova M, Chrastina A, Olive PL, Pastorekova S, Pastorek J, Lerman MI, Stanbridge EJ (2002) Lowered oxygen tension induces expression of the hypoxia marker MN/carbonic anhydrase IX in the absence of hypoxia-inducible factor 1 alpha stabilization: a role for phosphatidylinositol 3¢-kinase. Cancer Res 62:4469–4477 Kaluz S, Kaluzova M, Liao SY, Lerman M, Stanbridge EJ (2009) Transcriptional control of the tumor- and hypoxia-marker carbonic anhydrase 9: a one transcription factor (HIF-1) show? Biochim Biophys Acta 1795:162–172 Kaluz S, Kaluzova M, Opavsky R, Pastorekova S, Gibadulinova A, Dequiedt F, Kettmann R, Pastorek J (1999) Transcriptional regulation of the MN/CA 9 gene coding for the tumorassociated carbonic anhydrase IX. Identification and characterization of a proximal silencer element. J Biol Chem 274:32588–32595 Kappler M, Taubert H, Holzhausen HJ, Reddemann R, Rot S, Becker A, Kuhnt T, Dellas K, Dunst J, Vordermark D, Hansgen G, Bache M (2008) Immunohistochemical detection of HIF-1alpha and CAIX in advanced head-and-neck cancer. Prognostic role and correlation with tumor markers and tumor oxygenation parameters. Strahlenther Onkol 184:393–399 Karumanchi SA, Jiang L, Knebelmann B, Stuart-Tilley AK, Alper SL, Sukhatme VP (2001) VHL tumor suppressor regulates Cl-/HCO3- exchange and Na+/H+ exchange activities in renal carcinoma cells. Physiol Genomics 5:119–128 Kim JY, Shin HJ, Kim TH, Cho KH, Shin KH, Kim BK, Roh JW, Lee S, Park SY, Hwang YJ, Han IO (2006) Tumor-associated carbonic anhydrases are linked to metastases in primary cervical cancer. J Cancer Res Clin Oncol 132:302–308 Kim SJ, Rabbani ZN, Dewhirst MW, Vujaskovic Z, Vollmer RT, Schreiber EG, Oosterwijk E, Kelley MJ (2005) Expression of HIF-1alpha, CA IX, VEGF, and MMP-9 in surgically resected non-small cell lung cancer. Lung Cancer 49:325–335 Kim SJ, Rabbani ZN, Vollmer RT, Schreiber EG, Oosterwijk E, Dewhirst MW, Vujaskovic Z, Kelley MJ (2004) Carbonic anhydrase IX in early-stage non-small cell lung cancer. Clin Cancer Res 10:7925–7933 Kim SJ, Shin HJ, Jung KY, Baek SK, Shin BK, Choi J, Kim BS, Shin SW, Kim YH, Kim JS, Oosterwijk E (2007) Prognostic value of carbonic anhydrase IX and Ki-67 expression in squamous cell carcinoma of the tongue. Jpn J Clin Oncol 37:812–819 Kirkpatrick JP, Rabbani ZN, Bentley RC, Hardee ME, Karol S, Meyer J, Oosterwijk E, Havrilesky L, Secord AA, Vujaskovic Z, Dewhirst MW, Jones EL (2008) Elevated CAIX expression is associated with an increased risk of distant failure in early-stage cervical cancer. Biomark Insights 3:45–55 Kivela AJ, Saarnio J, Karttunen TJ, Kivela J, Parkkila AK, Pastorekova S, Pastorek J, Waheed A, Sly WS, Parkkila TS, Rajaniemi H (2001) Differential expression of cytoplasmic carbonic
4 Molecular Mechanisms Regulating Expression
85
anhydrases, CA I and II, and membrane-associated isozymes, CA IX and XII, in normal mucosa of large intestine and in colorectal tumors. Dig Dis Sci 46:2179–2186 Klatte T, Seligson DB, Rao JY, Yu H, de Martino M, Kawaoka K, Wong SG, Belldegrun AS, Pantuck AJ (2009) Carbonic anhydrase IX in bladder cancer: a diagnostic, prognostic, and therapeutic molecular marker. Cancer 115:1448–1458 Kon-no H, Ishii G, Nagai K, Yoshida J, Nishimura M, Nara M, Fujii T, Murata Y, Miyamoto H, Ochiai A (2006) Carbonic anhydrase IX expression is associated with tumor progression and a poor prognosis of lung adenocarcinoma. Lung Cancer 54:409–418 Kopacek J, Barathova M, Dequiedt F, Sepelakova J, Kettmann R, Pastorek J, Pastorekova S (2005) MAPK pathway contributes to density- and hypoxia-induced expression of the tumorassociated carbonic anhydrase IX. Biochim Biophys Acta 1729:41–49 Korkeila E, Talvinen K, Jaakkola PM, Minn H, Syrjanen K, Sundstrom J, Pyrhonen S (2009) Expression of carbonic anhydrase IX suggests poor outcome in rectal cancer. Br J Cancer 100:874–880 Korkolopoulou P, Perdiki M, Thymara I, Boviatsis E, Agrogiannis G, Kotsiakis X, Angelidakis D, Rologis D, Diamantopoulou K, Thomas-Tsagli E, Kaklamanis L, Gatter K, Patsouris E (2007) Expression of hypoxia-related tissue factors in astrocytic gliomas. A multivariate survival study with emphasis upon carbonic anhydrase IX. Hum Pathol 38:629–638 Kowalewska M, Radziszewski J, Kulik J, Barathova M, Nasierowska-Guttmajer A, Bidzinski M, Pastorek J, Pastorekova S, Siedlecki JA (2005) Detection of carbonic anhydrase 9-expressing tumor cells in the lymph nodes of vulvar carcinoma patients by RT-PCR. Int J Cancer 116:957–962 Kozin SV, Shkarin P, Gerweck LE (2001) The cell transmembrane pH gradient in tumors enhances cytotoxicity of specific weak acid chemotherapeutics. Cancer Res 61:4740–4743 Kreis TE, Lodish HF (1986) Oligomerization is essential for transport of vesicular stomatitis viral glycoprotein to the cell surface. Cell 46:929–937 Kummola L, Hamalainen JM, Kivela J, Kivela AJ, Saarnio J, Karttunen T, Parkkila S (2005) Expression of a novel carbonic anhydrase, CA XIII, in normal and neoplastic colorectal mucosa. BMC Cancer 5:41 Kuo WH, Chiang WL, Yang SF, Yeh KT, Yeh CM, Hsieh YS, Chu SC (2003) The differential expression of cytosolic carbonic anhydrase in human hepatocellular carcinoma. Life Sci 73:2211–2223 Lam JS, Pantuck AJ, Belldegrun AS, Figlin RA (2005) G250: a carbonic anhydrase IX monoclonal antibody. Curr Oncol Rep 7:109–115 Lendahl U, Lee KL, Yang H, Poellinger L (2009) Generating specificity and diversity in the transcriptional response to hypoxia. Nat Rev Genet 10:821–832 Li Y, Wang H, Oosterwijk E, Selman Y, Mira JC, Medrano T, Shiverick KT, Frost SC (2009) Antibody-specific detection of CAIX in breast and prostate cancers. Biochem Biophys Res Commun 386:488–492 Liao SY, Aurelio ON, Jan K, Zavada J, Stanbridge EJ (1997) Identification of the MN/CA9 protein as a reliable diagnostic biomarker of clear cell carcinoma of the kidney. Cancer Res 57: 2827–2831 Liao SY, Brewer C, Zavada J, Pastorek J, Pastorekova S, Manetta A, Berman ML, DiSaia PJ, Stanbridge EJ (1994) Identification of the MN antigen as a diagnostic biomarker of cervical intraepithelial squamous and glandular neoplasia and cervical carcinomas. Am J Pathol 145: 598–609 Liao SY, Darcy KM, Randall LM, Tian C, Monk BJ, Burger RA, Fruehauf JP, Peters WA, Stock RJ, Stanbidge EJ (2010) Prognostic relevance of carbonic anhydrase-IX in high-risk, earlystage cervical cancer: A Gynecologic Oncology Group study. Gynecol Oncol 116:452–458 Loncaster JA, Harris AL, Davidson SE, Logue JP, Hunter RD, Wycoff CC, Pastorek J, Ratcliffe PJ, Stratford IJ, West CM (2001) Carbonic anhydrase (CA IX) expression, a potential new intrinsic marker of hypoxia: correlations with tumor oxygen measurements and prognosis in locally advanced carcinoma of the cervix. Cancer Res 61:6394–6399
86
J. Pastorek and S. Pastorekova
Malentacchi F, Simi L, Nannelli C, Andreani M, Janni A, Pastorekova S, Orlando C (2009) Alternative splicing variants of carbonic anhydrase IX in human non-small cell lung cancer. Lung Cancer 64:271–276 Mandriota SJ, Turner KJ, Davies DR, Murray PG, Morgan NV, Sowter HM, Wykoff CC, Maher ER, Harris AL, Ratcliffe PJ, Maxwell PH (2002) HIF activation identifies early lesions in VHL kidneys: evidence for site-specific tumor suppressor function in the nephron. Cancer Cell 1:459–468 Maresca A, Temperini C, Vu H, Pham NB, Poulsen SA, Scozzafava A, Quinn RJ, Supuran CT (2009) Non-zinc mediated inhibition of carbonic anhydrases: coumarins are a new class of suicide inhibitors (#). J Am Chem Soc 131:3057–3062 Maseide K, Kandel RA, Bell RS, Catton CN, O’Sullivan B, Wunder JS, Pintilie M, Hedley D, Hill RP (2004) Carbonic anhydrase IX as a marker for poor prognosis in soft tissue sarcoma. Clin Cancer Res 10:4464–4471 McKiernan JM, Buttyan R, Bander NH, Stifelman MD, Katz AE, Chen MW, Olsson CA, Sawczuk IS (1997) Expression of the tumor-associated gene MN: a potential biomarker for human renal cell carcinoma. Cancer Res 57:2362–2365 McMurtrie HL, Cleary HJ, Alvarez BV, Loiselle FB, Sterling D, Morgan PE, Johnson DE, Casey JR (2004) The bicarbonate transport metabolon. J Enzyme Inhib Med Chem 19:231–236 Morgan PE, Pastorekova S, Stuart-Tilley AK, Alper SL, Casey JR (2007) Interactions of transmembrane carbonic anhydrase, CAIX, with bicarbonate transporters. Am J Physiol Cell Physiol 293:C738–C748 Mori M, Staniunas RJ, Barnard GF, Jessup JM, Steele GD Jr, Chen LB (1993) The significance of carbonic anhydrase expression in human colorectal cancer. Gastroenterology 105: 820–826 Morsy SM, Badawi AM, Cecchi A, Scozzafava A, Supuran CT (2009) Carbonic anhydrase inhibitors. Biphenylsulfonamides with inhibitory action towards the transmembrane, tumor-associated isozymes IX possess cytotoxic activity against human colon, lung and breast cancer cell lines. J Enzyme Inhib Med Chem 24:499–505 Mulders P, Bleumer I, Debruyne F, Oosterwijk E (2004) Specific monoclonal antibody-based immunotherapy by targeting the RCC-associated antigen carbonic anhydrase-IX(G250/MN). Urologe A 43(Suppl 3):S146–S147 Nakao M, Ishii G, Nagai K, Kawase A, Kenmotsu H, Kon-No H, Hishida T, Nishimura M, Yoshida J, Ochiai A (2009) Prognostic significance of carbonic anhydrase IX expression by cancer-associated fibroblasts in lung adenocarcinoma. Cancer 115:2732–2743 Niemela AM, Hynninen P, Mecklin JP, Kuopio T, Kokko A, Aaltonen L, Parkkila AK, Pastorekova S, Pastorek J, Waheed A, Sly WS, Orntoft TF, Kruhoffer M, Haapasalo H, Parkkila S, Kivela AJ (2007) Carbonic anhydrase IX is highly expressed in hereditary nonpolyposis colorectal cancer. Cancer Epidemiol Biomarkers Prev 16:1760–1766 Oosterwijk E (2008a) Carbonic anhydrase IX/G250/MN: a molecule too good to be true? BJU Int 101:527–528 Oosterwijk E (2008b) Carbonic anhydrase IX: historical and future perspectives. BJU Int 101(Suppl 4):2–7 Opavsky R, Pastorekova S, Zelnik V, Gibadulinova A, Stanbridge EJ, Zavada J, Kettmann R, Pastorek J (1996) Human MN/CA9 gene, a novel member of the carbonic anhydrase family: structure and exon to protein domain relationships. Genomics 33:480–487 Pantuck AJ, Klatte T, Seligson D, Atkins M, Belldegrun A (2008) Carbonic anhydrase IX as a predictive biomarker for clear cell renal cell carcinoma. J Clin Oncol 26:3105–3107; author reply 3107–3109 Parkkila S, Innocenti A, Kallio H, Hilvo M, Scozzafava A, Supuran CT (2009) The protein tyrosine kinase inhibitors imatinib and nilotinib strongly inhibit several mammalian alphacarbonic anhydrase isoforms. Bioorg Med Chem Lett 19:4102–4106 Parkkila S, Rajaniemi H, Parkkila AK, Kivela J, Waheed A, Pastorekova S, Pastorek J, Sly WS (2000) Carbonic anhydrase inhibitor suppresses invasion of renal cancer cells in vitro. Proc Natl Acad Sci USA 97:2220–2224
4 Molecular Mechanisms Regulating Expression
87
Parkkila S, Lasota J, Fietcher JA, Ou WB, Kivelä AJ, Nuorva K, Parkkila AK, Ollikainen J, Sly WS, Waheed A, Pastorekova S, Pastorek J, Isola J, Miettinen M (2010). Carbonic anhydrase II. A novel biomarker for gastrointestinal stromal tumors. Mod Pathol 23:743–750 Pastorek J, Pastorekova S, Callebaut I, Mornon JP, Zelnik V, Opavsky R, Zat’ovicova M, Liao S, Portetelle D, Stanbridge EJ et al (1994) Cloning and characterization of MN, a human tumorassociated protein with a domain homologous to carbonic anhydrase and a putative helixloop-helix DNA binding segment. Oncogene 9:2877–2888 Pastorekova S, Casini A, Scozzafava A, Vullo D, Pastorek J, Supuran CT (2004a) Carbonic anhydrase inhibitors: the first selective, membrane-impermeant inhibitors targeting the tumorassociated isozyme IX. Bioorg Med Chem Lett 14:869–873 Pastorekova S, Parkkila S, Parkkila AK, Opavsky R, Zelnik V, Saarnio J, Pastorek J (1997) Carbonic anhydrase IX, MN/CA IX: analysis of stomach complementary DNA sequence and expression in human and rat alimentary tracts. Gastroenterology 112:398–408 Pastorekova S, Parkkila S, Pastorek J, Supuran CT (2004b) Carbonic anhydrases: current state of the art, therapeutic applications and future prospects. J Enzyme Inhib Med Chem 19: 199–229 Pastorekova S, Parkkila S, Zavada J (2006) Tumor-associated carbonic anhydrases and their clinical significance. Adv Clin Chem 42:167–216 Pastorekova S, Zatovicova M, Pastorek J (2008) Cancer-associated carbonic anhydrases and their inhibition. Curr Pharm Des 14:685–698 Pastorekova S, Zavadova Z, Kostal M, Babusikova O, Zavada J (1992) A novel quasi-viral agent, MaTu, is a two-component system. Virology 187:620–626 Peles E, Nativ M, Campbell PL, Sakurai T, Martinez R, Lev S, Clary DO, Schilling J, Barnea G, Plowman GD, Grumet M, Schlessinger J (1995) The carbonic anhydrase domain of receptor tyrosine phosphatase beta is a functional ligand for the axonal cell recognition molecule contactin. Cell 82:251–260 Potter C, Harris AL (2004) Hypoxia inducible carbonic anhydrase IX, marker of tumour hypoxia, survival pathway and therapy target. Cell Cycle 3:164–167 Pouyssegur J, Dayan F, Mazure NM (2006) Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature 441:437–443 Rafajova M, Zatovicova M, Kettmann R, Pastorek J, Pastorekova S (2004) Induction by hypoxia combined with low glucose or low bicarbonate and high posttranslational stability upon reoxygenation contribute to carbonic anhydrase IX expression in cancer cells. Int J Oncol 24: 995–1004 Raghunand N, Gatenby RA, Gillies RJ (2003) Microenvironmental and cellular consequences of altered blood flow in tumours. Br J Radiol 76 Spec No 1:S11–S22 Raval RR, Lau KW, Tran MG, Sowter HM, Mandriota SJ, Li JL, Pugh CW, Maxwell PH, Harris AL, Ratcliffe PJ (2005) Contrasting properties of hypoxia-inducible factor 1 (HIF-1) and HIF-2 in von Hippel-Lindau-associated renal cell carcinoma. Mol Cell Biol 25:5675–5686 Ruas JL, Poellinger L (2005) Hypoxia-dependent activation of HIF into a transcriptional regulator. Semin Cell Dev Biol 16:514–522 Saarnio J, Parkkila S, Parkkila AK, Haukipuro K, Pastorekova S, Pastorek J, Kairaluoma MI, Karttunen TJ (1998) Immunohistochemical study of colorectal tumors for expression of a novel transmembrane carbonic anhydrase, MN/CA IX, with potential value as a marker of cell proliferation. Am J Pathol 153:279–285 Saczewski F, Slawinski J, Kornicka A, Brzozowski Z, Pomarnacka E, Innocenti A, Scozzafava A, Supuran CT (2006) Carbonic anhydrase inhibitors. Inhibition of the cytosolic human isozymes I and II, and the transmembrane, tumor-associated isozymes IX and XII with substituted aromatic sulfonamides activatable in hypoxic tumors. Bioorg Med Chem Lett 16: 4846–4851 Sansone P, Storci G, Giovannini C, Pandolfi S, Pianetti S, Taffurelli M, Santini D, Ceccarelli C, Chieco P, Bonafe M (2007) p66Shc/Notch-3 interplay controls self-renewal and hypoxia survival in human stem/progenitor cells of the mammary gland expanded in vitro as mammospheres. Stem Cells 25:807–815
88
J. Pastorek and S. Pastorekova
Scott FL, Stec B, Pop C, Dobaczewska MK, Lee JJ, Monosov E, Robinson H, Salvesen GS, Schwarzenbacher R, Riedl SJ (2009) The Fas-FADD death domain complex structure unravels signalling by receptor clustering. Nature 457:1019–1022 Shin KH, Diaz-Gonzalez JA, Russell J, Chen Q, Burgman P, Li XF, Ling CC (2007) Detecting changes in tumor hypoxia with carbonic anhydrase IX and pimonidazole. Cancer Biol Ther 6:70–75 Schrijvers ML, van der Laan BF, de Bock GH, Pattje WJ, Mastik MF, Menkema L, Langendijk JA, Kluin PM, Schuuring E, van der Wal JE (2008) Overexpression of intrinsic hypoxia markers HIF1alpha and CA-IX predict for local recurrence in stage T1-T2 glottic laryngeal carcinoma treated with radiotherapy. Int J Radiat Oncol Biol Phys 72:161–169 Signoretti S, Regan M, Atkins M (2008) Carbonic anhydrase IX as a predictive biomarker of response to kidney cancer therapy. BJU Int 101(Suppl 4):31–35 Smaine FZ, Winum JY, Montero JL, Regainia Z, Vullo D, Scozzafava A, Supuran CT (2007) Carbonic anhydrase inhibitors: Selective inhibition of the extracellular, tumor-associated isoforms IX and XII over isozymes I and II with glycosyl-thioureido-sulfonamides. Bioorg Med Chem Lett 17:5096–5100 Smyth LG, O’Hurley G, O’Grady A, Fitzpatrick JM, Kay E, Watson RW (2010) Carbonic anhydrase IX expression in prostate cancer. Prostate Cancer Prostatic Dis 13:178–181 Stiti M, Cecchi A, Rami M, Abdaoui M, Barragan-Montero V, Scozzafava A, Guari Y, Winum JY, Supuran CT (2008) Carbonic anhydrase inhibitor coated gold nanoparticles selectively inhibit the tumor-associated isoform IX over the cytosolic isozymes I and II. J Am Chem Soc 130: 16130–16131 Stock C, Schwab A (2009) Protons make tumor cells move like clockwork. Pflugers Arch 458:981–992 Stubbs M, McSheehy PM, Griffiths JR, Bashford CL (2000) Causes and consequences of tumour acidity and implications for treatment. Mol Med Today 6:15–19 Supuran CT (2007) Carbonic anhydrases as drug targets – an overview. Curr Top Med Chem 7:825–833 Supuran CT (2008) Carbonic anhydrases: novel therapeutic applications for inhibitors and activators. Nat Rev Drug Discov 7:168–181 Svastova E, Hulikova A, Rafajova M, Zat’ovicova M, Gibadulinova A, Casini A, Cecchi A, Scozzafava A, Supuran CT, Pastorek J, Pastorekova S (2004) Hypoxia activates the capacity of tumor-associated carbonic anhydrase IX to acidify extracellular pH. FEBS Lett 577: 439–445 Svastova E, Zilka N, Zat’ovicova M, Gibadulinova A, Ciampor F, Pastorek J, Pastorekova S (2003) Carbonic anhydrase IX reduces E-cadherin-mediated adhesion of MDCK cells via interaction with beta-catenin. Exp Cell Res 290:332–345 Swietach P, Patiar S, Supuran CT, Harris AL, Vaughan-Jones RD (2009) The role of carbonic anhydrase 9 in regulating extracellular and intracellular ph in three-dimensional tumor cell growths. J Biol Chem 284:20299–20310 Swietach P, Wigfield S, Supuran CT, Harris AL, Vaughan-Jones RD (2008) Cancer-associated, hypoxia-inducible carbonic anhydrase IX facilitates CO2 diffusion. BJU Int 101 (Suppl 4): 22–4 Swinson DE, Jones JL, Richardson D, Wykoff C, Turley H, Pastorek J, Taub N, Harris AL, O’Byrne KJ (2003) Carbonic anhydrase IX expression, a novel surrogate marker of tumor hypoxia, is associated with a poor prognosis in non-small-cell lung cancer. J Clin Oncol 21: 473–482 Takacova M, Holotnakova T, Vondracek J, Machala M, Pencikova K, Gradin K, Poellinger L, Pastorek J, Pastorekova S, Kopacek J (2009) Role of aryl hydrocarbon receptor in modulation of the expression of the hypoxia marker carbonic anhydrase IX. Biochem J 419: 419–425 Takacova M, Holotnakova T, Barathova M, Pastorekova S, Kopacek J, Pastorek J (2010). Src induces expression of carbonic anhydrase IX via hypoxia-inducible factor 1. Oncol Rep 23:869–874
4 Molecular Mechanisms Regulating Expression
89
Tan EY, Yan M, Campo L, Han C, Takano E, Turley H, Candiloro I, Pezzella F, Gatter KC, Millar EK, O’Toole SA, McNeil CM, Crea P, Segara D, Sutherland RL, Harris AL, Fox SB (2009) The key hypoxia regulated gene CAIX is upregulated in basal-like breast tumours and is associated with resistance to chemotherapy. Br J Cancer 100:405–411 Teicher BA (2009) Acute and chronic in vivo therapeutic resistance. Biochem Pharmacol 77: 1665–1673 Teicher BA, Liu SD, Liu JT, Holden SA, Herman TS (1993) A carbonic anhydrase inhibitor as a potential modulator of cancer therapies. Anticancer Res 13:1549–1556 Temperini C, Innocenti A, Mastrolorenzo A, Scozzafava A, Supuran CT (2007) Carbonic anhydrase inhibitors. Interaction of the antiepileptic drug sulthiame with twelve mammalian isoforms: kinetic and X-ray crystallographic studies. Bioorg Med Chem Lett 17:4866–4872 Trastour C, Benizri E, Ettore F, Ramaioli A, Chamorey E, Pouyssegur J, Berra E (2007) HIF1alpha and CA IX staining in invasive breast carcinomas: prognosis and treatment outcome. Int J Cancer 120:1451–1458 Turner JR, Odze RD, Crum CP, Resnick MB (1997) MN antigen expression in normal, preneoplastic, and neoplastic esophagus: a clinicopathological study of a new cancer-associated biomarker. Hum Pathol 28:740–744 van den Beucken T, Koritzinsky M, Niessen H, Dubois L, Savelkouls K, Mujcic H, Jutten B, Kopacek J, Pastorekova S, van der Kogel AJ, Lambin P, Voncken W, Rouschop KM, Wouters BG (2009a) Hypoxia-induced expression of carbonic anhydrase 9 is dependent on the unfolded protein response. J Biol Chem 284:24204–24212 van den Beucken T, Ramaekers CH, Rouschop K, Koritzinsky M, Wouters BG (2009b) Deficient carbonic anhydrase 9 expression in UPR-impaired cells is associated with reduced survival in an acidic microenvironment. Radiother Oncol 92:437–442 Vermylen P, Roufosse C, Burny A, Verhest A, Bosschaerts T, Pastorekova S, Ninane V, Sculier JP (1999) Carbonic anhydrase IX antigen differentiates between preneoplastic malignant lesions in non-small cell lung carcinoma. Eur Respir J 14:806–811 Vleugel MM, Greijer AE, Shvarts A, van der Groep P, van Berkel M, Aarbodem Y, van Tinteren H, Harris AL, van Diest PJ, van der Wall E (2005) Differential prognostic impact of hypoxia induced and diffuse HIF-1alpha expression in invasive breast cancer. J Clin Pathol 58: 172–177 Vukovic V, Tannock IF (1997) Influence of low pH on cytotoxicity of paclitaxel, mitoxantrone and topotecan. Br J Cancer 75:1167–1172 Vullo D, Franchi M, Gallori E, Pastorek J, Scozzafava A, Pastorekova S, Supuran CT (2003) Carbonic anhydrase inhibitors: inhibition of the tumor-associated isozyme IX with aromatic and heterocyclic sulfonamides. Bioorg Med Chem Lett 13:1005–1009 Vullo D, Scozzafava A, Pastorekova S, Pastorek J, Supuran CT (2004) Carbonic anhydrase inhibitors: inhibition of the tumor-associated isozyme IX with fluorine-containing sulfonamides. The first subnanomolar CA IX inhibitor discovered. Bioorg Med Chem Lett 14: 2351–2356 Weber A, Casini A, Heine A, Kuhn D, Supuran CT, Scozzafava A, Klebe G (2004) Unexpected nanomolar inhibition of carbonic anhydrase by COX-2-selective celecoxib: new pharmacological opportunities due to related binding site recognition. J Med Chem 47:550–557 Wilkinson BL, Bornaghi LF, Houston TA, Innocenti A, Supuran CT, Poulsen SA (2006) A novel class of carbonic anhydrase inhibitors: glycoconjugate benzene sulfonamides prepared by “click-tailing”. J Med Chem 49:6539–6548 Willam C, Warnecke C, Schefold JC, Kugler J, Koehne P, Frei U, Wiesener M, Eckardt KU (2006) Inconsistent effects of acidosis on HIF-alpha protein and its target genes. Pflugers Arch 451: 534–543 Winum JY, Cecchi A, Montero JL, Innocenti A, Scozzafava A, Supuran CT (2005a) Carbonic anhydrase inhibitors. Synthesis and inhibition of cytosolic/tumor-associated carbonic anhydrase isozymes I, II, and IX with boron-containing sulfonamides, sulfamides, and sulfamates: toward agents for boron neutron capture therapy of hypoxic tumors. Bioorg Med Chem Lett 15:3302–3306
90
J. Pastorek and S. Pastorekova
Winum JY, Dogne JM, Casini A, de Leval X, Montero JL, Scozzafava A, Vullo D, Innocenti A, Supuran CT (2005b) Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/ membrane-associated carbonic anhydrase isozymes I, II, and IX with sulfonamides incorporating hydrazino moieties. J Med Chem 48:2121–2125 Winum JY, Pastorekova S, Jakubickova L, Montero JL, Scozzafava A, Pastorek J, Vullo D, Innocenti A, Supuran CT (2005c) Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/tumor-associated carbonic anhydrase isozymes I, II, and IX with bis-sulfamates. Bioorg Med Chem Lett 15:579–584 Winum JY, Vullo D, Casini A, Montero JL, Scozzafava A, Supuran CT (2003a) Carbonic anhydrase inhibitors. Inhibition of cytosolic isozymes I and II and transmembrane, tumor-associated isozyme IX with sulfamates including EMATE also acting as steroid sulfatase inhibitors. J Med Chem 46:2197–2204 Winum JY, Vullo D, Casini A, Montero JL, Scozzafava A, Supuran CT (2003b) Carbonic anhydrase inhibitors: inhibition of transmembrane, tumor-associated isozyme IX, and cytosolic isozymes I and II with aliphatic sulfamates. J Med Chem 46:5471–5477 Wykoff CC, Beasley NJ, Watson PH, Turner KJ, Pastorek J, Sibtain A, Wilson GD, Turley H, Talks KL, Maxwell PH, Pugh CW, Ratcliffe PJ, Harris AL (2000) Hypoxia-inducible expression of tumor-associated carbonic anhydrases. Cancer Res 60:7075–7083 Yin Q, Lin SC, Lamothe B, Lu M, Lo YC, Hura G, Zheng L, Rich RL, Campos AD, Myszka DG, Lenardo MJ, Darnay BG, Wu H (2009) E2 interaction and dimerization in the crystal structure of TRAF6. Nat Struct Mol Biol 16:658–666 Yoshiura K, Nakaoka T, Nishishita T, Sato K, Yamamoto A, Shimada S, Saida T, Kawakami Y, Takahashi TA, Fukuda H, Imajoh-Ohmi S, Oyaizu N, Yamashita N (2005) Carbonic anhydrase II is a tumor vessel endothelium-associated antigen targeted by dendritic cell therapy. Clin Cancer Res 11:8201–8207 Zat’ovicova M, Tarabkova K, Svastova E, Gibadulinova A, Mucha V, Jakubickova L, Biesova Z, Rafajova M, Ortova Gut M, Parkkila S, Parkkila AK, Waheed A, Sly WS, Horak I, Pastorek J, Pastorekova S (2003) Monoclonal antibodies generated in carbonic anhydrase IX-deficient mice recognize different domains of tumour-associated hypoxia-induced carbonic anhydrase IX. J Immunol Methods 282:117–134 Zatovicova M, Sedlakova O, Svastova E, Ohradanova A, Ciampor F, Arribas J, Pastorek J, Pastorekova S (2005) Ectodomain shedding of the hypoxia-induced carbonic anhydrase IX is a metalloprotease-dependent process regulated by TACE/ADAM17. Br J Cancer 93:1267–1276 Zatovicova M, Jeleuska L, Hulikova A, Csaderova L, Ditte P, Goliasova T, Pastorek J, Pastorekova S (in press). Carbonic anhydrase IX as an anticancer target: preclinical evaluation of internalizing monoclonal antibody directed to catalytic domain. Curr Pharm Design Zavada J, Zavadova Z, Pastorek J, Biesova Z, Jezek J, Velek J (2000) Human tumour-associated cell adhesion protein MN/CA IX: identification of M75 epitope and of the region mediating cell adhesion. Br J Cancer 82:1808–1813 Zavada J, Zavadova Z, Pastorekova S, Ciampor F, Pastorek J, Zelnik V (1993) Expression of MaTu-MN protein in human tumor cultures and in clinical specimens. Int J Cancer 54: 268–274 Zavada J, Zavadova Z, Zat’ovicova M, Hyrsl L, Kawaciuk I (2003) Soluble form of carbonic anhydrase IX (CA IX) in the serum and urine of renal carcinoma patients. Br J Cancer 89: 1067–1071 Zhang J (2008) Recent advances in preoperative imaging of renal tumors. Curr Opin Urol 18: 111–115 Zheng J, Avvaru BS, Tu C, McKenna R, Silverman DN (2008) Role of hydrophilic residues in proton transfer during catalysis by human carbonic anhydrase II. Biochemistry 47: 12028–12036 Zhou GX, Ireland J, Rayman P, Finke J, Zhou M (2010) Quantification of carbonic anhydrase IX expression in serum and tissue of renal cell carcinoma patients using enzyme-linked immunosorbent assay: prognostic and diagnostic potentials. Urology 75:257–261
Chapter 5
Glycolytic Pathway as a Target for Tumor Inhibition Weiqin Lu and Peng Huang
Abstract It has been known for decades that cancer cells exhibit elevated aerobic glycolysis, a phenomenon first observed by Otto Warburg. The imaging technique 18 fluoro-deoxyglucose-positron emission tomography (FDG-PET) widely used in clinical diagnosis of cancer is based on the increased glucose uptake by cancer cells, likely due to a significant increase in glucose flow into the glycolytic pathway to generate ATP and conversion to other metabolic intermediates and building blocks for cell growth and proliferation. Furthermore, the increased glucose uptake in cancer tissues seems correlated with tumor aggressiveness and poor prognosis. Such profound metabolic alterations suggest that cancer cells may prefer to use glycolysis for their proliferative and survival advantage. Recent studies have begun to elucidate the molecular mechanisms underlying these metabolic alterations, and provide important new insights into the mechanistic links between oncogenic signals and metabolic regulation. Importantly, targeted inhibition of glycolysis and its regulatory pathways may provide exciting opportunities for the development of therapeutic strategies to preferentially kill cancer cells. This chapter will summarize our current understanding of glycolytic alterations in cancer cells and the relevant regulatory mechanisms, and discuss possible therapeutic strategies that exploit the metabolic abnormalities in cancer to preferentially kill malignant cells.
Introduction Glucose is a major source of fuel for ATP generation. It can be converted into two molecules of pyruvate in the cytosol with a net gain of 2 ATP and a reduction of 2 NAD+ to 2 NADH per glucose through a sequence of reactions known as glycolysis.
P. Huang (*) Department of Molecular Pathology, The University of Texas MD Anderson Cancer Center, 1515 Holcombe Boulevard, Houston, TX 77030, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_5, © Springer Science+Business Media, LLC 2010
91
92
W. Lu and P. Huang Glucose GLUT
Glucose
ROS
HK
PPP (oxidative)
G6P PFK2
NADPH anabolic process
G6PD
GPI
Ribose-5-P
F6P
F2,6-BP
TKL/TAL
PFK1
F1,6-BP Aldolase
Dihydroxyacetone-P
Nucleotide synthesis
PPP (non-oxidative) GA3P
Mitochondria
GAPDH
1,3-BPG 3-PG 2-PG
PDH Enolase
PEP
ATP Succinate
OAA
TCA cycle SDH
Pyr PK
Malate Pyr
Lactate
NADH
Ac-CoA
PGM
ATP
MRC
NAD+ Citrate
PGK
Fumarate Fumerase
LDH
MCT
Lactate
Fig. 5.1 Overview of glycolysis, PPP, and the TCA cycle. Glucose is metabolized through the glycolytic pathway and the TCA cycle in the mitochondria in conjunction with mitochondrial respiratory to generate ATP. Glucose can also be metabolized through the PPP (in dashed line) to generate ribose-5-phosphate (ribose-5-P) as a precursor for the synthesis of nucleic acids, and NADPH for anabolic synthesis of fat acids or as a reducing equivalent to combat ROS. Notes: Dihydroxyacetone-P, Dihydroxyacetone-phosphate; F6P, fructose-6-phosphate; F1,6-BP, fructose1,6-bisphosphate; G6PD, glucose-6-phosphate dehydrogenase; GA3P, glyceraldehyde-3-phosphate; G6P, glucose-6-phoaphate; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; GLUT, glucose transporter; GPI, glucose-phosphate isomerase; LDH, lactate dehydrogenase; HK, hexokinases; MCT, monocarboxylate transporter; MRC, mitochondrial respiratory chain (complexes I-V); PFK1, phosphofructokinase 1; PFK2, phosphofructokinase 2; PGK, phosphoglycerate kinase; PGM, phosphoglycerate mutase; PK, pyruvate kinase; Pyr, pyruvate; Ribose-5-P, ribose-5-phosphate; Ac-CoA, acetyl-CoA; OAA, oxaloacetate; PDH, pyruvate dehydrogenase; PPP, pentose phosphate pathway; ROS, reactive oxygen species; SDH, succinate dehydrogenase; TAL, transaldolases; TKL, transketolase
As illustrated in Fig. 5.1, glucose can be actively transported from blood stream into cells by Na+-dependent glucose specific transporters (GLUTs). Once it enters the cell, glucose is phosphorylated into glucose-6-phosphate (G6P) by a family of enzymes called hexokinases (HKs), the first rate-limiting enzyme of the glycolysis. G6P is a key metabolic intermediate at the branching point of two major metabolic pathways, namely glycolysis and pentose phosphate pathway (PPP). G6P can be
5 Glycolytic Pathway as a Target for Tumor Inhibition
93
dehydrogenated by glucose 6-phosphate dehydrogenase (G6PD) in the PPP to generate ribose 5-phosphate and NADPH when the cellular ratio of NADP+/ NADPH is increased. If cellular ATP level decreases, G6P will be actively converted to fructose-6-phosphate (F6P) by glucose phosphate isomerase (GPI) and thus favoring the glycolytic pathway. With ATP energy expenditure, F6P is then converted to fructose-1,6-bisphosphate (F1,6-BP) catalyzed by another rate-limiting enzyme phosphofructokinase 1 (PFK1). This reaction is energetically favored. The glycolytic process appears irreversible beyond this point, thus this is the key regulatory step in glycolysis. Aldolase further splits F1,6-BP into two triose sugars, dihydroxyacetone phosphate and glyceraldehyde-3-phosphate (GA3P). The former can be converted to GA3P by triosephosphate isomerase and proceeds further into glycolysis. GA3P is then dehydrogenated to 1, 3-bisphosphoglycerate (1,3-BPG) by an oxidoreductase glyceraldehyde dehydrogenase (GAPDH). This step requires NAD+ as a hydrogen carrier accepting H+ to generate NADH. Thus, the presence of sufficient amount of NAD+ is important to support a sustained level of glycolysis. 1,3BPG is then further metabolized by phosphoglycerate kinase (PGK) to 3-phosphoglycerate (3-PG), which is converted to 2-phosphoglycerate (2-PG) by phosphoglycerate mutase (PGM) and further converted to phosphoenolpyruvate (PEP) by enolase. Finally, pyruvate is produced from PEP by pyruvate kinase, one of the three key rate-limiting enzymes in glycolysis. If pyruvate is not further utilized in the mitochondria, it will be converted into lactate by lactate dehydrogenase (LDH) in the cytosol, generating NAD+ as substrate for the GAPDH-catalyzed reaction to allow continued glycolysis. NAD+/NADH ratio is important in regulating LDH activity. Because cells can obtain ATP supply through the oxygen-independent glycolytic pathway in cytosol and the oxygen-dependent oxidative phosphorylation (OXPHOS) pathway in mitochondria, the availability of oxygen and the mitochondrial functional status of the cells play important roles in determining the proportion of each pathway in ATP generation. In the presence of oxygen and functional mitochondria, pyruvate can be transported into the mitochondria and decarboxylated into acetyl CoA by the pyruvate dehydrogenase (PDH) complex in the mitochondria. The oxidation of acetyl CoA is achieved by the tricarboxylic acid (TCA) cycle generating CO2, H2O, GTP and other bioenergetic products including NADH and FADH2 (Krebs 1970a,b). The electrons from NADH and FADH2 are transported through the respiratory chain complexes (I–IV) and donated to oxygen to generate water, and the proton gradient generated through this process is used as the energy source for ATP production in complex V, where the ATP generation is highly efficient (Mitchell and Moyle 1967; Skulachev 1994). Under normal physiological conditions, mitochondria play an important role in OXPHOS generating most of the ATP for cellular function (Newmeyer and Ferguson-Miller 2003). However, in cancer cells the glucose metabolism seems significantly altered, and such alterations may serve as a biochemical basis to selectively kill malignant cells. A major goal of chemotherapy is to selectively kill cancer cells without significant toxicity to normal cells. Tremendous progress has been made recently in our
94
W. Lu and P. Huang
understanding of the biochemical and metabolic characteristics of cancer cells and the molecular mechanisms that govern cancer development and progression. It has been recognized that cancer cells exhibit a high degree of heterogeneity, adaptability, and genetic instability, making it difficult to eliminate the entire cancer cell population using target-specific agents. However, despite the different genetic or epigenetic alterations in cancer cells, aerobic glycolysis seems to be a common feature of the malignant cells as originally observed by Otto Warburg (1956a). Compared to normal adult cells, which mainly use OXPHOS to generate ATP, cancer cells have a higher tendency to generate ATP by converting glucose to lactate through the glycolytic pathway even in the presence of adequate oxygen (an aerobic glycolytic process known as the Warburg effect). It was estimated that more than 90% of metastatic tumors tested are highly glycolytic (Czernin and Phelps 2002). Such widespread metabolic alterations in cancer suggests that the malignant cells may prefer to use aerobic glycolysis rather than the highly efficient OXPHOS pathway to generate energy for their survival advantages, and perhaps facilitating tumor invasion and metastasis through acidification of the tumor microenvironment that in turn select the cancer cells with more aggressive malignant behaviors (Gatenby and Gawlinski 1996; Gatenby et al. 2006; Schornack and Gillies 2003; Wallace 2005). The increase in glucose uptake by cancer cells has been used as a metabolic basis in FDG-PET scan, which uses the glucose analogue 18 F-deoxyglucose as a tracer, for diagnosis of cancer and monitoring of tumor progression or therapeutic response (Czernin and Phelps 2002; Seemann 2004). The sustained activation of glycolysis observed in cancer cells may play a crucial role in cancer cell transformation and promotion of malignant cell behaviors (Gatenby 1995). With further understanding of the tumor micro-environment related to hypoxia and hypoxia-inducible factor (HIF) function, and the roles of Myc, Akt, Ras, p53, and AMPK in the glycolytic regulation, the metabolic diffe rences between cancer and normal cells have gained broad attention recently, and study in this area is likely to have profound influence on our understanding on cancer biology and cancer therapeutics.
Alterations of Glucose Metabolism in Cancer More than 80 years ago, Warburg pioneered the research on mitochondrial respiratory chain defects in the context of cancer, and was the first to recognize that tumors are highly glycolytic. His decades of work in this area was perhaps best illustrated in his landmark publication in Science more than 50 years ago (Warburg 1956b). It is estimated that over 50% of ATP in many tumor cells is derived from glycolysis to meet the needs for rapid cell growth and proliferation, although some of the highly proliferating tumor cells do not have obvious defects in oxidative metabolism. FDG-PET scan of tumor tissues has unequivocally demonstrated the increase in glucose uptake as a hallmark of human cancer, showing a profound alteration of glucose metabolism in tumors (Haberkorn et al. 1991; Weber 2006).
5 Glycolytic Pathway as a Target for Tumor Inhibition
95
It should be pointed out, however, that an increase in glucose uptake dose not necessarily indicate that this increase is solely due to increased glycolysis. The increase in glycolysis in cancer cells is likely to provide proliferative and survival advantages. In addition to ATP generation, the glycolysis also produces important metabolic intermediates for the biosynthesis of macromolecules and avoids ROS generation. Some of the glycolytic enzymes also have anti-apoptotic function. The metabolic switch or “reprogramming” from aerobic mitochondrial metabolism to glycolysis in tumor cells might be achieved through a number of putative mechanisms, including up-regulation of the glycolytic enzymes, suppression of the mitochondrial function due to mitochondrial mutation and/or defects in respiratory chain function at various points of the electron transport complexes, decreased pyruvate utilization by mitochondria, impaired TCA cycle, activation of oncogenes, mutation or loss of tumor suppressors, and adaptation to hypoxic tumor environment. Mitochondrial dysfunction may promote cancer cell survival and drug resistance through alterations in cell survival pathways such as activation of the Akt pathway and attenuation of the mitochondria-mediated apoptotic process (Carew and Huang 2002; Ohta 2006).
Overexpression of Glycolytic Enzymes in Cancer Favors Aerobic Glycolysis A large body of evidence suggests that tumor cells have enhanced enzyme activities or/and increased transcription of genes of glycolytic enzymes such as HKII, PFKs, LDH-A, and GLUTs when compared with non-tumorigenic cells (Brown et al. 2002; Mathupala et al. 1997; Minchenko et al. 2002; Smith 2000). Interestingly, some of the elevated glycolytic enzymes were found to have anti-apoptotic function and may confer resistance to anticancer agents. GLUT1 The entry of glucose into the cells relies on a family of membrane transporters that facilitate the cellular uptake of glucose. GLUTs are an important family of membraneassociated proteins with isoform-specific tissue distribution, some of which are over-expressed in cancer cells. Among the 12 isoforms of GLUTs (GLUT1-GLUT12), GLUT1, 2, and 5 are found to be expressed in tumor cells (Godoy et al. 2006). GLUT1, which transfers glucose from bloodstream into the cells, seems to play an important role in tumor development. It has a high affinity for glucose and transports glucose at a high rate. Compared to normal tissues, tumor tissues having higher 18fluoro-deoxyglucose (FDG) uptake by PET scan are associated with a higher GLUT1 expression. HIF-1 was found to bind to the promoter of GLUT1 in human HepG2 cells to stimulate glucose uptake, and deregulation of GLUT1 by c-Myc was also reported (Dombrauckas et al. 2005). Activated AKT can also
96
W. Lu and P. Huang
enhance GLUT function. Thus, it is likely that GLUT1 up-regulation and increased glucose uptake in cancer are associated with oncogene regulation. HKII HKs, the first rate-limiting enzymes in glycolysis, catalyze the conversion of glucose to G6P. There are four isoforms of HKs with different tissue expression patterns and regulation. Besides catalyzing the phosphorylation of glucose, HKII has been suggested to promote cell survival by preventing apoptosis through association with the mitochondrial permeability transition pore component. It has been shown that HKII may interact with the voltage-dependent anion channel (VDAC) to induce permeability transition pore closure, and prevents BAX/BAK oligomerization on the mitochondrial membrane, blocking cytochrome c release from the mitochondria (GolshaniHebroni and Bessman 1997; Mathupala et al. 2006). As such, the mitochondriaassociated HK (HKII) is regarded as a guardian of the mitochondria to prevent apoptosis (Robey and Hay 2005). Enhanced transcription of HKII occurs in many cancers, especially in breast cancer, suggesting a possibility of HKII promoter activation during tumorigenesis (Kim and Dang 2006). The promoter region of HKII has putative binding sites for p53, HIF, and c-Myc. Interestingly, p53 was found to downregulate HKII while c-Myc and HIF could upregulate HKII mRNA transcription (Mathupala et al. 1997; Pastorino et al. 2005). Therefore, mutation or loss of p53, activation of c-Myc, and/or stabilization of HIF may lead to upregulation of HKII expression and result in hyperpolarization of the mitochondria and suppression of apoptosis. AKT activation can also signal HKII to bind to the mitochondrial membrane and remove the G6P inhibitory regulation (Majewski et al. 2004). Thus, it is conceivable that increased HKII due to oncogene activation or tumor suppressor dysfunction can stimulate glycolysis and promote cancer cell survival and drug resistance. Previous study in mitochondrial defective cells (r0 cells) showed that inhibition of glycolysis using the HKII inhibitor 3-bromopyruvate was effective in killing cancer cells with high glycolytic activity or under hypoxic conditions (Xu et al. 2005b). PFK1 PFK1, a rate-limiting enzyme under a tight regulation during glycolysis, catalyzes the phosphorylation of F6P to F1,6BP. Interestingly, fructose-2,6-biphosphate (F2,6,BP) is an extremely potent allosteric activator of PFK1 and thus exerts a major regulatory control over the rate of glycolysis and glucose utilization. The enzyme 6-phosphofructo-2-kinase/fructose-2,6-biphosphase (PFKBF) catalyzing the reversible reaction to produce F2,6,BP plays an essential role in the regulation of glycolysis. F1,6-BP level and PFK1 activity are significantly increased in most tumors (Sanchez-Martinez and Aragon 1997). Through the transcriptional regulation of TIGAR (TP53-induced glycolysis and apoptosis regulator), p53 inhibits F2,6,BP production and therefore lowers PFK1 activity and downregulates
5 Glycolytic Pathway as a Target for Tumor Inhibition
97
g lycolysis (Bensaad et al. 2006). AKT was shown to phosphorylate and activate PFK1 to upregulate glycolysis (Rathmell et al. 2003). Several other oncoproteins such as Myc and Ras have been reported to activate PFK1 in immortalized cells (Kole et al. 1991; Osthus et al. 2000). PKM2 Pyruvate kinase, another key rate-limiting enzyme in glycolysis, has four isoforms (L, R, M1, and M2). Proliferating cancer cells of various histological types seem to mainly express the human embryonic M2 isoform PKM2 (Altenberg and Greulich 2004). When PKM2 forms a highly active tetramer, this enzyme favors the formation of pyruvate and ATP. However, when PKM2 is in a less active dimer form, it impairs pyruvate formation and causes an accumulation of the upstream intermediate F1,6BP. PKM2 can selectively bind to tyrosine-phosphorylated peptides, resulting in the release of its allosteric activator F1,6BP and thus induces a diversion of glucose metabolites from energy production to the non-oxidative PPP pathway leading to anabolic processes. A recent study showed that a knockdown of PKM2 in tumor cells by shRNAs and a replacement with isoform PKM1 led to reduced lactate production, increased oxygen consumption, and reduced tumor formation in a mouse xenograft model (Christofk et al. 2008). In proliferating cancer cells, expression of PKM2 enhances the cells’ ability to use glucose for anabolic processes and this metabolic phenotype seems to provide a selective growth advantage for the tumor cells (Vander Heiden et al. 2009). LDH-A LDH catalyzes the conversion of pyruvate to lactate and provides NAD+ to support glycolysis. LDH-A is an enzyme that favors the conversion of pyruvate to lactate in anaerobic condition, while LDH-B seems to favor the conversion of lactate to pyruvate in aerobic condition. Since NAD+ is required for the glycolytic reaction catalyzed by GAPDH, regeneration of NAD+ by LDH is important to maintain glycolysis, especially when mitochondrial respiration is compromised. It is reported that cancer cells produce more lactate and have increased LDH-A activity than normal cells (Hove et al. 1993; Schwickert et al. 1995). LDH-A plays an important role in cancer cell survival, invasion, and metastasis not only by regenerating NAD+ to maintain glycolysis but also by causing cellular acidosis due to the generation of the metabolic product lactate, which may exit cancer cells through a lactate carrier monocarboxylate transporter 4 (MCT4), and induce tissue acidosis. LDH-A gene transcription can be induced by hypoxia through the activation of HIF-1. Interestingly, MCT4 also appears to be a downstream target of HIF-1. Lactate was reported to stimulate HIF-1 expression regardless of hypoxia and lactate-induced acidosis often precedes angiogenesis (Lu et al. 2002; Shi et al. 2001). Thus, through transcriptional activation of LDH-A and MCT4, HIF1 stimulates glycolysis and lactate exit out of the cells. Lactate generation
98
W. Lu and P. Huang
due to an increased LDH-A activation in turn stimulates HIF1 expression and creates a cycle of sustained enhancement of glycolysis. LDH-A is also a transcriptional target of c-Myc and its overexpression is required for c-Myc-mediated transformation, since lowering its expression by LDH-A-antisense reduces the transformation capacity of c-Myc (Shim et al. 1997). A recent study showed that a knockdown of LDH-A in tumor cells by shRNAs led to a stimulation of mitochondrial OXPHOS, a decrease of cell proliferation under hypoxia, and a suppression of tumorigenicity (Fantin et al. 2006). Acidification of tumor microenvironment by lactate may facilitate a selection of more malignant cancer cells with highly aggressive behaviors, and thus promote tumor development, invasion, and metastasis (Gatenby and Gillies 2004).
The PPP The biological roles of glucose are not only to function as a major fuel source for ATP generation but also to serve as a metabolic precursor to provide carbon source for the biosynthesis of macromolecules in proliferating cells. For the fast-growing tumor cells, production of sufficient amount of such metabolic intermediates is particularly important for the timely replication of their genome, synthesis of amino acids, membrane lipids, and the package of daughter cells. As such, only a portion of the glucose utilized by the tumor cells is used to convert to lactate during aerobic glycolysis, while a significant portion of glucose is metabolized through the PPP (Kuo et al. 2000; Poulsen and Frederiksen 1981). PPP is a major branch of metabolic flow derived from glycolysis, with G6PD as the rate-limiting enzyme locating at the branching point of PPP (Fig. 5.1). This pathway is essential for generating metabolic intermediates pentose-5-phosphate for nucleotide synthesis and the reducing equivalent NADPH for the reductive biosynthesis of lipids and fats. Significantly, NADPH is also an important reducing equivalent that is essential for the maintenance of redox homeostasis (Tian et al. 1999). The sugar moiety of the nucleotide precursor ribose-5-phosphate can be synthesized either from the oxidative or non-oxidative part of PPP (for review, see Tong et al. 2009). Both the oxidative portion (including steps catalyzed by G6PD and 6-phosphogluconate dehydrogenase enzymes) and the non-oxidative steps (including those catalyzed by transketolase-like-1 and transaldolases) of PPP have been reported to be activated or over-expressed in cancer cells (Langbein et al. 2006; Ramos-Montoya et al. 2006). Transketolase and transaldolase seem to play an important role in the inter-conversion of the metabolic intermediates between PPP and glycolysis. When the requirement for NADPH exceeds the need for pentose-phosphate in biosynthesis, the excess amount of pentose-phosphate can be converted to the glycolytic intermediates F6P and GA3P by the metabolic steps catalyzed by transaldolase and transketolase. On the other hand, when pentose-phosphate is in high demand due to fast cell growth and proliferation, the generation of a large amount of pentosephosphate can be met by conversion from F6P and GA3P catalyzed by transaldolase and transketolase. Targeted inhibition of PPP is reported to be effective in preventing tumor metastasis (Langbein et al. 2008). The exact roles of oxidative PPP and nonoxidative PPP in different stages of tumorigenesis remain to be elucidated.
5 Glycolytic Pathway as a Target for Tumor Inhibition
99
Mitochondrial Dysfunction and Increased Glycolysis in Cancer Mitochondria consume the majority of the oxygen utilized in cellular metabolism, and are responsible for aerobic ATP generation from glucose through OXPHOS (Newmeyer and Ferguson-Miller 2003; Wallace 1999). Embedded in the inner mitochondrial membrane, the mitochondrial respiratory chain is composed of five multi-enzyme complexes, including NADH dehydrogenase (complex I), succinate dehydrogenase (complex II), cytochrome bc1 complex (complex III), cytochrome c oxidase (complex IV), and the ATP synthase (complex V) (for review see Mandavilli et al. 2002). One unique feature of mitochondria is that these cellular organelles have their own genetic materials, mitochondrial DNA (mtDNA). It should be noted that mtDNA lacks histone protection, has limited DNA repair capacity, and is located in a close proximity to the electron transport chain, the major site for ROS generation. As such, mtDNA is more vulnerable to damage compared to nuclear DNA. Since mitochondrial genome is devoid of introns, mutations in mtDNA are more likely to have significant functional consequences, leading to dysfunction in respiration and alterations in energy metabolism. Thus, mitochondrial genomic integrity is essential for maintaining normal ATP production and metabolic homeostasis (Wallace 1999). Because mitochondrial transcription is polycistronic, any deletion or insertion of a nucleo tide may cause a frame-shifting of the codon and thus affect the downstream genes encoding for the respiratory chain components (Zastawny et al. 1998). It was reported that mtDNA mutation rates are 10–20 times greater than nuclear DNA (Olgun et al. 2002), and increased mutations in mtDNA have been reported in various cancer cells (Verma et al. 2003; Wallace 1999). Patients bearing mtDNA mutations show refractory to conventional therapeutic agents (Carew et al. 2003; Ohta 2006). ROS-induced mtDNA mutations may play a key role in many diseases including cancer, ageing, and neurodegenerative diseases. mtDNA mutations, which may directly cause defective mitochondrial respiration to generate ATP through the coupled electron-transport chain and OXPHOS, have been found to alter cellular energy metabolism and induce higher glycolysis in a large spectrum of human cancers, suggesting that mitochondrial respiratory dysfunction may be mechanistically linked with increased glycolysis in human cancer (Cavalli and Liang 1998; Halabe Bucay 2007). In cancer cells with altered mitochondrial metabolism, mitochondria seem to export a large amount of citrate, which may be utilized as a metabolic intermediate for the biosynthesis of cholesterol and fatty acids (Parlo and Coleman 1986). Interestingly, citrate has been observed to increase in certain subsets of cancers (Kuhajda et al. 1994; Ristow 2006). The impaired activity of mitochondrial aconitase in cancer cells may negatively affect the TCA cycle, leading to elevated fatty acid generation and a coupling of increased glycolysis with lipogenesis. In addition, impairment of the mitochondrial Fe-S complex activity and a decrease in OXPHOS can lead to altered energy metabolism and tumor formation (Thierbach et al. 2005). ATP synthase, an enzyme required for OXPHOS, was found to be decreased in certain cancer types (Isidoro et al. 2004). Recently it was
100
W. Lu and P. Huang
reported that cells bearing mutations in mtDNA encoding for the NADH dehydrogenase complex component ND6 may cause defect in complex I, leading to elevated ROS production and increased metastatic potential (Ishikawa et al. 2008). The mitochondrial fumarate hydratase and succinate dehydrogenase (SDH), two enzymes involved in the TCA cycle and encoded by the nuclear DNA, have recently been reported to have tumor suppressor function (Baysal et al. 2000; Tomlinson et al. 2002). SDH is also an integral component of the electron transport chain complex II. Patients with inherited mutations in succinate dehydrogenase and fumarate hydratase are more likely to develop hereditary paraganglioma and papillary renal cell carcinoma (Baysal et al. 2000; Tomlinson et al. 2002). This may be due to the accumulation of succinate or fumarate, resulting in the inhibition of prolyl hydroxylase function that otherwise promotes HIF degradation. These tumors display high HIF activity even in the presence of oxygen (Isaacs et al. 2005; Pollard et al. 2005; Selak et al. 2005). Further study in fumarate hydratase-deficient cells showed an increase in LDH-A expression due to HIF-1a stabilization. Interestingly, inhibition of LDH-A in these cells led to ROS-mediated apoptosis. A knockdown of LDH-A resulted in a significant reduction in tumor growth in xenograft mouse model (Xie et al. 2009). Recently, it is reported that SDH5, a gene required for the interaction and flavination of SDH, is mutated in paraganglioma, a neuroendocrine tumor previously linked to mutations in genes encoding for SDH subunits (Hao et al. 2009). In addition, cells with increased HIF-1 stabilization or p53 mutation also show elevated aerobic glycolysis and reduced mitochondrial respiration. Such metabolic shift is characteristic of the Warburg effect and seems to play an important role in malignant transformation, not just a bystander in cancer (Frezza and Gottlieb 2009).
Tumor Microenvironment and Selection of Highly Glycolytic Cancer Cells HIF-1 and Glycolysis Cancer cells in a tumor mass often endure hypoxia, especially when the tumor mass has overgrown the capacity of blood supply. The ability of cancer cells to adapt and survive in hypoxic environment is a crucial step in tumor progression (Dang and Semenza 1999). Under hypoxic conditions, mitochondrial respiration may be limited and ATP generation through OXPHOS is decreased, forcing the cancer cells to utilize alternative mechanisms such as increased glycolysis to maintain sufficient cellular ATP for cell growth and proliferation. During adaptation to hypoxic microenvironment, the increase in HIF-1 plays an essential role (Sharp and Bernaudin 2004). HIF-1 is a heterodimer of HIF1a and HIF1b and functions as a transcription factor that regulates a number of genes critical for glycolysis, angiogenesis, cell proliferation, mitochondrial function, and cell sur-
5 Glycolytic Pathway as a Target for Tumor Inhibition
101
vival. Under physiologic conditions, the half-life (protein stability) and the transcriptional activity of HIF1a are negatively regulated by the O2-dependent prolyl and asparaginyl hydroxylation, respectively (Semenza 2003). Prolyl hydroxylase is a key enzyme capable of modifying HIF-1a at proline sites and marking it for Von Hippel Lindau protein (VHL)-mediated degradation by the 26S proteasome (Cockman et al. 2000; Thomas et al. 2003). The stabilized HIF-1a may be transported into the nucleus and forms a heterodimer with HIF-1b. The HIF1 heterodimer can then recruit p300/CBP and bind to the promoter region of the specific DNA sequence known as the hypoxia responsive element (HRE) to promote the expression of its targeted genes (Dang and Semenza 1999). Examples of HIF-1 downstream genes include a number of molecules involved in glycolysis such as GLUT1, HKII, aldolase A, enolase 1, LDHA, PFK, PGK, and pyruvate kinase (Dang and Semenza 1999; Denko 2008). The increased glycolytic gene expression due to HIF-1a stabilization in turn leads to enhanced glycolysis (Lu et al. 2002; Safran and Kaelin 2003). Clinically, cancer patients whose tumors with a greater hypoxic fraction seem to be associated with high HIF-1 expression and resistance to chemotherapy and radiotherapy (Hockel et al. 1993; Hockel and Vaupel 2003). Furthermore, due to the lack of pVHL tumor suppressor in some clear cell renal carcinoma, HIF-1 is consistently expressed and its targeted glycolytic genes are upregulated even under normoxic condition, resulting in the increased glucose uptake, increased lactate production, and decreased mitochondrial respiration (Semenza 2007). HIF-1 can also be activated by other non-hypoxic stimuli such as oncogenes and ROS, leading to increased glycolysis.
HIF-1 and Mitochondria In addition to promoting glycolysis, HIF-1 can also suppress aerobic respiration by enhancing the expression of pyruvate dehydrogenase kinase 1 (PDK1), which inhibits the activity of PDH. As illustrated in Fig. 5.2, an increased expression of PDK1 promotes the phosphorylation of PDH, leading to its inactivation, and thus reduces the conversion of pyruvate to acetyl CoA. As such, an increase in HIF-1 due to hypoxic microenvironment may limit the entry of pyruvate into the TCA cycle in the mitochondria, and thus suppresses the utilization of glucose as the energy supply via OXPHOS pathway (Papandreou et al. 2006). This suppression of substrate entry into the TCA cycle would further enhance the dependence of cells on glycolysis for ATP generation. The ability of HIF-1 to enhance glycolysis has been observed in many tumor cells. As described above, glycolysis not only serves as a rapid energetic pathway to generate ATP but also provides the building blocks for the synthesis of other biomolecules in cancer cells. For instance, the metabolic intermediates of the glycolytic pathway can be converted to ribose for the synthesis of nucleic acids via the PPP. Thus, it is possible that
102
W. Lu and P. Huang Glucose GLUT1
p53
Glucose
AMPK
LKB -1
Bioenergetic stress
HK
TIGAR
AMP
G6P PFK2
H 2O
PGI
F6P
F2,6-BP
O2
SCO2 Respiration Chain
PFK1
F1,6-BP Aldolase
NADH FADH2
Dihydroxyacetone-P + GA3P GAPDH
Pyr
1,3-BPG
Myc
Ac-CoA PDH
TCA cycle Mito.
PGK
3-PG
PDK1
PGM
2-PG Enolase
PEP
PK-M2
Pyr Lactate
LDH-A
AKT HIF-1
PIP3 Ras
PTEN
PI3K
MCT 4
Lactate
Fig. 5.2 Potential metabolic regulatory mechanisms contributing to the Warburg effect. Several important molecules including the oncogenic molecule Myc, the tumor suppressor p53, the transcription factor HIF-1, and the kinase Akt play essential roles in the regulation of cellular metabolism and contribute to the development of Warburg phenomenon in cancer cells. Activation of Myc oncogene upregulates the expression of many glycolytic molecules including the glucose transporter GLUT1, HK, and LDH, leading to increased glycolysis (orange arrows). Similarly, stabilization of HIF-1 due to hypoxia in tumor microenvironment or oncogenic signaling (e.g. Ras) may transactivate many glycolytic genes including GLUT1, HKII, PK-M2, and LDH-A, and enhances glycolytic flux (red arrows). HIF-1 also activates PDK1 gene expression, which in turn inhibits PDH and thus limits pyruvate entry into the mitochondria as fuel to generate ATP. Akt promotes plasma membrane association of GLUT1 to increase glucose uptake, and enhance HK association with the mitochondria and prevent apoptosis (blue arrows). On the contrary, the tumor suppressor p53 inhibits glycolysis by regulating the expression of HKII, PGM, and TIGAR, and stimulates mitochondrial respiration through transcriptional upregulation of SCO2 (green arrows). The loss of p53 function in cancer cells will lead to an increase in glycolysis and a decrease in respiration. The AMP-dependent kinase (AMPK) regulates glycolysis in part through its activation of p53
even in cancer cells without significant mitochondrial structural defects, the abnormal increase of HIF-1 may block normal mitochondrial function through transcriptional activation of PDK1, leading to a decrease in OXPHOS and an increase in glycolysis (Lu et al. 2002; Safran and Kaelin 2003). As such, the suppression of PDH by elevated HIF-1 in hypoxic tissue microenvironment may
5 Glycolytic Pathway as a Target for Tumor Inhibition
103
significantly contribute to the decrease in mitochondrial function and increase in glycolysis in cancer cells, characteristic of the Warburg effect.
Mutations of Tumor Suppressor Genes and Metabolic Alterations p53 Regulation of Glycolysis and Mitochondrial Respiration The tumor suppressor p53 plays major roles in a number of important cellular processes including stress responses (Vogelstein et al. 2000). Through transcriptional activation of its target genes, p53 involves in the regulation of cell cycle, DNA damage response, apoptosis, senescence, and metabolism. Through the transcriptional regulation of HK, PGM, TIGAR, and SCO2, p53 seems to prevent a switch from aerobic respiration to glycolysis (Green and Chipuk 2006; Shaw 2006). As illustrated in Fig. 5.2, HKII is the first rate-limiting enzyme of the glycolysis. The wildtype p53 protein can bind to the HKII promoter region and down-regulate HKII gene expression, limiting the entry of glucose into the glycolytic pathway. Similarly, p53 down-regulates the function of PGM, a glycolytic enzyme that converts 3-PG to 2-PG, and thus modulates the overall flux of glycolysis (Kondoh et al. 2005; Smith et al. 2006). The ability of p53 to modulate glycolysis has been further substantiated by the demonstration that p53 promotes the expression of TIGAR, a novel isoform of phosphofructose kinase (PFK2) that only contains the FBPase domain but lacks the kinase domain. It functions to lower the intracellular levels of fructose-2,6-bisphosphate. Because fructose-2,6-bisphosphate is a key allosteric activator of PFK1 (a key rate-limiting glycolytic enzyme), the elevated expression of TIGAR in cells with functional p53 would have low fructose-2, 6-bisphosphate levels and thus a lower flow of glucose metabolism through the step catalyzed by PFK1 (Bensaad et al. 2006). This causes a shift toward the PPP shunt, leading to increased production of NADPH, the major reducing equivalent for the generation of reduced glutathione and other cellular antioxidants, which in turn promote the scavenging of ROS and prevent mutational events associated with cancer. PPP can also produce ribose-5-phosphate, which could contribute to nucleotide biosynthesis and DNA repair. Therefore, through upregulation of TIGAR, p53 sets a glycolytic checkpoint to suppress glycolysis, to enhance ROS-scavenging capacity, and to promote genomic stability (Green and Chipuk 2006). A loss of p53 function in cancer cells, on the other hand, would promote glycolysis, leading to the Warburg phenomenon. Recently, p53 has been found to directly regulate mitochondrial energy production through transcriptional regulation of the synthesis of cytochrome c oxidase 2 (SCO2), a regulator of mitochondrial respiration and the oxygen utilization in the eukaryotic cells. SCO2 encodes a copper-binding protein required for the assembly of cytochrome c oxidase subunit II into the mitochondrial respiratory chain complex IV. Disruption of the SCO2 gene in human cancer cells with wild-type p53 recapitulated
104
W. Lu and P. Huang
the metabolic switch toward glycolysis similar to that observed in p53-deficient cells (Bensaad et al. 2006; Matoba et al. 2006). Therefore, p53 is able to regulate glycolysis and mitochondrial respiration in the concerted manner through HK, PGM, TIGER, and SCO2. Because p53 mutations and deletion are common in human cancer, it is likely that the loss of p53 function may contribute significantly to the switch from aerobic respiration to glycolysis observed in many cancer types (Shaw 2006), and provide a molecular link between genetic alterations and the Warburg effect. AMPK and Glycolytic Regulation AMP-activated protein kinase (AMPK) is a master metabolic regulator that plays an important role in cellular energy homeostasis. AMPK is a heterotrimeric protein complex consisting of an a catalytic subunit and b, g regulatory subunits (Kahn et al. 2005). It functions as a serine–threonine kinase at a regulatory pathway that integrates signals from energy metabolism (Fig. 5.2). As cellular ATP decreases and AMP level increases, AMPK is allosterically activated by increased cellular AMP levels, a condition that occurs under energetic stress such as cellular ATP depletion or hypoxia. Through binding to the regulatory g-subunit of the AMPK complex, AMP triggers a conformational change of the g-subunit, leading to the exposure of the active site Thr-172 on the a catalytic subunit. This in turn inhibits dephosphorylation and facilitates phosphorylation at Thr-172 by the upstream kinases (Suter et al. 2006). LKB-1, functioning as a tumor suppressor, is a primary upstream kinase of AMPK (Hardie 2007). LKB1 is required for AMPK activation under bioenergetic stress such as glucose withdrawal (Hawley et al. 2003; Shaw et al. 2004b; Woods et al. 2003). Activation of AMPK promotes phosphorylation of p53 at Ser-15 to regulate glycolysis. AMPK can also stimulate GLUT1,3mediated glucose uptake, and therefore enhances glycolytic pathway (Cidad et al. 2004). In addition, AMPK can phosphorylate the TSC1-TSC2 complex or/and block TORC1 activity to inhibit mammalian target of rapamycin (mTOR)-dependent protein translation in an LKB-dependent manner (Inoki et al. 2003). AMPK has been reported to down-regulate lipid synthesis through inhibition of acetylCoA carboxylase-1 (ACC1) (Hardie and Pan 2002). Mutations in LKB-1 have been associated with Peutz–Jeghers Syndrome (Scott et al. 2008). Mice with heterozygous disruption of LKB-1 develop neoplasia similar to Peutz–Jeghers Syndrome (Bardeesy et al. 2002). Loss of LKB-1 results in unchecked cell growth and the development of aggressive cancer (Gurumurthy et al. 2008; Ji et al. 2007).
Activation of Oncogenes and Increased Glycolysis c-Myc c-Myc is an oncogenic transcription factor with the helix-loop-helix leucine zipper structure. Through dimerization with its partner Max, c-Myc binds to specific DNA sequences known as E box, and stimulates the transcription of many genes
5 Glycolytic Pathway as a Target for Tumor Inhibition
105
involved in cell proliferation, metabolism, and apoptosis (Dang 1999). Max can form heterodimer with the Mad family proteins to repress transcription and therefore antagonize c-Myc function and achieve tight c-Myc regulation. Deregulated expression of the c-Myc gene is a common feature of human malignancies (Little et al. 1983). c-Myc expression enhances aerobic glycolysis by directly upregulating the expression of glycolytic genes such as enolase A, HK II, LDH-A, PFK, GAPDH, and GLUT1 independent of hypoxia (Kim et al. 2004). Many of the genes regulated by c-Myc are known to be regulated by HIF-1 as well (Dang et al. 2008). This suggests a possibility that in normal cells the low expression of glycolytic genes may be due to the tightly controlled Myc and HIF-1 expression, while in cancer cells, the increased expression of glycolytic enzymes may be attributed to the constitutive activation of c-Myc and/or elevated HIF-1 in response to the hypoxic tumor microenvironment. Genes of mitochondrial biogenesis and function such as cytochrome c and TFAM (transcription factor A of mitochondria) constitute another group of molecules upregulated by c-Myc overexpression (Dang et al. 2005). It is suggested that imbalanced expression of nuclear encoded versus mitochondrial encoded proteins could result in defective respiratory chain complex as noted in prostate cancer progression (Herrmann et al. 2003). The ability of c-Myc to upregulate glycolysis in the normoxic condition suggests that Myc may play a role in a switch to glycolytic metabolism during tumorigenesis.
Ras Ras is a guanine nucleotide triphosphatase that functions as a molecular switch in a large network of intracellular signaling pathways. Two major downstream signal pathways, MAPK and PI3K/Akt, are found to be upregulated in many cancer cells and may collaboratively promote the malignant phenotype of cancer cells. In particular, it has been shown that PI3K is required for both the development and maintenance of certain tumors driven by mutant H-Ras (Lim and Counter 2005). Alteration in the Ras-PI3K-Akt signal transduction pathway can result in the upregulation of HKII and elevation of glycolytic activity. Ras activation can also upregulate HIF1, NF-kB, PFK1 and thus enhances tumor glycolysis (Kole et al. 1991). PI3K/Akt Pathway Phosphatidylinositol 3-kinase (PI3K) is a heterodimer consisting of a catalytic subunit p110 and a regulatory subunit p85. Binding of growth factor molecules to their receptor tyrosine kinase activates PI3K, resulting in the phosphorylation of phosphatidylinositol on the plasma membrane and the activation of its downstream pathways such as Akt and mTOR. Tumor suppressor phosphatase and tensin homolog (PTEN) is a critical negative counterpart of PI3K by dephospho-
106
W. Lu and P. Huang
rylation PIP3 (phosphatidylinositol 3,4,5-triphosphate) to PIP2, leading to an inhibition of the Akt signal pathway, which is a serine/threonine kinase regulated by PIP3. Akt activation may result from PI3K activation or loss of PTEN function. Other oncogenes, such as Ras, HER2/neu, and Bcr-Abl, can also activate Akt. The PI3K/Akt kinase pathway is a central regulator of cell metabolism, proliferation, and survival and is deregulated in many cancer types. A recent paper reports that Akt stimulates glucose consumption in transformed cells and Akt-expressing cells are more susceptible to death after glucose withdrawal than control cells (Elstrom et al. 2004). Activation of the PI3K/Akt pathway results in coordinated localization of glucose transporters to the cell surface and upregulation of glycolytic enzymes such as PFK2 and enhances glucose capture through glucose phosphorylation by HK (Pozuelo Rubio et al. 2003). Akt can also inactivate p53 through stabilization of Mdm2 (Gottlieb et al. 2002). Thus, Akt activation seems to be able to cause a metabolic switch from OXPHOS to aerobic glycolysis in cancer cells. As such, Akt is considered as “Warburg kinase” by some investigators (Robey and Hay 2009). In addition to its role in glycolysis, Akt is also involved in other metabolic processes. For instance, Akt phosphorylates and inhibits glycogen synthase kinase-3, leading to the dephosphorylation of glycogen synthase by the glycogenassociated form of protein phosphatase 1, and a stimulation of glycogen synthesis. Activation of Akt stimulates the transport of both glucose and amino acids, which in turn support the mTOR-dependent promotion of protein translation. On the contrary, the LKB-1 tumor suppressor negatively regulates mTOR signaling (Shaw et al. 2004a). Many human tumors display an increased expression of lipogenic enzymes including ATP citrate lyase (ACL, an enzyme that converts cytosolic citrate to acetyl-CoA for fatty acid synthesis), fatty acid synthase, and acetyl-CoA carboxylase (Milgraum et al. 1997; Rossi et al. 2003; Swinnen et al. 2000; Yahagi et al. 2005). Besides, some tumors have a high rate of de-novo fatty acid synthesis from carbohydrate precursors as shown by 14C-glucose studies (Kuhajda 2000; Ookhtens et al. 1984). Such an increase of lipogenic activity in cancer cells may in part be attributed to Akt activation, which stimulates the synthesis and nuclear accumulation of a key lipogenic transcription factor sterol regulatory element binding protein-1, resulting in the upregulation of lipogenic enzymes (Porstmann et al. 2005). Akt also directly phosphorylates and activates ACL (Swinnen et al. 2006). Thus, it appears that the PI3K/Akt/mTOR axis is a major regulatory pathway that governs glycolysis and cellular biosynthesis (DeBerardinis et al. 2008a).
Glycolytic Pathway as a Target for Tumor Inhibition Compelling evidence shows that cancer cells have profound alterations in their energy metabolism, with an increase in aerobic glycolysis as one of the most prominent metabolic changes. As described above, multiple factors including mitochondrial dysfunction, oncogenic signals, loss of tumor suppressor function, and aberrant expression of certain regulatory molecules and enzymes can all contribute
5 Glycolytic Pathway as a Target for Tumor Inhibition
107
to the highly glycolytic phenotype of cancer cells. Several key regulators of cancer metabolism and their influence on the major metabolic steps are illustrated in Fig. 5.2. Although whether an increase in glycolysis is a cause or a “symptom” of malignant transformation still remains as a matter of debate, the profound metabolic alterations in cancer cells are likely to have significant therapeutic implications. In particular, the increased dependency of cancer cells on the glycolysis for energy supply and for metabolic intermediates as the building blocks for cell proliferation provides a biochemical basis to preferentially inhibit tumor growth by targeting the glycolytic pathway. Recent studies suggest that it is indeed possible to use small compounds or siRNA to abrogate certain key steps in the glycolytic pathway by suppression of the key enzymes such as HK and LDH, leading to selective killing of cancer cells (Fantin et al. 2006; Pelicano et al. 2006; Xu et al. 2005b). Theoretically, many of the enzymes and regulatory molecules shown in Figs. 5.1 and 5.2 may serve as potential targets for such a metabolic intervention. The following sections provide a brief summary of several important compounds that have been shown to inhibit glycolysis with potential use in cancer therapeutics.
2-Deoxyglucose 2-Deoxyglucose is a glucose molecule which has the 2-hydroxyl group replaced by hydrogen. Upon transport into the cells, 2-deoxyglucose is phosphorylated by HK to 2-deoxyglucose-phosphate, which cannot be further metabolized by phosphohexose isomerase. Thus, 2-deoxyglucose-phosphate accumulates in the cells and interferes with glycolytic metabolism by competing with glucose for glycolytic enzymes such as HK and phosphoglucose isomerase. Because of the presence of abundant glucose in the cell culture medium and in the blood circulation, high concentrations of 2-deoxyglucose in the millimolar range are required to be effective to inhibit cancer cell growth. An in vivo study showed that this compound alone exhibited limited anticancer activity, but could significantly enhance the therapeutic efficacy of doxorubicin and taxol in animal tumor models (Maschek et al. 2004). This compound has been evaluated clinically in phase I trials to determine its safety profile and pharmacokinetics in 34 patients with solid tumors who had relapsed after chemotherapy with other agents. It appears that a combination of 2-deoxyglucose and docetaxel was safe with no evidence of pharmacokinetic interaction (http://www.thresholdpharm.com/sec/pipe_2deoxyglucose). The clinical therapeutic activity of this compound, alone or in combination with other anticancer agents, still remains unclear and requires further evaluation.
3-Bromopyruvate The pyruvate analog 3-bromopyruvate (3-BrPA) has been shown to inhibit HK and cause severe depletion of cellular ATP, leading to massive cancer cell death, especially in cells with mitochondrial respiratory dysfunction or under hypoxic conditions
108
W. Lu and P. Huang
(Geschwind et al. 2004; Ko et al. 2001; Xu et al. 2005b). Although depletion of cellular ATP is thought to be a key mechanism by which 3-BrPA exerts its cytotoxic effect, a recent study also showed that this compound could covalently modify HKII protein and cause its dissociation from mitochondria. Dissociation of HKII from mitochondria could in turn cause the release of apoptosis-inducing factor and subsequent cell death (Chen et al. 2009). It is worth noting that cancer cells with mitochondrial dysfunction or in hypoxic environment usually become less sensitive to many anticancer drugs currently used in clinic, due in part to the accumulation of HIF-1 and activation of Akt survival pathways. These cancer cells, however, are particularly sensitive to the cytotoxic action of 3-BrPA because of their high dependency on glycolysis. Furthermore, ATP depletion by 3-BrPA may disable the ATPdependent drug-exporting pump and thus overcome multi-drug resistance associated with increased expression of MDR pump (Xu et al. 2005b). Thus, the unique mechanism of action of 3-BrPA makes this compound potentially useful in killing cancer cells in hypoxic tumor and overcoming multi-drug resistance. Interestingly, combination of glycolytic inhibition with the mTOR inhibitor rapamycin seems to have synergistic anticancer effect (Xu et al. 2005a). It should be pointed out, however, that 3-BrPA is an alkylating agent, which may potentially attack other cellular molecules, and the cytotoxic action of this compound may not be entirely due to its inhibition of HKII. Although there is compelling evidence showing that 3-BrPA has promising anticancer activity both in vitro and in vivo in several animal tumor models, this compound is still in its pre-clinical development, and has not yet entered clinical trials.
Lonidamine Lonidamine (1-(2,4-dichlorobenzyl)-1H-indazole-3-carboxylic acid) is a derivative of indazole-3-carboxylic acid known to inhibit glycolysis in tumor cells. This compound is an orally administered small molecule that has been evaluated in phase II/III clinical trials for the treatment of various solid tumors (Di Cosimo et al. 2003). The mechanism by which lonidamine suppresses glycolysis appears through the inhibition of the mitocondrially bound HK (Floridi et al. 1998). Interestingly, studies using murine cells showed that this compound inhibits aerobic glycolysis in tumor cells but enhances glycolysis in the normal cells (Floridi et al. 1998). The mechanism responsible for this differential effect on glycolysis in normal and cancer cells still remains unclear. Further studies by the same group showed that lonidamine may also inhibit mitochondrial respiration, as evidenced by the suppression of oxygen consumption in both normal and cancer cells (Floridi et al. 1998). The anticancer activity of this compound may be ascribed to its overall ability to abrogate energy metabolism in cancer cells. In clinical trials, lonidamine appeared to be able to enhance the therapeutic activity of several major anticancer agents, including paclitaxel and cisplatin in the treatment of advanced ovarian cancer (De Lena et al. 2001).
5 Glycolytic Pathway as a Target for Tumor Inhibition
109
Oxythiamine and 6-Aminonicotinamide The PPP is important for cancer cell survival because this pathway provides the metabolic intermediates ribose-5-phosphate for the synthesis of nucleic acids and NADPH for scavenging of reactive oxygen species (ROS). In addition, the non-oxidative phase of PPP also converts 5-carbon metabolic intermediates to 6-carbon and 3-carbon compounds (e.g. F6P and glyceraldehydes-3-phosphate), which can then be used in the glycolytic pathway. Thus, PPP is an important sugar metabolic pathway that generates the key building blocks for cell proliferation, reducing equivalents for redox regulation, and metabolic intermediates for energy metabolism. The enzymes transketolase and transaldolase are essential for the non-oxidative phase of PPP, and seem to be increased in tumor cells (Coy et al. 2005; Heinrich et al. 1976). Oxythiamine is a thiamine antagonist that inhibits transketolase and PDH, which use thiamine pyrophosphate (TPP) as a cofactor for their enzymatic activities. Oxythiamine has been shown to have anticancer activity both in vitro and in vivo (Comin-Anduix et al. 2001; Rais et al. 1999). Specific inhibition of TKTL1 by RNAi can inhibit cancer cell proliferation and compromise cell viability (Hu et al. 2007; Zhang et al. 2007). The compound 6-aminonicotinamide (6-AN) is an inhibitor of G6PD, an enzyme that catalyzes the conversion of G-6-P to 6-phosphogluconolactone at the first step of PPP. This compound exhibits anticancer activity, causes oxidative stress, and sensitizes cells to anticancer agents (Budihardjo et al. 1998; Varshney et al. 2003). However, potential neurotoxicity is a concern for using 6-AN as a therapeutic agent.
Dichloroacetate Dichloroacetate (DCA) is a small compound capable of penetrating most tissues after oral administration, and has been used in the treatment of human lactic acidosis and inherited mitochondrial diseases. Although this compound could lower blood lactate levels, it showed no significant clinical benefit in a controlled clinical trial (Stacpoole and Greene 1992). However, a recent study showed that DCA exhibited significant anticancer activity by inhibiting pyruvate dehydrogenase kinase and thus activating PDH and enhancing the flux of pyruvate into the mitochondria (Bonnet et al. 2007). As such, DCA is able to decrease aerobic glycolysis and promote glucose oxidation and mitochondrial function, leading to a reversion of the Warburg effect in cancer cells. Through promoting cytochrome c release and ROS generation, DCA was shown to induce apoptosis in various human cancer cells (Bonnet et al. 2007; Cao et al. 2008), and enhance the effectiveness of hypoxiaspecific cytotoxic chemotherapy in solid tumor (Cairns et al. 2007). Administration of DCA decreases lactate level, activates PDH activity, and decreases tumor growth in xenograft models (Bonnet et al. 2007; Parolin et al. 2000). With its apparent advantages in drug delivery and anticancer activity/selectivity, DCA is currently in phase I trial in Canada for the treatment of recurrent or metastatic solid tumors.
110
W. Lu and P. Huang
Other Metabolic Modulators Other metabolic modulators, which either involve in the direct inhibition of glycolysis or target the relevant regulatory pathways, are also under investigation in pre-clinical and clinical settings (Chen et al. 2007; Lopez-Lazaro 2008; Pelicano et al. 2006). For instance, metformin is an inhibitor of gluconeogenesis and has been used for clinical treatment of diabetes. This compound is able to activate AMPK and attenuate the mTOR pathway, thus impacting glycolysis and inhibiting tumor growth. Interestingly, a large case–control study suggests that the use of metformin may significantly reduce the risk of developing pancreatic cancer (Li et al. 2009). A potent inhibitor of the glycolytic enzyme PGM was identified as an effective agent against breast cancer in a chemical screening study (Evans et al. 2005). Selective LDH inhibitors such as dihydroxynaphthoic acids have been developed (Yu et al. 2001), and their potential anticancer activity should be further tested. The MCT, which functions to export lactate out of the cells, is upregulated in various human tumor and may be considered as a potential target for tumor inhibition (Sonveaux et al. 2008). NVP-BEZ235, a small molecule that inhibits PI3K and mTOR, is able to decrease glucose uptake and cause marked tumor regression as measured by FDG-PET (Engelman et al. 2008). Disruption of glucose transport may also be a strategy to inhibit glucose metabolism in cancer. Imatinib, a BCRABL tyrosine kinase inhibitor, was found to reduce the expression of GLUT1 and decrease glycolysis in leukemic cells (Gottschalk et al. 2004).
Summary Increased aerobic glycolysis is a hallmark of cancer metabolism. Recent studies suggest that this metabolic alteration may confer proliferative and survival advantages to cancer cells, and help the selection of aggressive malignant cells in the tumor microenvironment. Although OXPHOS in the mitochondria is highly efficient in ATP generation, glycolysis can rapidly generate ATP and, in conjunction with the PPP, can provide the building bricks for the synthesis of nucleic acids, lipids, and proteins to meet the needs of cancer cell proliferation (Bui and Thompson 2006). On the contrary, normal cells mainly use mitochondrial OXPHOS to oxidize glucose to CO2 and H2O and generate a large amount of ATP to fulfill the energy needs of cellular activity with little or low requirements for proliferation in welldifferentiated adult tissues. This profound metabolic difference between cancer and normal cells provides a biochemical basis for therapeutic intervention to preferentially kill cancer cells by inhibition of glycolysis. In addition, beyond their roles in catalyzing glucose breakdown and providing metabolic intermediates for the anabolic reactions, the elevated glycolytic enzymes in cancer cells may also contribute to tumor adaptation and survival through their non-glycolytic functions. For example, HKII exerts its anti-apoptotic function through association with the mitochondria
5 Glycolytic Pathway as a Target for Tumor Inhibition
111
and preventing cytochrome c release. Sustained aerobic glycolysis may also protect cancer cell from ROS-induced cell damage (Ahmad et al. 2005). The metabolic transformation of cancer cells can be achieved by upregulation of the key glycolytic enzymes and glucose transporters through various regulatory mechanisms, including oncogene activation and loss of tumor suppressors. Cancer cells are able to overproduce lactic acid aerobically, whereas normal cells undergo anaerobic glycolysis only when deprived of oxygen. The increase in aerobic glycolysis and elevation of lactate may confer cancer cells an advantage to invade surrounding tissues and facilitate their invasion and metastasis. Thus, inhibition of glycolysis may also be effective in suppressing cancer metastasis. It should be pointed out, however, that cancer cells are genetically unstable with a high degree of heterogeneity in gene regulation and metabolism. Some cancer cells may be more dependent on glutamine metabolism instead of glycolysis for energy supply (Deberardinis et al. 2008b), and inhibition of glycolysis may not be effective in killing these cancer cells. Furthermore, because certain normal tissues such as heart and brain also use glucose as an important fuel source, inhibition of glucose utilization in these tissues might compromise their normal functions. Thus, it is extremely important to understand the exact differences between normal and cancer cells in their metabolic regulation, and specifically target the unique metabolic alterations in cancer cells to improve therapeutic selectivity. With the new discoveries of metabolic regulatory mechanisms and novel glycolytic inhibitors with improved pharmacokinetic and pharmacodynamic properties, it is possible to develop more effective therapeutic strategies with high selectivity against cancer.
References Ahmad IM, Aykin-Burns N, Sim JE, Walsh SA, Higashikubo R, Buettner GR, Venkataraman S, Mackey MA, Flanagan SW, Oberley LW et al (2005) Mitochondrial O2*- and H2O2 mediate glucose deprivation-induced stress in human cancer cells. J Biol Chem 280:4254–4263. Altenberg B, Greulich KO (2004) Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics 84:1014–1020. Bardeesy N, Sinha M, Hezel AF, Signoretti,S, Hathaway NA, Sharpless NE, Loda,M, Carrasco DR, DePinho RA (2002) Loss of the Lkb1 tumour suppressor provokes intestinal polyposis but resistance to transformation. Nature 419:162–167. Baysal BE, Ferrell RE, Willett-Brozick JE, Lawrence EC, Myssiorek D, Bosch A, van der Mey A, Taschner PE, Rubinstein WS, Myers EN et al (2000) Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science (New York, NY) 287:848–851. Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, Bartrons R, Gottlieb E, Vousden KH (2006) TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126:107–120. Bonnet S, Archer SL, Allalunis-Turner J, Haromy A, Beaulieu C, Thompson R, Lee CT, Lopaschuk GD, Puttagunta L, Bonnet S et al (2007) A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 11: 37–51. Brown RS, Goodman TM, Zasadny KR, Greenson JK, Wahl RL (2002) Expression of hexokinase II and Glut-1 in untreated human breast cancer. Nuclear Med Biol 29:443–453.
112
W. Lu and P. Huang
Budihardjo II, Walker DL, Svingen PA, Buckwalter CA, Desnoyers S, Eckdahl S, Shah GM, Poirier GG, Reid JM, Ames MM et al (1998) 6-Aminonicotinamide sensitizes human tumor cell lines to cisplatin. Clin Cancer Res 4:117–130. Bui T, Thompson CB (2006) Cancer’s sweet tooth. Cancer Cell 9:419–420. Cairns RA, Papandreou I, Sutphin PD, Denko NC (2007) Metabolic targeting of hypoxia and HIF1 in solid tumors can enhance cytotoxic chemotherapy. Proc Natl Acad Sci USA 104: 9445–9450. Cao W, Yacoub S, Shiverick KT, Namiki K, Sakai Y, Porvasnik S, Urbanek C, Rosser CJ (2008) Dichloroacetate (DCA) sensitizes both wild-type and over expressing Bcl-2 prostate cancer cells in vitro to radiation. Prostate 68:1223–1231. Carew JS, Huang P (2002) Mitochondrial defects in cancer. Mol Cancer 1:9. Carew JS, Zhou Y, Albitar M, Carew JD, Keating MJ, Huang P (2003) Mitochondrial DNA mutations in primary leukemia cells after chemotherapy: clinical significance and therapeutic implications. Leukemia 17:1437–1447. Cavalli LR, Liang BC (1998) Mutagenesis, tumorigenicity, and apoptosis: are the mitochondria involved? Mutat Res 398:19–26. Chen Z, Lu W, Garcia-Prieto C, Huang P (2007) The Warburg effect and its cancer therapeutic implications. J Bioenerg Biomembr 39:267–274. Chen Z, Zhang H, Lu W, Huang P (2009) Role of mitochondria-associated hexokinase II in cancer cell death induced by 3-bromopyruvate. Biochim et Biophys Acta 1787:553–560. Christofk HR, Vander Heiden MG, Wu N, Asara JM, Cantley LC (2008) Pyruvate kinase M2 is a phosphotyrosine-binding protein. Nature 452:181–186. Cidad P, Almeida A, Bolanos JP (2004) Inhibition of mitochondrial respiration by nitric oxide rapidly stimulates cytoprotective GLUT3-mediated glucose uptake through 5’-AMP-activated protein kinase. Biochem J 384:629–636. Cockman ME, Masson N, Mole DR, Jaakkola P, Chang GW, Clifford SC, Maher ER, Pugh CW, Ratcliffe PJ, Maxwell PH (2000) Hypoxia inducible factor-alpha binding and ubiquitylation by the von Hippel-Lindau tumor suppressor protein. J Biol Chem 275:25733–25741. Comin-Anduix B, Boren J, Martinez S, Moro C, Centelles JJ, Trebukhina R, Petushok N, Lee WN, Boros LG, Cascante M (2001) The effect of thiamine supplementation on tumour proliferation. A metabolic control analysis study. Eur J Biochem/FEBS 268:4177–4182. Coy JF, Dressler D, Wilde J, Schubert P (2005) Mutations in the transketolase-like gene TKTL1: clinical implications for neurodegenerative diseases, diabetes and cancer. Clin Lab 51:257–273. Czernin J, Phelps ME (2002) Positron emission tomography scanning: current and future applications. Annu Rev Med 53:89–112. Dang CV (1999) c-Myc target genes involved in cell growth, apoptosis, and metabolism. Mol Cell Biol 19:1–11. Dang CV, Kim JW, Gao P, Yustein J (2008) The interplay between MYC and HIF in cancer. Nat Rev 8:51–56. Dang CV, Li F, Lee LA (2005) Could MYC induction of mitochondrial biogenesis be linked to ROS production and genomic instability? Cell Cycle 4:1465–1466. Dang CV, Semenza GL (1999) Oncogenic alterations of metabolism. Trends Biochem Sci 24:68–72. De Lena M, Lorusso V, Latorre A, Fanizza G, Gargano G, Caporusso L, Guida M, Catino A, Crucitta E, Sambiasi D et al (2001) Paclitaxel, cisplatin and lonidamine in advanced ovarian cancer. A phase II study. Eur J Cancer 37:364–368. DeBerardinis RJ, Lum JJ, Hatzivassiliou G, Thompson CB (2008a) The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab 7:11–20. Deberardinis RJ, Sayed N, Ditsworth D, Thompson CB (2008b) Brick by brick: metabolism and tumor cell growth. Curr Opin Genet Dev 18:54–61. Denko NC (2008) Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat Rev 8: 705–713. Di Cosimo S, Ferretti G, Papaldo P, Carlini P, Fabi A, Cognetti F (2003) Lonidamine: efficacy and safety in clinical trials for the treatment of solid tumors. Drugs Today (Barc) 39:157–174.
5 Glycolytic Pathway as a Target for Tumor Inhibition
113
Dombrauckas JD, Santarsiero BD, Mesecar AD (2005) Structural basis for tumor pyruvate kinase M2 allosteric regulation and catalysis. Biochemistry 44:9417–9429. Elstrom RL, Bauer DE, Buzzai M, Karnauskas R, Harris MH, Plas DR, Zhuang H, Cinalli RM, Alavi A, Rudin CM et al (2004) Akt stimulates aerobic glycolysis in cancer cells. Cancer Res 64:3892–3899. Engelman JA, Chen L, Tan X, Crosby K, Guimaraes AR, Upadhyay R, Maira M, McNamara K, Perera SA, Song Y et al (2008) Effective use of PI3K and MEK inhibitors to treat mutant Kras G12D and PIK3CA H1047R murine lung cancers. Nat Med 14:1351–1356. Evans MJ, Saghatelian A, Sorensen EJ, Cravatt BF (2005) Target discovery in small-molecule cell-based screens by in situ proteome reactivity profiling. Nat Biotechnol 23:1303–1307. Fantin VR, St-Pierre J, Leder P (2006) Attenuation of LDH-A expression uncovers a link between glycolysis, mitochondrial physiology, and tumor maintenance. Cancer Cell 9:425–434. Floridi A, Bruno T, Miccadei S, Fanciulli M, Federico A, Paggi MG (1998) Enhancement of doxorubicin content by the antitumor drug lonidamine in resistant Ehrlich ascites tumor cells through modulation of energy metabolism. Biochem Pharmacol 56:841–849. Frezza C, Gottlieb E (2009) Mitochondria in cancer: not just innocent bystanders. Semin Cancer Biol 19:4–11. Gatenby RA (1995) The potential role of transformation-induced metabolic changes in tumor-host interaction. Cancer Res 55:4151–4156. Gatenby RA, Gawlinski ET (1996) A reaction-diffusion model of cancer invasion. Cancer Res 56:5745–5753. Gatenby RA, Gawlinski ET, Gmitro AF, Kaylor B, Gillies RJ (2006) Acid-mediated tumor invasion: a multidisciplinary study. Cancer Res 66:5216–5223. Gatenby RA, Gillies RJ (2004) Why do cancers have high aerobic glycolysis? Nat Rev 4: 891–899. Geschwind JF, Georgiades CS, Ko YH, Pedersen PL (2004) Recently elucidated energy catabolism pathways provide opportunities for novel treatments in hepatocellular carcinoma. Expert Rev Anticancer Ther 4:449–457. Godoy A, Ulloa V, Rodriguez F, Reinicke K, Yanez AJ, Garcia Mde L, Medina RA, Carrasco M, Barberis S, Castro T et al (2006) Differential subcellular distribution of glucose transporters GLUT1-6 and GLUT9 in human cancer: ultrastructural localization of GLUT1 and GLUT5 in breast tumor tissues. J Cell Physiol 207:614–627. Golshani-Hebroni SG, Bessman SP (1997) Hexokinase binding to mitochondria: a basis for proliferative energy metabolism. J Bioenerg Biomembr 29:331–338. Gottlieb TM, Leal JF, Seger R, Taya Y, Oren M (2002) Cross-talk between Akt, p53 and Mdm2: possible implications for the regulation of apoptosis. Oncogene 21:1299–1303. Gottschalk S, Anderson N, Hainz C, Eckhardt SG, Serkova NJ (2004) Imatinib (STI571)-mediated changes in glucose metabolism in human leukemia BCR-ABL-positive cells. Clin Cancer Res 10:6661–6668. Green DR, Chipuk JE (2006) p53 and metabolism: inside the TIGAR. Cell 126:30–32. Gurumurthy S, Hezel AF, Sahin E, Berger JH, Bosenberg MW, Bardeesy N (2008) LKB1 deficiency sensitizes mice to carcinogen-induced tumorigenesis. Cancer Res 68:55–63. Haberkorn U, Strauss LG, Reisser C, Haag D, Dimitrakopoulou A, Ziegler S, Oberdorfer F, Rudat V, van Kaick G (1991) Glucose uptake, perfusion, and cell proliferation in head and neck tumors: relation of positron emission tomography to flow cytometry. J Nucl Med 32: 1548–1555. Halabe Bucay A (2007) The biological significance of cancer: mitochondria as a cause of cancer and the inhibition of glycolysis with citrate as a cancer treatment. Med Hypotheses 69: 826–828. Hao HX, Khalimonchuk O, Schraders M, Dephoure N, Bayley JP, Kunst H, Devilee P, Cremers CW, Schiffman JD, Bentz BG et al (2009) SDH5, a gene required for flavination of succinate dehydrogenase, is mutated in paraganglioma. Science (New York, NY) 325:1139–1142. Hardie DG (2007) AMP-activated/SNF1 protein kinases: conserved guardians of cellular energy. Nat Rev Mol Cell Biol 8:774–785.
114
W. Lu and P. Huang
Hardie DG, Pan DA (2002) Regulation of fatty acid synthesis and oxidation by the AMP-activated protein kinase. Biochem Soc Trans 30:1064–1070. Hawley SA, Boudeau J, Reid JL, Mustard KJ, Udd L, Makela TP, Alessi DR, Hardie DG (2003) Complexes between the LKB1 tumor suppressor, STRAD alpha/beta and MO25 alpha/beta are upstream kinases in the AMP-activated protein kinase cascade. J Biol 2:28. Heinrich PC, Morris HP, Weber G (1976) Behavior of transaldolase (EC 2.2.1.2) and transketolase (EC 2.2.1.1) activities in normal, neoplastic, differentiating, and regenerating liver. Cancer Res 36:3189–3197. Herrmann PC, Gillespie JW, Charboneau L, Bichsel VE, Paweletz CP, Calvert VS, Kohn EC, Emmert-Buck MR, Liotta LA, Petricoin EF, 3rd (2003) Mitochondrial proteome: altered cytochrome c oxidase subunit levels in prostate cancer. Proteomics 3:1801–1810. Hockel M, Knoop C, Schlenger K, Vorndran B, Baussmann E, Mitze M, Knapstein PG, Vaupel P (1993) Intratumoral pO2 predicts survival in advanced cancer of the uterine cervix. Radiother Oncol 26:45–50. Hockel M, Vaupel P (2003) Oxygenation of cervix cancers: impact of clinical and pathological parameters. Adv Exp Med Biol 510:31–35. Hove H, Rye Clausen M, Brobech Mortensen P (1993) Lactate and pH in faeces from patients with colonic adenomas or cancer. Gut 34:625–629. Hu LH, Yang JH, Zhang DT, Zhang S, Wang L, Cai PC, Zheng JF, Huang JS (2007) The TKTL1 gene influences total transketolase activity and cell proliferation in human colon cancer LoVo cells. Anticancer Drugs 18:427–433. Inoki K, Zhu T, Guan KL (2003) TSC2 mediates cellular energy response to control cell growth and survival. Cell 115:577–590. Isaacs JS, Jung YJ, Mole DR, Lee S, Torres-Cabala C, Chung YL, Merino M, Trepel J, Zbar B, Toro J et al (2005) HIF overexpression correlates with biallelic loss of fumarate hydratase in renal cancer: novel role of fumarate in regulation of HIF stability. Cancer Cell 8:143–153. Ishikawa K, Takenaga K, Akimoto M, Koshikawa N, Yamaguchi A, Imanishi H, Nakada K, Honma Y, Hayashi J (2008) ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science (New York, NY) 320:661–664. Isidoro A, Martinez M, Fernandez PL, Ortega AD, Santamaria G, Chamorro M, Reed JC, Cuezva JM (2004) Alteration of the bioenergetic phenotype of mitochondria is a hallmark of breast, gastric, lung and oesophageal cancer. Biochem J 378:17–20. Ji H, Ramsey MR, Hayes DN, Fan C, McNamara K, Kozlowski P, Torrice C, Wu MC, Shimamura T, Perera SA et al (2007) LKB1 modulates lung cancer differentiation and metastasis. Nature 448:807–810. Kahn BB, Alquier T, Carling D, Hardie DG (2005) AMP-activated protein kinase: ancient energy gauge provides clues to modern understanding of metabolism. Cell Metab 1:15–25. Kim JW, Dang CV (2006) Cancer’s molecular sweet tooth and the Warburg effect. Cancer Res 66: 8927–8930. Kim JW, Zeller KI, Wang Y, Jegga AG, Aronow BJ, O’Donnell KA, Dang CV (2004) Evaluation of myc E-box phylogenetic footprints in glycolytic genes by chromatin immunoprecipitation assays. Mol Cell Biol 24:5923–5936. Ko YH, Pedersen PL, Geschwind JF (2001) Glucose catabolism in the rabbit VX2 tumor model for liver cancer: characterization and targeting hexokinase. Cancer Lett 173:83–91. Kole HK, Resnick RJ, Van Doren M, Racker E (1991) Regulation of 6-phosphofructo-1-kinase activity in ras-transformed rat-1 fibroblasts. Arch Biochem Biophys 286:586–590. Kondoh H, Lleonart ME, Gil J, Wang J, Degan P, Peters G, Martinez D, Carnero A, Beach D (2005) Glycolytic enzymes can modulate cellular life span. Cancer Res 65:177–185. Krebs HA (1970a) The history of the tricarboxylic acid cycle. Perspect Biol Med 14:154–170. Krebs HA (1970b) Rate control of the tricarboxylic acid cycle. Adv Enzyme Regul 8: 335–353. Kuhajda FP (2000) Fatty-acid synthase and human cancer: new perspectives on its role in tumor biology. Nutrition 16:202–208.
5 Glycolytic Pathway as a Target for Tumor Inhibition
115
Kuhajda FP, Jenner K, Wood FD, Hennigar RA, Jacobs LB, Dick JD, Pasternack GR (1994) Fatty acid synthesis: a potential selective target for antineoplastic therapy. Proc Natl Acad Sci USA 91:6379–6383. Kuo W, Lin J, Tang TK (2000) Human glucose-6-phosphate dehydrogenase (G6PD) gene transforms NIH 3T3 cells and induces tumors in nude mice. Int J Cancer 85:857–864. Langbein S, Frederiks WM, Hausen AZ, Popa J, Lehmann J, Weiss C, Alken P, Coy JF (2008) Metastasis is promoted by a bioenergetic switch: new targets for progressive renal cell cancer. Int J Cancer 122:2422–2428. Langbein S, Zerilli M, Zur Hausen A, Staiger W, Rensch-Boschert K, Lukan N, Popa J, Ternullo MP, Steidler A, Weiss C et al (2006) Expression of transketolase TKTL1 predicts colon and urothelial cancer patient survival: Warburg effect reinterpreted. Br J Cancer 94:578–585. Li D, Yeung SC, Hassan MM, Konopleva M, Abbruzzese JL (2009) Antidiabetic therapies affect risk of pancreatic cancer. Gastroenterology 137:482–488. Lim KH, Counter CM (2005) Reduction in the requirement of oncogenic Ras signaling to activation of PI3K/AKT pathway during tumor maintenance. Cancer Cell 8:381–392. Little CD, Nau MM, Carney DN, Gazdar AF, Minna JD (1983) Amplification and expression of the c-myc oncogene in human lung cancer cell lines. Nature 306:194–196. Lopez-Lazaro M (2008) The warburg effect: why and how do cancer cells activate glycolysis in the presence of oxygen? Anticancer Agents Med Chem 8:305–312. Lu H, Forbes RA, Verma A (2002) Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. J Biol Chem 277:23111–23115. Majewski N, Nogueira V, Bhaskar P, Coy PE, Skeen JE, Gottlob K, Chandel NS, Thompson CB, Robey RB, Hay N (2004) Hexokinase-mitochondria interaction mediated by Akt is required to inhibit apoptosis in the presence or absence of Bax and Bak. Mol Cell 16:819–830. Mandavilli BS, Santos JH, Van Houten B (2002) Mitochondrial DNA repair and aging. Mutat Res 509:127–151. Maschek G, Savaraj N, Priebe W, Braunschweiger P, Hamilton K, Tidmarsh GF, De Young LR, Lampidis TJ (2004) 2-Deoxy-d-glucose increases the efficacy of adriamycin and paclitaxel in human osteosarcoma and non-small cell lung cancers in vivo. Cancer Res 64:31–34. Mathupala SP, Ko YH, Pedersen PL (2006) Hexokinase II: cancer’s double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene 25:4777–4786. Mathupala SP, Rempel A, Pedersen PL (1997) Aberrant glycolytic metabolism of cancer cells: a remarkable coordination of genetic, transcriptional, post-translational, and mutational events that lead to a critical role for type II hexokinase. J Bioenerg Biomembr 29:339–343. Matoba S, Kang JG, Patino WD, Wragg A, Boehm M, Gavrilova O, Hurley PJ, Bunz F, Hwang PM (2006) p53 regulates mitochondrial respiration. Science 312:1650–1653. Milgraum LZ, Witters LA, Pasternack GR, Kuhajda FP (1997) Enzymes of the fatty acid synthesis pathway are highly expressed in in situ breast carcinoma. Clin Cancer Res 3:2115–2120. Minchenko A, Leshchinsky I, Opentanova I, Sang N, Srinivas V, Armstead V, Caro J (2002) Hypoxia-inducible factor-1-mediated expression of the 6-phosphofructo-2-kinase/fructose-2, 6-bisphosphatase-3 (PFKFB3) gene. Its possible role in the Warburg effect. J Biol Chem 277:6183–6187. Mitchell P, Moyle J (1967) Chemiosmotic hypothesis of oxidative phosphorylation. Nature 213:137–139. Newmeyer DD, Ferguson-Miller S (2003) Mitochondria: releasing power for life and unleashing the machineries of death. Cell 112:481–490. Ohta S (2006) Contribution of somatic mutations in the mitochondrial genome to the development of cancer and tolerance against anticancer drugs. Oncogene 25:4768–4776. Olgun A, Akman S, Serdar MA, Kutluay T (2002) Oxidative phosphorylation enzyme complexes in caloric restriction. Exp Gerontol 37:639–645. Ookhtens M, Kannan R, Lyon I, Baker N (1984) Liver and adipose tissue contributions to newly formed fatty acids in an ascites tumor. Am J Physiol 247:R146–R153.
116
W. Lu and P. Huang
Osthus RC, Shim H, Kim S, Li Q, Reddy R, Mukherjee M, Xu Y, Wonsey D, Lee LA, Dang CV (2000) Deregulation of glucose transporter 1 and glycolytic gene expression by c-Myc. J Biol Chem 275:21797–21800. Papandreou I, Cairns RA, Fontana L, Lim AL, Denko NC (2006) HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab 3:187–197. Parlo RA, Coleman PS (1986) Continuous pyruvate carbon flux to newly synthesized cholesterol and the suppressed evolution of pyruvate-generated CO2 in tumors: further evidence for a persistent truncated Krebs cycle in hepatomas. Biochim Biophys Acta 886:169–176. Parolin ML, Spriet LL, Hultman E, Matsos MP, Hollidge-Horvat MG, Jones NL, Heigenhauser GJ (2000) Effects of PDH activation by dichloroacetate in human skeletal muscle during exercise in hypoxia. Am J Physiol 279:E752–E761. Pastorino JG, Hoek JB, Shulga N (2005) Activation of glycogen synthase kinase 3beta disrupts the binding of hexokinase II to mitochondria by phosphorylating voltage-dependent anion channel and potentiates chemotherapy-induced cytotoxicity. Cancer Res 65:10545–10554. Pelicano H, Martin DS, Xu RH, Huang P (2006) Glycolysis inhibition for anticancer treatment. Oncogene 25:4633–4646. Pollard PJ, Briere JJ, Alam NA, Barwell J, Barclay E, Wortham NC, Hunt T, Mitchell M, Olpin S, Moat SJ et al (2005) Accumulation of Krebs cycle intermediates and over-expression of HIF1alpha in tumours which result from germline FH and SDH mutations. Hum Mol Genet 14:2231–2239. Porstmann T, Griffiths B, Chung YL, Delpuech O, Griffiths JR, Downward J, Schulze A (2005) PKB/Akt induces transcription of enzymes involved in cholesterol and fatty acid biosynthesis via activation of SREBP. Oncogene 24:6465–6481. Poulsen HS, Frederiksen P (1981) Glucose-6-phosphate dehydrogenase activity in human breast cancer. Lack of association with oestrogen receptor content. Acta Pathol Microbiol Scand 89:263–270. Pozuelo Rubio M, Peggie M, Wong BH, Morrice N, MacKintosh C (2003) 14-3-3s regulate fructose-2,6-bisphosphate levels by binding to PKB-phosphorylated cardiac fructose-2,6-bisphosphate kinase/phosphatase. EMBO J 22:3514–3523. Rais B, Comin B, Puigjaner J, Brandes JL, Creppy E, Saboureau D, Ennamany R, Lee WN, Boros LG, Cascante M (1999) Oxythiamine and dehydroepiandrosterone induce a G1 phase cycle arrest in Ehrlich’s tumor cells through inhibition of the pentose cycle. FEBS Lett 456:113–118. Ramos-Montoya A, Lee WN, Bassilian S, Lim S, Trebukhina RV, Kazhyna MV, Ciudad CJ, Noe V, Centelles JJ, Cascante M (2006) Pentose phosphate cycle oxidative and nonoxidative balance: a new vulnerable target for overcoming drug resistance in cancer. Int J Cancer 119: 2733–2741. Rathmell JC, Fox CJ, Plas DR, Hammerman PS, Cinalli RM, Thompson CB (2003) Akt-directed glucose metabolism can prevent Bax conformation change and promote growth factorindependent survival. Mol Cell Biol 23:7315–7328. Ristow M (2006) Oxidative metabolism in cancer growth. Curr Opin Clin Nutr Metab Care 9:339–345. Robey RB, Hay N (2005) Mitochondrial hexokinases: guardians of the mitochondria. Cell cycle (Georgetown, TX) 4:654–658. Robey RB, Hay N (2009) Is Akt the “Warburg kinase”?-Akt-energy metabolism interactions and oncogenesis. Semin Cancer Biol 19:25–31. Rossi S, Graner E, Febbo P, Weinstein L, Bhattacharya N, Onody T, Bubley G, Balk S, Loda M (2003) Fatty acid synthase expression defines distinct molecular signatures in prostate cancer. Mol Cancer Res 1:707–715. Safran M, Kaelin WG, Jr (2003) HIF hydroxylation and the mammalian oxygen-sensing pathway. J Clin Invest 111:779–783. Sanchez-Martinez C, Aragon JJ (1997) Analysis of phosphofructokinase subunits and isozymes in ascites tumor cells and its original tissue, murine mammary gland. FEBS Lett 409: 86–90.
5 Glycolytic Pathway as a Target for Tumor Inhibition
117
Schornack PA, Gillies RJ (2003) Contributions of cell metabolism and H+ diffusion to the acidic pH of tumors. Neoplasia (New York, NY) 5:135–145. Schwickert G, Walenta S, Sundfor K, Rofstad EK, Mueller-Klieser W (1995) Correlation of high lactate levels in human cervical cancer with incidence of metastasis. Cancer Res 55: 4757–4759. Scott R, Crooks R, Meldrum C (2008) Gene symbol: STK11. Disease: Peutz-Jeghers syndrome. Hum Genet 124:300. Seemann MD (2004) PET/CT: fundamental principles. Eur J Med Res 9:241–246. Selak MA, Armour SM, MacKenzie ED, Boulahbel H, Watson DG, Mansfield KD, Pan Y, Simon MC, Thompson CB, Gottlieb E (2005) Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell 7:77–85. Semenza GL (2003) Targeting HIF-1 for cancer therapy. Nat Rev Cancer 3:721–732. Semenza GL (2007) HIF-1 mediates the Warburg effect in clear cell renal carcinoma. J Bioenerg Biomembr 39:231–234. Sharp FR, Bernaudin M (2004) HIF1 and oxygen sensing in the brain. Nat Rev Neurosci 5: 437–448. Shaw RJ (2006) Glucose metabolism and cancer. Curr Opin Cell Biol 18:598–608. Shaw RJ, Bardeesy N, Manning BD, Lopez L, Kosmatka M, DePinho RA, Cantley LC (2004a) The LKB1 tumor suppressor negatively regulates mTOR signaling. Cancer Cell 6:91–99. Shaw RJ, Kosmatka M, Bardeesy N, Hurley RL, Witters LA, DePinho RA, Cantley LC (2004b) The tumor suppressor LKB1 kinase directly activates AMP-activated kinase and regulates apoptosis in response to energy stress. Proc Natl Acad Sci USA 101:3329–3335. Shi Q, Le X, Wang B, Abbruzzese JL, Xiong Q, He Y, Xie K (2001) Regulation of vascular endothelial growth factor expression by acidosis in human cancer cells. Oncogene 20: 3751–3756. Shim H, Dolde C, Lewis BC, Wu CS, Dang G, Jungmann RA, Dalla-Favera R, Dang CV (1997) c-Myc transactivation of LDH-A: implications for tumor metabolism and growth. Proc Natl Acad Sci USA 94:6658–6663. Skulachev VP (1994) Chemiosmotic concept of the membrane bioenergetics: what is already clear and what is still waiting for elucidation? J Bioenerg Biomembr 26:589–598. Smith TA (2000) Mammalian hexokinases and their abnormal expression in cancer. Br J Biomed Sci 57:170–178. Smith TA, Sharma RI, Thompson AM, Paulin FE (2006) Tumor 18F-FDG incorporation is enhanced by attenuation of P53 function in breast cancer cells in vitro. J Nucl Med 47: 1525–1530. Sonveaux P, Vegran F, Schroeder T, Wergin MC, Verrax J, Rabbani ZN, De Saedeleer CJ, Kennedy KM, Diepart C, Jordan BF et al (2008) Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J Clin Investig 118:3930–3942. Stacpoole PW, Greene YJ (1992) Dichloroacetate. Diabetes Care 15:785–791. Suter M, Riek U, Tuerk R, Schlattner U, Wallimann T, Neumann D (2006) Dissecting the role of 5’-AMP for allosteric stimulation, activation, and deactivation of AMP-activated protein kinase. J Biol Chem 281:32207–32216. Swinnen JV, Brusselmans K, Verhoeven G (2006) Increased lipogenesis in cancer cells: new players, novel targets. Curr Opin Clin Nutr Metab Care 9:358–365. Swinnen JV, Vanderhoydonc F, Elgamal AA, Eelen M, Vercaeren I, Joniau S, Van Poppel H, Baert L, Goossens K, Heyns W et al (2000) Selective activation of the fatty acid synthesis pathway in human prostate cancer. Int J Cancer 88:176–179. Thierbach R, Schulz TJ, Isken F, Voigt A, Mietzner B, Drewes G, von Kleist-Retzow JC, Wiesner RJ, Magnuson MA, Puccio H et al (2005) Targeted disruption of hepatic frataxin expression causes impaired mitochondrial function, decreased life span and tumor growth in mice. Hum Mol Genet 14:3857–3864. Thomas DD, Miranda KM, Colton CA, Citrin D, Espey MG, Wink DA (2003) Heme proteins and nitric oxide (NO): the neglected, eloquent chemistry in NO redox signaling and regulation. Antioxid Redox Signal 5:307–317.
118
W. Lu and P. Huang
Tian WN, Braunstein LD, Apse K, Pang J, Rose M, Tian X, Stanton RC (1999) Importance of glucose-6-phosphate dehydrogenase activity in cell death. Am J Physiol 276:C1121–C1131. Tomlinson IP, Alam NA, Rowan AJ, Barclay E, Jaeger EE, Kelsell D, Leigh I, Gorman P, Lamlum H, Rahman S et al (2002) Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Nat Genet 30:406–410. Tong X, Zhao F, Thompson CB (2009) The molecular determinants of de novo nucleotide biosynthesis in cancer cells. Curr Opin Genet Dev 19:32–37. Vander Heiden MG, Cantley LC, Thompson CB (2009) Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science (New York, NY) 324:1029–1033. Varshney R, Adhikari JS, Dwarakanath BS (2003) Contribution of oxidative stress to radiosensitization by a combination of 2-DG and 6-AN in human cancer cell line. Ind J Exp Biol 41:1384–1391. Verma M, Kagan J, Sidransky D, Srivastava S (2003) Proteomic analysis of cancer-cell mitochondria. Nat Rev 3:789–795. Vogelstein B, Lane D, Levine AJ (2000) Surfing the p53 network. Nature 408:307–310. Wallace DC (1999) Mitochondrial diseases in man and mouse. Science (New York, NY) 283:1482–1488. Wallace DC (2005) Mitochondria and cancer: Warburg addressed. Cold Spring Harb Symp Quant Biol 70:363–374. Warburg O (1956a) On respiratory impairment in cancer cells. Science 124:269–270. Warburg O (1956b) On the origin of cancer cells. Science (New York, NY) 123:309–314. Weber WA (2006) Positron emission tomography as an imaging biomarker. J Clin Oncol 24:3282–3292. Woods A, Johnstone SR, Dickerson K, Leiper FC, Fryer LG, Neumann D, Schlattner U, Wallimann T, Carlson M, Carling D (2003) LKB1 is the upstream kinase in the AMP-activated protein kinase cascade. Curr Biol 13:2004–2008. Xie H, Valera VA, Merino MJ, Amato AM, Signoretti S, Linehan WM, Sukhatme VP, Seth P (2009) LDH-A inhibition, a therapeutic strategy for treatment of hereditary leiomyomatosis and renal cell cancer. Mol Cancer Ther 8:626–635. Xu RH, Pelicano H, Zhang H, Giles FJ, Keating MJ, Huang P (2005a) Synergistic effect of targeting mTOR by rapamycin and depleting ATP by inhibition of glycolysis in lymphoma and leukemia cells. Leukemia 19:2153–2158. Xu RH, Pelicano H, Zhou Y, Carew JS, Feng L, Bhalla KN, Keating MJ, Huang P (2005b) Inhibition of glycolysis in cancer cells: a novel strategy to overcome drug resistance associated with mitochondrial respiratory defect and hypoxia. Cancer Res 65:613–621. Yahagi N, Shimano H, Hasegawa K, Ohashi K, Matsuzaka T, Najima Y, Sekiya M, Tomita S, Okazaki,H, Tamura,Y et al (2005) Co-ordinate activation of lipogenic enzymes in hepatocellular carcinoma. Eur J Cancer 41:1316–1322. Yu Y, Deck JA, Hunsaker LA, Deck LM, Royer RE, Goldberg E, Vander Jagt DL (2001) Selective active site inhibitors of human lactate dehydrogenases A4, B4, and C4. Biochem Pharmacol 62:81–89. Zastawny TH, Dabrowska M, Jaskolski T, Klimarczyk M, Kulinski L, Koszela A, Szczesniewicz M, Sliwinska M, Witkowski P, Olinski R (1998) Comparison of oxidative base damage in mitochondrial and nuclear DNA. Free Radical Biol Med 24:722–725. Zhang S, Yang JH, Guo CK, Cai PC (2007) Gene silencing of TKTL1 by RNAi inhibits cell proliferation in human hepatoma cells. Cancer Lett 253:108–114.
Part II
Malignant Cells
Chapter 6
Aberrant DNA Methylation in Cancer Cells Toshikazu Ushijima
Abstract DNA methylation is inherited upon somatic cell division, and methylation of a promoter CpG island silences its downstream gene. Methylation alterations in cancer cells are characterized by aberrant methylation of specific promoter CpG islands, which can cause inactivation of both tumor-suppressor genes (driver) and other genes (passenger), and by global hypomethylation. Recent studies revealed that aberrant methylation is present even in non-cancerous tissues, and its level is associated with cancer risk (an epigenetic field for cancerization). Quantification of methylation revealed that aberrant methylation can be induced much more frequently than mutations, and it has already been indicated that methylation alterations are involved in epithelial-mesenchymal transition. The high frequency of methylation alterations also suggested that they could be involved in phenotypic changes of stromal cells, and thus formation of cancer microenvironment. DNA methylation alterations are likely to be important players not only in transformation of epithelial cells but also in the formation of cancer microenvironment by stromal cells.
Introduction Epigenetic modifications are responsible for long-term cellular memory of gene expression, and play critical roles in development, differentiation, and reprogramming (Bird 2007). To play these roles, epigenetic modifications must be faithfully replicated upon DNA replication. DNA methylation and histone modifications are known to have such characteristics, and DNA methylation is especially known for its high fidelity in replication (Ushijima et al. 2003; Laird et al. 2004; Riggs and Xiong 2004). It is generally considered that, during developmental processes, epigenetic modifications of the genome (epigenome) are established as a T. Ushijima (*) National Cancer Center Research Institute, 5-1-1 Tsukiji, Chuo-ku, Tokyo, Japan e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_6, © Springer Science+Business Media, LLC 2010
121
122
T. Ushijima
c onsequence of interactions between cellular lineage and environmental input (Gan et al. 2007). It was previously believed that aberrant DNA methylation occurs rarely, but that, once induced in critical genes, it can be causally involved in development of diseases, such as cancers. However, recent studies indicate that DNA methylation alterations of specific genes can be induced relatively frequently under specific conditions. Epigenetic changes were shown to be physiologically involved in determination of lymphocyte differentiation (Kioussis and Georgopoulos 2007). When gastric epithelial cells are exposed to a specific pathological condition, several to several 10% of them can have aberrant methylation of specific genes (Maekita et al. 2006). These widespread occurrences of epigenetic changes in somatic cells indicate a possibility that they can be responsible for formation of tumor microenvironments, although cells involved in it are polyclonal. This chapter will introduce the basic characteristics of DNA methylation and methylation alterations in cancer cells, and finally discuss possible involvement of altered methylation in formation of tumor microenvironments.
Characteristics of DNA Methylation DNA is physiologically methylated at the 5 position of cytosines at some CpG sites (Fig. 6.1a). This physiological methylation is different from DNA methylations at O6 and N7 positions of guanines, which are abnormal adducts produced by alkylating agents. DNA methylation is characterized by its stable main tenance upon somatic cell division, and critical roles in regulation of gene transcription.
Maintenance of DNA Methylation Statuses When a CpG site is methylated, cytosines on both strands are methylated (Fig. 6.1b). At DNA replication, a newly synthesized DNA strand does not contain methyl groups, and hemi-methylated CpG sites are temporarily formed. However, a maintenance methylase, DNA methyltransferase 1 (DNMT1), associated with a replication fork (Hermann et al. 2004b), restores those hemi-methylated CpG sites into fully methylated CpG sites. DNMT1 has much lower activity on unmethylated CpG sites, and unmethylated CpG sites are kept unmethylated. Therefore, DNA methylation patterns are replicated at somatic DNA replication with a high fidelity (~99.9%), especially in CpG islands (Ushijima et al. 2003; Laird et al. 2004; Riggs and Xiong 2004).
6 Aberrant DNA Methylation in Cancer Cells
a
123
NH2 N
O
3
NH2
1 N
N
O
dR
dR
Cytosine
b 5' 3'
CH3
N
5
5-methylcytosine
M CG
M CG
GC M
GC M
CG
CG
3'
GC
GC
5'
DNA replication
5'
M CG
M CG
CG
CG
3'
3'
GC
GC
GC
GC
5'
5'
CG
CG
CG
CG
3'
3'
GC M
GC M
GC
GC
5'
DNMT1
5' 3'
5' 3'
M CG
M CG
GC M M CG GC M
CG
CG
3'
GC M M CG
GC
GC
5'
CG
CG
3'
GC M
GC
GC
5'
Fig. 6.1 DNA methylation and its maintenance upon DNA replication. (a) Chemical structures of cytosine and 5-methylcytosine. (b) Maintenance of methylated and unmethylated statuses of CpG sites. Upon DNA replication, hemimethylated CpG sites are temporarily formed, but DNMT1 at replication fork restores hemimethylated CpG sites specifically into fully methylated CpG sites
124
T. Ushijima
Regulation of Gene Transcription by DNA Methylation DNA methylation of a CpG island in a gene promoter region has been known to be consistently associated with transcriptional repression of its downstream gene (Baylin and Ohm 2006; Ushijima 2005). Also, in recent genome-wide analysis of DNA methylation and gene expression, the consistent association was clearly observed (Weber et al. 2007; Rauch et al. 2009; Yamashita et al. in press). In contrast, methylation of CpG islands in gene body regions is often, but not consistently, associated with increased gene transcription (Hellman and Chess 2007; Rauch et al. 2009; Yamashita et al. in press). The consistent association between promoter methylation and transcriptional repression is now known to be due to the presence of a cause–consequence relationship, mediated by several mechanisms. First, an approximately 200-bp region just upstream of a transcription start site usually lacks a nucleosome, designated as a nucleosome-free region (NFR) (Fig. 6.2a) (Lee et al. 2004; Li et al. 2007; Ozsolak et al. 2007). However, if the NFR is methylated, a nucleosome is formed, and transcription from the region is severely repressed (Fig. 6.2b) (Lin et al. 2007). Second, DNA methylation is recognized by methylated DNA binding proteins, such as MeCP2 and MBDs, and these proteins recruit histone deacetylases (Richards and Elgin 2002) and a histone methyltransferase, SUV39H1 (Fujita et al. 2003). Deacetylated histones are known to be positively charged and to associate tightly with DNA, inhibiting accession of transcription complexes to DNA. SUV39H1 is a
b
Unmethylated CpG site Methylated CpG site
Fig. 6.2 Transcriptional repression by DNA methylation of a CpG island in a promoter region. (a) A NFR is present when DNA is kept unmethylated. (b) Nucleosome is formed when DNA in the NFR is methylated, and transcription is strongly repressed
6 Aberrant DNA Methylation in Cancer Cells
125
known to be involved in formation of a heterochromatin structure (Stewart et al. 2005). Thirdly, even in naked DNA, some transcription factors, such as CTCF, are known to have much less binding capacity with methylated CpG sites (Bird 2002).
Maintenance and De Novo DNA Methylases DNA methyltransferases are essential machineries to establish and maintain DNA methylation, and individual DNA methyltransferases are known as maintenance methylase, DNMT1, and de novo methylases, DNMT3A, and DNMT3B, respectively (Hermann et al. 2004a). As mentioned above, DNMT1 plays a major role in maintaining DNA methylation upon DNA replication. Homozygous knockout of Dnmt1 is lethal in midgestation (Li et al. 1992). Dnmt3a and Dnmt3b have the activity of methylating CpG sites without any preference for hemimethylated CpG sites (Okano et al. 1998). While Dnmt3a cannot methylate nucleosomal DNA, Dnmt3b can (Takeshima et al. 2006). Homozygous knockout of Dnmt3a causes lethality after birth (Okano et al. 1999), and Dnmt3a is essential in establishment of genomic imprinting (Kaneda et al. 2004b). Homozygous knockout of Dnmt3b causes lethality before birth, and germline mutations of DNMT3B cause a recessive inherited disorder, ICF syndrome, in humans (Okano et al. 1999).
Methylation Alterations in Cancer Cells DNA methylation patterns of the genome (methylome) in normal cells are precisely regulated according to developmental stages (Meissner et al. 2008; Rauch et al. 2009). In brief, the vast majority of CpG islands are kept unmethylated, and repetitive sequences, which consist of more than 40% of the genome (Lander et al. 2001), are heavily methylated. In cancer cells, the presence of altered DNA methylation has been long known, and the alteration is characterized by “genome-overall hypomethylation and regional hypermethylation” (Fig. 6.3).
Genome-Overall Hypomethylation The “genome-overall” hypomethylation was discovered in the early 1980s, and the phenomenon is now observed in almost all types of cancer cells (Feinberg and Tycko 2004). The decrease of 5-methylcytosine content in the genome is mainly due to hypomethylation of repetitive sequences (Feinberg and Tycko 2004; Kaneda et al. 2004a), but can involve demethylation of normally methylated CpG islands. Demethylation of normally methylated promoter CpG islands leads to aberrant transcription of their downstream genes, and a well known example is aberrant
126
T. Ushijima Normal cell Cancer cell Unmethylated Methylated Repetitive sequence CpG island
Fig. 6.3 DNA methylation changes in cancer cells. In normal cells, the vast majority of CpG islands are kept unmethylated, and repetitive sequences are kept methylated. In cancer cells, a limited number of CpG islands are aberrantly methylated, and some repetitive sequences are demethylated
expression of cancer-testis antigen genes, such as melanoma antigen genes (MAGEs) (de Smet et al. 1999). Also, hypomethylation of a differentially methylated region (DMR) of IGF2 is known for its loss of imprinting, leading to its increased expression and tumor development (Cui et al. 2002). Despite the long history, there are still limited findings on the causal role of genome-overall hypomethylation in carcinogenesis. A mouse strain with genomeoverall hypomethylation, due to introduction of a hypomorphic allele of Dnmt1, demonstrated increased rates of chromosomal loss (Chen et al. 1998). The mouse strain showed increased incidences of lymphomas, colonic microadenomas and liver tumors (Chen et al. 1998; Eden et al. 2003; Yamada et al. 2005). At the same time, the systemic low activity of Dnmt1 led to suppression of macroscopic tumors of the intestine (Laird et al. 1995; Yamada et al. 2005).
Aberrant DNA Methylation of CpG Islands “Regional hypermethylation” denotes methylation of CpG islands that are normally unmethylated. If such methylation is induced in promoter CpG islands of tumor-suppressor genes, it leads to their permanent inactivation, and is causally involved in cancer development and progression (Baylin and Ohm 2006; Jones and Baylin 2007). After the RB tumor-suppressor gene was shown to be inactivated by its promoter methylation (Ohtani-Fujita et al. 1993), many tumorsuppressor genes involved in various cellular processes, such as WNT signalling (SFRP family and CDH1), DNA repair (MLH1 and MGMT), and cell cycle regulation (CDKN2A), were shown to be inactivated by promoter methylation (Baylin and Ohm 2006). In some cancer types, such as gastric cancers, tumor-suppressor genes are inactivated more frequently by promoter methylation than by mutations (Ushijima and Sasako 2004). CpG islands aberrantly methylated in cancers are not limited to CpG islands in promoter regions, and methylation of CpG islands in gene bodies are often associated with increased gene expression (Ushijima 2005; Rauch et al. 2009; Yamashita
6 Aberrant DNA Methylation in Cancer Cells
127
et al. in press). However, the biological meaning of physiological and pathological DNA methylation of gene bodies remains unclear.
Driver Methylation and Passenger Methylation Recent genome-wide screening revealed that several hundred to one thousand promoter CpG islands are methylated in cancer cells (Rauch et al. 2009; Yamashita et al. in press). Most of the genes methylated in cancers are not expressed or have only low expression in normal counterpart cells (Takeshima et al. 2009). These clearly show that promoter CpG islands aberrantly methylated in cancers are not always causally involved in cancer development or progression. According to tradition in the field of mutations, methylation causally involved in cancer development and progression is designated as “driver methylation”, and methylation that simply accompanies the process is designated as “passenger methylation”. Although the distinction between driver and passenger is conceptually clear, actual distinctions between the two types are often difficult, especially when methylation is present in a promoter CpG island of a gene. Although most genes with low or no transcription in normal cells have limited functions, tumor suppressor genes can have only low transcription levels in normal cells when not induced. Although an important tumor-suppressor gene is methylated with a high incidence in specific types of cancers, we cannot conclude that genes with methylation at low incidences are not important. At least, to demonstrate that a gene with promoter methylation in cancers is a driver, its expression in normal counterpart cells and gene function at physiological expression levels should be examined. If a gene is inactivated by mutations, the gene is likely to be a tumor-suppressor gene.
Possible Involvement of Altered Methylation in Tumor Microenvironments Traditionally, aberrant DNA methylation was considered to be similar to mutations in terms of its frequency, target genes, and inducers. However, it was recently indicated that DNA methylation of specific genes can be induced in a significant fraction of cells even in polyclonal tissue, and there is a possibility that altered methylation statuses can be involved in formation of tumor microenvironments.
Unique Natures of Aberrant DNA Methylation, in Contrast with Mutations Recent findings showed unique natures of DNA methylation in contrast with mutations. First, as described above, aberrant methylation of promoter CpG islands, which inactivate their downstream genes, is induced in several hundred to
128
T. Ushijima
one thousand genes in cancer cells (Rauch et al. 2009; Yamashita et al. in press) while a cancer cell is estimated to harbor mutations of 90 genes (Sjoblom et al. 2006). Second, aberrant methylation of a specific gene can be present in several 10% of cells in non-cancerous tissues (Maekita et al. 2006; Ushijima 2007) while mutations of a specific gene are quite infrequent, being less than 1 out of 105 cells (Nagao et al. 2001). Third, aberrant methylation is induced in specific genes (Nakajima et al. 2009; Oka et al. 2009) while mutations are considered to occur in random genes. Fourth, chronic inflammation is an important inducer of aberrant methylation, while mutagenic chemicals and radiations induce mutations.
Field for Cancerization and DNA Methylation The presence of aberrant DNA methylation in non-cancerous tissues is important from the viewpoint of the field cancerization. It has long been known that some cancers tend to have multiple independent occurrences, suggesting that there is a field for cancerization produced by exposure to environmental and other carcinogenic stimuli (Braakhuis et al. 2003). It was recently clarified that the degree of accumulation of aberrant methylation in gastric mucosae correlates with risk of developing gastric cancers (Maekita et al. 2006; Nakajima et al. 2006). Similar associations between methylation in non-cancerous tissues and cancer development were obtained in various cancers, such as liver, colorectal, breast, and urothelial cancers (Kondo et al. 2000; Shen et al. 2005; Yan et al. 2006; Arai et al. 2006), and such accumulation is considered to represent an “epigenetic field for cancerization” (epigenetic field defect) (Fig. 6.4) (Ushijima 2007). The epigenetic field for cancerization can be utilized for cancer risk diagnosis that takes account of an individual’s own life history.
Epithelial-Mesenchymal Transition and DNA Methylation Epithelial-mesenchymal transition (EMT) refers to a phenomenon in which epithelial cells lose their epithelial characteristics and acquire mesenchymal features (Polyak and Weinberg 2009). EMT has been known to be critically involved in embryonic development and tumor progression (Polyak and Weinberg 2009; Dumont et al. 2008). In addition, it was recently demonstrated that epithelial cells acquire stem cell features when they undergo EMT, and that stem-like cells in normal epithelium and cancer tissues have EMT markers (Mani et al. 2008). Physiological EMT, which takes place in a group of cells and cannot be attributed to a mutation in a cell, is likely to be due to epigenetic changes (Polyak and Weinberg 2009). Also, EMT involved in tumor progression, which can be due to a mutation or an epigenetic alteration, was recently shown to be mediated by DNA methylation (Dumont et al. 2008; Yang et al. 2009).
6 Aberrant DNA Methylation in Cancer Cells
Noncancerous predisposed tissue
Normal tissue 1
cells
2 3
129
Cancer development
Exposure to carcinogenic factors
4 5 6
Methylated CpG island Unmethylated CpG island
Fig. 6.4 Epigenetic field for cancerization. In a normal tissue, the vast majority of CpG islands are unmethylated. When it is exposed to some carcinogenic stimulus, such as Helicobacter pylori infection, various but specific genes, most of which are passengers, are methylated in various cells. If further genetic or epigenetic alterations take place, clinical cancer will develop. The amount of aberrant DNA methylation accumulated in a tissue can be used as a cancer risk marker
Tumor Microenvironments and DNA Methylation Not only is aberrant DNA methylation present in non-cancerous tissues, but also the degree can be severe, reaching several 10%. If a specific gene is inactivated in a significant fraction of cells in a tissue, it can affect the overall function of the tissue. Functional change of a tissue due to the presence of somatic mutations has been considered very unlikely because the population of cells with mutation of a specific gene is very small. This unique nature of aberrant DNA methylation suggests that it can be involved in pathological conditions without any clonal expansion, including formation of cancer microenvironments and any diseases that involve persistent gene repression. In cancer stromal cells, persistent changes in gene expression have been already reported (Finak et al. 2008; Hu et al. 2005; Lin et al. 2008), and involvement of mutations in such changes was shown to be unlikely (Qiu et al. 2008). Moreover, epigenetic silencing of specific genes due to changes in histone modification has been reported in tumor endothelial cells (Hellebrekers et al. 2007). Mesenchymal cells produced by EMT of tumor cells may contribute to tumor
130
T. Ushijima
microenvironments. Clarification of DNA methylation changes in cancer stromal cells is becoming an important issue.
Epilogue Unique natures of DNA methylation, in sharp contrast with mutations, indicate that altered DNA methylation is involved in formation of tumor microenvironments. Although I did not describe technological aspects of methylation analysis in this chapter, all the technologies for genome-wide screening are not capable of identifying methylation changes present only in a fraction of cells. Therefore, innovation in technology or approach seems to be necessary to identify critical DNA methylation alterations in cancer stromal cells. Even without such an innovation, sensitive analysis of candidate genes might greatly contribute to clarification of genes critically involved in tumor microenvironments. Acknowledgments The author is grateful to Dr. Eriko Okochi for her critical reading of the manuscript. The author’s research described here was supported by a Grant-in-Aid for the 3rd-term Comprehensive Cancer Control Strategy from Ministry of Health, Labor, and Welfare, Japan.
References Arai E, Kanai Y, Ushijima S et al (2006) Regional DNA hypermethylation and DNA methyl transferase (DNMT) 1 protein overexpression in both renal tumors and corresponding nontumorous renal tissues. Int J Cancer 119:288–296 Baylin SB, Ohm JE (2006) Epigenetic gene silencing in cancer – a mechanism for early oncogenic pathway addiction? Nat Rev Cancer 6:107–116 Bird A (2002) DNA methylation patterns and epigenetic memory. Genes Dev 16:6–21 Bird A (2007) Perceptions of epigenetics. Nature 447:396–398 Braakhuis BJ, Tabor MP, Kummer JA et al (2003) A genetic explanation of Slaughter’s concept of field cancerization: evidence and clinical implications. Cancer Res 63:1727–1730 Chen RZ, Pettersson U, Beard C et al (1998) DNA hypomethylation leads to elevated mutation rates. Nature 395:89–93 Cui H, Onyango P, Brandenburg S et al (2002) Loss of imprinting in colorectal cancer linked to hypomethylation of H19 and IGF2. Cancer Res 62:6442–6446 de Smet C, Lurquin C, Lethe B et al (1999) DNA methylation is the primary silencing mechanism for a set of germ line- and tumor-specific genes with a CpG-rich promoter. Mol Cell Biol 19: 7327–7335 Dumont N, Wilson MB, Crawford YG et al (2008) Sustained induction of epithelial to mesenchymal transition activates DNA methylation of genes silenced in basal-like breast cancers. Proc Natl Acad Sci USA 105:14867–14872 Eden A, Gaudet F, Waghmare A et al (2003) Chromosomal instability and tumors promoted by DNA hypomethylation. Science 300:455 Feinberg AP, Tycko B (2004) The history of cancer epigenetics. Nat Rev Cancer 4:143–153 Finak G, Bertos N, Pepin F et al (2008) Stromal gene expression predicts clinical outcome in breast cancer. Nat Med 14:518–527
6 Aberrant DNA Methylation in Cancer Cells
131
Fujita N, Watanabe S, Ichimura T et al (2003) Methyl-CpG binding domain 1 (MBD1) interacts with the Suv39h1-HP1 heterochromatic complex for DNA methylation-based transcriptional repression. J Biol Chem 278:24132–24138 Gan Q, Yoshida T, McDonald OG et al (2007) Concise review: epigenetic mechanisms contribute to pluripotency and cell lineage determination of embryonic stem cells. Stem Cells 25:2–9 Hellebrekers DM, Melotte V, Vire E et al (2007) Identification of epigenetically silenced genes in tumor endothelial cells. Cancer Res 67:4138–4148 Hellman A, Chess A (2007) Gene body-specific methylation on the active X chromosome. Science 315:1141–1143 Hermann A, Gowher H, Jeltsch A (2004a) Biochemistry and biology of mammalian DNA methyltransferases. Cell Mol Life Sci 61:2571–2587 Hermann A, Goyal R, Jeltsch A (2004b) The Dnmt1 DNA-(cytosine-C5)-methyltransferase methylates DNA processively with high preference for hemimethylated target sites. J Biol Chem 279:48350–48359 Hu M, Yao J, Cai L et al (2005) Distinct epigenetic changes in the stromal cells of breast cancers. Nat Genet 37:899–905 Jones PA, Baylin SB (2007) The epigenomics of cancer. Cell 128:683–692 Kaneda A, Tsukamoto T, Takamura-Enya T et al (2004a) Frequent hypomethylation in multiple promoter CpG islands is associated with global hypomethylation, but not with frequent promoter hypermethylation. Cancer Sci 95:58–64 Kaneda M, Okano M, Hata K et al (2004b) Essential role for de novo DNA methyltransferase Dnmt3a in paternal and maternal imprinting. Nature 429:900–903 Kioussis D, Georgopoulos K (2007) Epigenetic flexibility underlying lineage choices in the adaptive immune system. Science 317:620–622 Kondo Y, Kanai Y, Sakamoto M et al (2000) Genetic instability and aberrant DNA methylation in chronic hepatitis and cirrhosis – a comprehensive study of loss of heterozygosity and microsatellite instability at 39 loci and DNA hypermethylation on 8 CpG islands in microdissected specimens from patients with hepatocellular carcinoma. Hepatology 32:970–979 Laird PW, Jackson-Grusby L, Fazeli A et al (1995) Suppression of intestinal neoplasia by DNA hypomethylation. Cell 81:197–205 Laird CD, Pleasant ND, Clark AD et al (2004) Hairpin-bisulfite PCR: assessing epigenetic methylation patterns on complementary strands of individual DNA molecules. Proc Natl Acad Sci USA 101:204–209 Lander ES, Linton LM, Birren B et al (2001) Initial sequencing and analysis of the human genome. Nature 409:860–921 Lee CK, Shibata Y, Rao B et al (2004) Evidence for nucleosome depletion at active regulatory regions genome-wide. Nat Genet 36:900–905 Li E, Bestor TH, Jaenisch R (1992) Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 69:915–926 Li B, Carey M, Workman JL (2007) The role of chromatin during transcription. Cell 128: 707–719 Lin JC, Jeong S, Liang G et al (2007) Role of nucleosomal occupancy in the epigenetic silencing of the MLH1 CpG island. Cancer Cell 12:432–444 Lin HJ, Zuo T, Lin CH et al (2008) Breast cancer-associated fibroblasts confer AKT1-mediated epigenetic silencing of Cystatin M in epithelial cells. Cancer Res 68:10257–10266 Maekita T, Nakazawa K, Mihara M et al (2006) High levels of aberrant DNA methylation in Helicobacter pylori-infected gastric mucosae and its possible association with gastric cancer risk. Clin Cancer Res 12:989–995 Mani SA, Guo W, Liao MJ et al (2008) The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell 133:704–715 Meissner A, Mikkelsen TS, Gu H et al (2008) Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature 454:766–770 Nagao M, Ochiai M, Okochi E et al (2001) LacI transgenic animal study: relationships among DNA-adduct levels, mutant frequencies and cancer incidences. Mutat Res 477:119–124
132
T. Ushijima
Nakajima T, Maekita T, Oda I et al (2006) Higher methylation levels in gastric mucosae significantly correlate with higher risk of gastric cancers. Cancer Epidemiol Biomarkers Prev 15:2317–2321 Nakajima T, Yamashita S, Maekita T et al (2009) The presence of a methylation fingerprint of Helicobacter pylori infection in human gastric mucosae. Int J Cancer 124:905–910 Ohtani-Fujita N, Fujita T, Aoike A et al (1993) CpG methylation inactivates the promoter activity of the human retinoblastoma tumor-suppressor gene. Oncogene 8:1063–1067 Oka D, Yamashita S, Tomioka T et al (2009) The presence of aberrant DNA methylation in noncancerous esophageal mucosae in association with smoking history: a target for risk diagnosis and prevention of esophageal cancers. Cancer 115:3412–3426 Okano M, Xie S, Li E (1998) Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat Genet 19:219–220 Okano M, Bell DW, Haber DA et al (1999) DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99:247–257 Ozsolak F, Song JS, Liu XS et al (2007) High-throughput mapping of the chromatin structure of human promoters. Nat Biotechnol 25:244–248 Polyak K, Weinberg RA (2009) Transitions between epithelial and mesenchymal states: acquisition of malignant and stem cell traits. Nat Rev Cancer 9:265–273 Qiu W, Hu M, Sridhar A et al (2008) No evidence of clonal somatic genetic alterations in cancerassociated fibroblasts from human breast and ovarian carcinomas. Nat Genet 40:650–655 Rauch TA, Wu X, Zhong X et al (2009) A human B cell methylome at 100-base pair resolution. Proc Natl Acad Sci USA 106:671–678 Richards EJ, Elgin SC (2002) Epigenetic codes for heterochromatin formation and silencing: rounding up the usual suspects. Cell 108:489–500 Riggs AD, Xiong Z (2004) Methylation and epigenetic fidelity. Proc Natl Acad Sci USA 101: 4–5 Shen L, Kondo Y, Rosner GL et al (2005) MGMT promoter methylation and field defect in sporadic colorectal cancer. J Natl Cancer Inst 97:1330–1338 Sjoblom T, Jones S, Wood LD et al (2006) The consensus coding sequences of human breast and colorectal cancers. Science 314:268–274 Stewart MD, Li J, Wong J (2005) Relationship between histone H3 lysine 9 methylation, transcription repression, and heterochromatin protein 1 recruitment. Mol Cell Biol 25:2525–2538 Takeshima H, Suetake I, Shimahara H et al (2006) Distinct DNA methylation activity of Dnmt3a and Dnmt3b towards naked and nucleosomal DNA. J Biochem 139:503–515 Takeshima H, Yamashita S, Shimazu T et al (2009) The presence of RNA polymerase II, active or stalled, predicts epigenetic fate of promoter CpG islands. Genome Res 19:1974–1982 Ushijima T (2005) Detection and interpretation of altered methylation patterns in cancer cells. Nat Rev Cancer 5:223–231 Ushijima T (2007) Epigenetic field for cancerization. J Biochem Mol Biol 40:142–150 Ushijima T, Sasako M (2004) Focus on gastric cancer. Cancer Cell 5:121–125 Ushijima T, Watanabe N, Okochi E et al (2003) Fidelity of the methylation pattern and its variation in the genome. Genome Res 13:868–874 Weber M, Hellmann I, Stadler MB et al (2007) Distribution, silencing potential and evolutionary impact of promoter DNA methylation in the human genome. Nat Genet 39:457–466 Yamada Y, Jackson-Grusby L, Linhart H et al (2005) Opposing effects of DNA hypomethylation on intestinal and liver carcinogenesis. Proc Natl Acad Sci USA 102:13580–13585 Yamashita S, Hosoya K, Gyobu K et al (2009) Development of a novel output value for quantitative assessment in methylated DNA immunoprecipitation-CpG island microarray analysis. DNA Research 16:275–286 Yan PS, Venkataramu C, Ibrahim A et al (2006) Mapping geographic zones of cancer risk with epigenetic biomarkers in normal breast tissue. Clin Cancer Res 12:6626–6636 Yang X, Pursell B, Lu S et al (2009) Regulation of beta 4-integrin expression by epigenetic modifications in the mammary gland and during the epithelial-to-mesenchymal transition. J Cell Sci 122:2473–2480
Chapter 7
DNA Repair and Redox Signaling Mark R. Kelley, Millie M. Georgiadis, and Melissa L. Fishel
Abstract DNA repair pathways have long been recognized as important in protecting the genome from damage caused by endogenous and exogenous DNA damaging agents. However, the role of reduction–oxidation (redox) signaling as a regulatory mechanism in these pathways is a comparatively recent discovery. We have attempted to cover the general aspects of DNA repair, while focusing more attention on the interface between redox regulation and DNA repair activity and response in mammalian cells. The primary focus of this chapter is on APE1, the only DNA repair protein currently known to serve a dual role as a DNA repair enzyme and redox signaling factor. APE1 has been shown to regulate a number of downstream transcription factors (TF) such as HIF-1a, Egr-1, AP-1, NFkB, CREB, p53 and an increasing number of other TFs and other proteins. Additionally, this redox signaling function of APE1 indirectly alters the activity of other DNA repair pathways through its regulation of TF binding to DNA. Included in this chapter is also an overview of general redox systems. We also discuss the role small molecule inhibitors play in the modulation of APE1 redox activity and the potential for chemotherapeutic development via targeting redox regulation of DNA repair. Abbreviations 8-oxoG AAG AGT Ang II AP AP-1
8 oxoguanine alkyladenine DNA glycosylase O-6-alkylguanine-DNA methyltransferase angiotension II apurinic/apyrimidinic activator protein 1
M.R. Kelley (*) Department of Pediatrics, Section of Hematology/Oncology, Herman B Wells Center for Pediatric Research, Indiana University, Indianapolis, IN 46202, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_7, © Springer Science+Business Media, LLC 2010
133
134
M.R. Kelley et al.
APE1 apurinic/apyrmidinic endonuclease 1 ATF activating transcription factor ATFa activating transcription factor a ATF2 activating transcription factor 2 ATF3 activating transcription factor 3 ATF4 activating transcription factor 4 ATM ataxia telangiectasia mutated protein BER base excision repair BKca Large-conductance Ca2+-activated K+channels BLM Bloom’s syndrome gene product BRCA1 breast cancer 1, early onset BRCA2 breast cancer 2, early onset BSO buthionine sulfoximine CD40L CD40 ligand CHO Chinese hamster ovary cells CREB cyclic AMP response-element binding protein CSA Cockayne’s syndrome protein A CSB Cockayne’s syndrome protein B CTL cytotoxic T lymphocyte DDB1 DNA damage binding protein 1 DDB2 DNA damage binding protein 2 DNA-PKcs DNA-dependent kinase catalytic subunit DSBs double-strand breaks DR direct repair dRP deoxyribophosphate dRPase deoxyribophosphodiesterase E3330 2E-3-[5-(2, 3 dimethoxy-6-methyl-1, 4-benzoquinolyl)]-2-nonyl-2propenoic acid Egr-1 early growth response protein 1 ERCC1 excision repair cross-complementation group 1 ERCC3 excision repair cross-complementation group 3 ERCC5 excision repair cross-complementation group 5 flavin adenine dinucleotide dehydrogenase FADH2 FEN1 flap endonuclease 1 Flk-1/KDR fetal liver kinase 1/kinase insert domain receptor FosB FBJ murine osteosarcoma viral oncogene homolog B FPG formamidopyrimidine DNA glycosylase Fra-1 fos related antigen 1 Fra-2 fos related antigen 2 GAPDH glyceraldehyde-3-phosphate dehydrogenase GGR global genome repair GRX glutaredoxin GSH glutathione GSSG glutathione disulfide GTPase Guanosine-5’-triphosphate
7 DNA Repair and Redox Signaling
HEK293 HeLa HIF-1 HIF-1a HR HRE HR32B HUVES IL-2 LP-BER Lys6 Lys7 MeSeH MGMT MiTF MLH1 MMR MMS MPG MRN MSH2 MSH3 MSH6 MSI MYH NAC NADPH Nbs1 NEIL NER NFkB NHEJ NK NTH O6-meG OGG1 P2Y PARP1 PCEC PCNA PEBP-2 PMS2 PRX Pol b Pol d
Human embryonic kidney 293 cells Henrietta Lacks cells hypoxia-inducible factor-1 hypoxia-inducible factor-1 alpha homologous recombination hypoxia response elements human homologue of the yeast RAD23 protein human umbilical vein endothelial cells interleukin 2 long-patch base excision repair lysine 6 lysine 7 methylselenol O6-methylguanine-DNA methyltransferase Microphthalmia-associated transcription factor MutL homolog 1 mismatch repair methyl methane sulfonate N-methylpurine DNA glycosylase Mre11/Rad50/Nbs1 MutS homolog 2 MutS homolog 3 MutS homolog 6 microsatellite instability MutY homolog N-acetyl-L-cysteine Nicotinamide adenine dinucleotide phosphate Nijmegen breakage syndrome 1 fgp/nei family DNA glycosylase nucleotide excision repair nuclear factor kappa B non-homologous end joining natural killer homolog of E. coli endonuclease III (nth) O6 methyl guanine oxoguanine glycosylase 1 purinergic receptor poly ADP ribose polymerase 1 pancreatic cancer-associated endothelial cells proliferating cell nuclear antigen phosphatidylethanolamine-binding protein 2 postmeiotic segregation increased 2 peroxiredoxin DNA polymerase beta DNA polymerase delta
135
136
M.R. Kelley et al.
Pol e DNA polymerase epsilon PMS1 post-meiotic-segregation increased-1 protein PMS2 post-meiotic-segregation increased-2 protein PTEN Phosphatase and tensin homolog deleted on chromosome 10 Ref-1 redox effector factor 1 RF-C replication factor C rRNA ribosomal RNA RNA Pol II RNA polymerase II ROS reactive oxygen species RPA replication protein A SeMet selenomethionine SETMAR SET domain and mariner transposase fusion siRNA small interfering RNA SMUG1 mammalian 5-formyluracil DNA glycosylase SP-BER short-patch base excision repair ssDNA single-strand DNA TCR transcription-coupled repair TDG thymine-DNA glycosylase TFIIH transcription factor IIH TNF-a tumor necrosis factor alpha Trx thioredoxin TrxR thioredoxin reductase TSH thyroid stimulating hormone TTF-1 thyroid transcription factor 1 UNG uracil-DNA glycosylase UV-DDB UV-damaged DNA-binding protein WRN Werner protein, deficient in Werner’s syndrome XPB xeroderma pigmentosum complementary group B protein XPC xeroderma pigmentosum complementary group C protein XPD xeroderma pigmentosum complementary group D protein XPF1 xeroderma pigmentosum complementary group F protein XPG xeroderma pigmentosum group XRCC1 X-ray repair complementing defective repair in Chinese hamster cells 1 XRCC4 X-ray repair complementing defective repair in Chinese hamster cells 4
Introduction Genomic stability relies on efficient, effective methods of repairing DNA. Unrepaired damage can cause cell cycle arrest, apoptosis or accumulation of mutations (Eker et al. 2009; Hofmann et al. 2002). To protect DNA and keep the genome intact, a number of DNA repair pathways exist. Although much has been written about these pathways, the role of redox regulation in DNA repair is a relatively recent discovery. This chapter discusses the mechanisms for human DNA repair,
7 DNA Repair and Redox Signaling
137
redox signaling and its regulation of DNA repair transcription factors, the unique APE1 protein that participates in both categories of activities, how DNA repair differs in tumor cells and implications for emerging cancer therapies.
Overview of DNA Repair Pathways DNA repair is a never-ending task of all eukaryotic cells, as endogenous and exogenous agents constantly assault DNA. Damage from endogenous agents include oxidation by reactive oxygen species (ROS) from normal metabolic processes, alkylation by agents such as S-adenosylmethionine, adduct formation caused by reactive carbonyl species created during lipid peroxidation, hydrolytic depurination that forms abasic sites and deamination of bases, primarily cytidine (and, to a lesser extent, adenine) (De Bont and van Larebeke 2004). Environmental insults from exogenous agents include chemicals, carcinogens and ultraviolet light, as well as chemotherapeutic agents and ionizing radiation damage (Christmann et al. 2003; Fleck and Nielsen 2004). Failure to repair DNA damage can cause a host of problems that ultimately lead to genomic instability (Arcangeli et al. 1997; Kelley and Fishel 2008). For example, improper DNA damage response in mitotic cells can halt cell cycle progress. In post-mitotic cells, DNA damage may result in cell cycle activation without proper coordination of cell cycle machinery, leading to deleterious events, including potentially lethal DNA double-strand breaks (Kruman et al. 2004; Schwartz et al. 2007). However, eukaryotic cells possess a cadre of DNA repair pathways to correct the damage: direct repair (DR), base excision repair (BER), mismatch repair (MMR), nucleotide excision repair (NER), non-homologous end joining (NHEJ) and homologous recombination (HR) (Hoeijmakers 2001, 2001) (Fig. 7.1). The number of known DNA repair proteins and other factors involved in the cellular response to DNA damage keeps growing as we learn more details about each repair pathway and induction of regulatory networks when the damage persists (Wood et al. 2005). The type of DNA damage dictates which pathway is activated to perform repairs. However, some overlap exists among the pathways. Although they are not redundant, certain pathways can function as “backup systems” to others. In addition, some mechanisms overlap. (See Table 7.1 for examples.) These overlaps and interactions provide the most efficient and effective defense net for preserving the cell genome, whereas reduced repair capacity can lead to genomic modification.
DR The DR pathway, sometimes called direct reversal, repairs damaged bases by removing the alteration to the base – as opposed to removing the damaged base itself. This unique mechanism may be the most efficient of all repair pathways
138
M.R. Kelley et al.
Direct Repair
Base Excision
MGMT
Glycosylase
Mismatch
MSH2/6 MSH2/3
TCR
RNA PoLII, CSA, CSB, XPAB2
APE1
Long Patch
Nucleotide Excision
Short Patch
Non-homologous End Joining (NHEJ)
Homologous Recombination (HR)
GGR
DDB-XPE, XPC, HR23β
Ku70, Ku80
ATM
MLH1-PMS2 MLH1-PMS1 DNA-PK RPA, XPA, XPCTFIIH, XPB, XPC, XPD, XPG, XPFERCC1
EX01, RFC, PCNA, RPA, NA Polδ, Ligase
RFC, PCNA, Polδ/ε, Ligase I
MRN
MRN, Rad51, Rad52, Rad54, BRCA1, BRCA2, PCNA, Polδ/ε, Ligase I
DNA Polµ, XRCC4, Ligase 4, Artemis
Fig. 7.1 Schematic overview of DNA repair pathway. Several DNA repair pathways are involved in maintaining cell genomic stability; these include DR, BER, NER, MMR, HR, and NHEJ. More than 150 proteins are involved. Only selected genes of each pathway are shown here. (Adapted from: Fishel et al. (Fishel and Kelley 2007))
Table 7.1 Examples of overlap in DNA repair pathways Overlap Mechanism DR/MMR In the DR pathway, if O6-methylguanine-DNA methyltransferase (AGT) is unsuccessful in removing O6-methylguanine, the MMR pathway can recognize and fix O6-meG mispairs (Fishel and Wilson 1997) BER/NER The BER pathway is primarily responsible for repairing oxidative DNA damage, but NER can serve as a backup for repairing some minor damage (Farmer et al. 2005) BER/HR If BER does not repair single-strand DNA breaks (SSBs), they may lead to double-strand breaks (DSBs), which HR can repair HR/NHEJ HR also can repair DNA DSBs that the NHEJ pathway fails to process (Essers et al. 2000)
(Kaina et al. 2007). O6-alkylguanine-DNA methyltransferase, or AGT, removes an alkyl group from the O6position of guanine on the damaged base, the alkyl group is transferred to the active site of AGT and AGT is subsequently degraded. AGT performs this function to a lesser extent on the O4position of thymine. This is a stoichiometric reaction, as AGT can work only once before being inactivated. It is essential to repair O6-meG adducts, as they cause thymine mispairing errors during replication, leading to G:C-to-A:T transitions or a strand break.
7 DNA Repair and Redox Signaling
139
Clinical application: small-molecule inhibitors of AGT Indication for utility Increased AGT levels in tumors correlate with tumor resistance How it works
A chemotherapy agent generates lesions at the O 6 position of guanine, causing interstrand crosslink formation, which blocks DNA replication. An AGT inhibitor would allow crosslink formation by a chemotherapeutic agent by blocking AGT’s removal of the O6lesion NOTE: AGT was one of the first DNA repair targets identified for cancer therapeutics
Anti-AGT agents in use
• BG (O6-benzylguanine) • PARP1 inhibitors
Additional applications
A mutant AGT protein is in development to protect the bone marrow from myelosuppression while inhibiting tumors with anti-AGT therapies
BER The BER pathway is responsible for repairing DNA damage caused by oxidation, alkylation, deamination and ionizing radiation (Christmann et al. 2003; Altieri et al. 2008; Evans et al. 2000). This pathway repairs single-base lesions including N-alkylated purines (N3-methyladenine, N7-methylguanine and N3-methylguanine), 8-oxo-7,8-dihydroguanine(8-OxoG), thymine glycols, 5-OH and 6-OH dihydrothymine, uracil glycol, 5-hydroxycytosine, urea residues and other adducts (Aiyer et al. 2008; Bindra and Glazer 2007; Chen and Olkowski 1994). The process starts when a DNA glycosylase recognizes and excises the damaged base, leaving an abasic site. The type of DNA lesion dictates which glycosylase performs this activity. For example, methylpurine DNA glycosylase (MPG, AAG), 8-oxoguanine DNA glycosylase (OGG1), uracil DNA glycosylase (UNG), thymine DNA glycosylase (TDG) and others remove specific lesions as their names suggest (Essers et al. 2000). In addition to acting on a specific target, glycosylases may be monofunctional or bifunctional. Monofunctional glycosylases excise the damaged base to generate an apurinic/apyrimidinic (AP) or abasic site. Bifunctional glycosylases do the same, but also perform a lyase function (Chen et al. 1991; David and Williams 1998) that nicks the phosphodiester backbone 3¢ to the AP site. Monofunctional glycosylases repair alkylated DNA damage, while bifunctional glycosylases repair oxidative DNA damage. In contrast to the plethora of glycosylases that can remove damaged bases in BER, only one enzyme processes BER abasic sites: APE1/Ref-1 (apurinic apyrimidinic endonuclease redox effector factor-1, or APE1). APE1 hydrolyzes the phosphodiester backbone 5¢ to the AP site, creating a normal 3¢ hydroxyl terminus and an abasic 5¢ deoxyribose phosphate (5¢ dRP) terminus. In addition, APE1 recruits other proteins to assist with the repair, which includes processing by various polymerases and ligases. At this stage, DNA repair can progress via the short-patch BER (SP-BER) pathway or the long-patch BER (LP-BER) pathway. These branches of BER differ in the proteins involved, as well as the length of DNA synthesized during the repair process. The mechanisms by which one pathway is chosen over the other are still
140
M.R. Kelley et al.
being investigated; however, the type of damage and the glycosylase that removes the damage appear to be two contributing factors. The predominant BER pathway is SP-BER. SP-BER preferentially repairs normal AP sites, while LP-BER preferentially repairs oxidized and reduced AP sites (Bapat et al. 2009). SP-BER removes single-base damage by DNA polymerase b (Pol b) acting upon APE1’s endonuclease activity. Pol b removes the 5¢ dRP moiety and inserts the correct base at the 3¢ hydroxyl termimus that APE1 creates. DNA ligase III/ XRCC1 closes the nick to finish the repair. LP-BER repairs damage to two or more bases. In LP-BER, flap endonuclease 1 (FEN1) displaces a segment of 3–8 nucleotides surrounding the AP site. A DNA polymerase (d, e, or b), along with the DNA clamp PCNA (proliferating cell nuclear antigen) and replication factor-C (RF-C) inserts the correct nucleotides; then DNA ligase I finishes the repair (Bapat et al. 2009). APE1 is responsible for 95% of the endonuclease activity in the cell; in addition, its ability to recruit repair proteins to the site is critical to both BER sub-pathways (Demple and Harrison 1994; Doetsch and Cunningham 1990) (See Fig. 7.3). APE1’s reduction/oxidation signaling also influences DNA repair indirectly via modulation of transcription factors, as discussed later in this chapter. APE1’s multifunctional abilities are unique, and there is effectively no backup to any of APE1’s intracellular activities. Clinical application: small-molecule inhibitors of APE1 Indication for APE1 is responsible for almost all endonuclease activity in both branches utility of the BER pathway. Because it performs direct DNA repair functions and also influences DNA transcription proteins through redox regulation, APE1 is an attractive target for chemotherapeutics How it works Potentially two ways: • Inhibiting APE1 can sensitize tumor cells to alkylating and oxidizing agents • Some chemotherapeutic agents, including platinum-based drugs and anthracyclines, generate ROS as a by-product, which could inhibit APE1’s redox functions. Keeping transcription factors in an oxidized state adversely affects gene transcription Anti-BER agents None yet in use A dozen APE1-inhibiting compounds are under investigation. However, many challenges exist in creating a usable inhibitory APE1 agent. Some compounds have proven to be non-specific, binding to the AP site instead of to APE1; others have difficulty penetrating the cell membrane or reaching APE1’s location in the nucleus. To date, E3330 is the sole compound that can selectively inhibit APE1’s redox functions only
MMR The MMR pathway recognizes and repairs single base-pair mismatches (A-G, T-C) or misaligned short nucleotide repeats (e.g., small-loop insertions). Base–base mismatches can be due to endogenous causes: spontaneous deamination of 5-methylcytosine to
7 DNA Repair and Redox Signaling
141
thymine that causes a guanine-to-thymine mismatch, cytosine deamination to uracil resulting in a guanine-to-uracil mismatch, damage to the cellular nucleotide pool or incorrect incorporation by a DNA polymerase. The process starts when MSH2 (MutS homolog 2) combines with MSH3 or MSH6, depending on which type of mismatch is present. MSH2:MSH6 recognizes both insertion–deletion mispairs and single-base mismatches; the MSH2:MSH3 dimer recognizes only insertion–deletion mispairs. After identifying the type of mismatch, MSH proteins recruit MutL homolog 1 (MLH1) and its binding partners, post-meiotic-segregation increased-1 protein (PMS1) and PMS2, to determine the specific strand error. An exonuclease removes the DNA lesion; a DNA polymerase (d or e) synthesizes a new strand, and a DNA ligase completes the repair (Jiricny 2006; Lopez de Saro et al. 2006). Clinical application: small-molecule inhibitors of MMR Indication for Unclear. Genetic diseases linked to MMR defects (e.g., hereditary utility non-polyposis colon cancer; HNPCC) are deterrents to researching MMR inhibitors. Defective MMR repair functions actually increase tumor resistance to cisplatin How it works Inhibiting MMR enzymes is probably not an effective approach for developing new chemotherapeutics. However, reactivation of MMR enzymes may re-sensitize tumors to chemotherapy Anti-MMR agents None yet in use Experimentally, MLH1 reactivation by decitabine can increase cisplatin sensitivity – but the actual mechanism by which it does this is unknown and may not be specific to MLH reactivation
NER The NER pathway is responsible for repairing bulky DNA lesions: cyclobutane– pyrimidine dimers induced by ultraviolet light, adducts induced by polycyclic aromatic hydrocarbons and other adducts induced by cross-linking agents and basedamaging chemical carcinogens. The NER and MMR pathways are presumed to be the major pathways for repairing cross-link damage, which has significant implications for tumor survival. More than 25 proteins participate in the NER pathway, which can be divided into two sub-pathways: global genome repair (GGR) and transcription-coupled repair (TCR). The distinction is predicated on which complex recognizes the helix-distorting damage (Altieri et al. 2008; Friedberg 2001). After the initial recognition step, both sub-pathways operate the same way. A multi-protein complex binds at the damaged site. Then incisions are made on either side of the damaged strand, several nucleotides away from the actual damage. Both the 5¢ and 3¢ sides are cut. The damage-containing oligonucleotide is removed; a DNA polymerase fills the gap, and a ligase completes the repair. In TCR, recognition occurs when RNA polymerase II (RNA Pol II) stalls at a site of DNA damage. TCR-specific factors, including the Cockayne’s Syndrome
142
M.R. Kelley et al.
proteins, CSA and CSB, are recruited at the site of transcription arrest. The GGR sub-pathway starts when the heterodimer XPC/HR23B recognizes the damage, binds to the damaged DNA (distorting it so other complexes can bind) and recruits other proteins to the site. In addition, the UV-DNA damage binding protein (UV-DDB) is required for recruitment when NER repairs UV-induced cyclobutane pyrimidine dimers. Following recognition, both TCR and GGR use the same proteins to repair the damaged DNA. Transcription factor IIH (TFIIH) is recruited to the site of DNA damage. This large complex includes two helicases, XPB and XPD (xeroderma pigmentosum group B and D proteins), which use ATP energy to unwind the DNA strand on either side of the damage. The helicases work in complementary fashion: XPB unwinds DNA in the 3¢ to 5¢ direction; XPD unwinds DNA in the opposite direction. Then, XPA and RPA (replication protein A) stabilize the exposed single-strand DNA. XPA fine-tunes the location of normal versus abnormal DNA, while RPA protects both of the separated strands in the open complex. Endonucleases XPG, ERCC1 (excision repair cross-complementing group 1), XPF1 and XPC cleave the 27- to 30-nucleotide fragment and 3¢ and 5¢ of the lesion. Finally, DNA polymerase d or e, along with PCNA, RPA and replication factor C (RFC), fill the gap, using the undamaged DNA strand as a template (Kelley and Fishel 2008). Research in clinical applications for exploiting various steps in the NER pathway to develop new chemotherapeutics points to the fact that we still have much to learn about pathway signaling proteins. As we elucidate the actions of each step and the implications of inhibiting selected proteins, the opportunity for developing targeted therapies will become more apparent.
Clinical application: small-molecule inhibitors of NER Indication for utility Unclear. Genetic diseases linked to NER defects (xeroderma pigmentosum, Cockayne’s Syndrome) are a deterrent to researching NER inhibitors How it works Efforts in modulating NER response primarily have focused on potentiating the effects of platinum-based drugs, so any smallmolecule inhibitors developed would be chemotherapeutic adjuncts. Inhibitors to two NER proteins (XPA, RPA) and a NER enzyme (ERCC1) are currently under investigation • A global cyclin-dependent kinase inhibitor (UCN-01; 7-hydroxystaurosporine) blocks the interaction between XPA and ERCC1, which can affect the damage-recognition and incision steps of NER • An RPA inhibitor targets RPA’s ability to bind to single-strand DNA • ERCC1 expression may be modulated by a number of small molecules, including gefitinib (an epidermal growth factor receptor tyrosine kinase) and vascular epidermal growth factor (VEGF). Whether their action is specific to NER is unclear Anti-NER agents in use None yet
7 DNA Repair and Redox Signaling
143
NHEJ Repair NHEJ is the main repair pathway for DNA double-stranded breaks (DSBs) in mammalian cells. DSBs are the most severe form of DNA damage, as they can cause deletion or translocation of chromosomal DNA. Ionizing radiation, chemotherapeutic drugs, cleavage during V(D)J-recombination, meiotic recombination or the collapse of replication forks can cause DSBs. NHEJ rejoins DSBs at the ends of the broken DNA, regardless of sequence homology. In NHEJ, the break ends are ligated directly without requiring a homologous template – hence, the name of the pathway. Furthermore, NHEJ can repair DNA double-strand breaks during any phase of the cell cycle (Kelley and Fishel 2008). NHEJ begins when Ku70 and Ku80 proteins bind to the ends of the DNA breaks. Then, the Ku proteins recruit the catalytic subunit of DNA-dependent protein kinase (DNA-PKcs), which creates a nick to mark the spot for repair. The Ku/DNAPKcscomplex recruits additional proteins to make the actual repair. DNA Ligase IV and XRCC4 function as a complex to ligate the nick and complete the process (Burma et al. 2006). Discovery of new proteins involved in NHEJ includes Metnase (also known as SETMAR, histone–lysine N-methyltransferase, SET domain and mariner transposase fusion gene-containing protein) (Lee et al. 2005), which has been shown to interact with DNA Ligase IV to enhance the efficiency and accuracy of NHEJ (Hromas et al. 2008). Because repairing DNA DSBs is critical for tumor cell survival, the search for inhibitors to this pathway is an ongoing pursuit (Altieri et al. 2008; Aiyer et al. 2008; Akamatsu et al. 1997; Akterin et al. 2006). Currently, DNA-PKcsis the most highly researched target in the NHEJ pathway. Clinical application: small-molecule inhibitors of NHEJ Indication for utility DNA-protein kinase is a member of the phosphatidylinostitol 3-kinase (PI3-K) superfamily (Shinohara et al. 2005), which plays a key role in repairing non-homologous doublestranded DNA breaks How it works Inhibition of DNA-PKcsmakes cells hypersensitive to ionizing radiation and numerous DNA cross-linking agents by blocking re-ligation of DNA DSBs Anti-NHEJ agents in use None yet Several compounds in development (NU7441, IC60211, IC86621, AMA37 and others) are being tested in animal models
HR HR is also involved in repairing DNA DSBs. In contrast to NHEJ, HR operates only during the cell cycle’s S and G2 phases during DNA replication. The HR pathway is aptly named, as a sister chromatid must be available nearby to provide a repair template.
144
M.R. Kelley et al.
Repair starts when the MRN complex (Mre11/Rad50/Nbs1) recognizes the DSB and recruits ATM (ataxia telangiectasia mutated protein) to monitor the cell cycle’s checkpoints and analyze the break (Aiyer et al. 2008). MRN’s nuclease activity creates single-strand DNA (ssDNA). When ATM sees the block in replication, it activates numerous downstream proteins via phosphorylation to create a complex that protects the ends of the ssDNA and enables the repair process. ATM-activated BRCA1 (breast cancer 1, early onset) attracts BRCA2 and RAD51 to bind to the ssDNA ends, allowing the RAD52/RAD54 complex to join to and form a larger complex with BLM (Bloom’s syndrome gene product) and WRN (Werner) proteins. These large protein complexes attach to the strand break, direct the pairing of processed DNA with a homologous region on the sister chromatid and initiate strand exchange (Sung and Klein 2006). Clinical application: small-molecule inhibitors of HR Indication for utility ATM regulates cell cycle checkpoints. It interrupts the cell cycle when it detects DNA damage, especially double-strand breaks How it works Inhibiting signaling through ATM can impair the cells’ response to DNA damage, which can make cells more sensitive to chemotherapeutic agents. Ionizing radiation, ROS and (indirectly) alkylating agents can easily kill ATM-deficient cells Anti-HR agents in use None yet, as previously investigated. ATM inhibitors were too non-specific and produced too many off-target effects. Under investigation is KU55933, which blocks phosphorylation of numerous ATM targets and CP466722. Early evidence indicates that it can potentiate the effects of camptothecin, doxorubicin and etoposide
Overview of Redox Signaling Oxidation–reduction (redox) signaling maintains cellular homeostasis and regulates numerous cellular processes, including metabolism, growth, reproductive biosynthesis and apoptosis. Two systems are primarily responsible for general redox regulation: the thioredoxin (Trx) and glutaredoxin/glutathione (GRX/GSH) systems. Both perform their redox and cellular regulation functions through a thiol– redox exchange mechanism (Holmgren 1995; Nakamura et al. 1997). Thiol-based redox reactions rely on binding activities of cysteine (Cys) residues. Special properties of Cys allow them to adopt 10 different sulfur oxidation states from +6 to -2 (the latter being the fully reduced state) (Giles et al. 2003; Jacob et al. 2003), enabling them to exist in many forms (Jacob et al. 2003). The thiol/disulfide configurations in Cys drive human redox system functions in general (Giles et al. 2003). In a thiol/disulfide exchange, a reduced redox protein recognizes an oxidized protein containing a disulfide bond; the redox protein donates two hydrogen molecules to create an intermediate mixed disulfide bond on the oxidized protein.
7 DNA Repair and Redox Signaling
145
In doing so, the redox protein becomes oxidized. A reductase restores the redox protein to a reduced state, using cofactors such as FADH2and NADPH (Giles et al. 2003; Jacob et al. 2003). A specific Cys acts as a nucleophile in this type of reaction (Giles et al. 2003; Jacob et al. 2003). Because the proper folding and functionality of many proteins relies in part on the reversible formation of inter- or intramolecular disulfide bonds, redox systems have far-reaching effects on protein-mediated cellular activities (Berndt et al. 2008).
The Thioredoxin (Trx) System This system includes thioredoxin (Trx), thioredoxin reductase (TrxR) and NADPH (Jacob et al. 2003; Holmgren 1989). Thioredoxins (Trxs) are a large family of structurally conserved oxidoreductases that reduce disulfide bonds in a variety of proteins through a thiol/disulfide exchange mechanism (Nakamura 2005). The system has a built-in “recycle-and-reuse” mechanism: thioredoxin reductase, a flavoprotein containing a selenocysteine, reduces oxidized thioredoxin in a reaction using NADPH as an electron donor. Thus, thioredoxin may be used repeatedly (Jacob et al. 2003). Although all thioredoxins contain a similar active-site motif of Cys-X-X-Cys, their most defining structural feature is the “Trx fold:” a four-stranded b-sheet surrounded by three a-helices (Holmgren 1995; Lillig and Holmgren 2007; Powis and Montfort 2001). Cys-X-X-Cys is located on the loop connecting b-sheet 1 and a-helix 1. The N-terminal Cys32 residue in the active site of human Trx is surface exposed and has a low estimated pKavalue of 6.3 (Forman-Kay et al. 1992), while the C-terminal Cys is buried in the molecule and has a much higher pKavalue. The low pKavalue of the N-terminal Cys may be due to the partial positive charge from the dipole moment associated with a-helix 2 (Hol 1985), or its hydrogen bond to the C-terminal Cys (Weichsel et al. 1996). Regardless of the cause, the low pKaof the thiolate group of the Cys increases its selective reactivity. The proposed reaction mechanism of disulfide reduction by thioredoxin is as follows: the N-terminal cysteine thiolate of Trx acts as a nucleophile and attacks the target disulfide on the oxidized protein, resulting in a transient mixed disulfide intermediate. This, in turn, is reduced by the C-terminal active site Cys residue, generating a dithiol in the target protein and a disulfide in thioredoxin (Holmgren 1995; Lillig and Holmgren 2007; Kondo et al. 2006). The resulting disulfide in the active site of Trx can be reduced by TrxR. The mechanism by which it regenerates Trx back to the dithiol involves the formation of a selenylsulfide on the active site of Trx (Datta et al. 2004). A second redox active site located in the other subunit of the dimeric TrxR contains two thiols that reduce the selenylsulfide back to a thiol and selenol, resulting in the formation of a disulfide bond. Electron transfer from FADH2reduces this disulfide; the resulting FAD is then reduced by electron transfer from NADPH (Jacob et al. 2003) to complete the catalytic cycle.
146
M.R. Kelley et al.
The Glutaredoxin/Glutathione (GRX/GSH) System The glutaredoxin system includes glutaredoxin, glutathione, NADPH and the flavoprotein glutathione reductase (Holmgren 1989; Kondo et al. 2006; Lillig et al. 2008). Like the Trx system, this system also works through a cascade of disulfide oxidation and reduction. Glutaredoxins (GRXs) are small redox enzymes (100 amino acids) that use glutathione as a cofactor. Structurally, glutaredoxins are very similar to thioredoxins, exhibiting the same fold and active sites. However, the active site of GRXs includes Cys-X-X-Cys or Cys-X-X-Ser. GRXs catalyze the reversible reduction of substrate protein disulfides, resulting in the oxidation of the GRXs. Oxidized GRXs are reduced non-enzymatically by glutathione (glutamyl-cysteinyl-glycine, or GSH). Then, the oxidized glutathionine disulfide group (GSSG) is returned to a reduced state by the action of glutathione reductase, at the expense of NADPH (Lillig et al. 2008). Although the thioredoxin system and glutaredoxin system share many functions, they are not redundant systems. Trx and GRX act on different substrates (Lillig et al. 2008; Fernandes and Holmgren 2004). For example, Trx – but not GRX – is implicated in the reduction of APE1, a protein exhibiting both redox and DNA repair functions (Akamatsu et al. 1997; Hirota et al. 1997; Huang and Adamson 1993; Tell et al. 2005).
Roles of Redox Systems The thioredoxin and glutaredoxin systems scavage ROS to reduce intracellular oxidative stress. A reduced state is important for optimizing many cell functions including growth and differentiation, biosynthesis and apoptosis. The systems’ collective efforts to maintain a reduced intracellular environment not only facilitate correct protein folding and efficient work of protein chaperones but they also prevent oxidative modifications to DNA, including base changes, changes to the deoxyribose backbone and direct breakage of a strand (Berndt et al. 2008). Both the DNA damage response and the wide-ranging influence of redox systems include activation and control of numerous transcription proteins (Powis and Montfort 2001; Arner and Holmgren 2000; Yamawaki and Berk 2005). The “dark side” of the two redox systems is that they can contribute to the pathophysiology of various human diseases, including cancers, viral infections, neurodegenerative diseases (including Alzheimer’s) and others (Akterin et al. 2006; Arner and Holmgren 2006; Burke-Gaffney et al. 2005; Raffel et al. 2003). For example, APE1 plays a role in the accumulation of amyloid b protein in Alzheimer’s plaques (Bhakat et al. 2008). Thus, inhibition of redox systems is a vital drug target for many therapeutic areas (Biaglow and Miller 2005; Chew et al. 2008; Liu et al. 2009; Liu 2008; Mukherjee and Martin 2008).
7 DNA Repair and Redox Signaling
147
Clinical application: small-molecule inhibitors of Trx (Tonissen and Di Trapani 2009) Indication for utility Increased levels of Trx and TrxR are present in many cancers. This also correlates with increased resistance to chemotherapeutics The Trx system may assist early cancer development via its growth-promoting and anti-apoptotic functions. Later, the Trx system may influence angiogeneis and metastasis How it works All redox-inhibiting agents ultimately force cancer cells to die by creating an oxidative intracellular environment that exceeds the cell’s redox damage–repair abilities; this, in turn, triggers one of the apoptotic pathways. Redox-inhibiting agents often target the selenothiol of TrxR to accomplish this Anti-Trx agents in use • SAHA (a histone deactylase inhibitor): upregulates an endogenous inhibitor of Trx • Platinum compounds: Although they are most well known for forming irreversible DNA-platinum adducts, these compounds also affect both the Trx and GRX redox systems • ATO (arsenic trioxide): inhibits TrxR via a largely unknown mechanism of action Emerging treatments • PX-12 (1-methyl-propyl-2-imidazolyl disulfide): binds to Cys73 of Trx, irreversibly alkylating it, so that TrxR cannot regenerate it. Also decreases expression of VEGF (which Trx indirectly influences)· Gold compounds: inhibit TrxR in a mechanism still being elucidated. Mode of action may be similar to that of platinum compounds • MGd (motexafin gadolinium): inhibits Trx and generates ROS, including superoxides and hydrogen peroxide. Also inhibits an essential enzyme for DNA synthesis • DCNB (nitroaromatic compounds): inhibits TrxR and induces superoxide generation • Curcumin & flavonoids: convert TrxR to a pro-oxidant form that stimulates ROS production
The Redox Activity of APE1 APE1 is the only DNA repair protein known to have a role in redox regulation, which affects the expression of a number of DNA transcription factors. The protein’s dual functions are reflected in the history behind its full designation: APE1/Ref-1. The multifunctional protein was first identified as being responsible for reducing transcription factor AP-1; therefore, it was named “redox effector factor 1,” or “Ref-1” (Xanthoudakis and Curran 1992, 1996). Since this initial discovery, further studies revealed that APE1 is responsible for regulating many transcription factors including NFkB, HIF-1a, p53, PAX and others (Hirota et al. 1997; Cao et al. 2002; Ema et al. 1999; Hirota et al. 1999; Lando et al. 2000; Tell et al. 2000; Ueno et al. 1999) (Fig. 7.3a). And, as discussed in the first section of this chapter, APE1 performs critical DNA endonuclease repair functions. That spawned its other designation: “APE1,” or apurinic/apyrimidinic endonuclease. Today, the complete name of this protein (APE1/Ref-1) is usually truncated to “APE1.”
148
M.R. Kelley et al.
How APE1 Performs Its Redox Functions APE1’s uniqueness is manifested both in its breadth of function and how it functions, which is still somewhat of a mystery (Georgiadis et al. 2008; Walker et al. 1993). Animal and bacterial APE1 proteins contribute to DNA repair but lack any redox abilities. However, human APE1 does both, making it unique in that regard. How APE1 performs its redox activity remains a controversy, as only the Cys65 residue in APE1 has been identified as critical for its redox reactions (Table 7.1). Cys65 is buried near the N-terminus on the first beta strand in the fold, which is part of a beta sheet in the core of the protein. Normally, two Cys residues are required for thiol/disulfide exchange reactions; however, APE1 does not contain two Cys residues within a C-X-X-C motif that are structurally close enough (~2.2Å) to form the disulfide bond required to drive a redox reaction. APE1’s Cys residues that are in closest proximity to each other are buried in APE1’s core and are on opposite sides of the beta sheet (Fig. 7.3a, b) (Luo et al. 2009). Given these apparent obstacles, four possibilities have been proposed regarding the mechanism for APE1’s redox abilities. The first speculation is that a yet undiscovered Cys residue is involved. A second possibility is that an undiscovered Ser reside is involved, as is found in glutaredoxins. However, to function properly, the Ser would require a pKathat is unusually low for its proposed location. A third postulation is that APE1 undergoes a conformational change to restore its reduced state. This type of activity is seen in another group of redox proteins called peroxiredoxins, which are responsible for catalyzing intracellular hydrogen peroxide (Hofmann et al. 2002). Peroxiredoxins lack a C-X-X-C motif, but they do include two Cys residues that are required for activity. Peroxiredoxins exist as fully folded dimeric structures, where the nucleophilic thiolate is sequestered and the resolving thiolate is located near the C-terminus of the molecule. One Cys from each monomer forms the active site (Wood et al. 2003). Being approximately 9Å from each other, one would wonder how they could form an active site; however, a local unfolding places them in close proximity to detoxify hydrogen peroxide. APE1 could have a similar requirement of a conformational change to complete its catalytic cycle. Further support for this theory can be found in the way that GAPDH interacts with APE1. GADPH acts upon APE1’s Cys152 active site, causing a conformational change that enables APE1 to perform its endonuclease activity (Luo et al. 2009). A fourth possibility exists for explaining APE1’s redox functionality: the involvement of two APE1 molecules. This would provide the second Cys or Ser residue needed to resolve the redox reaction. Further studies are needed to elucidate this important feature of APE1, as it has many implications for future clinical therapeutics.
APE1-Regulated Transcription Factors and Their Link to DNA Damage Repair Because APE1 is involved in both DNA damage repair and the redox regulation of a number of stress-inducible transcription factors, APE1 clearly plays pivotal roles
7 DNA Repair and Redox Signaling
149
in both redox signaling that modulates DNA damage response and direct DNA repair functions. APE1 is a reductive activator for several transcription factors, including p53, AP-1 and HIF1-a (Fig. 7.3a, b). Analysis of APE1’s effector action on these proteins underscores additional unique features about APE1. These transcription factors are unrelated multidomain proteins; that is, APE1 does not recognize any common structural motif when acting upon them. Furthermore, the redox-regulated DNA-binding domains of these factors are structurally distinct: p53, a single immunoglobulin-like domain; NFkB (discussed here in conjunction with p53), a dimer of two-domain immunoglobulin-like subunits; AP-1, a heterodimeric bZip family protein; and HIF-1a, a basic helix-loop-helix domain. Redox regulation of transcription factors is one of several mechanisms that control sequence-specific DNA binding and thereby, gene expression. APE1’s influence on p53, AP-1 and HIF-1a affects the DR, BER, HR, MMR and GGR pathways, as discussed on the following pages (Table 7.1 and Fig. 7.2).
p53 Often called “the guardian of the genome” (Zaky et al. 2008), p53 is a transcription regulatory protein that helps preserve genomic integrity by its participation in stress–response pathways and DNA repair pathways (Lando et al. 2000; Helton and Chen 2007; Sengupta and Harris 2005; Zurer et al. 2004). More than half of all cancers contain mutant or inactive p53, underscoring its importance as a tumor suppressor. Two overall strategies that p53 employs in maintaining the genome are to arrest the cell cycle or to induce apoptosis. For example, hypoxia in tumor cells prompts p53 to induce apoptosis (Jayaraman et al. 1997). Under other circumstances, p53 can interrupt the cell cycle to allow sufficient time to complete DNA repairs (Jiang et al. 2009). The far-reaching influence of p53 also includes complex interactions with other proteins and effector genes. Initial investigation of redox regulation of p53 was based on the finding that oxidized p53 has little binding affinity for DNA (Hainaut and Milner 1993). This led to the discovery that APE1 was responsible for enhancing the DNA-binding activity of wild-type p53 by reducing it (Jayaraman et al. 1997). The mechanisms by which APE1 enhances this activity of p53 are still being discovered, but they appear to include both a redox-dependent activation of the DNAbinding domain (through direct reduction of a disulfide bond), as well as a redox-independent mechanism (Jayaraman et al. 1997). APE1 reduces p53 (Jayaraman et al. 1997). This is essential, but not sufficient, for optimizing p53’s DNA-binding function (Jayaraman et al. 1997). Recent evidence shows that APE1 also facilitates tetramerization of p53 by interacting with its C-terminal regulatory domain in a redox-independent manner, producing p53’s optimal configuration for binding to DNA. Thus, APE1 appears to enhance sequence-specific DNA-binding activity of p53 through two independent, coordinated mechanisms (Hanson et al. 2005). Redox signaling involving p53 depends upon APE1 plus a general redox factor as well. Trx has been shown to enhance stimulation of p53-dependent expression
150
M.R. Kelley et al.
APE1 has two major functions: redox regulatory and DNA repair activity
DNA damage
Redox Role
(reduced) APE1 (oxidized)
T A
AP 1 p53 NFkB HLF HIF 1a CREB ATF Egr 1 NF Y PAX
(oxidized)
(reduced)
DNA Repair G C
A T
T A
C G
T G A T C A C T T A G
*
T A
G A T C C T T A G
T A
G C
APE1
5’ P
T G A C
b
T
Glycosylase (Ogg1, Nth1)
Redox control of Transcripon factors
DNA Repair gene expression altered
Alteraon of gene expression
A
A T C T T A G
A
T A
C G
G A A T C T T A
C G
T
A T
DNA Ligase T A
Transcription Factor
H2N
S-S
Oxidized AP-1, HIF-1α, NFkB
COOH
DBD
APE1RED
DNA
H2N
APE1OXIDIZED
COOH HS
DBD
SH
Target gene expression
Regulatory region
Fig. 7.2 Role of APE1 in regulating expression of other DNA repair proteins and pathway through its redox signaling function
7 DNA Repair and Redox Signaling
151
of p21 by APE1, suggesting a link between Trx and APE1 in cellular response to oxidative stress (Ueno et al. 1999). Because p53 influences many DNA repair pathways and APE1 constitutively influences p53, a short discussion of their effects on each pathway is warranted. p53’s regulation of DNA repair is complex, involving both transactivation-dependent and -independent mechanisms that affect the HR, BER, NER, MMR and NHEJ pathways (Sengupta and Harris 2005). The BER pathway, the main mechanism for repairing oxidative and alkylating base modifications, is initiated by highly specialized DNA glycosylases that cleave the DNA base to create an AP site. p53 regulation of BER varies according to the type of genotoxic stress placed on cells. For example, exposure to nitric oxide species causes p53 to down-regulate the specific glycosylase that would begin repairs in response to this threat. Interestingly, NO does not alter APE1 activity. In this respect, p53 acts as a “brake,” stopping repairs to prevent the possibility of engendering a mutagenic phenotype. However, in response to g-ray treatment, p53 promotes activation of AAG (Zurer et al. 2004). The choice of response also may be specific to the cell line involved. Recent studies with human colorectal cancer line HCT116 show that wild-type p53 (but not mutant p53) downregulates APE1 by indirect recruitment of its promoter, possibly blocking binding of SP1 (Zaky et al. 2008). Thus, in this example, APE1 regulates DNA binding of p53; and p53, in turn, regulates both expression of APE1 as well as certain proteins that bind to the promoter of APE1. This system of checks and balances maintains equilibrium between p53’s pro-apoptopic activity and APE1’s pro-survival activity. In addition, p53 regulates DNA polymerase b (Pol b), which replaces excised bases in the SP-BER pathway. The single-nucleotide replacement mechanism of Pol b requires dRP lyase activity for the repair (Allinson et al. 2001). This lyase activity occurs after APE1 activity and is often rate limiting in BER. Pol b expression is down-regulated in p53-deficient cells (Cao et al. 2002), which increases the likelihood of the accumulation of cytotoxic repair intermediates that effectively can halt DNA replication. Additionally, in vitro studies with various cell lines show that CREB (which recruits and stabilizes Pol b) can upregulate Pol b gene expression in response to DNA alkylating exposure (He et al. 2003). CREB (cAMP response element binding) is under redox control for its DNA binding and activating transcription. Other proteins in the DNA repair arsenal are under redox control, and this control regulates DNA repair response (Fig. 7.2). p53 differentially affects the two NER sub-pathways, TCR and GGR. Several studies found that p53 selectively affects GGR, but not TCR. In GGR, sublethal UV radiation causes p53 to arrest the cell cycle in G1, to gain the needed time for DNA repair before progressing to S phase (Smith and Seo 2002; Li et al. 2001). In addition, p53 also regulates GGR downstream effectors, DDB2 and XPC, the main proteins involved in DNA damage recognition (Adimoolam and Ford 2002; Hwang et al. 1999; Tan and Chu 2002). Inactive or absent p53 results in defective GGR repair activity and appears to lead to genomic instability (Xu and Morris 1999). Studies of knockout mice support this: in one study, 100% of XPC-/-mice developed lung cancer and 100% of DDB2-/-mice developed skin tumors (Hollander
152
M.R. Kelley et al.
et al. 2005). PCNA, a progressivity factor in the NER pathway, is also under p53 regulation (Xu and Morris 1999). The MMR pathway removes and repairs DNA mismatches because of the DNA polymerase errors. MSH2, which must be complexed with MSH6 or MSH3 to recognize DNA damage, is under p53 regulation. Low levels of p53 prevent MSH2:MSH6 recognition of single-base mismatches and short insertion/deletion mispairs, and MSH2:MSH3 recognition of larger loops of unpaired nucleotides. MSH2, MLH1, and PMS2 all are under p53 regulation, similar to the way that DDB2 and XPC are in NER (Chen and Sadowski 2005; Scherer et al. 2000). Both the HR and NHEJ pathways address reparation of double-strand breaks. DSBs are the most severe threats to genomic stability because they facilitate deletion and/or translocation of chromosomal DNA. As a result, both pathways are highly regulated. Once again, p53 appears as an important player in the repair of DSBs through regulation of both the HR and NHEJ repair pathways. Mice that are deficient in wild-type p53 exhibit increased levels of HR activity (Bishop et al. 2003; Lu et al. 2003). p53 inhibits HR through repression of RAD51 expression (Arias-Lopez et al. 2006). (Conversely, absent or mutated p53 increases RAD51 DNA repair activity, which contributes to chemotherapeutic resistance.) In the NHEJ pathway, p53 downregulates expression of WRN and RecQ4, two RecQ helicases (Sengupta et al. 2005; Yamabe et al. 1998). Finally, in the DR pathway, studies on murine fibroblast cell lines have demonstrated that AGT is under p53 regulation (Grombacher et al. 1998; Rafferty et al. 1996). Interactions between p53 and AGT are complex and depend on many variables, including the type of cellular stress present. For example, p53 appears to induce AGT in response to ionizing radiation. AGT also appears to be regulated by NFkB, a transcription factor mediating immune and inflammatory responses. For example, overexpression of NFkB in HEK293 cells (human embryonic kidney 293 cells) increases AGT expression (Lavon et al. 2007). NFkB is under redox control by APE1. All these studies point to the possible clinical relevance of redox-controlled DNA repair responses through p53. If reduced p53 is required to bind DNA and either activate or repress the transcription of DNA repair genes, then it follows that APE1 plays an important role not only in the redox modulation of p53, but also in the regulation of DNA repair.
AP-1 Activator protein-1 (AP-1) refers to a family of proteins that recognize AP-1 sites. The AP-1 family is also referred to as tetradecanoylphorbol-13-acetate (TPA)responsive elements. AP-1 mediates diverse cellular functions including proliferation, differentiation, apoptosis and transformation; it also responds to environmental changes including stress, radiation or growth factor signals (Hess et al. 2004). The AP-1 family consists of structurally and functionally related basic leucine zipper
7 DNA Repair and Redox Signaling
153
proteins (bZIP) that intermix to form heterodimeric sequence-specific DNAbinding proteins. These include Jun proteins (c-Jun, JunB and JunD), Fos proteins (c-Fos, FosB, Fra-1 and Fra-2) and some ATF family members (ATFa, ATF-2 and ATF-3) (Hess et al. 2004). Studies of AP-1 activity provided the backdrop for early investigations into APE1’s redox activity. In fact, APE1 was first identified as a result of efforts to determine the nuclear factor responsible for reducing AP-1 (c-Jun/c-Fos) and thereby enhancing its DNA-binding activity (Xanthoudakis and Curran 1992; Abate et al. 1990). Redox regulation of AP-1 was found to result from oxidation/ reduction of conserved Cys residues within the basic DNA-binding domains of c-Jun and c-Fos, which enables them to dimerize (Abate et al. 1990). APE1 copurifies with AP-1 (Xanthoudakis and Curran 1992), which is in stark contrast to APE1’s interaction with p53. The former is a stable interaction; the latter is transient. Trx has also been identified as a factor modulating transcriptional activation of AP-1 through its reduction of APE1 (Hirota et al. 1997). AP-1 is rapidly induced in response to many cellular stimuli, and its activation regulates the expression of several proteins involved primarily in NER and MMR DNA repair pathways. Numerous studies of AP-1’s influence on DNA repair proteins have revealed more than 20 genes whose promoters are bound following cisplatin treatment. The proteins that AP-1 regulates in repairing DNA–cisplatin adducts include RAD23B, XPA, ERCC3, XPF, ERCC1 and ERCC3 in the NER pathway and MSH2, MSH6, MLH1 and PMS2 in the MMR pathway. Phosphorylated c-Jun or ATF2 activates these proteins to increase their DNA-binding capacity (Luo et al. 2009; Luo et al. 2008). Because AP-1 must be converted from an oxidized to a reduced state in order to bind to its target sequence, redox control of this protein has significant implications. APE1 is implicated as a potential point of control for regulating the DNAbinding activity of AP-1, thus modulating the expression of DNA repair proteins (Table 7.1and Fig. 7.2).
HIF-1a Hypoxia inducible factor-1 (HIF-1), as its name suggests, responds to low-oxygen conditions in the cell. Increasing evidence reveals that hypoxic stress in the tumor microenvironment can cause genetic instability in cancer cells. HIF-1 is a heterodimeric transcription factor composed of two subunits; the HIF1a subunit is crucial for regulating cellular response to hypoxia and is frequently overexpressed in human cancers. Under hypoxic conditions, HIF-1a translocates to the nucleus and dimerizes with HIF-1b, forming HIF-1. HIF-1, along with coactivators, binds hypoxia response elements (HRE) within promoters and regulates the expression of their downstream genes, including vascular endothelial growth factor (VEGF) (Forsythe et al. 1996; Jiang et al. 1996). Redox-dependent stabilization of the HIF-1a protein is required for activation of HIF-1 (Huang et al. 1996), and
154
M.R. Kelley et al.
redox signaling regulates the DNA-binding activity of HIF-1. These combined activities suggest a role for APE1 as a redox regulator of HIF-1a. Furthermore, APE1 has been implicated in its role of helping to form the hypoxia-inducible transcriptional complex that includes HIF-1 and transcriptional coactivators p300 and CREB. More recently, HIF-1a has been shown to play a role in down-regulating mRNA and APE1 levels under hypoxic conditions in human microvascular endothelial cells (Loboda et al. 2009). Thus, HIF-1a regulates the expression levels of APE1 and is itself regulated by APE1. Hypoxia downregulates the expression of key genes within two DNA repair pathways. In HR, several critical mediators including BRCA1, BRCA2 and RAD51 are inhibited, resulting in significant genetic instability. In addition, HIF-1a decreases the expression of NBS1, a member of the Mre11-RAD50-NBS1 (MRN) complex that initially recognizes DSBs (To et al. 2006). In MMR, MLH1 and MSH2 are inhibited (Bindra and Glazer 2007; Bindra et al. 2004; Bindra et al. 2005, 2007; Bindra and Glazer 2007; Bristow and Hill 2008; Koshiji et al. 2005; Meng et al. 2005; Mihaylova et al. 2003). The MMR genes appear to be repressed through a mechanism involving c-Myc (Bindra and Glazer 2007). Interactions between HIF-1a and these genes are still being determined, but some studies show that HIF-1a represses MSH2:MSH6 in a p53-dependent manner (Koshiji et al. 2005). Regulation of DNA repair is an integral part of the hypoxic response; and HIF1a, a critical response mediator, influences the DNA repair response (Fig. 7.2). As APE1 is required for redox regulation of HIF-1a, it then also plays a role in the regulation of DNA repair genes controlled by HIF-1a. APE1 also exhibits redox-independent transcriptional regulatory functions that are subject to post-translational modifications including phosphorylation, acetylation and nitrosation. The purposes of these modifications are still being elucidated. For example, acetylation of APE may influence its subcellular localization or transacting functions (Bhakat et al. 2008). Review articles by Bhakat and others describe these implications further (Bhakat et al. 2008; Tell et al. 2009).
Other Global Influences of APE1 APE1 is a critical “master regulator” for maintaining genomic stability (Fig. 7.3). Earlier sections of this chapter discussed APE’1s apurinic/apyrimidinic endonuclease activity that is essential to BER’s DNA repair activities (Bhakat et al. 2008). In addition, APE1 performs redox functions that include modulation of transcription factors. APE1 controls the redox status of either ubiquitous (AP-1, Egr-1, NF-kB, p53, CREB, HIF-1a) or tissue-specific transcription proteins (PEBP-2, Pax-5 and -8 TTF-1) (Fig. 7.3a) (Akamatsu et al. 1997; Huang and Adamson 1993; Xanthoudakis and Curran 1992; Cao et al. 2002; Ema et al. 1999; Hirota et al. 1999; Lando et al. 2000; Tell et al. 2000; Ueno et al. 1999; Xanthoudakis et al. 1992).
7 DNA Repair and Redox Signaling
155
Fig. 7.3 APE1 has dual roles in redox and DNA repair. APE1 possesses two major functions: Redox regulatory/signaling and DNA repair. Through its redox function, APE1 regulates gene expression by modifying the redox status of some transcription factors which are involved in variety of cancer processes. (Adapted from: Luo et al. (Luo et al. 2008))
Cell Survival APE1 is also essential for cell survival and organism development, as evidenced by the fact that no viable knockout model of APE1 exists. The absence of APE1 in mice causes post-implantation embryonic lethality between days E5 and E9 (Xanthoudakis et al. 1996). Conditional knockout and knock-down strategies also confirm the crucial role of this protein in cells (Fishel et al. 2008; Fung and Demple 2005; Izumi et al. 2005). Studies have demonstrated that altering APE1 levels leads to changes in cell growth, survival and sensitivity (Chen and Olkowski 1994; Fishel et al. 2008; Bobola et al. 2005; Fishel et al. 2007; Herring et al. 1998; Lau et al. 2004; McNeill and Wilson 2007; Ono et al. 1994; Walker et al. 1994; Wang et al. 2004). Experiments with normal cells show that siRNA directed against APE1 results in decreased cell proliferation, increased AP sites and increased apoptosis (Fishel et al. 2007). While the extent of APE1’s activities for cell survival and overall growth are still unknown, recent studies show that APE1 acts as a quality control monitor of RNA. During the S phase of the cell cycle (when rRNA is synthesized), APE1 localizes in the nucleus and enhances de novo protein synthesis and cell proliferation by methods only partially understood. The N-terminus of APE1 binds to proteins involved in generating ribosomes and processing RNA (Barnes et al. 2009; Vascotto et al. 2009). APE1 also performs an RNA “cleansing” process that includes cleaving abasic RNA. Perhaps RNA’s single-strand construction makes it more susceptible to oxidative insults than DNA; therefore, APE1’s “cleansing” function affects overall cellular viability. APE1’s nuclear localization corresponds to increased
156
M.R. Kelley et al.
κ
α
β α
κ
Fig. 7.4 Redox control of DNA repair. In its role as a redox factor, APE1 reduces a number of transcription factors including AP-1, Egr-1, NF-kB, p53, CREB, and HIF-1a. Thus, APE1 controls the redox status of several transcription factors that in turn regulate expression of APE1
intracellular resistance to ROS, g radiation, and a number of other cytotoxic agents (Bhakat et al. 2008) Fig. 7.4. Other recent studies propose that APE1’s intracellular redox control may also affect the cell cycle by inducing nucleus–cytoplasm redistribution of p21, which regulates the G1 phase (Merluzzi et al. 2008).
Angiogenesis A mosaic of APE1’s redox influence on angiogenesis can be seen by observing its known functions in normal endothelial cells, plus examining in vitro and preclinical studies of its inhibition. In normal endothelial cells, APE1’s redox activity helps (1) maintain cell differentiation, (2) maintain vascular tone by regulating NO levels (Zou et al. 2009), (3) enable differentiation of angiogenic progenitor cells and (4) prevent apoptosis by repressing pro-apoptotic TNFa signaling and upregulating pro-survival NFkB signaling. Intracellular hypoxia is a feature of many cancers, including pancreatic and prostate carcinomas. Thus, decreased activity of APE1 influences downstream genes that induce endothelial cell death (Bhakat et al. 2009). In vitro studies hint at additional influences of APE1 on endothelial cells. Studies show that inhibition of APE1 attenuates the formation of retinal vascular endothelial cells (RVECs), and, to a lesser extent, their proliferation (Luo et al. 2008). In a
7 DNA Repair and Redox Signaling
157
d ifferent realm of endothelial health, Tat-mediated protein transduction of APE1 decreases vascular cell adhesion molecule-1 expression, blunting the response of monocyte adhesion factors to atherosclerotic injury (Song et al. 2008). Preclinical studies of E3330, a unique quinone that selectively inhibits APE1’s redox functions, allude to APE1’s influence on endothelial cells in the tumor microenvironment. Tumors will not grow larger than a few millimeters unless they have constant access to a blood supply. Migration of endothelial progenitor cells to the tumor microenvironment is essential for neoangiogenesis. Preclinical utility studies with E3330 demonstrate that it inhibits the growth of pancreatic cancer cell lines (Zou and Maitra 2008) as well as in pancreatic cancer-associated endothelial (PCECs) and endothelial progenitor cells (Zou et al. 2009). Because intracellular hypoxia is a common feature of many cancers, E3330’s influence on HIF-1 has been studied as well. E3330 inhibits HIF-1’s DNA-binding activity, which is consistent with E3330’s ability to inhibit APE1’s redox functions. Thus, inhibition of APE1 induces endothelial cell growth arrest via a cascade of signaling upstream from HIF-1a. Additional studies of E3330’s effect on human bone marrow cells, pancreatic cancer cells and human umbilical vein endothelial cells show that E3330 can reduce tumor endothelial VEGF secretion – and, at the same time, downregulate levels of the cognate receptor Flk-1/KDR on PCECs, blocking a potentially critical angiogenic ligand–receptor interaction in the tumor microenvironment. In addition, E3330 can block the differentiation of bone-marrow hemangioblasts (Zou et al. 2009).
Inflammation Long-standing inflammation is a risk factor for tumorgenesis. For example, patients who live with ulcerative colitis (UC) for more than 10 years have a 20- to 30-fold increased risk of developing colorectal cancer. Chromosomal instabilities can be detected even in early-stage dysplastic UC tissues (Guo et al. 2008). Sources of inflammation such as ROS and toxic agents transiently increase intracellular levels of APE1 (Evans et al. 2000; Luo et al. 2009; Tell et al. 2009). Prolonged stimulation and response of APE to intracellular stresses from inflammation can cause genomic instability. It is well known that many DNA repair intermediates are cytotoxic and potentially mutagenic – including abasic sites and single-strand breaks (Bapat et al. 2009). If DNA repairs stall or are performed incorrectly, they can lead to microsatellite instability (MSI) – which is present in the aforementioned dysplastic colon cells. Inflamed areas of UC lesions contain significantly greater amounts of APE1 and AAG than tissues in non-inflamed areas, suggesting that long-term adaptive increases in these proteins contribute to the production of MSI. Many other sporadic tumors also contain MSI, which provides motivation to study BER and other DNA repair pathways in other chronic inflammatory diseases. Overexpression of
158
M.R. Kelley et al.
AAG and APE1 is likely just one example of how long-term dysregulation of a DNA repair pathway can contribute to tumorgenicity (Guo and Loeb 2003).
DNA Repair in the Tumor Microenvironment Little is known about DNA repair in the tumor microenvironment, except that it is abnormal. If DNA repair is imperfect or incomplete, it can result in DNA damage that is genomically unstable, which is a key step for initiation and progression of cancer (Kruman et al. 2004; Schwartz et al. 2007). Tumors can be expected to be defective or deficient in one DNA repair pathway, but they have an astonishing ability to compensate and adapt by using other repair mechanisms. Crossover, interaction and compensation within and among DNA repair pathways often allow cancer cells to avoid or escape apoptosis (Reed et al. 2009). For example, cancer cells deficient in the proteins of the HR pathway may compensate for this deficiency by repairing DSBs through the NHEJ or BER pathway (Christmann et al. 2003; Hess et al. 2004; Beckman and Loeb 2006; Handel et al. 1995). A tumor’s DNA repair abilities are directly related to its unchecked cellular proliferation and its intrinsic or acquired cellular resistance to clinical DNAdamaging agents (Fishel and Wilson 1997; Holmgren 1989; Arner and Holmgren 2006; Bravard et al. 2009). So, in theory, the very defect that makes cancer cells highly mutable should be able to be exploited – to create a drug that makes the original defect lethal to cancer cells without harming normal cells. Therefore, inhibiting specific proteins in DNA damage repair pathways is a promising strategy for developing targeted cancer treatments (Fishel and Wilson 1997; Hromas et al. 2008; Holmgren 1989; Arner and Holmgren 2006; Bravard et al. 2009). It may sound paradoxical to inhibit DNA repair pathway proteins because cancer promotion and deregulated cellular growth are aided by deficient DNA repair pathways. However, a fine balance exists between the induction of DNA damage and its efficient repair (Hromas et al. 2008; Arner and Holmgren 2006). Inhibiting specific DNA repair proteins can selectively sensitize cancer cells to chemotherapeutic agents (Hromas et al. 2008; Holmgren 1989). Great clinical interest exists for combining DNA repair inhibition with current chemotherapy regimens (Beckman and Loeb 2006) to develop targeted therapies. Redox regulation plays an important role in DNA repair, and increasing attention is being focused on redox activities as potential targets for new chemotherapeutic agents. APE1 is a major player in redox signaling to downstream transcription factors that regulate stress and DNA repair responses. Increased levels of APE1 have been linked to tumor promotion, progression, poorer prognosis, shorter time to progression and poorer outcomes from chemotherapy (Hanson et al. 2005). Abnormal APE1 levels are speculated to be one cause of resistance to chemotherapy. Although APE1 largely is restricted to the nucleus in normal cells, its increased level in the cytoplasm has been associated with more aggressive lung, thyroid, breast and ovarian cancers (Zaky et al. 2008; Di Maso et al. 2007). The cause of
7 DNA Repair and Redox Signaling
159
this phenomenon is yet unknown, but it may provide another clue in understanding how APE1 redox activity is altered in the tumor microenvironment. Ample data demonstrate that downregulation or inhibition of APE1 using RNA interference and DNA antisense oligonucleotide techniques can sensitize cancer cells to laboratory and clinical chemotherapeutic agents (Bapat et al. 2009; Lillig et al. 2008; Hainaut and Milner 1993; Herring et al. 1998; Barnes et al. 2009; Langie et al. 2007; Madhusudan and Hickson 2005; Mathers et al. 2007). Thus, regulation of APE1 is emerging as a major player in manipulating the tumor microenvironment and modulating immune response. Redox regulation clearly plays an important role in DNA repair. APE1’s multifunctional abilities in DNA repair and redox signaling of transcription proteins are unique; no backup exists for its critically important activities. Therefore, identifying selective inhibitors of APE1 make it an excellent target for both single-agent and combination therapeutic approaches. Future research in this area will bring additional insights into the importance of redox regulation in DNA repair and will likely result in the identification of other key redox proteins.
Modulating APE1’s Activities as a Cancer Therapeutic Approach Elevated APE1 levels are typically associated with aggressive proliferation, increased resistance to therapeutic agents and poor prognosis (Evans et al. 2000; Wang et al. 2004; Moore et al. 2000; Xu et al. 1997) in various cancers. Altered or elevated levels of APE1 have been observed in breast and ovarian cancers, gliomas, sarcomas (osteosarcomas and rhabdomyosarcomas) and multiple myelomas, among others (Chen and Olkowski 1994; Hainaut and Milner 1993; Hanson et al. 2005; Koshiji et al. 2005; Langie et al. 2007; Madhusudan et al. 2005). Decreasing APE1 levels can block tumor cell growth and increase cellular sensitivity to DNA-damage agents (Fishel et al. 2008, 2007; Lau et al. 2004; Ono et al. 1994; Wang et al. 2004). However, current methods for modulating APE1 alter all its functions. Because of APE1’s diversity, it is important to distinguish and characterize which of APE1’s function(s) are involved in different biological events – especially those that may differ in normal versus pathologic cells such as cancer cells. The relative importance of APE1’s redox vs. repair function in cancer is unknown. Efforts to determine the differential therapeutic effects of inhibiting one or the other are ongoing. Blocking of the redox function of APE1 would be expected to impact the activity of a number of downstream transcription factors and the genes products that they regulate. Exploration of novel targets clearly merits pursuit, particularly in cancers for which current treatments are ineffective. Many small-molecule inhibitors in development are specific to proteins in DNA damage–response. Designed as potentiators of traditional chemotherapeutics, their goal is to inhibit targeted proteins while adding minimal toxicity to the treatment (Reed et al. 2009).
160
M.R. Kelley et al.
There has been considerable debate in the literature regarding the wisdom of inhibiting essential DNA repair enzymes (Kelley and Fishel 2008; Bapat et al. 2009; Madhusudan and Hickson 2005; Damia and D’Incalci 2007; Fishel and Kelley 2007). Targeting a specific protein (particularly one that plays an important role in cellular response to stress) by chemical knockout through the use of a smallmolecule inhibitor may have unintended consequences. However, small-molecule inhibitors have been identified for several DNA repair enzymes including MGMT (AGT), poly ADP-ribose polymerase (PARP1), ataxia-telangiectasia mutated kinase (ATM kinase), APE1 and DNA PKcs (Madhusudan and Hickson 2005; Damia and D’Incalci 2007; Fishel and Kelley 2007). The potential for small-molecule inhibitors to attenuate one or more of APE1’s functions offers potential for reversing drug resistance (Bhakat et al. 2008). Targeting the redox function of APE1 represents a novel approach in the development of a cancer therapeutic agent, as selectively blocking this function seems to lead to cytostatic growth without causing cell death. Finding specific small-molecule inhibitors that can block either of APE1’s repair or redox functions, but not both, will help determine how to best modify its function in treating different diseases. A unique compound called E3330 has single-agent cancer-cell killing abilities in cancer cell lines including ovarian, colon, lung, breast, brain, pancreatic, prostate and multiple myeloma cancers. It is significant that E3330 does not produce cell killing in normal cells, which implies a particular redox role of APE1 in cancer survival, but not “normal” cell survival. This suggests that a “therapeutic window” exists, as the results of the studies of E3330 demonstrate that it has a greater effect on differentiated endothelial tumor cells versus bone marrow stem cells (Zou et al. 2009). The collective data on E3330 suggest that APE1’s redox function is a promising target for cancer treatment Tables 7.2 and 7.3. Table 7.2 APE1’s relationship between redox activities and DNA repair APE1’s Influence on Redox Regulation of DNA Repair Pathways PATHWAY WHAT APE1 ACTS UPON DR Reduces p53, which dictates AGT’s level of activity, depending on the type of cytotoxic stress present Reduces NFkB, which appears to regulate AGT BER Reduces p53 to optimize its DNA-binding ability; which, in turn, influences DNA polymerase b activity NER Reduces p53 so it can activate DDB2 and XPC to recognize DNA damage in the GGR sub-pathway Reduces AP-1 so it can regulate the expression of several proteins, including those that repair DNA-cisplatin adducts MMR Reduces p53 so it can enable MSH2 to complex and recognize DNA damage Reduction of p53 also enables it to activate MLH1 and PMS2 Reduces AP-1 so it can regulate the expression of several proteins, including those that repair DNA-cisplatin adducts Stabilizes HIF-1a so it can activate HIF-1, which, in turn, inhibits NLH1 and MSH2 during hypoxia HR Reduces p53 so it can modulate RAD51 expression Stabilizes HIF-1a so it can activate HIF-1, which, in turn, inhibits BRCA1, BRCA2, and RAD51 during hypoxia NHEJ Reduces p53 so it can regulate the expression of RecQ helicases WRN and RecQ4
Is reduced by
Reduce(s)
Proximity of residues How reaction resolution occurs
TrxR+FADH2+NADPH
N-terminal cys is the nucleophile; attacks the target protein disulfide; creates a transient mixed disulfide intermediate, this is reduced by the C-terminal active site Cys, generat-ing a dithiol in the target protein and a disulfide in thioredoxin Disulfide bonds of signaling and regulatory proteins
Unknown
Disulfide bonds of signaling and regulatory proteins Thioredoxin (for redox activity; GADPH reduces APE1 for its endonuclease activity)
Close
Far apart
Table 7.3 Comparison of APE1’s redox features to other redox systems Redox system Feature APE1 Trx Cys65 Cys32+Cys35 Residues required for activity Location of active Near the N-terminus; Surface accessible; near the C-terminus residue(s) buried
Structural unfolding in the area of the thiolate near the C-terminus places it in close proximity to the nucleophilic thiolate
Hydrogen peroxide
Reduces the target protein disulfide; creates a transient mixed disulfide intermediate, GRX is oxidized; dithiol created in the protein Disulfide bonds of signaling and regulatory proteins Glutathione+glutathione reductase+NADPH
Sulfiredoxin
1 near the N-terminus; other near the C-terminus Far apart
N-terminus
Close
Peroxiredoxins 1 or 2 Cys (depends on which Prx type)
GRX/GSH Cys22+Cys25
7 DNA Repair and Redox Signaling 161
162
M.R. Kelley et al.
Conclusions Redox regulation has an important role in DNA repair, not only directly with the APE1 protein but also through other redox signaling and cellular redox status changes. APE1 has been a major focus of this chapter since it is clearly the DNA repair protein most directly involved in redox control and signaling. Additionally, we discuss recent data demonstrating a new role of APE1 with its involvement in angiogenesis and potential role in the tumor microenvironment. We expect that future research in this area will bring additional insights into the importance or redox regulation in DNA repair, particularly as more interest if brought to bear on this topic. These new studies will likely result in the identification of other proteins that also play important roles in redox regulation and DNA repair. Acknowledgments Financial support for this work was provided by the National Institutes of Health, National Cancer Institute CA106298, CA114571 and CA121168 to M.R.K., CA114571 to M.M.G. and the Riley Children’s Foundation to M.R.K., and National Cancer Institute CA122298 and to M.L.F.
References Abate C, et al (1990) Expression and purification of the leucine zipper and DNA-binding domains of Fos and Jun: both Fos and Jun contact DNA directly. Proc Natl Acad Sci USA 87(3):1032–1036. Abate C, et al (1990) Redox regulation of fos and jun DNA-binding activity in vitro. Science 249(4973):1157–1161. Adimoolam S, Ford JM(2002) p53 and DNA damage-inducible expression of the xeroderma pigmentosum group C gene. Proc Natl Acad Sci USA 99(20):12985–12990. Aiyer HS, Kichambare S, Gupta RC (2008) Prevention of oxidative DNA damage by bioactive berry components. Nutr Cancer 60(Suppl 1):36–42. Akamatsu Y, et al (1997) Redox regulation of the DNA binding activity in transcription factor PEBP2. J Biol Chem 272:14497–14500. Akterin S, et al (2006) Involvement of glutaredoxin-1 and thioredoxin-1 in beta-amyloid toxicity and Alzheimer’s disease. Cell Death Differ 13(9):1454–1465. Allinson SL, Dianova II, Dianov GL (2001) DNA polymerase beta is the major dRP lyase involved in repair of oxidative base lesions in DNA by mammalian cell extracts. EMBO J 20(23):6919–6926. Altieri F, et al (2008) DNA damage and repair: from molecular mechanisms to health implications. Antioxid Redox Signal 10(5):891–938. Arcangeli L, et al (1997) Site- and strand-specific mismatch repair of human H-ras genomic DNA in a mammalian cell line. Carcinogenesis 18(7):1311–1318. Arias-Lopez C, et al (2006) p53 modulates homologous recombination by transcriptional regulation of the RAD51 gene. EMBO Rep 7(2):219–224. Arner ES, Holmgren A (2000) Physiological functions of thioredoxin and thioredoxin reductase. Eur J Biochem 267(20):6102–6109. Arner ES, Holmgren A (2006) The thioredoxin system in cancer. Semin Cancer Biol 16(6):420–426. Bapat A, et al (2009) Going ape as an approach to cancer therapeutics. Antioxid Redox Signal 11(3):651–668.
7 DNA Repair and Redox Signaling
163
Barnes T, et al (2009) Identification of Apurinic/apyrimidinic endonuclease 1 (APE1) as the endoribonuclease that cleaves c-myc mRNA. Nucl Acids Res 37(12):3946–3958. Beckman RA, Loeb LA (2006) Efficiency of carcinogenesis with and without a mutator mutation. Proc Natl Acad Sci USA 103(38):14140–14145. Berndt C, Lillig CH, Holmgren A (2008) Thioredoxins and glutaredoxins as facilitators of protein folding. Biochim Biophys Acta (BBA): Mol Cell Res 1783(4):641–650. Bhakat KK, Mantha AK, Mitra S (2009) Transcriptional regulatory functions of mammalian AP-endonuclease (APE1/Ref-1), an essential multifunctional protein. Antioxid Redox Signal 11:621–638. Biaglow JE, Miller RA (2005) The thioredoxin reductase/thioredoxin system: novel redox targets for cancer therapy. Cancer Biol Ther 4(1):6–13. Bindra RS, Glazer PM (2007) Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene 26(14):2048–2057. Bindra RS, Glazer PM (2007), Co-repression of mismatch repair gene expression by hypoxia in cancer cells: role of the Myc/Max network. Cancer Lett 252(1):93–103. Bindra RS, et al (2004) Down-regulation of Rad51 and decreased homologous recombination in hypoxic cancer cells. Mol Cell Biol 24(19):8504–8518. Bindra RS, et al (2005) Alterations in DNA repair gene expression under hypoxia: elucidating the mechanisms of hypoxia-induced genetic instability. Ann N Y Acad Sci 1059:184–195. Bindra RS, Crosby ME, Glazer PM (2007) Regulation of DNA repair in hypoxic cancer cells. Cancer Metastasis Rev 26(2):249–260. Bishop AJ, et al (2003) Atm-, p53-, and Gadd45a-deficient mice show an increased frequency of homologous recombination at different stages during development. Cancer Res63(17):5335–5343. Bobola MS, et al (2005) Apurinic/apyrimidinic endonuclease activity is associated with response to radiation and chemotherapy in medulloblastoma and primitive neuroectodermal tumors. Clin Cancer Res 11(20):7405–7414. Bravard A, et al (2009) Oxidation status of human OGG1-S326C polymorphic variant determines cellular DNA repair capacity. Cancer Res 69(8):3642–3649. Bristow RG, Hill RP (2008) Hypoxia and metabolism. Hypoxia, DNA repair and genetic instability. Nat Rev Cancer 8(3):180–192. Burke-Gaffney A, Callister ME, Nakamura H (2005) Thioredoxin: friend or foe in human disease? Trends Pharmacol Sci 26(8):398–404. Burma S, Chen BP, Chen DJ (2006) Role of non-homologous end joining (NHEJ) in maintaining genomic integrity. DNA Repair (Amst) 5(9–10):1042–1048. Cao X, et al (2002) Oxidoreductive modification of two cysteine residues in paired domain by Ref-1 regulates DNA-binding activity of Pax-8. Biochem Biophys Res Commun 297(2):288–293. Chen DS, Olkowski ZL (1994) Biological responses of human apurinic endonuclease to radiationinduced DNA damage. Ann N Y Acad Sci 726:306–308. Chen DS, Herman T, Demple B (1991) Two distinct human DNA diesterases that hydrolyze 3’-blocking deoxyribose fragments from oxidized DNA. Nucl Acids Res 19(21):5907–5914. Chen J, Sadowski I (2005) Identification of the mismatch repair genes PMS2 and MLH1 as p53 target genes by using serial analysis of binding elements. Proc Natl Acad Sci USA 102(13):4813–4818. Chew EH, et al (2008) Thioredoxin reductase inhibition by antitumor quinols: a quinol pharmacophore effect correlating to antiproliferative activity. FASEB J 22(6):2072–2083. Christmann M, et al (2003) Mechanisms of human DNA repair: an update. Toxicology 193(1–2):3–34. Damia G, D’Incalci M (2007) Targeting DNA repair as a promising approach in cancer therapy. Eur J Cancer 43(12):1791–1801. Datta K, et al (2004) Sensitizing glioma cells to cisplatin by abrogating the p53 response with antisense oligonucleotides. Cancer Gene Ther 11(8):525–531. David SS, Williams SD (1998) Chemistry of glycosylases and endonucleases involved in baseexcision repair. Chem Rev 98(3):1221.
164
M.R. Kelley et al.
De Bont R, van Larebeke N (2004) Endogenous DNA damage in humans: a review of quantitative data. Mutagenesis 19(3):169–185. Demple B, Harrison L (1994) Repair of oxidative damage to DNA: enzymology and biology. Ann Rev Biochem 63:915–948. Di Maso V, et al (2007) Subcellular localization of APE1/Ref-1 in human hepatocellular carcinoma: possible prognostic significance. Mol Med 13(1–2):89–96. Doetsch PW, Cunningham RP (1990) The enzymology of apurinic/apyrimidinic endonucleases. Mutat Res 236(2–3):173–201. Eker AP, et al (2009) DNA repair in mammalian cells: Direct DNA damage reversal: elegant solutions for nasty problems. Cell Mol Life Sci 66(6):968–980. Ema M, et al (1999) Molecular mechanisms of transcription activation by HLF and HIF1alpha in response to hypoxia: their stabilization and redox signal-induced interaction with CBP/p300. EMBO J 18(7):1905–1914. Essers J, et al (2000) Homologous and non-homologous recombination differentially affect DNA damage repair in mice. EMBO J 19(7):1703–1710. Evans AR, Limp-Foster M, Kelley MR (2000) Going APE over ref-1. Mutat Res 461(2):83–108. Farmer H, et al (2005) Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434(7035):917–921. Fernandes AP, Holmgren A (2004) Glutaredoxins: glutathione-dependent redox enzymes with functions far beyond a simple thioredoxin backup system. Antioxid Redox Signal 6(1):63–74. Fishel ML, Kelley MR (2007) The DNA base excision repair protein Ape1/Ref-1 as a therapeutic and chemopreventive target. Mol Aspects Med 28(3–4):375–395. Fishel ML, et al (2007) Manipulation of base excision repair to sensitize ovarian cancer cells to alkylating agent temozolomide. Clin Cancer Res 13(1):260–267. Fishel ML, et al (2008) Knockdown of the DNA repair and redox signaling protein Ape1/Ref-1 blocks ovarian cancer cell and tumor growth. DNA Repair (Amst) 7(2):177–186. Fishel R, Wilson T (1997) MutS homologs in mammalian cells. Curr Opin Genet Dev 7(1):105–113. Fleck O and O. Nielsen (2004) DNA repair. J Cell Sci 117(Pt 4):515–517. Forman-Kay JD, Clore GM, Gronenborn AM (1992) Relationship between electrostatics and redox function in human thioredoxin: characterization of pH titration shifts using two-dimensional homo- and heteronuclear NMR. Biochemistry 31(13):3442–3452. Forsythe JA, et al (1996) Activation of vascular endothelial growth factor gene transcription by hypoxia-inducible factor 1. Mol Cell Biol 16(9):4604–4613. Friedberg EC (2001) How nucleotide excision repair protects against cancer. Nat Rev Cancer 1(1):22–33. Fung H, Demple B (2005) A vital role for Ape1/Ref1 protein in repairing spontaneous DNA damage in human cells. Mol Cell 17(3):463–470. Georgiadis M, et al.(2008) Evolution of the redox function in mammalian apurinic/apyrimidinic. Mutat Res 643(1–2):54–63. Giles NM, Giles GI, Jacob C (2003) Multiple roles of cysteine in biocatalysis. Biochem Biophys Res Commun 300(1):1–4. Grombacher T, Eichhorn U, Kaina B (1998) P53 is involved in regulation of the DNA repair gene O-6-methylguanine-DNA methyltransferase (MGMT) by DNA damaging agents. Oncogene 17(7):845–851. Guo HH, Loeb LA (2003) Tumbling down a different pathway to genetic instability. J Clin Invest 112(12):1793–1795. Guo Y, et al (2008) Granzyme K degrades the redox/DNA repair enzyme Ape1 to trigger oxidative stress of target cells leading to cytotoxicity. Mol Immunol 45(8):2225–2235. Hainaut P, Milner J (1993) Redox modulation of p53 conformation and sequence-specific DNA binding in vitro. Cancer Res 53(19):4469–4473. Handel ML, et al (1995) Inhibition of AP-1 binding and transcription by gold and selenium involving conserved cysteine residues in Jun and Fos. Proc Natl Acad Sci USA 92(10):4497–4501. Hanson S, Kim E, Deppert W (2005) Redox factor 1 (Ref-1) enhances specific DNA binding of p53 by promoting p53 tetramerization. Oncogene 24(9):1641–1647.
7 DNA Repair and Redox Signaling
165
He F, et al (2003) DNA polymerase beta gene expression: the promoter activator CREB-1 is upregulated in Chinese hamster ovary cells by DNA alkylating agent-induced stress. Biol Chem 384(1):19–23. Helton ES, Chen X (2007) p53 modulation of the DNA damage response. J Cell Biochem 100(4):883–896. Herring CJ, et al (1998) Levels of the DNA repair enzyme human apurinic/apyrimidinic endonuclease (APE1, APEX, Ref-1) are associated with the intrinsic radiosensitivity of cervical cancers. Br J Cancer 78(9):1128–1133. Hess J, Angel P, Schorpp-Kistner M (2004) AP-1 subunits: quarrel and harmony among siblings. J Cell Sci 117(Pt 25):5965–5973. Hirota K, Matsui M, Yodoi J (1997) AP-1 transcriptional activity is regulated by a direct association between thioredoxin and Ref-1. Proc Natl Acad Sci USA 94(8):3633–3638. Hirota K, et al (1999) Distinct roles of thioredoxin in the cytoplasm and in the nucleus. A two-step mechanism of redox regulation of transcription factor NF-kappaB. J Biol Chem 274(39):27891–27897. Hoeijmakers JH (2001) DNA repair mechanisms. Maturitas 38(1):17–22; discussion 22–23. Hoeijmakers JH (2001) Genome maintenance mechanisms for preventing cancer. Nature 411(6835):366–374. Hofmann B, Hecht HJ, Flohe L (2002) Peroxiredoxins. Biol Chem 383(3–4):347–364. Hollander MC, et al (2005) Deletion of XPC leads to lung tumors in mice and is associated with early events in human lung carcinogenesis. Proc Natl Acad Sci USA 102(37):13200–13205. Holmgren A (1989) Thioredoxin and glutaredoxin systems. J Biol Chem 264(24):13963–13966. Holmgren A (1995) Thioredoxin structure and mechanism: conformational changes on oxidation of the active-site sulfhydryls to a disulfide. Structure 3(3):239–243. Hol WG (1985) The role of the alpha-helix dipole in protein function and structure. Prog Biophys Mol Biol 45(3):149–195. Hromas R, et al (2008) The human set and transposase domain protein Metnase interacts with DNA Ligase IV and enhances the efficiency and accuracy of non-homologous end-joining. DNA Repair (Amst) 7(12):1927–1937. Huang LE, et al (1996) Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its alpha subunit. J Biol Chem 271(50):32253–32259. Huang RP, Adamson ED (1993) Characterization of the DNA-binding properties of the early growth response-1 (Egr-1) transcription factor: evidence for modulation by a redox mechanism. DNA & Cell Biol 12(3):265–273. Hwang BJ, et al (1999) Expression of the p48 xeroderma pigmentosum gene is p53-dependent and is involved in global genomic repair. Proc Natl Acad Sci USA 96(2):424–428. Izumi T, et al (2005) Two essential but distinct functions of the mammalian abasic endonuclease. Proc Natl Acad Sci USA 102(16):5739–5743. Jacob C, et al (2003) Sulfur and selenium: the role of oxidation state in protein structure and function. Angew Chem Int Ed Engl 42(39):4742–4758. Jayaraman L, et al (1997) Identification of redox/repair protein Ref-1 as a potent activator of p53. Genes and Develop 11:558–570. Jiang BH, et al (1996) Dimerization, DNA binding, and transactivation properties of hypoxiainducible factor 1. J Biol Chem 271(30):17771–17778. Jiang Y, et al (2009) Role of APE1 in differentiated neuroblastoma SH-SY5Y cells in response to oxidative stress: use of APE1 small molecule inhibitors to delineate APE1 functions. DNA Repair 8:1273–1282. Jiricny J (2006) The multifaceted mismatch-repair system. Nat Rev Mol Cell Biol 7(5):335–346. Kaina B, et al (2007) MGMT: key node in the battle against genotoxicity, carcinogenicity and apoptosis induced by alkylating agents. DNA Repair (Amst) 6(8):1079–1099. Kelley MR, Fishel ML (2008) DNA repair proteins as molecular targets for cancer therapeutics. Anticancer Agents Med Chem 8(4):417–425. Kondo N, et al (2006) Redox regulation of human thioredoxin network. Antioxid Redox Signal 8(9–10):1881–1890.
166
M.R. Kelley et al.
Koshiji M, et al (2005) HIF-1alpha induces genetic instability by transcriptionally downregulating MutSalpha expression. Mol Cell 17(6):793–803. Kruman, II, et al (2004) Cell cycle activation linked to neuronal cell death initiated by DNA damage. Neuron 41(4):549–561. Lando D, et al (2000) A redox mechanism controls differential DNA binding activities of hypoxiainducible factor (HIF) 1alpha and the HIF-like factor. J Biol Chem 275(7):4618–4627. Langie SA, et al (2007) The role of glutathione in the regulation of nucleotide excision repair during oxidative stress. Toxicol Lett 168(3):302–309. Lau JP, et al (2004) Effects of gemcitabine on APE/ref-1 endonuclease activity in pancreatic cancer cells, and the therapeutic potential of antisense oligonucleotides. Br J Cancer 91(6):1166–1173. Lavon I, et al (2007) Novel Mechanism whereby nuclear factor {kappa}B mediates DNA damage repair through regulation of O6-methylguanine-DNA-methyltransferase. Cancer Res 67(18):8952–8959. Lee SH, et al (2005) The SET domain protein Metnase mediates foreign DNA integration and links integration to nonhomologous end-joining repair. Proc Natl Acad Sci USA 102(50):18075–18080. Lillig CH, Holmgren A (2007) Thioredoxin and related molecules – from biology to health and disease. Antioxid Redox Signal 9(1):25–47. Lillig CH, Berndt C, Holmgren A (2008) Glutaredoxin systems. Biochim Biophys Acta 1780(11):1304–1317. Li SL, et al (2001) Transcriptional regulation of the p67phox gene: role of AP-1 in concert with myeloid-specific transcription factors. J Biol Chem 276(42):39368–39378. Liu F, Fu Y, Meyskens FL, Jr (2008) MiTF regulates cellular response to reactive oxygen species through transcriptional regulation of APE-1//Ref-1. J Invest Dermatol 129:422–431. Liu Z, et al (2009) Inhibition of thioredoxin reductase by mansonone F analogues: Implications for anticancer activity. Chem Biol Interact 117(1):48–57. Loboda A, et al (2009) HIF-1 attenuates Ref-1 expression in endothelial cells: Reversal by siRNA and inhibition of geranylgeranylation. Vascul Pharmacol 51(2–3):133–139. Lopez de Saro FJ, et al (2006) The beta sliding clamp binds to multiple sites within MutL and MutS. J Biol Chem 281(20):14340–14349. Luo M, et al (2008) Role of the multifunctional DNA repair and redox signaling protein Ape1/ Ref-1 in cancer and endothelial cells: Small molecule inhibition of Ape1’s redox function. Antioxid Redox Signal 10(11):1853–1867. Luo M, et al (2010) Redox regulation of DNA repair: implications for human health and cancer therapeutic development. Antioxid Redox Signal 12:1247–1269. Lu X, Lozano G, Donehower LA (2003) Activities of wildtype and mutant p53 in suppression of homologous recombination as measured by a retroviral vector system. Mutat Res 522(1–2):69–83. Madhusudan S, Hickson ID (2005) DNA repair inhibition: a selective tumour targeting strategy. Trends Mol Med 11(11):503–511. Madhusudan S, et al (2005) Isolation of a small molecule inhibitor of DNA base excision repair. Nucleic Acids Res 33(15):4711–4724. Mathers J, Coxhead J, Tyson J (2007) Nutrition and DNA Repair – potential molecular mechanisms of action. Curr Cancer Drug Targets 7(5):425–431. McNeill DR, Wilson DM 3rd (2007) A dominant-negative form of the major human abasic endonuclease enhances cellular sensitivity to laboratory and clinical DNA-damaging agents. Mol Cancer Res 5(1):61–70. Meng AX, et al (2005) Hypoxia down-regulates DNA double strand break repair gene expression in prostate cancer cells. Radiother Oncol 76(2):168–176. Merluzzi S, et al (2008) APE/Ref-1 makes fine-tuning of CD40-induced B cell proliferation. Mol Immunol 45(14):3731–3739. Mihaylova VT, et al (2003) Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol Cell Biol 23(9):3265–3273.
7 DNA Repair and Redox Signaling
167
Moore DH, et al (2000) Alterations in the expression of the DNA repair/redox enzyme APE/ref-1 in epithelial ovarian cancers. Clin Cancer Res 6(2):602–609. Mukherjee A, Martin SG (2008) The thioredoxin system: a key target in tumour and endothelial cells. Br J Radiol 81(1):S57–S68. Nakamura H (2005) Thioredoxin and its related molecules: update 2005. Antioxid Redox Signal 7(5–6):823–828. Nakamura H, Nakamura K, Yodoi J (1997) Redox regulation of cellular activation. Annu Rev Immunol 15:351–369. Ono Y, et al (1994) Stable expression in rat glioma cells of sense and antisense nucleic acids to a human multifunctional DNA repair enzyme, APEX nuclease. Mutat Res 315(1):55–63. Powis G, Montfort WR (2001) Properties and biological activities of thioredoxins. Annu Rev Pharmacol Toxicol 41:261–295. Raffel J, et al (2003) Increased expression of thioredoxin-1 in human colorectal cancer is associated with decreased patient survival. J Lab Clin Med 142(1):46–51. Rafferty JA, et al (1996) Induction of murine O6-alkylguanine-DNA-alkyltransferase in response to ionising radiation is p53 gene dose dependent. Oncogene 12(3):693–697. Reed AM, Fishel ML, Kelley MR (2009) Small-molecule inhibitors of proteins involved in base excision repair potentiate the anti-tumorigenic effect of existing chemotherapeutics and irradiation. Future Oncol 5(5):713–726. Scherer SJ, et al (2000) p53 and c-Jun functionally synergize in the regulation of the DNA repair gene hMSH2 in response to UV. J Biol Chem 275(48):37469–37473. Schwartz EI, et al (2007) Cell cycle activation in postmitotic neurons is essential for DNA repair. Cell Cycle 6(3):318–329. Sengupta S, Harris CC (2005) p53: traffic cop at the crossroads of DNA repair and recombination. Nat Rev Mol Cell Biol 6(1):44–55. Sengupta S, et al (2005) Tumor suppressor p53 represses transcription of RECQ4 helicase. Oncogene 24(10):1738–1748. Shinohara ET, et al (2005) DNA-dependent protein kinase is a molecular target for the development of noncytotoxic radiation-sensitizing drugs. Cancer Res 65(12):4987–4992. Smith ML, Seo YR (2002) p53 regulation of DNA excision repair pathways. Mutagenesis 17(2):149–156. Song YJ, et al (2008) Tat-APE1/ref-1 protein inhibits TNF-[alpha]-induced endothelial cell activation. Biochem Biophys Res Commun 368(1):68–73. Sung P, Klein H (2006) Mechanism of homologous recombination: mediators and helicases take on regulatory functions. Nat Rev Mol Cell Biol 7(10):739–750. Tan T, Chu G (2002) p53 Binds and activates the xeroderma pigmentosum DDB2 gene in humans but not mice. Mol Cell Biol 22(10):3247–3254. Tell G, et al (2000) An ‘environment to nucleus’ signaling system operates in B lymphocytes: redox status modulates BSAP/Pax-5 activation through Ref-1 nuclear translocation. Nucleic Acids Res 28(5):1099–1105. Tell G, et al (2005) The intracellular localization of APE1/Ref-1: more than a passive phenomenon? Antioxid Redox Signal 7(3–4):367–384. Tell G, et al (2009) The many functions of APE1/Ref-1: not only a DNA repair enzyme. Antioxid Redox Signal 11(3):601–620. To KK, et al (2006) The phosphorylation status of PAS-B distinguishes HIF-1alpha from HIF2alpha in NBS1 repression. EMBO J 25(20):4784–4794. Tonissen KF, Di Trapani G (2009) Thioredoxin system inhibitors as mediators of apoptosis for cancer therapy. Mol Nutr Food Res 53(1):87–103. Ueno M, et al (1999) Thioredoxin-dependent redox regulation of p53-mediated p21 activation. J Biol Chem 274(50):35809–35815. Vascotto C, et al (2009) APE1/Ref-1 interacts with NPM1 within nucleoli and plays a role in the rRNA quality control process. Mol Cell Biol 29(7):1834–1854.
168
M.R. Kelley et al.
Walker LJ, et al (1993) Identification of residues in the human DNA repair enzyme HAP1 (Ref-1) that are essential for redox regulation of Jun DNA binding. Mol Cell Biol 13(9):5370–5376. Walker LJ, et al (1994) A role for the human DNA repair enzyme HAP1 in cellular protection against DNA damaging agents and hypoxic stress. Nucl Acids Res 22(23):4884–4889. Wang D, Luo M, Kelley MR (2004) Human apurinic endonuclease 1 (APE1) expression and prognostic significance in osteosarcoma: enhanced sensitivity of osteosarcoma to DNA damaging agents using silencing RNA APE1 expression inhibition. Mol Cancer Ther 3(6):679–686. Weichsel A, et al (1996) Crystal structures of reduced, oxidized, and mutated human thioredoxins: evidence for a regulatory homodimer. Structure 4(6):735–751. Wood RD, Mitchell M, Lindahl T (2005) Human DNA repair genes, 2005. Mutat Res 577(1–2):275–283. Wood ZA, Poole LB, Karplus PA (2003) Peroxiredoxin evolution and the regulation of hydrogen peroxide signaling. Science 300(5619):650–653. Xanthoudakis S, Curran T (1992) Identification and characterization of Ref-1, a nuclear protein that facilitates AP-1 DNA-binding activity. EMBO J 11(2):653–665. Xanthoudakis S, Curran T (1996) Redox regulation of Ap-1: a link between transcription factor signaling and DNA repair. Adv Exp Med Biol 387:387. Xanthoudakis S, et al (1992) Redox activation of Fos-Jun DNA binding activity is mediated by a DNA repair enzyme. EMBO J 11(9):3323–3335. Xanthoudakis S, et al (1996) The redox/DNA repair protein, Ref-1, is essential for early embryonic development in mice. Proc Natl Acad Sci USA 93(17):8919–8923. Xu J, Morris GF (1999) p53-mediated regulation of proliferating cell nuclear antigen expression in cells exposed to ionizing radiation. Mol Cell Biol 19(1):12–20. Xu Y, et al (1997) The apurinic/apyrimidinic endonuclease (APE/ref-1) DNA repair enzyme is elevated in premalignant and malignant cervical cancer. Anticancer Res 17(5B):3713–3719. Yamabe Y, et al (1998) Sp1-mediated transcription of the Werner helicase gene is modulated by Rb and p53. Mol Cell Biol 18(11):6191–6200. Yamawaki H, Berk BC (2005) Thioredoxin: a multifunctional antioxidant enzyme in kidney, heart and vessels. Curr Opin Nephrol Hypertens 14(2):149–153. Zaky A, et al (2008) Regulation of the human AP-endonuclease (APE1/Ref-1) expression by the tumor suppressor p53 in response to DNA damage. Nucl Acids Res 36(5):1555–1566. Zou G-M, Maitra A (2008) Small-molecule inhibitor of the AP endonuclease 1/REF-1 E3330 inhibits pancreatic cancer cell growth and migration. Mol Cancer Ther 7(7):2012–2021. Zou GM, et al (2009) The Ape-1/Ref-1 redox antagonist E3330 inhibits the growth of tumor endothelium and endothelial progenitor cells: therapeutic implications in tumor angiogenesis. J Cell Physiol 219(1):209–218. Zurer I, et al (2004) The role of p53 in base excision repair following genotoxic stress. Carcinogenesis 25(1):11–19.
Chapter 8
Cancer Stem Cells and Microenvironment Mario Federico and Antonio Giordano
Abstract The theory of the cancer stem cell (CSC) is fairly recent and has both challenged and disrupted the previous understandings of cancer biology. From the initial findings of cancer-driving cellular sub-populations, the interest in the CSC theory has flourished. Here we discuss the biology behind both embryonic and adult stem cells and how this biology is the basis for our understanding of CSCs. Furthermore, we elaborate on findings demonstrating the importance of the microenvironment on the stem cell and how changes in the microenvironment alone may direct stem cell differentiation or possibly tumorigenesis. Cancer research has taken a new fork in the road towards the involvement of CSCs and this road has been productive. However, the challenge remains in understanding and identifying CSCs independently from the embryonic and adult stem cell models of the past. Moving beyond the preconceived understanding of stem cells will allow researchers to be able to fully explore the complexity and role of CSCs in cancer and hopefully provide the knowledge necessary to combat cancer from that first initiating cell.
Introduction The traditional model explaining tumor progression assumes that the tumor cell, through a series of stochastic events, acquires increased proliferative capability because of the deregulation of oncogenes and/or tumor suppressor genes. Once the cancerous phenotype is established, every cell within the tumor has the same capacity to proliferate extensively with no differences in the intrinsic tumorigeneity. The occurrence of additional stochastic genetic events (conferring to the tumor cell an increased malignant potential) will lead to the substitution of the former cancer clone with a more aggressive clone, otherwise known as clonal selection. A. Giordano (*) Biology Department, Sbarro Health Research Organization, Temple University, Philadelphia, PA, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_8, © Springer Science+Business Media, LLC 2010
169
170
M. Federico and A. Giordano
In 1997 Bonnet and Dick (1997) challenged this paradigm. They isolated a subpopulation of CD34+/CD38− leukemic precursors from human AML and demonstrated that only those cells were able to initiate leukemia when engrafted into NOD/SCID mice, whereas the other fractions of leukemic blasts representing the vast majority of all leukemic cells (CD34+, CD38+, or CD34−) were unable to do so. Moreover, they found that engrafted CD34+/CD38− leukemic precursors could be serially transplanted into secondary recipients, thus providing evidence of an intrinsic self-renewal capability. After this observation, several other groups also found tumorigenic self-renewal stem-like populations in a wide spectrum of solid tumors, and nice reviews have been published so far (Kakarala and Wicha 2008; Dirks 2008; Maitland and Collins 2008; Sell and Leffert 2008; Boman and Huang 2008; Takaishi et al. 2008; Peacock and Watkins 2008; Huff and Matsui 2008; Lee et al. 2008; Prince and Ailles 2008). Self-renewal and differentiation capability are the hallmarks of staminality, a concept that traditionally has been associated with totipotent embryonic cells (embryonic stem cells, ESCs), which possess characteristics such as asymmetric cellular division and the ability to differentiate into all three germ layers originating all the different cell types within the body. Stem-like cellular populations also exist in adult organisms and are defined adult stem cells (ASCs). ASCs are immature, undifferentiated, and functionally unspecialized cells, which exist in small numbers in nearly all post-embryonic tissues. ASC populations are thought to give rise to multiple, more differentiated cell types, maintaining a long-term capacity for selfrenewal, and so regulating homeostatic tissue regeneration. In contrast to ESCs, ASCs lack the capacity to differentiate into the three germ layers, are capable of regenerating a cellular population of a specific tissue type, and maintain asymmetric cellular division. ASCs are characterized as being in a state of relative proliferative quiescence, with a high dependence on the microenvironment in which they reside: the so-called stem cell niche. Under the proper conditions, they can exit from the quiescent state to obtain the proliferative potential necessary for tissue regeneration or wound repair. Bonnet and Dick sustained the hypothesis that tumors, similar to normal adult tissues, are constituted with a heirarchy of cell types (Fig. 8.1). Stem or stem-like cells (heirin referred to as cancer stem cells or CSCs) are a relatively small population within this heirarchy and are the only population that maintains the ability to initiate tumorigenesis by undergoing self-renewal and differentiation. The differentiation of this specialized population leads to the loss in self-renewal capacity and physically creates the vast majority of cells found within a tumor. This change of landscape is a radical shift from the traditional stochastic model. First of all because it assumes that not any cell can originate a tumor but just a stem cell, which through an acquired deregulation of the self-renewal process becomes tumorigenic. The second reason is that it postulates a histological hierarchy within the tumor, further implying (by analogy with ESCs and ASCs) a series of mechanisms that control CSCs along their differentiation path. Since stem cells generally have a higher resistance to chemo- and radio-therapy than committed cells, CSC
8 Cancer Stem Cells and Microenvironment
171
Fig. 8.1 The CSC hypothesis
theory has been advocated to explain clinical relapses after cancer therapy. In fact, chemo- and radio-therapy normally produce a significant shrinkage of the tumor mass acting mainly on the “mature” cancer cells with limited effect on CSCs or tumor progenitors. The latter is postulated to be the engine of tumor re-formation, leading to subsequent clinical relapse. This hypothesis has been recently validated in a breast cancer clinical trial, in which patients receiving cytotoxic chemotherapy displayed a marked increase in breast CSC frequency in the residual tumor, while a targeted anti-CSC therapeutic stabilized the CSC population (Li et al. 2008). Finally, from the cellular biology point of view, the CSC paradigm challenges our knowledge of the molecular mechanisms that underlie tumorigeneity. What are the molecular mechanisms behind self-renewal and how is this maintained over time? What stimuli are necessary for differentiation to occur and how are these stimuli limited to only a portion of the population? And how does the environment around the stem cell population affect these events? In the last few years, an exponentially growing number of publications have addressed these issues; however, the majority of knowledge on the general biological mechanisms of staminality still derives from the studies on ESCs and ASCs. Even if we can “phenomenologically” define CSCs through their attitude toward self-renewal and differentiation into a “mature” cancer phenotype, we are still dependent upon the ESC and ASC knowledge to guide us through this novel research field in oncology.
172
M. Federico and A. Giordano
ESCs: A Prototype Model of Stem Cell Biology Thomson and Gearhart isolated and characterized the initial human ESCs (hESCs) from the inner cell mass (ICM) of the blastocyst, noting the self-renewing and differentiation capacity of the cells in vitro (Shamblott et al. 1998; Thomson et al. 1998). These cells, through further characterization, were found to express cell surface markers typical of undifferentiated nonhuman primate ESCs (pESCs) and human embryonic carcinoma cells as was originally described (Shamblott et al. 1998; Thomson et al. 1998; Martin and Evans 1974, 1975). These specific markers included stage-specific embryonic antigen (SSEA)-1, SSEA-3, SSEA-4, TRA-1– 60, TRA-1–81, alkaline phosphatase activity, and high levels of telomerase activity. Telomerase is a ribonucleoprotein enzyme that preserves the telomeric regions at the ends of chromosomes by de novo oligonucleotide synthesis (Greider and Blackburn 1987) and is not present in normal diploid somatic cells, which incur shorted telomeres with age, leading to replicative senescence after a finite number of replications (Hayflick 1965; Hayflick and Moorhead 1961; Harley et al. 1990; Allsopp et al. 1992). It has been shown that TRA-1–60 and TRA-1–81 are specific epitopes of a larger membrane-bound protein podocalyxin, which undergoes retinoic acid modification when ESCs differentiate losing its reactivity with the TRA-1–60 and TRA-1–81 antibodies (Schopperle and DeWolf 2007). These characterizations remain to be used to identify stem cells today, along with the expression of the intrinsic transcription factor Oct-4 and in mouse ESCs (mESCs) the constitutive ability to receive extrinsic signals from the cytokine leukemia inhibitory factor (LIF) (Niwa et al. 1998; Nichols et al. 1998; Fuchs and Segre 2000). The specific characteristic of ESCs is to differentiate during development into many different “mature” phenotypes, realizing a sophisticated balance between pluripotency and commitment that allows the timely and spatially appropriate embryo development. Although not completely elucidated, such a unique characteristic relies upon at least four general different mechanisms: a distinctive transcriptional hierarchy, a balanced epigenetic state, a particular cell cycle regulation, and a tight crosstalk with the microenvironment. Transcriptional regulators contribute to the maintenance of pluripotency and three major players have been identified so far in ESCs: POU5FI, NANOG, and SOX2. POU5FI is a member of the POU transcription factor family and is expressed in the mouse embryo from the four to the eight cells stage until the epiblast starts to differentiate. Its presence is essential for the initial development of pluripotency in the ICM and high levels of this transcription factor are restricted to pluripotent lineages (Wobus and Boheler 2005). Mouse embryos lacking POU5FI protein die following implantation because of the lack of ICM (Nichols et al. 1998). POU5FI represses differentiation to the trophoblast; however, appropriate levels are necessary to maintain pluripotency. In fact, a twofold increase in its expression causes differentiation into primitive endoderm and mesoderm, while loss of POU5FI induces the formation of trophectoderm and concomitant loss of pluripotency (Niwa et al. 2000). NANOG homeodomain protein in pre-implantation embryos is restricted to the inner cells of the compacted morula and blastocyst
8 Cancer Stem Cells and Microenvironment
173
and conversely is absent in differentiated cells (Chambers et al. 2003). High expression levels of NANOG confer constitutive self-renewal capacity to mESCs in the absence of LIF stimulation and can be propagated in serum-free media in the absence of otherwise needed stimuli, such as bone morphogenetic protein (Ying et al. 2003). NANOG knockout mice develop to the blastocyst stage, but when cultured, the ICM differentiates to parietal endoderm-like cells. From these findings, NANOG appears to be involved not only in pluripotency maintenance but also in the inhibition of the transition to primitice endoderm. Finally, the SOX2 gene is expressed in the pluripotent lineages of early mouse embryos and functions as a transcription factor with an HMG-DNA-binding domain. Sox2−/− mice fail to develop beyond implantation (Avilion et al. 2003) and down regulation of Sox2 is associated with differentiation. A ChIP-on-chip analysis revealed that all three of these factors co-occupy a substantial portion of target genes (Boyer et al. 2005) with an evident overlapping function. The generation of induced pluripotent stem cells using PUI5FI, SOX2, NANOG, and LIN28 along with the generation of pluripotent cells using POU5FI and SOX2 alone with the histone deacetylase inhibitor, valproic acid (Huangfu et al. 2008), have further demonstrated the role of these transcription factors in maintaining pluripotency. The second crucial mechanism that regulates ESC development relies on a precise epigenetic control of gene expression patterns. The term epigenetics specifically refers to heritable genotypic changes that are caused by mechanisms other than changes in the primary DNA sequence. Central to epigenetic control is chromatin, a complex of DNA, histones, and other proteins that make up chromosomes (Misteli 2004, 2007; Meshorer 2007). Histones are subject to several post-translational modifications that occur primarily in their amino terminal tails including methylation, acetylation, phosphorylation, sumoylation, and ubiquitination (Kouzarides 2007). During differentiation, genes regulating pluripotency and self-renewal become progressively silenced, whereas genes controlling cell-type specific functions are gradually turned on. These alterations in gene expression patterns are orchestrated by defined epigenetic programs that are capable of indirectly altering the expression of hundreds to thousands of genes, proving the efficiency of this type of control in establishing cellular identities (Lunyak and Rosenfeld 2008; Cheng et al. 2005; Shukla et al. 2008; Wu and Sun 2006). Based on a genome-wide analyses of chromatin patterns of multiple cell types, several observations have emerged: acetylation of lysine residues on histones H3 and H4 and methylation of lysine 4 of histone H3 (H3K4me), in general, are associated with transcriptionally active chromatin, whereas methylation of histone H3 lysine 9 (H3K9me) and 27 (H3K27me) are hallmarks of transcriptionally silenced chromatin (Kouzarides 2007). Several studies demonstrated that the differentiation of ESCs is accompanied with an increase in histone H3K9me3 and decreased acetylation of histone H3 and H4 (Lee et al. 2004; Meshorer et al. 2006). However, further studies will be necessary to elucidate the role of epigenetics in stem cell regulation. The third major difference between stem cells and somatic cells is found in the basic regulation of the cell cycle. In somatic cells, the cell cycle is controlled mainly by Rb-E2F family complexes, cyclin–cyclin-dependent kinases (Cdks), and
174
M. Federico and A. Giordano
Cdk inhibitors through the INK4a/ARF pathway. Undulations in expression and post-translational modifications of the proteins involved in these pathways result in the control and regulation of the cell cycle. Similarly, mutations or de-regulation of these proteins can lead to uncontrolled cell proliferation, aneuploidy, and genomic instability (Haas et al. 1997; Mumberg et al. 1996). The cell cycle regulatory mechanisms, which differ between somatic cells and ESCs, have been determined using the mESC model in combination with mESCs, representing a pluripotent lineage (mEPLC) (Rathjen et al. 1999). mESCs of late pre-implantation and early post-implantation embryos proliferate at an unusually rapid rate (Solter et al. 1971). Between 4.5 and 6.0 dpc (days post coitum), the epiblast expands with a generation time of approximately 10 h (Hogan et al. 1994). This increases between 6.5 and 7.0 dpc, where mean generation times are found to be approximately 4.4 h (Hogan et al. 1994; Power and Tam 1993). The cell cycle in mESCs and mEPLCs has been found to curtail G1 and G2 phases with an increased proportion of the cycle, approximately 50–60%, spent in S phase (Savatier et al. 1994; Stead et al. 2002). Under normal somatic cell cycle conditons, Rb/p105, in the hypophosphorylated state, interacts with E2F transcription factors inhibiting the transcription of genes necessary for the progression of the cell cycle through the restriction point (R point). The phosphorylation levels of Rb/p105 are dependent upon the CDK activity present in the cell. Mitogen signaling through the Ras/Raf/mitogen activated protein kinase (MAPK) pathway activates the cyclin D–CDK4/6 complexes, which are believed to initially activate Rb/p105 activity by hypophosphorylating the unphosphorylated protein. To pass the R point of the cell cycle, cyclin E/CDK2 hyperphosphorylates Rb/p105 inhibiting the protein from binding to E2F transcription factors, thus initiating the transcription of genes required in the S phase of the cell cycle. To obtain a cell cycle that is less influenced by mitogen variations, the stem cells appear to adopt a different regulation mechanism as depicted in Fig. 8.2 (Giacinti and Giordano 2006; Kasten 1998). Along with shortened gap phases in the ESC cell cycle, the R point does not seem to regulate the G1–S transition. Stead and collegues found that in both mESCs and mEPLCs, there was a precocious cell cycle-independent expression of CDK2, cyclin A and cyclin E kinase activity (Kasten and Giordano 1998). Furthermore, when CDK2 was suppressed, they found a significant decrease in cell proliferation rate. Instead of CDC2, cyclin B, essential to G2 – M transition, was the only CDK activity that was found to be cell cycle-dependent and E2F target genes were constitutively expressed throughout the cell cycle (Stead et al. 2002). Evidence has also shown a lack in hypophosphorylated Rb/p105, instead findings support the presence of hyperphosphorylated Rb/p105 in mESCs and mEPLCs (Savatier et al. 1994; Fraichard et al. 1995). of cell cycle. There is cell cycle-independent expression of cyclin E – CDK2 maintaining the hyperphosphorylated levels of Rb family member proteins. This results in cell cycle-independent expression of E2F-regulated genes. Cyclin B – CDC2 is the only CDK activity that appears to be regulated by the cell cycle. ESCs have shortened gap phases and an elongated S phase of the cell cycle, with an apparent lack in the R point for G1-S transition
8 Cancer Stem Cells and Microenvironment
175
Fig. 8.2 Specific characteristics of pluripotent cells a) Interplay between the key transcriptional regulators POU5F1, NANOG and SOX2 and their effect on the formation of primitive endoderm and trophectoderm. b) Schematic representation of a nucleosome model and major post-translational modifications which play essential roles in gene expression regulation. c) Cell cycle in somatic cells vs. ESCs. Cell cycle regulation in somatic cells: mitogen signaling through MAPK pathway activates cyclin D – CDK4/6 kinase activity hypophosphorylating Rb family member proteins. Hypophosphorylated Rb family member proteins bind to E2F transcription factors blocking the transcription of E2F-regulated genes. To surpass the R point cyclin E – CDK2 kinase activity is activated hyperphosphorylating Rb family member proteins. Hyperphosphorylated Rb family member proteins are unable to interact with E2F factors, allowing them to activate transcription of genes necessary in the progression of cell cycle. Cell cycle regulation in ESCs as is currently understood: Mitogen signaling through MAPK pathways seems to be irrelevant in the progression
176
M. Federico and A. Giordano
Given the cell-cycle independent expression of cyclin E and CDK2, it would be logical that Rb/p105 would be found in the hyperphosphorylated state, further supporting the absence of the R point in ESC cell cycle progression (Fig. 8.1). Mitogen signaling through the MAPK pathway normally stimulates cell division in somatic cells; however, this signaling when prolonged is a potent inducer of differentiation. mESCs appear to avoid this stimulation by maintaining low levels of cyclin D expression and almost no detectable CDK4 kinase activity (Fluckinger et al. 2006). This corresponds to the lack in hypophosphorylated Rb/p105 levels previously detected in mESCs. These findings support the absence of early G1 in mESCs, allowing them to avoid the differentiation-inducing effects of MAPK signaling as is found in other cell types. Although the majority of studies thus far have been performed in mESCs, hESCs similarly show a truncation of the G1 phase of the cell cycle; however, not much else is known about cell cycle regulation in hESCs. Interestingly, pESCs behave similarly to mESCs in having cell cycle-independent expression of cyclin E, constitutive hyperphosphorylation of Rb/p105 and serum and MAPK-independent cell cycle progression (Fluckinger et al. 2006; Becker et al. 2006). Therefore, it could be inferred that through conserved evolution, hESCs may regulate the cell cycle in a similar fashion. Taken together, these data lead to the hypothesis that the ESC cell cycle is rate-dependent upon high levels of CDK activity, is not regulated by Rb/p105 or E2F gene expression, and lacks the G1 check point and the traditional periodicity found during the somatic cell cycle.
Stem Cells and Microenvironment Soon after the initial isolations and characterizations of hESCs, interest shifted toward understanding the factors involved in their differentiation. For example, if all cells are derived from initial progenitor cells, what directs the timely and spatially appropriate differentiation toward various mature cell types (i.e., glial cells versus adipocytes)? In this sense, it was immediately clear that a crucial role in differentiation was played by the microenvironment. Brüstle et al. (1999) were among the first to demonstrate in vitro controlled differentiation of hESCs using a series of growth factor combinations, which successfully elicited a reactivity to a monoclonal antibody specific for a membrane epitope typically found on the membranes of glial precursors. They initially grew ES cells in a media that favored the growth of neural precursors. They then exposed cells to the following series of growth factors: (1) basic fibroblast growth factor (FGF2), (2) FGF2 and epidermal growth factor (EGF), and (3) FGF2 and platelet-derived growth factor (PDGF) (Brüstle et al. 1999). The cells maintained in the final growth factor-supplemented media could be stored and kept in culture without further differentiation for many passages. However, as growth factors were removed, cells further differentiated into more specific neural cell types such as oligodendrocytes and astrocytes (Brüstle et al. 1999). Today, it is well established that ESC microenvironment
8 Cancer Stem Cells and Microenvironment
177
provides a dynamic regulatory niche for stem cell function constituting a finely regulated network of cell–cell communication that is critical to cell behavior and fate determination throughout embryogenesis. The stem cell niche can be defined as the production of local signal(s) and spatial organization of cells receptive to those signals to generate location-dependent control over contrasting cell-fate decisions such as self-renewal and differentiation. The concept of niche is also crucial in the regulation of ASCs as demonstrated by the pioneering studies on Drosophila (Deng and Lin 1997). Deng showed that the female fly has cells named “cap cells” at the tip of the ovary to which germ line stem cells adhere. During mitosis, daughter stem cells divided in such a way that they no longer adhered to the cap cells and started differentiating and growing into cystoblasts. On the contrary, daughter stem cells that stayed attached to the cap cells after cell division remained germ line stem cells. These results suggested that the cap cells provided a special environment for the germ line stem cells and played a role as a stem cell niche in the fly ovary. Today, human ASC niches have been characterized in several body regions; however, how they function as stem cell niches is far from being elucidated. In ESCs molecular cues in the embryonic milieu, originating from various cell types, possess various signaling characteristics with diverse effects on specific target cells or tissues. Certain stimulatory factors such as Activin, Nodal, and FGF maintain pluripotency in ESCs, while other factors, such as ngn3, Notch, and ASK-1 (endocrine, hematopoietic, and neuronal lineages, respectively), dictate a differentiated fate (Serafimidis et al. 2007; Radtke et al. 2004; Elmi et al. 2007; Vallier et al. 2005). These signals regulate the phenotype of stem cells or their progeny, are at times antagonistic and may be separated in a spatial and/or temporal fashion. These processes are extremely complex, and their sophistication is far to be fully understood (Constantinescu 2000; Lensch et al. 2006). It has to be underlined that the stem niche environment is not only determined by soluble endocrine or paracrine growth factors through cell–cell cross-talk but also on the contrary is regulated by a more complex interplay of different factors such as cell–matrix adhesion and the physical and physical–chemical environment. Peerani et al. (2007) in a series of sophisticated experiments demonstrated that the modulation of ESCs’ density per se can be used to generate distinct patterns of selfrenewal and differentiation. It has been demonstrated that in the osteoblastic niche, hematopoietic stem cells are predominantly located at the lowest end of an oxygen gradient in the bone marrow (Parmar et al. 2007); therefore, it is reasonable to think that physiochemical properties of the niche (pH, oxygen abundance, and size) are important factors regulating stem cell fate. Moreover, an increasing number of experimental evidence suggests that physical forces acting on stem cells through extracellular matrix or physical–chemical parameters, such as hypoxia (pH), can commit their fate. For example, the shear forces exerted by the blood flow have been found to direct isolated ESCs through an endothelial differentiation (Yamamoto et al. 2005) and slight mechanical deformations can induce mesenchymal stem cells (MSCs) to differentiate into smooth muscle cells (Kurpinski et al. 2006). A wide range of in vitro studies confirm the relevance of the so-called mechanobiology to
178
M. Federico and A. Giordano
stem cell differentiation. Undifferentiated MSCs grown on polymer gels mimicking the extracellular matrix elasticity of a given tissue were found to express precursor proteins for the cell type typically present in that tissue in the absence of specific growth factors. Cells grown on gels of elasticity comparable to brain express neuronal markers and their phenotypes and are not contractile. Conversely, MSCs cultured on more rigid gels (mimicking muscular tension) begin to express early skeletal muscle proteins. Finally, more rigid substrates that mimic pre-calcified bone yield MSCs that resemble bone cells and exert significant traction on their environment (Reilly and Engler 2009). Recently, to highlight the contribution of extracellular matrix to stem cell behavior, the physical parameters of the microenvironment have been proposed to act also as homing signals to migrating stem cells. As seen from the use of various gel matrices with controlled elasticity and non-limiting ligand density, most, if not all, cells are found to adhere and to anchor more strongly to stiff substrates compared with soft substrates. In a gradient of elasticity, cells therefore accumulate on stiffer substrates in a process called durotaxis, which might constitute a biophysical basis for which MSCs home to sites of injury, scar formation, and fibrosis (Disher et al. 2009).
CSCs and the Microenvironment Some of the general mechanisms, mentioned above, that regulate stem cell selfrenewal and differentiation are also crucial in CSC biology. The analogies are so many that maybe cancer could be considered a disease of unregulated self-renewal, in which mutations convert normal stem cell self-renewal pathways into the promoters of neoplastic proliferation (Pardal et al. 2003; Al-Hajj and Clarke 2004; Beachy et al. 2004). The Bmi-1 (B-cell specific Moloney murineleukemia virus integration site 1), Notch, Wnt, and Sonic hedgehog pathways (Park et al. 2003; Lessard and Sauvageau 2003; Duncan et al. 2005; Bhardwaj et al. 2001/2002) were initially identified to play a role in carcinogenesis and were later related to stem cell self-renewal (Taipale and Beachy 2001; Weng and Aster 2004). Studies on these pathways have revealed close links between cancer cells and normal ASCs (Pardal et al. 2003; Reya et al. 2001; Reya and Clevers 2005). Uncontrolled activation of these pathways may result in specific cancers, possibly as an attempt to recapitulate normal embryonic organogenesis (Molofsky et al. 2004; Ruiz i Altaba et al. 2004). Aberrant Notch signaling has been detected in several cancers and has recently been strongly connected to T-cell acute lymphoblastic leukemia (Grabher et al. 2006). Aberrant activation of the Wnt pathway has been found in various human tumors and is strongly associated with colorectal cancer (Taipale and Beachy 2001). Increased Hedgehog signaling has been linked not only to a small subset of tumors in the brain, skin, and muscle but also recently to cancers in the lungs, gastrointestinal tract, and pancreas (Pasca di Magliano and Hebrok 2003). Bmi-1 is an oncogene that is found overexpressed in several human cancers (Valk-Lingbeek et al. 2004) as for instance in the majority
8 Cancer Stem Cells and Microenvironment
179
of medulloblastomas (Leung et al. 2004; Marino 2005). As for ESCs and ASCs, also for CSCs the spatial definition of these signaling pathways in a particular microenvironment is crucial and the concept of CSC niche is rapidly emerging. A study of stem cells in brain tumors has made significant progress in this area. Calabrese and colleagues showed that brain tumor cells co-expressing Nestin and CD133 (the fraction believed to contain the CSCs) were found near the capillaries in the brain tumor (Calabrese et al. 2007). When these cells were co-cultured, the CSCs selectively adhered to the endothelial cells, similar to the cap cells in drosophila. This suggested that endothelial cells secrete factors necessary to maintain brain tumor CSCs. Moreover, it has been shown that the CD133-positive cells derived from human medulloblastoma together with endothelial cells developed brain tumors when xenografted into the brain of a recipient nude mouse. Comparably, Krause and colleagues (2006) showed that there was impaired induction of chronic myeloid leukemia (CML)-like myeloproliferative disease among recipient mice that received transplanted BCR–ABL1transduced CML progenitors from CD44-mutant donors. These results indicate that cell–cell interactions are crucial in providing a homing signal to CSCs; however, they are also able to interfere with engraftment in their niche within the recipient organism. Tumor hypoxia is a well-established factor of aggressiveness that since the 1980s and recent studies on ASC and CSC microenvironments offer a molecular explanation of the former clinical observation (Overgaard 1989). In fact, as for some ASC populations, hypoxia is thought to play a critical role also in CSC regulation. Soeda colleagues (2009) clearly demonstrated that human-derived CSCs when cultured in hypoxic conditions are promped toward self-renewal and strongly retain the undifferentiated phenotype, whereas CSCs cultured in normoxia do not. Moreover, it has been recently reported (Li et al. 2009) that physiological oxygen levels differentially induce hypoxia inducible factor-2a (HIF2a) levels in CSCs. HIF1a is known to function in proliferation and survival of all cancer cells, while HIF2 seems to be essential only in CSCs and was not expressed by normal neural progenitors. Heddleston et al. (2009) extended these studies to examine the role of hypoxia in regulating tumor cell plasticity and found that hypoxia promotes the self-renewal capability of the stem and non-stem populations. Furthermore, hypoxia induces a more stem-like phenotype in the non-stem population with a marked increased in neurosphere formation as well as upregulation of important stem cell factors, such as OCT4, NANOG, and c-MYC. The importance of HIF2a was further supported as forced expression of non-degradable HIF2a-induced expression of a CSC marker and augmented the tumorigenic potential of the non-stem population. It has to be underlined that the influence of the microenvironment on the tumor progenitor is eventually even greater. Several groups have reported the ability of the microenvironment to be able to dramatically reprogram the cancer phenotype. In the early seventies, Mintz and Illmensee reported that teratocarcinoma cells injected into murine blastocysts aquired totipotency and a reversion to a normal phenotype (Mintz and Illmensee 1975; Illmensee and Mintz 1976). This model clearly established a non-mutational basis for the transformation of these teratocarcinoma cells
180
M. Federico and A. Giordano
into a malignant phenotype, and demonstrated the unique ability of an embryonic microenvironment to reprogram these cells to a normal phenotype. In more recent years, other researchers (Lee et al. 2005; Kulesa et al. 2006) demonstrated that the chick and zebrafish embryonic microenvironments not only repress tumor formation when implanted with aggressive human melanoma cells but induce lineagespecific genes and migratory behaviors associated with a more normal phenotypes, representative of an epigenetic reprogramming event. The same group with an elegant in vitro model showed that apart from any cell–cell interaction, the extracellular matrix when conditioned by specific cell types can clearly influence the phenotype of aggressive cancers (melanoma) in a reversible (epigenetic) way (Abbott et al. 2008). To further demonstrate a specific influence of tumor microenvironment on stem cells, Taylor et al. (2008) combined human ES cells with cancer-associated fibroblasts (CAFs) from patients with prostate cancer. When grafted into the kidney capsule of host mice, human ES cells form teratomas that rapidly enlarge and kill the host. When combined with CAFs, the teratoma formation is significantly abrogated, demonstrating that the stromal tumor niche is able to influence human ES cell differentiation. Furthermore, the CAF-derived niche differs from that derived from nonmalignant prostatic fibroblasts that were unable to alter or impede teratoma formation. Further data in support of the relevance of the environmental influence on CSC plasticity comes from Knutson et al. (2006; Santisteban et al. 2009), who demonstrated that in vivo the induction of an immune response against an epithelial breast cancer led to the T-cell-dependent outgrowth of a tumor, the cells of which had undergone epithelial to mesenchymal transition (EMT). The resulting mesenchymal tumor cells had a CD24/low/CD44+ phenotype, consistent with breast CSCs. EMT induced by CD8 T cells produced tumors that had characteristics of breast CSCs, including potent tumorigenicity, ability to re-establish an epithelial tumor, and enhanced resistance to drugs and radiation. These data in contrast to the hierarchical CSC hypothesis suggest that breast cancer arises from the transformation of a resident tissue stem cell and that EMT can produce the breast CSC phenotype. These findings have two important implications: first, the observed phenotypic plasticity suggests that a dynamic equilibrium may exist between CSCs and nonCSCs within tumors; second, this equilibrium may be shifted in one direction or another by contextual signals within the tumor microenvironment influencing the probability for there to be interchanged between the CSC and non-CSC compartments and thus regulating the overall tumor homeostasis.
Concluding Remarks CSCs represent a budding field of research that has yet to emerge on its own. We are still completely reliant on our understanding of embryonic and ASC biology to explain the role of stem cells in cancer. However, it is clear from the initial studies that CSCs play a critical role in tumorigenisis and resistance to cancer therapies;
8 Cancer Stem Cells and Microenvironment
181
therefore, our continued efforts are essential in elucidating the biological role of these tumor sub-populations. Here we have given an overview of what is known so far in CSC biology, from the initial documentations of CSCs to the role of specific growth factors and transcriptional regulators in CSC biology. We have also discussed major findings in the ESC and ASC research fields, which we retain necessary for the proper physiological understanding of CSCs. It is important to note that these correlations are assumed because of the similarities between CSCs and ESCs or ASCs, respectively. It will be imparitive in the future to better characterize CSCs on their own to understand their origin and their unique role in cancer biology.
References Abbott DE, Bailey CM, Postovit LM et al. (2008) The epigenetic influence of tumor and embryonic microenvironments: how different they are? Cancer Microenvironment 1:13–21. Al-Hajj M, Clarke MF (2004) Self-renewal and solid tumor stem cells. Oncogene 23:7274–7282. Allsopp RC, Vaziri H, Patterson C, Goldstein S, Younglai EV, Futcher B, Greider CW, Harley CB (1992) Telomere length predicts replicative capacity of human fibroblasts. Proc Natl Acad Sci U S A 89:10114–10118. Avilion AA, Nicolis SK, Pevny LH, Perez L, Vivian N, Lovell-Badge R (2003) Multipotent cell lineages in early mouse development depend on SOX2 function. Genes Dev 17:126–140. Beachy PA, Karhadkar SS, Berman DM (2004) Tissue repair and stem cell renewal in carcinogenesis. Nature 432 :324–331, 639–648. Becker KA, Ghule PN, Therrien JA, Lian JB, Stein JL, van Wijnen AJ, Stein GS (2006) Selfrenewal of human embryonic stem cells is supported by a shortened G1 cell cycle phase. J Cell Physiol 209:883–893. Bhardwaj G, Murdoch B, Wu D et al. (2001/2002) Sonic hedgehog induces the proliferation of primitive human hematopoietic cells via BMP regulation. Nat Immunol 2:172–180, 542–545. Boman BM, Huang E (2008) Human colon cancer stem cells: a new paradigm in gastrointestinal oncology. J Clin Oncol June:2828–2838. Bonnet D, Dick J (1997) Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med 3:730–737. Boyer LA, Lee TI, Cole MF et al. (2005) Core transcriptional regulatory circuitry in human embryonic stem cells. Cell 122:947–956. Brüstle O, Jones KN, Learish RD, Karram K, Choudhary K, Wiestler OD, Duncan ID, McKay RDG (1999). Embryonic stem cell-derived glial precursors: a source of myelinating transplants. Science 285:754–756. Calabrese C, Poppleton H, Kocak M et al. (2007) A perivascular niche for brain tumor stem cells. Cancer Cell 11:69–82. Cheng LC, Tavazoie M, Doetsch F (2005) Stem cells: from epigenetics to microRNAs. Neuron 46:363–367. Constantinescu SN (2000) Stem cell generation and choice of fate: role of cytokines and cellular microenvironment. J Cell Mol Med 4(4):233–248. Deng W, Lin H (1997) Spectrosomes and fusomes anchor mitotic spindles during asymmetric germ cell divisions and facilitate the formation of a polarized microtubule array for oocyte specification in Drosophila. Dev Biol 189:79–94. Dirks PB (2008) Brain tumor stem cells: bringing order to the chaos of brain. J Clin Oncol June:2916–2924.
182
M. Federico and A. Giordano
Disher D, Mooney DJ, Zandstra PW (2009) Growth factors, matrices and forces combine and control stem cells. Science 324:1673–1677. Duncan AW, Rattis FM, Dimascio LN, Congdon KL et al. (2005) Integration of Notch and Wnt signaling in hematopoietic stem cell maintenance. Nat Immunol 6:314–322. Elmi M et al. (2007) Mechanism of MASH1 induction by ASK1 and ATRA in adult neural progenitors. Mol Cell Neurosci 36(2):248–259. Fluckinger AC, Marcy G, Marchand M, Négre D, Cosset FL, Mitalipov S, Wolf D, Savatier P, Dehay C (2006) Cell cycle features of primate embryonic stem cells. Stem Cells 24:547–556. Fraichard A, Chassande O, Bilbaut G, Dehay C, Savatier P, Samarut J (1995) In vitro differentiation of embryonic stem cells into glial cells and functional neurons. J Cell Sci 108:3181–3188. Fuchs E, Segre JA (2000) Stem cells: a new lease on life. Cell 100:143–155. Giacinti C, Giordano A (2006) RB and cell cycle progression. Oncogene 25:5220–5227. Grabher C, von Boehmer H, Look AT (2006) Notch 1 activation in the molecular pathogenesis of T-cell acute lymphoblastic leukaemia. Nat Rev Cancer 6:347–359. Greider CW, Blackburn EH (1987) The telomere terminal transferase of tetrahymena is a ribonucleoprotein enzyme with two kinds of primer specificity. Cell 51:887–898. Haas K, Johannes C, Geisen C, Schmidt T, Karsunky H, Blass-Kampmann S, Obe G, Möröy T (1997). Malignant transformation by cyclin E and Ha-Ras correlates with lower sensitivity towards induction of cell death but requires functional Myc and CDK4. Oncogene 15:2615–2623. Harley CB, Futcher AB, Greider CW (1990) Telomeres shorten during aging of human fibroblasts. Nature 345:458–460. Hayflick L (1965) The limited in vitro lifetime of human diploid cell strains. Exp Cell Res 37:614–636. Hayflick L, Moorhead PS (1961) The serial cultivation of human diploid cell strains. Exp Cell Res 25:585–621. Heddleston JM, Li Z, McLendon RE et al. (2009) The hypoxic microenvironment maintains glioblastoma stem cells and promotes reprogramming towards a cancer stem cell phenotype. Cell Cycle 8(20):3274–3284. Hogan BLM, Beddington RSB, Constantini F, Lacy E (1994) Manipulating the mouse embryo, 2nd ed. Cold Spring Harbor Laboratory Press, New York. Huff CA, Matsui W (2008) Multiple myeloma cancer stem cells. J Clin Oncol 2895–2900. Huangfu DW, Osafune K et al. (2008) Induction of pluripotent stem cells from primary human fibroblasts with only Oct4 and Sox2. Nat Biotechnol 26:1269–1275. Illmensee K, Mintz B (1976) Totipotency and normal differentiation of single teratocarcinoma cells cloned by injection into blastocysts. Proc Natl Acad Sci U S A 73(2,9):549–553, 1998–2006. Kakarala M, Wicha MS (2008) Implications of the cancer stem-cell hypothesis for breast cancer prevention. J Clin Oncol June:2813–2820. Kasten MM, Giordano A (1998) pRB and the Cdks in apoptosis and the cell cycle. Cell Death Differ 5:132–140. Knutson KL, Lu H, Stone B et al. (2006). Immunoediting of cancers may lead to epithelial to mesenchymal transition. J Immunol 177:1526–1533. Kouzarides T (2007) Chromatin modifications and their function. Cell 128:693–705. Krause DS, Lazarides K, von Andrian UH et al. (2006) Requirement for CD44 in homing and engraftment of BCR-ABL-expressing leukemic stem cells. Nat Med 12:1175–1180. Kulesa PM, Kasemeier-Kulesa JC, Teddy JM et al. (2006) Reprogramming metastatic melanoma cells to assume a neural crest cell-like phenotype in an embryonic microenvironment. Proc Natl Acad Sci U S A 103(10):3752–3757. Kurpinski K, Chu J, Hashi C et al. (2006) Anisotropic mechanosensing by mesenchymal stem cells. Proc Natl Acad Sci U S A 103:16095. Lee CJ, Dosch J, Simeone DM (2008) Pancreatic cancer stem cells. J Clin Oncol June:2806–2812.
8 Cancer Stem Cells and Microenvironment
183
Lee JH, Hart SR, Skalnik DG (2004) Histone deacetylase activity is required for embryonic stem cell differentiation. Genesis 38:32–38. Lee LM, Seftor EA, Bonde G et al. (2005) The fate of human malignant melanoma cells transplanted into zebrafish embryos: assessment of migration and cell division in the absence of tumor formation. Dev Dyn 233(4):1560–1570. Lensch MW, Daheron L, Schlaeger TM (2006) Pluripotent stem cells and their niches. Stem Cell Rev 2(3):185–201, 773–789. Lessard J, Sauvageau G (2003) Bmi-1 determines the proliferative capacity of normal and leukaemic stem cells. Nature 423:255–260. Leung C, Lingbeek M, Shakhova O (2004). Bmi1 is essential for cerebellar development and is overexpressed in human medulloblastomas. Nature 428:337–341. Li X, Lewis MT, Huang J et al. (2008) Intrinsic resistance of tumorigenic breast cancer cells to chemotherapy. J Natl Cancer Inst 100:672–679. Li Z, Bao S, Wu Q et al. (2009) Hypoxia inducible factors regulate tumorigenic capacity of glioma stem cells. Cancer Cell 15:501–513. Lunyak VV, Rosenfeld MG (2008) Epigenetic regulation of stem cell fate. Hum Mol Genet 17:28–36. Maitland NJ, Collins AT (2008) Prostate cancer stem cells: a new target for therapy. J Clin Oncol June:2862–2870. Marino S (2005) Medulloblastoma: developmental mechanisms out of control. Trends Mol Med 11:17–22. Martin GR, Evans MJ (1974) The morphology and growth of a pluripotent teratocarcinoma cell line and its derivatives in tissue culture. Cell 2:163–172. Martin GR, Evans MJ (1975) Differentiation of clonal lines of teratocarcinoma cells: formation of embryoid bodies in vitro. Proc Natl Acad Sci U S A 72:1441–1445. Meshorer E (2007) Chromatin in embryonic stem cell neuronal differentiation. Histol Histopathol 22:311–319. Meshorer E, Yellajoshula D, George E, Scambler PJ, Brown DT, Misteli T (2006) Hyperdynamic plasticity of chromatin proteins in pluripotent embryonic stem cells. Dev Cell 10: 105–116. Mintz B, Illmensee K (1975) Normal genetically mosaic mice produced from malignant teratocarcinoma cells. Proc Natl Acad Sci U S A 72(9):3585–3589. Misteli T (2004) Spatial positioning: a new dimension in genome function. Cell 119:153–156. Misteli T (2007) Beyond the sequence: cellular organization of genome function. Cell 128:787–800. Molofsky AV, Pardal R, Morrison SJ (2004) Diverse mechanisms regulate stem cell self-renewal. Curr Opin Cell Biol 16:700–707. Mumberg D, Haas K, Möröy T, Niedenthal R, Hegemann JH, Funk M, Müller R (1996) Uncoupling of DNA replication and cell cycle progression by human cyclin E. Oncogene 13:2493–2497. Nichols J, Zevnik B, Anastassiadis K, Niwa H, Klewe-Nebenius D, Chambers I, Schöler H, Smith A (1998) Formation of pluripotent stem cells in the mammalian embryo depends on the POU transcription factor Oct4. Cell 95:379–391. Niwa H, Burdon T, Chambers I, Smith A (1998) Self-renewal of pluripotent embryonic stem cells is mediated via activation of STAT3. Genes Dev 12:2048–2060. Overgaard J (1989) Sensitization of hypoxic tumour cells; clinical experience. Int J Radiat Biol 56(5):801–811. Pardal R, Clarke MF, Morrison SJ (2003) Applying the principles of stem-cell biology to cancer. Nat Rev Cancer 3:895–902. Park IK, Qian D, Kiel M et al. (2003) Bmi-1 is required for maintenance of adult self-renewing haematopoietic stem cells. Nature 423:302–305. Parmar K, Mauch P, Vergilio JA et al. (2007) Distribution of hematopoietic stem cells in the bone marrow according to regional hypoxia. PNAS 104(13):5431–5436.
184
M. Federico and A. Giordano
Pasca di Magliano M, Hebrok M (2003) Hedgehog signalling in cancer formation and maintenance. Nat Rev Cancer 3:903–911. Peacock CD, Watkins DN (2008) Cancer stem cells and the ontogeny of lung cancer. J Clin Oncol June:2883–2889. Peerani R, Balaji MR, Bauwens C et al. (2007) Niche-mediated control of human embryonic stem cell self-renewal and differentiation. EMBO J 26:4744–4755. Power MA, Tam PP (1993) Onset of gastrulation, morphogenesis and somitogenesis in mouse embryos displaying compensatory growth. Anat Embryol (Berl) 187:493–504. Prince MEP, Ailles LE (2008) Cancer stem cells in head and neck squamous cell cancer. J Clin Oncol June:2871–2875. Radtke F et al. (2004) Notch regulation of lymphocyte development and function. Nat Immunol 5(3):247–253. Rathjen J, Lake JA, Bettess MD, Washington JM, Chapman G, Rathjen PD (1999) Formation of a primitive ectoderm like cell population, EPL cells, from ES cells in response to biologically derived factors. J Cell Sci 112:601–612. Reilly GC, Engler AJ (2010) Intrinsic extracellular matrix properties regulate stem cell differentiation. J Biomechanics. 43:55–62. Reya T, Clevers H (2005) Wnt signalling in stem cells and cancer. Nature 434:843–850. Reya T, Morrison SJ, Clarke MF et al. (2001) Stem cells, cancer and cancer stem cells. Nature 1; 414 (6859):105–11. Risbridger GP, Taylor RA (2008) Minireview: regulation of prostatic stem cells by stromal niche in health and disease. Endocrinology 149(9):4303–4306. Ruiz i Altaba A, Stecca B, Sanchez P (2004) Hedgehog–Gli signaling in brain tumors: stem cells and paradevelopmental programs in cancer. Cancer Lett 204:145–157. Santisteban M, Reiman MJ, Asiedu MK et al. (2009) Immune-induced epithelial to mesenchymal transition in vivo generates breast cancer stem cells. Cancer Res 69(7):2887–2895. Savatier P, Huang S, Szekely L, Wiman KG, Smaraut J (1994) Contrasting patterns of retinoblastoma protein expression in mouse embryonic stem cells and embryonic fibroblasts. Oncogene 9:809–818. Schopperle WM, DeWolf WC (2007) The TRA-1–60 and TRA-1–81 human pluripotent stem cell markers are expressed on podocalyxin in embryonal carcinoma. Stem Cells 25:723–730. Sell S, Leffert HL (2008) Liver cancer stem cells. J Clin Oncol June:2800–2805. Serafimidis I et al. (2007) Novel effectors of directed and ngn3 mediated differentiation of mouse embryonic stem cells into endocrine pancreas progenitors. Stem Cells 26:3–16. Shamblott MJ, Axelman J, Wang S, Bugg EM, Littlefield JW, Donovan PJ, Blumenthal PD, Huggins GR, Gearhart JD (1998). Derivation of pluripotent stem cells from cultured human primordial germ cells. Proc Natl Acad Sci U S A 95:13726–13731. Shukla V, Vaissiere T, Herceg Z (2008) Histone acetylation and chromatin signature in stem cell identity and cancer. Mutat Res 637:1–15. Soeda A, Park M, Lee D et al. (2009) Hypoxia promotes expansion of the CD133-positive glioma stem cells through activation of HIF1a. Oncogene. 28:3949–3459. Solter D, Skreb N, Damjanov I (1971) Cell cycle analysis in the mouse EGG-cylinder. Exp Cell Res 64:331–334. Stead E, White J, Faast R, Conn S, Goldstone S, Rathjen J, Dhingra U, Rathjen P, Walker D, Dalton S (2002) Pluripotent cell division cycles are driven by ectopic Cdk2, cyclin A/E and E2F activities. Oncogene 21:8320–8333. Taipale J, Beachy PA (2001) The Hedgehog and Wnt signalling pathways in cancer. Nature 411:349–354. Takaishi S, Okomura T, Wang TC (2008) Gastric cancer stem cells. J Clin Oncol June:2876–2882. Thomson JA, Itskovitz-Eldor J, Shapiro SS, Waknitz MA, Swiergiel JJ, Marshall VS, Jones JM (1998) Embryonic stem cell lines derived from human blastocysts. Science 282:1145–1147. Valk-Lingbeek ME, Bruggeman SW, van Lohuizen M (2004) Stem cells and cancer; the polycomb connection. Cell 118:409–418.
8 Cancer Stem Cells and Microenvironment
185
Vallier L, Alexander M, Pedersen RA (2005) Activin/Nodal and FGF pathways cooperate to maintain pluripotency of human embryonic stem cells. J Cell Sci 118(Pt 19)(17):4495–4509, 6322–6327. Weng AP, Aster JC (2004) Multiple niches for Notch in cancer: context is everything. Curr Opin Genet Dev 14:48–54. Wu H, Sun YE (2006) Epigenetic regulation of stem cell differentiation. Pediatr Res 59:21–25. Wobus AM, Boheler KR (2005) Embryonic stem cells: Prospects for developmental biology and cell therapy. Physiol Rev 85:635–678. Yamamoto K, Sokabe T, Watanabe T et al. (2005) Fluid shear stress induces differentiation of Flk-1-positive embryonic stem cells into vascular endothelial cells in vitro. Am J Physiol Heart Circ Physiol 288:H1915.
Chapter 9
Epithelial–Mesenchymal Transition in Development and Diseases Yadi Wu and Binhua P. Zhou
Abstract Epithelial–mesenchymal transition (EMT) is critical for appropriate embryo implantation, embryogenesis, wound healing, tissue regeneration, and organ development. During this EMT process, epithelial cells lose the adherent and tight junctions. They gain a mesenchymal cell phenotype that enables them to invade and migrate over long distances and resist apoptosis. A similar process is also detected in tumor metastasis, suggesting that the tumor cells hijack this developmental pathway for tumor progression. Here, we review three different types of EMT in physiological and pathological conditions with special focus on the new development on type 3 EMT in metastasis. We summarize the recent new findings on tumor microenvironment, signaling pathways, and mechanisms underlying the regulation of EMT at metastasis. Understanding the biology of EMT will open new avenues for controlling fibrosis and cancer progression.
Overview of EMT Epithelial–mesenchymal transition (EMT) is a phenotypic conversion during embryonic development when tissue remodeling and cell migration shape the future organism, such as those in embryonic development and wound healing. In pathological condition, a similar EMT process is also observed in fibrotic diseases and metastasis. During EMT, epithelial cells lose the adherent and tight junctions that keep them in contact with their neighbors; they gain a mesenchymal cell phenotype enabling them to break through the basal membrane and migrate over a long distance owing to profound changes in their cytoskeleton
B.P. Zhou (*) Markey Cancer Center, University of Kentucky School of Medicine, Lexington, KY, USA and Department of Molecular and Cellular Biochemistry, University of Kentucky School of Medicine, Lexington, KY, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_9, © Springer Science+Business Media, LLC 2010
187
188
Y. Wu and B.P. Zhou Type 1: formations of Mesoderm, Neural Crest, Cardiac Valve and Secondary palate
EMT types
Type 2: Fibrosis and re-epithelialization of wounded skin Type 3: tumor metastasis
Fig. 9.1 EMT is classified into three different subtypes (please see text for more detail)
architecture and gene expression profile (Kalluri and Neilson 2003). This concept was pioneered by the seminal study from Elizabeth Hay using chick primitive streak formation as a model in 1967 (Hay 1995). Hay realized that an epithelial phenotypic conversion was of crucial importance during gastrulation and cell migration in the early vertebrate embryo. She proposed that differentiated epithelial cells could undergo dramatic “transformation” into mesenchymal cells (Hay 1995; Greenburg and Hay 1998). However, since this “transformation” is reversible and mesenchymal cells can revert back to epithelial cells through a reverse process of mesenchymal–epithelial transition (MET), the term of “transition” is now used. The field of EMT has vastly expanded recently in both scope and understanding. EMT was divided into three subtypes based on the biological context under which they occur (Kalluri 2009; Kalluri and Weinberg 2009; Zeisberg and Neilson 2009) (Fig. 9.1). Type 1 EMT describes the transition events of epithelial cells to become motile mesenchymal cells during implantation, embryo formation, gastrulation, and neural crest migration. These primary mesenchymal cells can revert back and form secondary epithelia in mesodermal and endodermal organs through MET. Type 2 EMT involves the program of generating tissue fibroblasts from epithelial or endothelial cells during injury and chronic inflammation. This type of EMT will eventually lead to tissue or organ fibrosis and organ malfunction if inflammation persists. Type 3 EMT describes how tumor cells at the invasive front of primary tumors undergo a phenotypic transitioning in order to invade and metastasize through the blood stream or lymph node system and form a metastatic lesion through MET at distant tissues or organs. Although these three subtypes of EMT represent distinct biological processes, the genetic element and mechanism of regulation may be similar and well conserved. Studies of EMT often involve various model systems ranging from different epithelial cell types to assorted stimulations, it is important to use validated biomarkers to examine the phenotypic conversion during all three classes of EMT. Common biomarkers include cell-surface and extracellular molecules, cytoskeletal proteins, and specific transcription factors (Table 9.1). For example, downregulation of E-cadherin is a hallmark of EMT and loss of E-cadherin expression facilitates the induction of EMT (Huber et al. 2005). E-cadherin is a cell–cell adhesion molecule that participates in homotypic, calcium-dependent interactions to form epithelial adherent junctions (Cowin et al. 2005; Junghans et al. 2005).
9 Epithelial–Mesenchymal Transition in Development and Diseases
189
Table 9.1 EMT markers Cell-surface proteins
Cytoskeketal molecules
ECM proteins
Transcription factors (common)
Acquired markers N-cadherin ob-cadherin a5b1 integrin aVb6 integrin Syndecan-1 FSP1 a-SMA Vimentin b-catenin a1(I) collagen a1(III) collagen Fibronectin Laminin 5 Snail Slug Twist SIP1/ZEB2
Attenuated markers E-cadherin ZO-1 Claudins Occludin Desmoplakin Cytokeratin Plakophilin Crumbs3 HUGHL2 a1(IV) collagen Laminin 1
The cytoplasmic tail of E-cadherin links to cytoskeleton through a- and b-catenin to modulate cell–cell adhesion and cell migration. Cadherin switching from E-cadherin to N-cadherin is commonly found in EMT. N-cadherin is expressed in mesenchymal cells, fibroblasts, cancer cells, and neural tissue, and N-cadherin levels in these cells are closely associated with their invasiveness, motility, and metastasis. In addition, E-cadherin repressors, such as Snail, Slug, Twist, and ZEB1/2, are commonly used as EMT markers. Snail is the first described E-cadherin repressor and is also the common downstream target of various signaling pathways leading to EMT. Vimentin, an intermediate filament that is mainly expressed in fibroblasts, endothelial cells, and hematopoietic cells, is also commonly used as an indicator for type 3 EMT, as expression of vimentin in tumor cells correlates with their invasiveness and metastatic potential. Furthermore, differential expression of integrin is also used as a biomarker of EMT, as integrins modulate the interaction of cells with extracellular matrix (ECM). For example, increased expression of a5 integrin is commonly found in type 2 and type 3 EMT (Qian et al. 2005; Davidson et al. 2006; White et al. 2007).
Type 1 EMT in the Formation of Mesoderm and Neural Crest EMT is crucially important to the tissue morphogenetic events during embryonic development, such as in the mesoderm formation, neural crest formation, heart valve development, and secondary palate formation. In the absence of EMT,
190
Y. Wu and B.P. Zhou
development cannot proceed through the blastula stage. Mesoderm formation and neural crest development represent the major EMT programs that occur during early embryonic development; the resulting mesenchymal and neural crest cells act as progenitors and further differentiate into various cell types via MET program. For example, gastrulation EMT produces the mesoderm, giving rise to muscle, bone, and connective tissues, whereas neural crest delamination EMT gives rise to glial and neuronal cells, adrenal glandular tissues, pigment-containing cells of the epidermis, and skeletal and connective tissues. The heart valve development and secondary palate formation occur in relatively well-differentiated epithelial cells that are destined to become defined mesenchymal cells types.
Mesoderm Formation The classic example of EMT is the formation of mesoderm from the primitive ectoderm during gastrulation. Gastrulation is observed in all metazoans and is accompanied by drastic morphogenic changes from a single epithelial layer (the epiblast) into three embryonic germ layers, the ectoderm, mesoderm, and endoderm, to form a complex three-dimensional multilayered embryo (Shook and Keller 2003). In chicken and mouse embryo, Wnt and TGFb signaling provide the initial induction signals for EMT while the FGF signal is necessary to maintain the EMT regulatory network during mesoderm formation. All these signaling events activate the expression of Snail, which represses the expression of E-cadherin and other tight junction components (such as claudins, occludins, and Crumbs) and promotes cell migration. In Snail knock-out mice, although mesoderm specification is not affected, the cells that form are unable to migrate. Gastrulation in Drosophila melanogaster initiates at the completion of the celluarization step in a stripe of ventral cells. The newly formed mesenchymal cells migrate dorsally along the ectoderm before differentiating into visceral and somatic mesoderm. Snail and Twist are expressed in the presumptive mesoderm and in the invaginating cells. These two molecules are under the control of Dorsal (p65 subunit of nuclear factor-kB), which is involved in the establishment of the dorso-ventral axis. Snail represses the transcription of Shotgun (the orthologue of vertebrate E-cadherin) in mesodermal precursor cells that undergo invagination at the ventral furrow (Oda et al. 1998). More importantly, Snail and Twist are involved in the control of rapid and orderly transition from the ventral epithelium into a mesenchymal state (Dawes-Hoang et al. 2005; Kolsch et al. 2007). Therefore, Dorsal, Twist and Snail act in concert to induce EMT and bring the internalization and formation of the Drosophila mesoderm. The zebrafish (Dabio rerio) organizer is the site for EMT of axial mesendoderm cells. Liv1, a seven-spanning transmembrane receptor that is involved in zinc transport is a downstream of signal transducer and activator of transcription 3 (STAT3) (Yamashita et al. 2004). STAT3 is essential for gastrulation movements in the organizer during zebrafish gastrulation. Interestingly, LIV1 is essential for the
9 Epithelial–Mesenchymal Transition in Development and Diseases
191
nuclear localization and repressive activity of Snail, which inhibits E-cadherin expression and promotes the up-regulation of N-Cadherin.
Neural Crest Formation Neural crest formation is another example of Type I EMT in embryogenesis where premigratory neural crest cells form at the border of the neural plate and non-neural ectoderm as a result of signals emanating from these two tissues. Interestingly, similar signaling pathways that operate during EMT at gastrulation are utilized in the neural crest formation. Indeed, a combination of Wnt, FGF, and TGFb (mainly BMP) induce the expression of Snail, Sox, and forkhead box D3 transcription factors (Villanueva et al. 2002). In addition, experimental evidence shows that Notch signaling plays an important role in neural crest formation through induction of Slug in frog and chick embryo (Nieto 2002). The combination of these transcription factors generates the full spectrum of phenotypic changes associated with EMT and primes the precursor cells to become migratory neural crest cells. These neural crest cells are equipped with the ability to migrate over extraordinarily long distances in the embryo prior to their reaggregation via MET for further differentiation.
MET EMT is a reversible process during the embryogenesis, tissue construction and tumor progression, and metastasis. The converted mesenchymal cells can revert to an epithelial cell state by passing through MET. This suggests that inter-conversion between epithelial and mesenchymal cell states may also occur under certain pathological conditions and be affected by signals from microenvironment. In fact, following metastatic spread and colonization, an MET often converts the disseminated mesenchymal cancer cells back to a more differentiated epithelial cell state to form secondary metastatic lesions.
Type 2 EMT in Tissue and Organ Fibrosis Implications of EMT in Fibrosis Re-epithelization, tissue regeneration, and organ fibrosis constitute type 2 EMT. Organ fibrosis is mediated by inflammatory cells and fibroblasts which deposit collagens, elastin, tenacin, and other matrix molecules. Fibrosis associated type 2 EMT specifically occurs in kidney, liver, lung and intestine (Zeisberg et al. 2007).
192
Y. Wu and B.P. Zhou
A series of typical experiments has shown that EMT is an important process during tissue injury leading to organ fibrosis. In terms of the EMT proteomes, fibroblastspecific protein 1 (FSP1, also known as S100A4 and MTS-1), a-SMA (smooth muscle actin), and collagen I are reliable markers to characterize the mesenchymal products generated by the EMTs in the development of fibrosis in various organs. TGFb1, as the major pro-fibrotic cytokine, induces many of the central processes involved in fibrosis, including fibroblast to myfibroblasat differentiation, ECM deposition and EMT. TGFb not only contributes to pulmonary and hepatic fibrosis but also plays a key role in cardiac fibrosis (Gressner et al. 2002; Willis and Borok 2007; Zeisberg et al. 2007). TGFb induces EMT via both a Smad2/3-dependent pathway and a MAPK-dependent pathway. The relevance of TGFb-induced EMT for progression of organ fibrosis was recently further elucidated using BMP-7 as an intracellular competitor of TGFb signaling in mouse models of kidney, liver, billiard tract, lung, and intestinal fibrosis (Zeisberg et al. 2003a, b). The function of TGFb in fibrosis is highlighted by the finding that Smad3–/– mice are resistant to the induction of several fibrotic diseases (Flanders 2004). TGFb levels are also overproduced and are associated with functional impairment in patients with fibrotic pulmonary diseases such as idiopathic pulmonary fibrosis (Salez et al. 1998). Clinical studies also demonstrated the correlation between fibrosis and EMT (Rastaldi et al. 2002). Using immunohistochemistry and in situ hybridization, an EMT was demonstrated with the expression of several markers of tubular phenotype transition such as cytokeratin, vimentin, a-SMA, and zona occludens (ZO-1) in 133 kidney biopsies (Rastaldi et al. 2002). Similarly, in patients with Crohn disease, an expression pattern of EMT was found in areas of fibrosis in the colon (Bataille et al. 2008).
Re-epithelialization of Wounded Skin Re-epithelialization recapitulates several aspects of EMT. Re-epithelialization requires epithelial cells at the edge of the wounded tissue to loosen their cell–cell and cell–ECM contacts and assume a migratory phenotype which are reminiscent of EMT. Slug has a crucial role in wound-healing which is expressed in keratinocytes at the boundary of wounds. Importantly, epithelial cell outgrowth from skin explants was markedly reduced in Slug knock-out mice, whereas overexpression of Slug in cultured human keratinocytes results in increased cell spreading and desmosomal disruption (Savagner et al. 2005). Arnoux et al. further found that EGF can activate Erk5 which specifically enhances Slug promoter activity and controls wound healing in keratinocyte-derived HacaT cells in vitro (Arnoux et al. 2008). However, it should be noted that not all features of EMT are seen. First, the migrating keratinocytes remain part of a cohesive cell sheet since they retain some intercellular junction. Second, the epithelial cells do not actually become mesenchymal, i.e., interstitial cells. They remain epithelial in characteristics. Once wound closure is completed, the involved epithelial cells revert to their tissue-specific differentiated state.
9 Epithelial–Mesenchymal Transition in Development and Diseases
193
Type 3 EMT in Cancer Metastasis Cancer metastasis is believed to consist of four distinct steps: invasion, intravasation, extravasation, and metastatic colonization (Chambers et al. 2002; Pantel and Brakenhoff 2004). During invasion, tumor cells lose cell–cell adhesion, gain mobility, and leave the site of the primary tumor to invade adjacent tissues. In intravasation, tumor cells penetrate through the endothelial barrier and enter the systemic circulation through blood and lymphatic vessels. In extravasation, cells that survive the anchorage-independent growth conditions in the bloodstream attach to vessels at distant sites and leave the bloodstream. Finally, in metastatic colonization, tumor cells form macrometastases in the new host environment (Chambers et al. 2002; Pantel and Brakenhoff 2004). All of these steps, from initial breakdown of tissue structure, through increased invasiveness, and ultimately distribution and colonization throughout the body, are characteristics of the developmental process at EMT/ MET. The similarity of genetic controls and biochemical mechanisms underlying the acquisition of the invasive phenotype and the subsequent systemic spread of the cancer cells highlights that tumor cells usurp this developmental pathway for their metastatic dissemination. We will discuss in detail on the environmental stimuli, the signaling pathways, and the mechanisms of regulation that control type 3 EMT in metastasis.
EMT Stimuli from Tumor Microenvironment EMT is a dynamic process that is controlled by signals that cells receive from their microenvironment. Consistent with this notion, EMT commonly occurs at the invasive front (tumor–stromal boundary) of many invasive carcinomas (Franci and Christofori 2006). These observations indicate that EMT is triggered by cellular signals from the tumor microenvironment. The tumor microenvironment as a complex system consisting of extracellular matrix (ECM) (such as collagen and hyaluronic acid), stromal cells (such as fibroblasts of various phenotypes/myofibroblasts, endothelial cells, pericytes and smooth-muscle cells), and infiltrated inflammatory cells (such as neutrophils, myeloid-derived suppressor cells (MDSC), lymphocytes, and macrophages). These infiltrated immune cells secrete cytokines, chemokines, and growth factors, such as TNFa, TGFb, IL-6, FGF, epidermal growth factor (EGF), and HGF (Joyce and Pollard 2009) and play an essential role for supporting tumor progression and metastasis. It is conceivable that the migratory and invasive ability of tumor cells at the invasive front is initiated and propelled by an inflammatory microenvironment through the induction of EMT. Consistent with this notion, a high content of inflammatory cells, particularly tumor-associated macrophages, is commonly found at the invasive fronts of advanced carcinoma (Condeelis and Pollard 2006). Macrophages are key cells in chronic inflammation. Tumor-associated macrophages (TAMs) are important
194
Y. Wu and B.P. Zhou
component of infiltrated leukocytes in most malignant tumors. They modulate inflammation and adaptive immunity and promote cell proliferation by producing growth factors and by enhancing angiogenesis, tissue remodeling, and repair. Macrophage directly influences the behavior and function of tumor cells and has been regarded as an “obligate partner for tumor-cell migration, invasion and metastasis” (Condeelis and Pollard 2006). For example, in PyMT-induced mammary tumors, macrophages are present in the areas of basement membrane breakdown during the development of “early-stage” metastatic lesions and systemic depletion of macrophages results in reduced formation of lung metastasis (Lin et al. 2001). Inflammatory macrophages increase peritoneal dissemination and metastatic spread of tumor cells both in vitro and in vivo in an ovarian cancer model (Mantovani et al. 2008). Clinical studies also indicate a correlation between TAM density and poor prognosis (Pollard 2004). TAMs produce a wide variety of growth factors and cytokines to stimulate the growth, motility, and invasiveness of tumor cells. Interestingly, TAM often cluster around blood vessels and the tumor margin and produce many proteases, ranging from uPA to a variety of matrix metalloproteinases (MMPs) to degrade the basement membrane for creating a channel for tumor cell invasion. TAMs have been shown to be a major source of MMP9 and urokinase-type plasminogen activator. In our recent study, we found that macrophage increase tumor invasive capacity in a TNFa–NF-kB–Snail dependent manner. TNFa-stabilized Snail is critical for inflammation-induced invasion of breast cancer cells. Knockdown of Snail expression significantly inhibited cell migration and invasion induced by inflammatory cytokines and suppressed inflammation-mediated breast cancer metastasis in an animal model (Wu et al. 2009). Thus, macrophages, the major inflammatory component of the stroma in malignancies, facilitate angiogenesis, extracellular matrix breakdown, invasion, and metastasis through multiple mechanisms. In addition to macrophages, fibroblasts/myofibroblasts comprise another major component of tumor stroma. The presence of large numbers of fibroblasts and myofibroblasts in tumor is a hallmark of advanced carcinomas. Tumor fibroblasts are derived not only from the mesenchyme that surrounds the tumor but also from the epithelial cells through EMT process (Radisky et al. 2007). Recent studies have demonstrated that cancer-activated fibroblasts (CAFs) are important in tumor cell migration and invasion. CAFs isolated from metastatic breast cancer produce elevated levels of IL-6 and enhance cancer cell invasiveness (Studebaker et al. 2008). In agreement with this, fibroblasts isolated from surgical colon cancer specimens can stimulate the invasive growth of breast and colon cancer cells (De Wever et al. 2004). Another study that systematically evaluated the effect of fibroblasts found that only melanoma cells that metastasized were influenced by fibroblasts (Cornil et al. 1991). This indicates that fibroblasts create a niche to promote metastasis. In addition, CAFs in pancreatic ductal adenocarcinoma are responsible for a poorly vascularized architecture that imposes a barrier for drug delivery and spurs metastasis (Olive et al. 2009). Furthermore, fibroblasts promote tumor cell proliferation and metastasis through the production of exogenous EMT stimuli
9 Epithelial–Mesenchymal Transition in Development and Diseases
195
such as growth factors, cytokines, chemokines, and MMPs. MMPs derived from tumor cells and stromal components are regarded as major players in assisting the metastasis of tumor cells. For example, transgenic expression of MMP3 stimulates expression of Snail through the increased cellular reactive oxygen species and thus induces down-regulation of E-cadherin and increased tumor progression (Radisky et al. 2005). Myeloid-derived suppressor cells (MDSC) are present in many cancer patients and mice with transplanted or spontaneously developed tumors (Young and Lathers 1999; Almand et al. 2001). MDSCs, characterized as CD11b+ Gr-1+ in mice, can be recruited and activated by multiple factors, such as VEGF, IL-1b, and IL-6, many of which are associated with chronic inflammation (Gabrilovich and Nagaraj 2009). Recruitment of MDSCs in turn further produces these pro-inflammatory factors, resulting in the amplification of the pro-inflammatory response. MDSCs not only suppress the adaptive immune responses but also regulate innate immune responses by modulating the cytokine production of macrophages (Sinha et al. 2007) and thus directly facilitate metastasis. Inflammatory chemo-attractants S100A8 and S100A9 regulate the accumulation of MDSC to pre-metastatic niche (Hiratsuka et al. 2006, 2008). Recent studies have shown a close correlation between the level of MDSCs and cancer stage, metastatic tumor burden, and responsiveness to chemotherapy (Diaz-Montero et al. 2009). MDSCs from mammary carcinoma can promote tumor invasion and metastasis (Bunt et al. 2007). In Tgfbr2-decificent mice, MDSCs are concentrated at the invasive tumor front and facilitate tumor cell invasion and metastasis through chemokine receptors CXCR2 and CXCR4 (Yang et al. 2008). The mesenchymal stem cells (MSCs) are a prominent component of the tumor microenvironment which can differentiate into osteoblasts, chondrocytes, adipocytes, and other cells of mesenchymal origin. MSCs can communicate with macrophage through secreting prostaglandin E2 (PGE2) to down-modulate inflammatory cytokines. PGE2 can also potentiate Wnt signaling during wound repair in various organs (Stappenbeck and Miyoshi 2009). The presence of MSCs at the primary site modulates the behavior of neighboring cancer cells (Karnoub et al. 2007). Co-mingling of bone-marrow-derived human MSCs with weakly metastatic human breast carcinoma cells significantly increases the dissemination of the cancer cells to the lung on a subcutaneous xenograft mouse model. The enhanced metastatic ability is reversible and is dependent on a paracrine loop involving MSC-supplied CCL5 and its receptor CCR5 on the breast cancer cell. The finding that MSCs concentrate in tumors from the circulation raises the possibility of using them as a cellular vehicle to deliver anticancer drug into the tumor. Other inflammatory cells, including lymphocytes, neutrophils, mast cells, T-regulatory cells, and platelets, contribute to the metastatic phenotype of cancers. All of these infiltrated inflammatory cells secrete different cytokines, chemokines, and other factors to promote the tumor cell migration and invasion, and affect the ability of tumor cells to colonize in the metastatic niche.
196
Y. Wu and B.P. Zhou
Molecular Regulation of EMT The hallmark of EMT is the loss of E-cadherin expression, an important caretaker of the epithelial phenotype. Loss of E-cadherin expression is often correlated with the tumor grade and stage because it results in the disruption of the cell–cell adhesion and an increase in the nuclear b-catenin (Cowin et al. 2005; Junghans et al. 2005). Several transcription factors have been implicated in the regulation of EMT, including zinc finger proteins of the Snail/Slug family, the basic helix– loop–helix factor Twist, E12/E47, Goosecoid, dEF1/ZEB1, and SIP1 (Nieto 2002; Yang et al. 2004; Hartwell et al. 2006). These factors act as a molecular switch of EMT program by repressing a subset of common genes that encode cadherins, claudins, integrins, mucins, plakophilin, occluding, and ZO1 to induce EMT. For example, the expression of Snail is associated with E-cadherin repression in metastasis and correlates with tumor recurrence and poor prognosis in various cancers (Elloul et al. 2005; Moody et al. 2005; Bruyere et al. 2009). In addition, extensive crosstalk among these transcription factors forms a signaling network that is responsible for establishing and maintaining mesenchymal cell phenotypes. Furthermore, some of these transcription factors, including Snail, play an important part in overcoming oncogene-induced senescence (Ansieau et al. 2008), inhibiting tumor immunosuppression (Kudo-Saito et al. 2009), and generating tumorigenic cancer stem cells (Mani et al. 2008). These transcription factors communicate and respond to the extracellular signals such as growth factors, cytokines and hypoxia from their microenvironment to induce EMT (Fig. 9.2).
κ
Fig. 9.2 An overview of signaling pathways and transcription factors involved in EMT
α
9 Epithelial–Mesenchymal Transition in Development and Diseases
197
Signaling Pathways TGFb is a primary inducer of EMT. It not only contributes to EMT during embryonic development but also induces EMT during tumor progression in vivo (Zavadil and Bottinger 2005). Overexpression of Smad2 and Smad3 results in increased EMT in a mammary epithelial model (Valcourt et al. 2005). Knockout of Smad3 blocks TGF-b-induced EMT in primary tubular epithelial cells; and the reduction of Smad2 and Smad3 function are associated with the decreased metastatic potential of breast cancer cell lines in a xenograft model (Zavadil et al. 2004). It is interesting that SMAD3 and SMAD4 interact and form a complex with Snail and target to the promoters of CAR (a tight-junction protein) and E-cadherin during TGFb-inducing EMT in breast epithelial cells (Vincent et al. 2009). Bos et al. (2009) identified that TGFb primed cancer cells for lung metastasis through angiopoietin-like 4 via Smad signaling pathway. On contrast, inhibition of TGFb or TGFb receptor reduces the invasive and metastatic activities of cancer cells. TGFb can also downregulate various epithelial molecules, including E-cadherin, ZO-1, and several specific keratins; and it also upregulates certain mesenchymal proteins such as fibronectin, fibroblast specific protein 1, a-smooth muscle actin and vimentin. In addition, TGFb cooperates with numerous kinases such as RAS, MAPK, p38MAP to promote EMT (Zavadil and Bottinger 2005; Buijs et al. 2007). More specifically, p38 MAPK and RhoA mediate an autocrine TGFb-induced EMT in NMuMG mouse mammary epithelial cells (Bhowmick et al. 2001). ECM molecules, such as integrin b1 and Fibulin-5, augment TGFb-induced EMT in a MAPK-dependent mechanism (Bhowmick et al. 2001; Lee et al. 2008). Constitutive activation of Raf enhances the function of TGFb in inducing EMT via MAPK in MDCK cells (Janda et al. 2002). TGFb also induces EMT through changes in the expression of certain cell polarity molecules. For example, TGFb can induce phosphorylation of Par6, which in turn stimulates binding of Par6 to E3 ligase Smurf1. The Par6–Smurf1 complex then mediates the localized ubiquination of RhoA to disrupt tight junctions during EMT (Ozdamar et al. 2005). TGFb can also downregulate the Par3 expression to destroy the cell polarity (Wang et al. 2008). It is interesting to note that Abl can inhibit TGFb-mediated EMT in normal and metastatic mammary epithelial cells (MECs) (Allington et al. 2009). Furthermore, TGFb can cooperate with other oncogenic pathways such as Notch, Wnt/b-catenin, and NF-kB to maintain the mesenchymal phenotype of invasive/metastatic tumor cells (Nawshad et al. 2005; Zavadil and Bottinger 2005; Neth et al. 2007). The Wnt/b-catenin pathway has a particularly tight link with EMT (Li et al. 2000). On one hand, b-catenin is an essential component of adherent junctions, where it provides the link between E-cadherin and a-catenin and modulates cell– cell adhesion and cell migration. On the other hand, b-catenin also functions as a transcription cofactor with T cell factor (TCF). Nuclear translocation of b-catenin can activate the expression of Slug and thus induces EMT. Expression of b-catenin in oocyte induces a premature EMT in the epiblast, concomitant with Snail transcription. Interestingly, Snail is a highly unstable protein and is dually regulated
198
Y. Wu and B.P. Zhou
by protein stability and cellular location. We showed that GSK-3b binds and phosphorylates Snail at two consensus motifs to dually regulate the function of this protein: phosphorylation at the first motif regulates its ubiquitination mediated by b-Trcp, whereas phosphorylation at the second motif controls its subcellular localization (Zhou et al. 2004). Thus, Wnt can suppress the activity of GSK-3b and stabilizes the protein level of Snail and b-catenin to induce EMT and cancer metastasis (Yook et al. 2005, 2006). Meanwhile, Snail can functionally interact with b-catenin to increase Wnt-dependent target gene expression to promote EMT (Stemmer et al. 2008). Increasing evidence indicates that Wnt signaling is strongly associated with human basal-like breast cancer. Inhibiting Wnt signaling through LRP6 reduces the capacity of cancer cells to self-renew and colonize in vivo, and results in the re-expression of breast epithelial markers and repression of EMT transcription factors Slug and Twist (DiMeo et al. 2009). How the synergistic activation of Snail and b-catenin by Wnt signaling pathway enhances EMT and metastasis remains further defined. Notch is an evolutionarily conserved signaling pathway that regulates cell fate specification, self-renewal and differentiation in embryonic and postnatal tissues. Four Notch (Notch 1−4) and five ligands (Jagged1, 2 and Deltalike1, 3, 4) have been identified. Notch signaling is normally activated followed by ligand–receptor binding between two neighboring cells, Notch undergoes intramembrane cleavage by g-secretase and its intracellular domain (NICD) is released and translocates to the nucleus to activate gene transcription by associating with Mastermind-like 1 (MAM) and histone acetyltransferase p300/CBP. Alteration of Notch signaling has been associated with various types of cancer in which Notch can act as an oncogene or as a tumor suppressor depending on the cellular context. The first observation that Notch pathway is required for EMT was derived from cardiac valve and cushion formation at heart development (Timmerman et al. 2004). This implies that Notch, acting through a similar mechanism, induces EMT during tumor progression and converts polarized epithelial cells into motile and invasive ones (Grego-Bessa et al. 2004). Indeed, overexpression of Jagged1 and Notch1 induces the expression of Slug and correlates with poor prognosis in various human cancers (Leong et al. 2007). Slug is essential for Notch-mediated EMT by repressing E-cadherin expression, which results in b-catenin activation and resistance to anoikis. Inhibition of Notch signaling in xenografted Slug-positive/E-cadherin-negative breast tumors promotes apoptosis and inhibits tumor growth and metastasis (Leong et al. 2007). In addition, Notch signaling deploys two distinct mechanisms that act in synergy to control the expression of Snail (Sahlgren et al. 2008). First, Notch directly upregulates Snail expression by recruitment of the Notch intracellular domain to the Snail promoter, and second, Notch potentiates hypoxia-inducible factor 1a (HIF-1a) recruitment to the lysyl oxidase (LOX) promoter and elevates the hypoxia-induced upregulation of LOX, which stabilizes the Snail protein. Thus, Notch signaling is required to convert the hypoxic stimulus into EMT and increases the invasiveness of tumor cells. In addition, Notch signaling pathway is involved in the acquisition of EMT phenotype of gemcitabine-resistant (GR) cells in pancreatic cancer (Wang et al. 2009). Downregulation of Notch signaling is associated with decreased invasive behavior of GR
9 Epithelial–Mesenchymal Transition in Development and Diseases
199
cells. Moreover, Notch signaling leads to the increased expression of vimentin, ZEB1, Slug, Snail, and NF-kB and results in EMT. Thus, inhibition of Notch signaling by novel therapeutic strategies can be clinically important in overcoming drug resistance and EMT phenotype of tumor cells. The Hedgehog (Hh) signaling pathway was first identified in a large screen for Drosophila genes required for patterning of the early embryo (Hooper and Scott 2005; Jacob and Lum 2007). The Hh ligands, Sonic-, Desert-, and Indian Hh in vertebrates and Hh in Drosophila, are secreted proteins that undergo several posttranslational modifications to gain full activity. Key effectors of Hh signaling include zinc-finger proteins of the Gli1-3 transcription factors. Hh signaling can initiate cell growth, cell division, lineage specification, and axon guidance and can also function as a survival factor. Activation of Hh signaling also leads to EMT. In mouse epidermal cells or in rat kidney epithelial cells immortalized with adenovirus E1A, Gli1 rapidly induces transcription of Snail and promotes EMT (Li et al. 2006, 2007). Targeted expression of Gli1 in the epithelial cells of mammary gland of mice induces the expression of Snail and thus results in the disruption of the mammary epithelial network and alveologenesis during pregnancy (Fiaschi et al. 2007). In addition, Hedgehog signals induce JAG2 up-regulation for Notch-CSLmediated Snail expression; on the other hand, Hedgehog induces TGFb1 secretion to induce ZEB1 and ZEB2 expression through TGFb and NF-kB pathways. Conversely, blockade of Hedgehog signaling by inhibitor cyclopamine suppresses pancreatic cancer invasion and metastasis through inhibiting EMT (Feldmann et al. 2007). The crosstalk between the Hh and EMT also presents in human esophageal squamous cell carcinoma (ESCC) (Isohata et al. 2009). Hh and EMT signaling genes are co-expressed on the undifferentiated esophageal epithelial cells and most ESCCs. These findings suggest that the mesenchymal gene expression is maintained or strengthened through Hh signaling in cancer cells.
Cytokines TNFa, a key inflammatory cytokine, plays a central role in the tumor progression. Constitutive expression of the TNFa from tumor microenvironment is a characteristic of many malignant tumors and its presence is often associated with poor prognosis. Several lines of evidence point to the tumor-promoting effects of TNFa in inflammation-driven tumorigenesis. First, overexpression of TNFa confers migratory and invasive properties of many tumor cell lines (Rosen et al. 1991). Second, TNFa and TNFa receptor 1 (TNFR1) knock-out mice are resistant to chemicalinduced carcinogenesis in skin and liver metastasis in an experimental colon cancer model (Knight et al. 2000; Arnott et al. 2004). Third, various tumor-promoting effects of TNFa are further confirmed in enhancing tumor cell motility, activating oncogenic pathways, and triggering EMT. TNFa can also promote breast cancer cell migration through up-regulation LOX (Liang et al. 2007). Endogenous TNFa contributes to the growth and invasiveness of primary pancreatic ductal adenocarcinoma
200
Y. Wu and B.P. Zhou
and anti-TNFa inhibit metastasis of these tumors (Egberts et al. 2008). Using RNA interference technology, Kulbe et al. (2007) demonstrated that tumor growth and dissemination were significantly inhibited when TNFa production was blocked. In addition, TNFa can up-regulate SELECTIN and VCAM1 on endothelial cells that promote tumor cell adhesion and migration (Mannel et al. 1994; Stoelcker et al. 1995). Furthermore, TNFa enhances the invasive property of cancer cells by inducing EMT through Snail or ZEB1/ZEB2 (Chua et al. 2007; Chuang et al. 2008). In our recent study, we found that inflammatory cytokine TNFa is the major signal to induce Snail stabilization and EMT induction (Wu et al. 2009). We showed that TNFa greatly enhanced the migration and invasion of tumor cells by inducing EMT program through NF-kB-mediated Snail stabilization. Knockdown of Snail expression not only inhibits TNFa-induced cancer cell migration and invasion in vitro but also suppresses LPS-mediated metastasis in vivo. Furthermore, knockdown of Snail expression not only blocks metastasis that is intrinsic to the metastatic breast cancer cells but also greatly suppresses inflammation-accelerated metastasis. Collectively, our study indicates that Snail stabilization and EMT induction mediated by the inflammatory cytokine TNFa are critical for metastasis. Our study provides a plausible molecular mechanism for tumor cell dissemination and invasion at the tumor invasive front. IL-6 is another important inflammatory cytokine linking inflammation and cancer. IL-6 transmits its signal through a common signaling receptor, gp130, expressed on many cell types. IL-6 binds to the sIL-6R receptor (gp80, present either on the cell surface or in solution), which then induces dimerization of gp130 chains resulting in activation of the associated Janus kinases (JAKs). JAKs phosphorylate gp130, leading to the recruitment and activation of the STAT3 and STAT1 transcription factors as well as other molecules (SHP2, Ras-MAPK, and PI3K) (Mumm and Oft 2008). The role of IL-6 in accelerating tumorigenesis is becoming clear as exogenous administration of IL-6 to mice during tumor initiation results in an increase in tumor burden and multiplicity (Grivennikov et al. 2009). IL-6 also enhances tumor proliferation in tumor-initiating intestinal epithelial cells (IECs) through NF-kB-IL-6-STAT3 cascade (Bollrath et al. 2009; Bromberg and Wang 2009; Grivennikov et al. 2009). IL-6 can also act as an inducer of EMT in breast cancer cells. Ectopic expressing of IL-6 in breast adenocarcinoma cells exhibits an EMT phenotype characterized by suppressing E-cadherin expression and inducting vimentin, N-cadherin, Snail, and Twist (Sullivan et al. 2009). In addition, IL-6 also synergizes with EGF in inducing EMT through the activation JNK2/STAT3 in ovarian carcinomas (Colomiere et al. 2009). The interleukin-1 (IL-1) also promotes inflammatory processes and augments metastasis. There are two forms of IL-1 protein, IL-1a and IL-1b, and one antagonistic protein IL-1 receptor antagonist (IL-1ra). IL-1b is active solely in its secreted form, whereas IL-1b is active mainly as an intracellular precursor. IL-1 is abundant at tumor sites, where it affects the process of carcinogenesis, tumor growth and invasiveness, and the patterns of tumor–host interactions (Apte et al. 2006). Genetic ablation of IL-1b in mice results in the absence of metastatic tumors in vivo (Voronov et al. 2003). Liver metastasis is almost completely inhibited in mice with
9 Epithelial–Mesenchymal Transition in Development and Diseases
201
deletion of interleukin-1b converting enzyme, which is required for the processing of IL-1b (Vidal-Vanaclocha et al. 2000). IL-1b also directly induces uPA expression and NF-kB activation, which results in the migration of A549 cells (Cheng et al. 2009).
Hypoxia A hypoxic condition has been detected in many human solid malignancies and it provides a strong selective advantage for tumor growth and survival by enhancing angiogenesis and metastasis. Under hypoxic conditions, cells respond by stabilizing hypoxia-inducible factor (HIF1a) that, in turn, forms the transcriptional complex with HIF1b and activates the transcription of downstream target genes through direct binding to the hypoxic response element located in the promoter of those genes that control glucose metabolism, angiogenesis, survival, and invasion. The well-recognized mechanism involved in hypoxia-induced metastasis is the activation of VEGF by HIF1a to promote angiogenesis and facilitate tumor metastasis. In addition, HIF1a can induce the expression of the chemokine receptor CXCR4 in renal cell carcinoma to promote organ specific metastatic dissemination (Staller et al. 2003). Although hypoxia and EMT are considered as crucial events favoring invasion and metastasis of many cancer cells, the two processes are long thought to involve few common molecular mechanisms. The link between hypoxia and EMT has been established recently as hypoxia induces the expression and coordinates the interplay of several EMT regulators. For example, HIF1a can regulate the expression of several EMT inducers such as Snail, Slug, Twist, ZEB1, and SIP1, directly or indirectly, and thus leads to the induction of EMT and enhances metastasis under hypoxic condition (Imai et al. 2003; Krishnamachary et al. 2003; Davidson and Sukumar 2005; Hotz et al. 2007; Yang et al. 2008). Interestingly, tumor hypoxia (measured by the expression of HIF1a) significantly correlated with both Snail and Twist expression and co-expression of these molecules correlated with the highest probability of metastasis and the worst prognosis in head and neck cancer (Yang et al. 2008). Hypoxia also induces the expression of uPA, uPAR, MMP-2, and MMP-9 to increase metastatic invasion both in vitro and in vivo (Lunt et al. 2009). Furthermore, hypoxia also induces the expression of lysyl oxidas (LOX), which promotes Snail stabilization and activity (Peinado et al. 2005a, b). LOX is required for the maturation of newly synthesized collagen fibrils and promotes metastasis through changes in focal adhesion kinase activity. These results indicate that Snail not only can repress E-cadherin but also cooperates with LOX to induce EMT under hypoxic condition. Interestingly, EMT mainly occurs at the tumor invasion front (tumor–stroma boundary) while hypoxia mainly operates in the intratumoral areas where the pO2 level is below 10 mmHg. How these two events communicate with each other in vivo remains to be further investigated. We speculate that the alterations within the tumor microenvironment as a consequence of hypoxia connect these two processes together.
202
Y. Wu and B.P. Zhou
EMT Generates Cancer Stem Cells In addition to its ability of enhancing cell migration during tissue morphogenesis and tumor metastasis, EMT creates cells that act as progenitors of many different tissues during development. For example, type 1 EMT generates mesoderm, which gives rise to a wide range of cell types, including muscle, bone, and connective tissues; type 1 EMT also creates neural crest delamination, which gives rise to glial and neuronal cells, adrenal glandular tissues, pigment-containing cells of the epidermis, and skeletal and connective tissues. These findings indicate that the genetic program of EMT possesses the ability to generate various cancer stem cells (CSC) in solid tumors. Interestingly, EMT characteristics are enriched in a subtype of basal-like breast cancer, which is highly metastatic and has CSC phenotype with the expression of high level of stem cell marker CD44+/CD24–. In addition, CD44+/ CD24– cells purified from normal and malignant breast cancer tissue exhibit features of an EMT, such as reduced expression of E-cadherin, significantly increased expression of fibronectin and vimentin, and robust levels of FOXC2, Twist, Snail, and Slug. Weinberg’s group demonstrated recently that EMT confers tumor cells with stem cell-like properties (Mani et al. 2007, 2008). Ectopic expression of Snail or Twist greatly increases the tumor cells’ ability to form tumorsphere, a property associated with mammary epithelial stem cells. In addition, TGFb, a potent inducer of EMT, also enhances the expression of CD44+/CD24– in tumor cells, implying that the surrounding tumor microenvironment is most likely the critical factor for metastasis by influencing the tumor cells undergoing EMT. Furthermore, breast tumors resistant to chemo- and endocrine therapies are enriched with cells bearing CSC signature and EMT markers in the post-treatment specimens. The findings that EMT is often seen at the invading edge of tumors and that EMT confers tumor cells with “metastatic cancer stem cell” traits highlight a notion that EMT is responsible for invasion and metastasis. Consistent with this notion, a distinct subset of highly migratory pancreatic stem cells co-expressing CD133+ and CXCR4 (CD133+/ CXC4+) found at the invasive edge of pancreatic carcinomas is essential for metastasis (Hermann et al. 2007). Interestingly, several signaling pathways that mediate stem cell self-renewal such as Wnt, Hh, and Notch also induce EMT. Thus, EMT endows cancer cells with stem cell properties for metastatic dissemination at primary tumor and self-renewal properties needed for initiating secondary tumor.
Genetic and Epigenetic Control of EMT Both EMT and epigenetic modification (DNA methylation and histone modifications) are dynamic and efficient processes during development, differentiation and carcinogenesis. Several lines of evidences indicate that gene expression involved in EMT is altered in tumors through genetic and epigenetic mechanisms. First, E-cadherin can be inactivated by genetic mutation or DNA hyper-methylation
9 Epithelial–Mesenchymal Transition in Development and Diseases
203
(epigenetic). Second, several EMT transcriptional factors are hyper-methylated in more differentiated tumor cells. For example, Bloushtain-Qimron et al. isolated stem cell-like CD44+/CD24– and more differentiated epithelial CD44–/CD24+ cells from either normal or neoplastic breast tissues and compared their methylation profiles. They found that transcriptional factors (such as FOXC2) implicated in the induction of EMT and stem cell functions are hypomethylated and highly expressed in CD44+/CD24– cells compared with their CD44–/CD24+ counterparts (Bloushtain-Qimron et al. 2008). In addition, sustained TGFb stimulation of immortalized human mammary epithelial cells induces EMT, which accompanies de novo DNA methylation on E-cadherin and estrogen receptor alpha (ERa) promoters, a prominent feature in human basal-like breast cancer (Dumont et al. 2008). Furthermore, TGFb induces de novo DNA methylation on the promoter of E-cadherin and b4 integrin, a receptor for laminins of the basement membrane in mammary gland cells. These DNA methylations are accompanied by the decrease in histone activation marks (H3K9Ac and H3K4me3) and an increased histone repression modification on H3K27me3 (Yang et al. 2009). Together, these studies indicate that signals from the tumor microenvironment can modulate the induction of EMT through an epigenetic mechanism.
Micro RNA for EMT MicroRNA appears as a powerful component of the cellular signaling regulator of EMT (Table 9.2). MicroRNAs are small 20- to 22-nucleotide long noncoding RNAs that modulate gene expression at the post-transcriptional level (Stadler and Ruohola-Baker 2008; Stefani and Slack 2008). MicroRNAs have been implicated in regulating diverse cellular pathways, such as cell differentiation, proliferation, and programmed cell death, and are commonly dysregulated in human cancers. Recent findings suggest that microRNAs also contribute to EMT as either an enhancer or a suppressor. For example, Twist induces microRNA-10b, which inhibits the translation of HOXD10 and results in elevated expression of RhoC, and thus facilitates metastasis of breast tumor cells (Ma et al. 2007). Acting in the opposite direction, miR-31 inhibits breast cancer metastasis via coordinated repression of a cohort of metastasis-promoting genes, including RhoA (Valastyan et al. 2009). Evidence for specific microRNAs in EMT regulation is further provided in that microRNA-200s, consisting of five members, are markedly downregulated in cells that have undergone EMT in response to TGFb (Burk et al. 2008; Gregory et al. 2008; Korpal et al. 2008). Because microRNA-200 directly targets the mRNA of ZEB1 and ZEB2/SIP1, expression of miR-200 induces upregulation of E-cadherin in cancer cell lines and suppresses their motility. Consistent with their role in regulating EMT, loss of microRNA-200 is commonly found in invasive breast cancer cell lines with mesenchymal phenotype and in regions of metaplastic breast cancer specimens lacking E-cadherin expression. Several recent reports overwhelmingly link the microRNA-200 family with EMT. These studies show that expression of
204
Y. Wu and B.P. Zhou
Table 9.2 MicroRNAs in EMT Name Function Targets miR-10b enhance HOXD10 miR-21 enhance PDCD4 TPM1 nmiR-29a enhance tristetraprolin miR-373 enhance CD44 miR-520c
enhance
CD44
miR-155 miR-126 miR-335 miR-205 miR-31 miR-146a miR-200
enhance inhibit inhibit inhibit inhibit inhibit inhibit
RhoA Crk, PI3K SOX4 and tenascin C ZEB1 and SIP1 RhoA Bcl-xl and STAT3 ZEB1 and SIP1
miR-138
inhibit
miR-221 miR-23b
inhibit inhibit
uPA and c-met
Cancer types Breast cancer Breast cancer Breast cancer Prostate, breast, and esophageal cancers, testicular germ cell tumors Prostate, breast, and esophageal cancers, testicular germ cell tumors Breast cancer Lung and colon cancer Breast cancer Breast cancer Breast cancer Breast, ovarian, bladder, pancreatic, and prostate cancers Head and neck squamous cell carcinoma Prostate cancer Hepatocellular carcinoma
microRNA-200 overcomes the resistance to EGFR therapy in bladder cancer cells and gemcitabine-resistant pancreatic cancer cells (Adam et al. 2009; Li et al. 2009). In addition, an EMT specific microRNA miR-21 is found in TGFb-induced EMT in human keratinocytes, a model of epithelial cell plasticity for epidermal injury and skin carcinogenesis (Zavadil et al. 2007). MiR-21 is abundantly expressed and associates with carcinogenesis. It targeted two tumor suppressors, tropomyosin 1 (TPM1) and programmed cell death-4 (PDCD4) to promote cell proliferation, microfilament organization, and anchorage-independent growth (Zhu et al. 2007; Asangani et al. 2008). MiR-155 also plays an important role in TGFb-induced EMT and cell migration and invasion through targeting RhoA (Kong et al. 2008).
Perspective Significant progress has been made in recent years regarding our understanding of EMT in development, fibrosis, and metastasis. Converging evidence from cell culture studies, transgenic mouse model, and documented immunohistochemical staining on human tumors, points out that EMT, amongst other ways, is the most important mechanism implicated in tumor metastasis and recurrence. It is clear that the induction of EMT is dependent on the signals that cells received from their microenvironment. Activation of several different EMT transcriptional factors, depending on the origin of the cells and the signals that cells received, induce a
9 Epithelial–Mesenchymal Transition in Development and Diseases
205
common genetic program of EMT. A recent link between EMT and cancer stem cells has sparked considerable interests and excitements, because it is widely accepted that only a minor population of tumor cells can initiate and support the development of tumors and that highly aggressive tumor cells share many of the characteristics of embryonic progenitor cells. The finding that EMT confers tumor cells with the CSC traits provides plausible molecular mechanism for tumor metastasis and recurrence. However, several challenges remain to be addressed regarding the fundamental mechanism and regulation of EMT. First, EMT provides a considerable degree of plasticity and reversibility for cells to differentiate or de-differentiate between cancer epithelial cells and mesenchymal cells (or CSC): why tumor cells have more plasticity than the normal differentiated epithelial cells? Do the mutation of tumor suppressors and/or activation of oncogenes during carcinogenesis endow tumor cells more intrinsic plasticity so that they are more sensitive to EMT inductive signals? Alternatively, does this plasticity come from stem or progenitor cells that actually initiate the tumor? And how the plasticity is regulated during tumor progress? Second, many studies are focused on EMT, however, the reversible event, the mesenchymal–epithelial transition, remains unclear. Uncovering the mechanism and signaling pathways of MET will help us to better understand the biology of EMT in development and metastasis. Third, EMT and MET are transient and dynamic events, particularly at the tumor invasive front, and a powerful imaging system that can capture this process will reveal the mystery of EMT. In summary, the association of EMT with tumor invasive front, disseminated cells, traits of CSC, and resistance to conventional therapies will place type 3 EMT in the spotlight for those wishing to control metastasis. A better understanding of this event will ultimately lead to new and improved therapies for patients with cancers. Acknowledgment We apologize to the many contributors to this field whose works are important while we are unable to cite here. Our study is supported by the grants from NIH (RO1CA125454), the Susan G Komen Foundation (KG081310), and the Mary Kay Ash Foundation (to B.P. Zhou).
References Adam L, et al (2009) miR-200 expression regulates epithelial-to-mesenchymal transition in bladder cancer cells and reverses resistance to epidermal growth factor receptor therapy. Clin Cancer Res 15(16):5060–5072. Allington TM, Galliher-Beckley AJ, Schiemann WP (2009) Activated Abl kinase inhibits oncogenic transforming growth factor-{beta} signaling and tumorigenesis in mammary tumors. FASEB J 23:4231–4243. Almand B, et al (2001) Increased production of immature myeloid cells in cancer patients: a mechanism of immunosuppression in cancer. J Immunol 166(1):678–689. Ansieau S, et al (2008) Induction of EMT by twist proteins as a collateral effect of tumor-promoting inactivation of premature senescence. Cancer Cell 14(1):79–89. Apte RN, et al (2006) Effects of micro-environment- and malignant cell-derived interleukin-1 in carcinogenesis, tumour invasiveness and tumour-host interactions. Eur J Cancer 42(6): 751–759.
206
Y. Wu and B.P. Zhou
Arnott CH, et al (2004) Expression of both TNF-alpha receptor subtypes is essential for optimal skin tumour development. Oncogene 23(10):1902–1910. Arnoux V, et al (2008) Erk5 controls Slug expression and keratinocyte activation during wound healing. Mol Biol Cell 19(11):4738–4749. Asangani IA, et al (2008) MicroRNA-21 (miR-21) post-transcriptionally downregulates tumor suppressor Pdcd4 and stimulates invasion, intravasation and metastasis in colorectal cancer. Oncogene 27(15):2128–2136. Bataille F, et al (2008) Evidence for a role of epithelial mesenchymal transition during pathogenesis of fistulae in Crohn’s disease. Inflamm Bowel Dis 14(11):1514–1527. Bhowmick NA, et al (2001) Transforming growth factor-beta1 mediates epithelial to mesenchymal transdifferentiation through a RhoA-dependent mechanism. Mol Biol Cell 12(1):27–36. Bloushtain-Qimron N, et al (2008) Cell type-specific DNA methylation patterns in the human breast. Proc Natl Acad Sci U S A 105(37):14076–14081. Bollrath J, et al (2009) gp130-mediated Stat3 activation in enterocytes regulates cell survival and cell-cycle progression during colitis-associated tumorigenesis. Cancer Cell 15(2):91–102. Bos PD, et al (2009) Genes that mediate breast cancer metastasis to the brain. Nature 459(7249): 1005–1009. Bromberg J, Wang TC (2009) Inflammation and cancer: IL-6 and STAT3 complete the link. Cancer Cell 15(2):79–80. Bruyere F, et al (2009) Snail expression is an independent predictor of tumor recurrence in superficial bladder cancers. Urol Oncol Epub ahead of print. Buijs JT, et al (2007) TGF-beta and BMP7 interactions in tumour progression and bone metastasis. Clin Exp Metastasis 24(8):609–617. Bunt SK, et al (2007) Reduced inflammation in the tumor microenvironment delays the accumulation of myeloid-derived suppressor cells and limits tumor progression. Cancer Res 67(20): 10019–10026. Burk U, et al (2008) A reciprocal repression between ZEB1 and members of the miR-200 family promotes EMT and invasion in cancer cells. EMBO Rep 9(6):582–589. Chambers AF, Groom AC, MacDonald IC (2002) Dissemination and growth of cancer cells in metastatic sites. Nat Rev Cancer 2(8):563–572. Cheng CY, et al (2009) IL-1 beta induces urokinase-plasminogen activator expression and cell migration through PKC alpha, JNK1/2, and NF-kappaB in A549 cells. J Cell Physiol 219(1):183–193. Christofori G (2006) New signals from the invasive front. Nature 441(7092):444–450. Chua HL, et al (2007) NF-kappaB represses E-cadherin expression and enhances epithelial to mesenchymal transition of mammary epithelial cells: potential involvement of ZEB-1 and ZEB-2. Oncogene 26(5):711–724. Chuang MJ, et al (2008) Tumor-derived tumor necrosis factor-alpha promotes progression and epithelial-mesenchymal transition in renal cell carcinoma cells. Cancer Sci 99(5):905–913. Colomiere M, et al (2009) Cross talk of signals between EGFR and IL-6R through JAK2/STAT3 mediate epithelial-mesenchymal transition in ovarian carcinomas. Br J Cancer 100(1): 134–144. Condeelis J, Pollard JW (2006) Macrophages: obligate partners for tumor cell migration, invasion, and metastasis. Cell 124(2):263–266. Cornil I, et al (1991) Fibroblast cell interactions with human melanoma cells affect tumor cell growth as a function of tumor progression. Proc Natl Acad Sci U S A 88(14):6028–6032. Cowin P, Rowlands TM, Hatsell SJ (2005) Cadherins and catenins in breast cancer. Curr Opin Cell Biol 17(5):499–508. Davidson LA, et al (2006), Integrin alpha5beta1 and fibronectin regulate polarized cell protrusions required for Xenopus convergence and extension. Curr Biol 16(9):833–844. Davidson NE, Sukumar S (2005) Of Snail, mice, and women. Cancer Cell 8(3):173–174. Dawes-Hoang RE, et al (2005) Folded gastrulation, cell shape change and the control of myosin localization. Development 132(18):4165–4178.
9 Epithelial–Mesenchymal Transition in Development and Diseases
207
De Wever O, et al (2004) Critical role of N-cadherin in myofibroblast invasion and migration in vitro stimulated by colon-cancer-cell-derived TGF-beta or wounding. J Cell Sci 117(Pt 20):4691–4703. Diaz-Montero CM, et al (2009) Increased circulating myeloid-derived suppressor cells correlate with clinical cancer stage, metastatic tumor burden, and doxorubicin-cyclophosphamide chemotherapy. Cancer Immunol Immunother 58(1):49–59. DiMeo TA, et al (2009) A novel lung metastasis signature links Wnt signaling with cancer cell self-renewal and epithelial-mesenchymal transition in basal-like breast cancer. Cancer Res 69(13):5364–5373. Dumont N, et al (2008) Sustained induction of epithelial to mesenchymal transition activates DNA methylation of genes silenced in basal-like breast cancers. Proc Natl Acad Sci U S A 105(39): 14867–14872. Egberts JH, et al (2008) Anti-tumor necrosis factor therapy inhibits pancreatic tumor growth and metastasis. Cancer Res 68(5):1443–1450. Elloul S, et al (2005) Snail, Slug, and Smad-interacting protein 1 as novel parameters of disease aggressiveness in metastatic ovarian and breast carcinoma. Cancer 103(8):1631–1643. Feldmann G, et al (2007) Blockade of hedgehog signaling inhibits pancreatic cancer invasion and metastases: a new paradigm for combination therapy in solid cancers. Cancer Res 67(5): 2187–2196. Fiaschi M, et al (2007) Targeted expression of GLI1 in the mammary gland disrupts pregnancyinduced maturation and causes lactation failure. J Biol Chem 282(49):36090–36101. Flanders KC (2004) Smad3 as a mediator of the fibrotic response. Int J Exp Pathol 85(2):47–64. Franci C, et al (2006) Expression of Snail protein in tumor-stroma interface. Oncogene 25(37):5134–5144. Gabrilovich DI, Nagaraj S (2009) Myeloid-derived suppressor cells as regulators of the immune system. Nat Rev Immunol 9(3):162–174. Greenburg G, Hay ED (1998) Cytoskeleton and thyroglobulin expression change during transformation of thyroid epithelium to mesenchyme-like cells. Development 102(3):605–622. Grego-Bessa J, et al (2004) Notch and epithelial-mesenchyme transition in development and tumor progression: another turn of the screw. Cell Cycle 3(6):718–721. Gregory PA, et al (2008) The miR-200 family and miR-205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nat Cell Biol 10(5):593–601. Gressner AM, et al (2002) Roles of TGF-beta in hepatic fibrosis. Front Biosci 7:d793–d807. Grivennikov S, et al (2009) IL-6 and Stat3 are required for survival of intestinal epithelial cells and development of colitis-associated cancer. Cancer Cell 15(2):103–113. Hartwell KA, et al (2006) The Spemann organizer gene, Goosecoid, promotes tumor metastasis. Proc Natl Acad Sci U S A 103(50):18969–18974. Hay ED (1995) An overview of epithelio-mesenchymal transformation. Acta Anat (Basel), 154(1):8–20. Hermann PC, et al (2007) Distinct populations of cancer stem cells determine tumor growth and metastatic activity in human pancreatic cancer. Cell Stem Cell 1(3):313–323. Hiratsuka S, et al (2006) Tumour-mediated upregulation of chemoattractants and recruitment of myeloid cells predetermines lung metastasis. Nat Cell Biol 8(12):1369–1375. Hiratsuka S, et al (2008) The S100A8-serum amyloid A3-TLR4 paracrine cascade establishes a pre-metastatic phase. Nat Cell Biol 10(11):1349–1355. Hooper JE, Scott MP (2005) Communicating with Hedgehogs. Nat Rev Mol Cell Biol 6(4): 306–317. Hotz B, et al (2007) Epithelial to mesenchymal transition: expression of the regulators snail, slug, and twist in pancreatic cancer. Clin Cancer Res 13(16):4769–4776. Huber MA, Kraut N, Beug H (2005) Molecular requirements for epithelial-mesenchymal transition during tumor progression. Curr Opin Cell Biol 17(5):548–558. Imai T, et al (2003) Hypoxia attenuates the expression of E-cadherin via up-regulation of SNAIL in ovarian carcinoma cells. Am J Pathol 163(4):1437–1447.
208
Y. Wu and B.P. Zhou
Isohata N, et al (2009) Hedgehog and epithelial-mesenchymal transition signaling in normal and malignant epithelial cells of the esophagus. Int J Cancer 125(5):1212–1221. Jacob L, Lum L (2007) Deconstructing the hedgehog pathway in development and disease. Science 318(5847):66–68. Janda E, et al (2002) Ras and TGF[beta] cooperatively regulate epithelial cell plasticity and metastasis: dissection of Ras signaling pathways. J Cell Biol 156(2):299–313. Joyce JA, Pollard JW (2009) Microenvironmental regulation of metastasis. Nat Rev Cancer 9(4): 239–252. Junghans D, Haas IG, Kemler R (2005) Mammalian cadherins and protocadherins: about cell death, synapses and processing. Curr Opin Cell Biol 17(5):446–452. Kalluri R (2009) EMT: when epithelial cells decide to become mesenchymal-like cells. J Clin Invest 119(6):1417–1419. Kalluri R, Neilson EG (2003) Epithelial-mesenchymal transition and its implications for fibrosis. J Clin Invest 112(12):1776–1784. Kalluri R, Weinberg RA (2009) The basics of epithelial-mesenchymal transition. J Clin Invest 119(6):1420–1428. Karnoub AE, et al (2007) Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 449(7162):557–563. Knight B, et al (2000) Impaired preneoplastic changes and liver tumor formation in tumor necrosis factor receptor type 1 knockout mice. J Exp Med 192(12):1809–1818. Kolsch V, et al (2007) Control of Drosophila gastrulation by apical localization of adherens junctions and RhoGEF2. Science 315(5810):384–386. Kong W, et al (2008) MicroRNA-155 is regulated by the transforming growth factor beta/Smad pathway and contributes to epithelial cell plasticity by targeting RhoA. Mol Cell Biol 28(22):6773–6784. Korpal M, et al (2008) The miR-200 family inhibits epithelial-mesenchymal transition and cancer cell migration by direct targeting of E-cadherin transcriptional repressors ZEB1 and ZEB2. J Biol Chem 283(22):14910–14914. Krishnamachary B, et al (2003) Regulation of colon carcinoma cell invasion by hypoxia-inducible factor 1. Cancer Res 63(5):1138–1143. Kudo-Saito C, et al (2009) Cancer metastasis is accelerated through immunosuppression during Snail-induced EMT of cancer cells. Cancer Cell 15(3):195–206. Kulbe H, et al (2007) The inflammatory cytokine tumor necrosis factor-alpha generates an autocrine tumor-promoting network in epithelial ovarian cancer cells. Cancer Res 67(2):585–592. Lee YH, et al (2008) Fibulin-5 initiates epithelial-mesenchymal transition (EMT) and enhances EMT induced by TGF-beta in mammary epithelial cells via a MMP-dependent mechanism. Carcinogenesis 29(12):2243–2251. Leong KG, et al (2007) Jagged1-mediated Notch activation induces epithelial-to-mesenchymal transition through Slug-induced repression of E-cadherin. J Exp Med 204(12):2935–2948. Liang M, Zhang P, Fu J (2007) Up-regulation of LOX-1 expression by TNF-alpha promotes transendothelial migration of MDA-MB-231 breast cancer cells. Cancer Lett 258(1):31–37. Lin EY, et al (2001) Colony-stimulating factor 1 promotes progression of mammary tumors to malignancy. J Exp Med 193(6):727–740. Li X, et al (2006) Snail induction is an early response to Gli1 that determines the efficiency of epithelial transformation. Oncogene 25(4):609–621. Li X, et al (2007) Gli1 acts through Snail and E-cadherin to promote nuclear signaling by betacatenin. Oncogene 26(31):4489–4498. Li Y, et al (2009) Up-regulation of miR-200 and let-7 by natural agents leads to the reversal of epithelial-to-mesenchymal transition in gemcitabine-resistant pancreatic cancer cells. Cancer Res 69(16):6704–6712. Li Y, Hively WP, Varmus HE (2000) Use of MMTV-Wnt-1 transgenic mice for studying the genetic basis of breast cancer. Oncogene 19(8):1002–1009. Lunt SJ, Chaudary N, Hill RP (2009) The tumor microenvironment and metastatic disease. Clin Exp Metastasis 26(1):19–34.
9 Epithelial–Mesenchymal Transition in Development and Diseases
209
Ma L, Teruya-Feldstein J, Weinberg RA (2007) Tumour invasion and metastasis initiated by microRNA-10b in breast cancer. Nature 449(7163):682–688. Mani SA, et al (2007) Mesenchyme Forkhead 1 (FOXC2) plays a key role in metastasis and is associated with aggressive basal-like breast cancers. Proc Natl Acad Sci U S A 104(24): 10069–10074. Mani SA, et al (2008) The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell 133(4):704–715. Mannel DN et al (1994) Mechanisms involved in metastasis enhanced by inflammatory mediators. Circ Shock 44(1):9–13. Mantovani A, et al (2008) Cancer-related inflammation. Nature 454(7203):436–444. Moody SE, et al (2005) The transcriptional repressor Snail promotes mammary tumor recurrence. Cancer Cell 8(3):197–209. Mumm JB, Oft M (2008) Cytokine-based transformation of immune surveillance into tumorpromoting inflammation. Oncogene 27(45):5913–5919. Nawshad A, et al (2005) Transforming growth factor-beta signaling during epithelial-mesenchymal transformation: implications for embryogenesis and tumor metastasis. Cells Tissues Organs 179(1–2):11–23. Neth P, et al (2007) The Wnt signal transduction pathway in stem cells and cancer cells: influence on cellular invasion. Stem Cell Rev 3(1):18–29. Nieto MA (2002) The snail superfamily of zinc-finger transcription factors. Nat Rev Mol Cell Biol 3(3):155–166. Oda H, Tsukita S, Takeichi M (1998) Dynamic behavior of the cadherin-based cell-cell adhesion system during Drosophila gastrulation. Dev Biol 203(2):435–450. Olive KP, et al (2009) Inhibition of Hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science 324(5933):1457–1461. Ozdamar B, et al (2005) Regulation of the polarity protein Par6 by TGFbeta receptors controls epithelial cell plasticity. Science 307(5715):1603–1609. Pantel K, Brakenhoff RH (2004) Dissecting the metastatic cascade. Nat Rev Cancer 4(6):448–456. Peinado H, et al (2005a) A molecular role for lysyl oxidase-like 2 enzyme in snail regulation and tumor progression. EMBO J 24(19):3446–3458. Peinado H, Portillo F, Cano A (2005b) Switching on-off Snail: LOXL2 versus GSK3beta. Cell Cycle 4(12):1749–1752. Pollard JW (2004) Tumour-educated macrophages promote tumour progression and metastasis. Nat Rev Cancer 4(1):71–78. Qian F, et al (2005) Interaction between integrin alpha(5) and fibronectin is required for metastasis of B16F10 melanoma cells. Biochem Biophys Res Commun 333(4):1269–1275. Radisky DC, et al (2005) Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature 436(7047):123–127. Radisky DC, Kenny PA, Bissell MJ (2007) Fibrosis and cancer: do myofibroblasts come also from epithelial cells via EMT? J Cell Biochem 101(4): 830–839. Rastaldi MP, et al (2002) Epithelial-mesenchymal transition of tubular epithelial cells in human renal biopsies. Kidney Int 62(1):137–146. Rosen EM, et al (1991) Tumor necrosis factor stimulates epithelial tumor cell motility. Cancer Res 51(19):5315–5321. Sahlgren C, et al (2008) Notch signaling mediates hypoxia-induced tumor cell migration and invasion. Proc Natl Acad Sci U S A 105(17):6392–6397. Salez F, et al (1998) Transforming growth factor-beta1 in sarcoidosis. Eur Respir J 12(4): 913–919. Savagner P, et al (2005) Developmental transcription factor slug is required for effective re-epithelialization by adult keratinocytes. J Cell Physiol 202(3):858–866. Shook D, Keller R (2003) Mechanisms, mechanics and function of epithelial-mesenchymal transitions in early development. Mech Dev 120(11):1351–1383. Sinha P, et al (2007) Cross-talk between myeloid-derived suppressor cells and macrophages subverts tumor immunity toward a type 2 response. J Immunol 179(2):977–983.
210
Y. Wu and B.P. Zhou
Stadler BM, Ruohola-Baker H (2008) Small RNAs: keeping stem cells in line. Cell 132(4): 563–566. Staller P, et al (2003) Chemokine receptor CXCR4 downregulated by von Hippel-Lindau tumour suppressor pVHL. Nature 425(6955):307–311. Stappenbeck TS, Miyoshi H (2009) The role of stromal stem cells in tissue regeneration and wound repair. Science 2009. 324(5935):1666–9. Stefani G, Slack FJ (2008) Small non-coding RNAs in animal development. Nat Rev Mol Cell Biol 9(3):219–230. Stemmer V, et al (2008) Snail promotes Wnt target gene expression and interacts with betacatenin. Oncogene 27(37):5075–5080. Stoelcker B, et al (1995) Role of adhesion molecules and platelets in TNF-induced adhesion of tumor cells to endothelial cells: implications for experimental metastasis. J Inflamm 46(3): 155–167. Studebaker AW, et al (2008) Fibroblasts isolated from common sites of breast cancer metastasis enhance cancer cell growth rates and invasiveness in an interleukin-6-dependent manner. Cancer Res 68(21):9087–9095. Sullivan NJ, et al (2009) Interleukin-6 induces an epithelial-mesenchymal transition phenotype in human breast cancer cells. Oncogene 28:2940–2947. Timmerman LA, et al (2004) Notch promotes epithelial-mesenchymal transition during cardiac development and oncogenic transformation. Genes Dev 18(1): p. 99–115. Valastyan S, et al (2009) A pleiotropically acting microRNA, miR-31, inhibits breast cancer metastasis. Cell 137(6):1032–1046. Valcourt U, et al (2005) TGF-beta and the Smad signaling pathway support transcriptomic reprogramming during epithelial-mesenchymal cell transition. Mol Biol Cell 16(4): 1987–2002. Vidal-Vanaclocha F, et al (2000) IL-18 regulates IL-1beta-dependent hepatic melanoma metastasis via vascular cell adhesion molecule-1. Proc Natl Acad Sci U S A 97(2):734–739. Villanueva S, et al (2002) Posteriorization by FGF, Wnt, and retinoic acid is required for neural crest induction. Dev Biol 241(2):289–301. Vincent T, et al (2009) A SNAIL1-SMAD3/4 transcriptional repressor complex promotes TGFbeta mediated epithelial-mesenchymal transition. Nat Cell Biol 11:943–950. Voronov E, et al (2003) IL-1 is required for tumor invasiveness and angiogenesis. Proc Natl Acad Sci U S A 100(5):2645–2650. Wang X, et al (2008) Downregulation of Par-3 expression and disruption of Par complex integrity by TGF-beta during the process of epithelial to mesenchymal transition in rat proximal epithelial cells. Biochim Biophys Acta 1782(1):51–59. Wang Z, et al (2009) Acquisition of epithelial-mesenchymal transition phenotype of gemcitabineresistant pancreatic cancer cells is linked with activation of the notch signaling pathway. Cancer Res 69(6):2400–2407. White LR, et al (2007) The characterization of alpha5-integrin expression on tubular epithelium during renal injury. Am J Physiol Renal Physiol 292(2):F567–F576. Willis BC, Borok Z (2007) TGF-beta-induced EMT: mechanisms and implications for fibrotic lung disease. Am J Physiol Lung Cell Mol Physiol 293(3):L525–L534. Wu Y, et al (2009) Stabilization of snail by NF-kappaB is required for inflammation-induced cell migration and invasion. Cancer Cell 15(5):416–428. Yamashita S, et al (2004) Zinc transporter LIVI controls epithelial-mesenchymal transition in zebrafish gastrula organizer. Nature 429(6989):298–302. Yang J, et al (2004) Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 117(7):927–939. Yang L, et al (2008) Abrogation of TGF beta signaling in mammary carcinomas recruits Gr-1+CD11b+ myeloid cells that promote metastasis. Cancer Cell 13(1):23–35. Yang MH, et al (2008) Direct regulation of TWIST by HIF-1alpha promotes metastasis. Nat Cell Biol 10(3):295–305.
9 Epithelial–Mesenchymal Transition in Development and Diseases
211
Yang X, et al (2009) Regulation of {beta}4-integrin expression by epigenetic modifications in the mammary gland and during the epithelial-to-mesenchymal transition. J Cell Sci 122(Pt 14): 2473–2480. Yook JI., et al (2005) Wnt-dependent regulation of the E-cadherin repressor snail. J Biol Chem 280(12):11740–11748. Yook JI, et al (2006) A Wnt-Axin2-GSK3beta cascade regulates Snail1 activity in breast cancer cells. Nat Cell Biol 8(12):1398–1406. Young MR, Lathers DM (1999) Myeloid progenitor cells mediate immune suppression in patients with head and neck cancers. Int J Immunopharmacol 21(4):241–252. Zavadil J, Bottinger EP (2005) TGF-beta and epithelial-to-mesenchymal transitions. Oncogene 24(37):5764–5774. Zavadil J, et al (2004) Integration of TGF-beta/Smad and Jagged1/Notch signalling in epithelialto-mesenchymal transition. EMBO J 23(5):1155–1165. Zavadil J, et al (2007) Transforming growth factor-beta and microRNA:mRNA regulatory networks in epithelial plasticity. Cells Tissues Organs 185(1–3):157–161. Zeisberg EM, et al (2007) Endothelial-to-mesenchymal transition contributes to cardiac fibrosis. Nat Med 13(8):952–961. Zeisberg M, et al (2003a) BMP-7 counteracts TGF-beta1-induced epithelial-to-mesenchymal transition and reverses chronic renal injury. Nat Med 9(7):964–968. Zeisberg M, et al (2003b) Bone morphogenic protein-7 inhibits progression of chronic renal fibrosis associated with two genetic mouse models. Am J Physiol Renal Physiol 285(6):F1060–F1067. Zeisberg M, Neilson EG (2009) Biomarkers for epithelial-mesenchymal transitions. J Clin Invest 119(6):1429–1437. Zhou BP, et al (2004) Dual regulation of Snail by GSK-3beta-mediated phosphorylation in control of epithelial-mesenchymal transition. Nat Cell Biol 6(10):931–940. Zhu S, et al (2007) MicroRNA-21 targets the tumor suppressor gene tropomyosin 1 (TPM1). J Biol Chem 282(19):14328–14336.
Chapter 10
Invasion and Metastasis Douglas M. Noonan, Giuseppina Pennesi, and Adriana Albini
Abstract In 1889, Stephen Paget recognized the key role that the local microenvironment must play in the formation of metastases by malignant tumor cells. Later, the interactions of the metastatic cell with the various barriers and within the final target organ were likened to a tumor cell decathlon, where the final metastasis was the result of the ability of a single tumor cell to perform series of steps needed to lodge and grow in a distant site. As the clinical oncologist finds metastatic disease much more difficult to manage than the primary tumor, these cells were considered to have attained additional mutations that confirmed a unique invasive and metastatic phenotype. Recent studies have challenged that hypothesis, thus what was once an accepted dogma is now on uncertain terms. The role of the tumor microenvironment has now come to the forefront in understanding the metastasis; here we review the mechanisms by which invasive cells may disseminate out of the local tumor microenvironment to colonize distant tissues.
Tumors as Tissues Numerous studies on the tumor microenvironment suggest that we need to revise the definition of the term “cancer”, currently defined in classical terms as a malignancy derived from hyperproliferating tumor cells, to that of a disease of a tissue. To control cancer, we need to regard carcinogenesis and tumors inherently as phenomena that occur within the whole tissue, not in individual cancer cells. In this new perspective, the microenvironment becomes an integral, essential part of the cancer and its metastasis. The tumor microenvironment is the complex society of many cell types, including the “stroma”, fibroblasts of various phenotypes,
A. Albini (*) Oncology Research, Science and Technology Park, IRCCS MultiMedica, Milan and Castellanza, Italy e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_10, © Springer Science+Business Media, LLC 2010
213
214
D.M. Noonan et al.
myofibroblasts, endothelial cells and their precursors, pericytes, smooth muscle cells, and immune cells: neutrophils and other granulocytes (eosinophils and basophils), mast cells, T, B and natural killer lymphocytes, and antigen-presenting cells such as macrophages and dendritic cells. All these cells can participate in tumor progression (Balkwill and Mantovani 2001; Brigati et al. 2002; Coussens and Werb 2002; De Visser et al. 2006; Hanahan and Weinberg 2000; Noonan et al. 2008; Orimo et al. 2005). If the process of carcinogenesis and its end result, invasive and metastatic cancer, are viewed as a maladaptive response of an entire tissue or organ to both genetic and epigenetic stress, then knowledge and control of the immediate microenvironment within a developing tumor become as important as the corresponding knowledge and control of the dysfunctional epithelial cells within that tumor (Albini and Sporn 2007; Balkwill and Mantovani 2001; Hanahan and Weinberg 2000).
Metastatic Disease Any clinical oncologist knows that, with current therapy, patients rarely die from their primary tumors. Benign tumors are easily cured by surgery, and a competent surgeon can treat most confined tumor masses. It is when surgery is not an option that cancer does become problematic. Local invasion, recurrence and, most dreaded, metastasis are also associated with resistance to chemotherapy, difficulty or inability in surgical management, and ensuing disease progression that is still often fatal. Traditionally, invasion and metastasis have been viewed from a tumor cell-centric aspect. The problem was considered to be the tumor cell itself; it was able to detach from the primary tumor, invade locally, access lymphatic or hematic vessels, and use these as a highway to distant sites (Fig. 10.1a). The role of the tumor microenvironment in favoring invasion and metastasis has returned into focus only more recently; it is now recognized as a key mechanism controlling invasion and metastasis, as well as a therapeutic and prevention target.
Metastatic Cascades The traditional view of the metastasis process was often likened to a tumor cell decathlon (Crissman 1986; Fidler 1978). The transformed cell had to acquire additional mutations that permitted it to perform all the various tasks associated with the road to malignancy at levels that were sufficient for each task. The end result was thought to be a malignant cell with extensive additional mutagenesis, often then associated with enhanced drug resistance as well. The decathlon largely consisted of discrete steps that could be assayed individually in vitro. These are (1) uncontrolled growth, (2) loss of cell–cell adhesion, (3) acquisition of migratory potential, (4) acquisition of invasive capacity, (5) exit from the primary tumor mass,
10 Invasion and Metastasis
a Classic
215
b Stromal
c Social
Tumor Cell
Tumor Cell
MMP producing Metastatic Tumor Cell or Cancer Stem Cell
MMP producing Stroma and Inflammatory Cells
2a
2b
2a
4
2a
4
Mobilized liz d Bone o e Marrow w Cells ell
1 2a
3
3
4
VEGF, GMCSF, IL-6…
2b
2b
3
5
Tumor Cell
1 2b
3
5
Proliferative Tumor T m r Cell C Migratory rat ry Tumor o Cell ll
1
1
d Remote Control
5
4
5
6
Fig. 10.1 Potential mechanisms of metastases: (a) Classic metastatic cascade. (b) The involvement of stromal tissues in facilitating invasion, typical of epithelial-derived tumors. (c) Social networks where highly proliferative but poorly invasive tumor cells cooperate with highly invasive but poorly proliferative tumor cells. (d) The involvement of generation of the pre-metastatic niche that facilitates tumor cell extravasation and organ colonization. In the four scenarios the primary tumors are proliferating, apoptosis-resistant cells. (1) The acquisition of the ability to detach from the primary tumor and migrate. Outlined in blue: cells producing matrix proteinases, in particular metalloprotinases, which permit tumor cells to breach the vessel basement membranes and enter into either the hematic (2a) or lymphatic (2b) systems, often moving as “cords” or emboli (3). (4) The tumor cells extravasate, again using proteinases (blue). (5) Tumor cells then proliferate at the distant site to form a metastasis. In some cases (5) the first cells to arrive are not tumor cells but local inflammation (b), specifically invasive but poorly proliferative tumor cells (c, 6) or the premetastatic niche is formed by bone marrow cells mobilized by products released from the primary tumor that act on bone-marrow-derived cells to form foci in specific organs (d). In many cases, tumor cells may extravasate but lack sufficient stimuli to form proliferating metastasis (6)
(6) breaching of the vascular basement membrane, (7) intravasation into a vessel, (8) survival of turbulence and immune surveillance in the circulatory system, (9) extravasation at a distant site to then (10) invade into the local tissue to reconstruct the uncontrolled tissue growth cycle (Fig. 10.1a). Each of these steps was considered to be a key point along the cascade and the ability of any one individual or clonal tumor cell to perform any one step could be measured, a deficiency at any point was to be considered a block of malignant progression. Assays were developed to measure each of these steps individually in vitro (Albini and Benelli 2007; Albini et al. 1987b), thus permitting analysis of mechanisms and testing of agents targeted the specific steps. These assays include (1) cell proliferation assays, (2) colony formation in soft agar to show resistance to anoikis, (3) adhesion assays to demonstrate the capacity to interact with specific, or complex, extracellular matrix components,
216
D.M. Noonan et al.
(4) migration assays to demonstrate motility, in particular directional motility (chemotaxis), and (5) invasion assays to demonstrate the acquisition of the ability to degrade and breach extracellular matrix barriers. These assays permitted assessment of the tumor cells’ capacity to perform the various tasks within the malignancy marathon. Further, this also permitted rapid development of agents targeting these individual steps that could potentially be employed in therapeutic approaches, theoretically preventing the oncology patient from developing metastatic disease. Recently, the infinite replication paradigm of tumor cells has been considered within the context of several cell death pathways, which sometimes counterbalance, but other times enhance malignancy. Resistance to apoptosis and apoptotic signals is a characteristic of most tumor cells (Hanahan and Weinberg 2000). In normal cells, when the replication machinery detects excessive or un-repairable errors in the genome, death by apoptosis is engaged; this system is frequently disengaged in tumor cells (Letai 2008). Tumor cells are also under constant cellular stress. Hypoxia is almost always present in tumors, and plays an important role in induction of vascularization as well as enhancing the plasticity of cancer cells (Pouyssegur et al. 2006). This is associated with alterations in metabolism (dependence on glycolysis) (Kroemer and Pouyssegur 2008), redox imbalances, and acidification of the microenvironment (Tosetti et al. 2009). Autophagy is one of the normal cellular reactions to adverse environments and lack of nutrients, the cell literally eats its own organelles to conserve energy and remove damaged organelles, it is also activated in cancer cells and is a potential therapeutic target (White and DiPaola 2009). A major focus has also been devoted to tumor cell resistance to anoikis, a sort of death by loneliness evoked when normal cells detach from their substrate (Simpson et al. 2008).
Migration, Invasion, and Metastasis Tumor cells adhere to many different extracellular matrices through numerous integrin receptors; a summary of this area is beyond the scope of this review, and will be dealt with in more detail in other chapters. We were among the first to report the enhanced chemotaxis of tumor cells (Mensing et al. 1984), as well as involvement of integrin engagement in migration (Albini et al. 1987a), concepts now widely accepted. A number of factors have been associated with tumor cell migration, from early “autocrine motility factors”, to extracellular matrix and coagulation cascade components and their fragments, to many growth factors, most of which also induce migration at even lower concentrations than that needed to induce cell proliferation (Albini and Benelli 2007). Finally, more recently the chemokines, a class of immune system migration factors, have also been implicated as key elements determining organ preferences in tumor metastasis (Balkwill 2004; Singh et al. 2007). Invasion is considered a key element in that it is a critical step in malignancy, has relative tumor cell specificity, and is potentially useful as a pharmacological
10 Invasion and Metastasis
217
target. Over 25 years ago, we developed one of the still most widely used invasion assays based on a tumor-derived matrix barrier (Matrigel or equivalents) within a porous membrane migration chamber (Albini et al. 1987b). One of the first molecular mechanisms postulated (Liotta et al. 1982), and later demonstrated (StetlerStevenson et al. 1989), was the existence of proteinases capable of degrading specific matrix barriers, such as the matrigel barrier, that would be associated with tumor cell invasion. Of these the matrix metalloproteinases (MMPs), a large family with over 20 members, and in particular those which are able to degrade type IV collagen – MMP2 and MMP9 – received a great deal of attention (Egeblad and Werb 2002), and will also be the focus of other chapters. Blockage of the MMPs with either endogenous inhibitors or, later, pharmacological MMP inhibitors demonstrated that this could drastically curb tumor cell invasiveness (Egeblad and Werb 2002). One of the “natural” inhibitors of MMPs and cell invasion is endogenous TIMP (Albini et al. 1991). These observations led to a campaign to develop MMP inhibitors for the clinic. Interestingly, it was precisely the MMPs that shed light on the role of the microenvironment in regulating tumor cell invasion and metastasis. The discovery that chemopreventive drugs, such as green tea flavonoids and N-acetyl-cystine (NAC), are metalloprotease inhibitors led to the initial suggestions of “angioprevention” mechanisms (Albini and Sporn 2007; Albini et al. 2005; Tosetti et al. 2002); see below. The extracellular matrix and the stromal cells provide tumor cells with the tensile scaffold necessary for appropriate assembly into three-dimensional macroscopic structures. In addition, it behaves as a reservoir of growth factors and immunomodulatory cytokines, for instance the numerous heparin-binding factors able to interact with heparan sulfate proteoglycans embedded within the matrix. So the matrix supplies to cancer cells both positional cues and signals to determine cellular fate, including growth, differentiation, survival, and movement (Egeblad and Werb 2002; Tasso et al. 2009). Several different scenarios of the metastatic cascade can be envisioned (Fig. 10.1), which will be discussed here at different points. The original is that of the “classic” metastatic cascade (Fig. 10.1a). As a small benign tumor expands, tumor cells acquire secondary “hits” or genetic alterations that improve the “fitness” of the cell for undergoing detachment from the primary mass, invasion, and metastatic dissemination (Fidler 1990). This cell must be able to (1) migrate and make matrix proteinases, in particular metalloprotinases, which permit it to (2) breach the vessel basement membranes and enter into either the lymphatic or hematic system. Recent studies have suggested that tumor cells migrate in “cords” (Friedl and Gilmour 2009; Friedl and Wolf 2008), and this “collective migration” may also contribute to the formation of emboli that contain coagulation factors, platelets, and other vehicles as well (2). The formation of these “emboli” may help the tumor cell to survive the turbulence found within the vascular system (3), as well as help hide it from eventual immune cells. Once in the distant site, the tumor cell adheres to the endothelium or an area of exposed vascular basement membrane, and reverses the intravasation process to extravasate into the target tissue (4). There the cell resumes proliferation to form a new tumor as a metastasis (5). Thus one could also assume that the metastasis,
218
D.M. Noonan et al.
c lonally derived from a highly invasively “fit” cell, would show a distinct molecular profile from that of the primary tumor, which in contrast was selected mostly on the basis of proliferative capacity alone. This also led to the hunt for genes that characterized metastatic cells, as well as the anti-metastatic genes that prevented cellular invasion and diffusion (Eccles and Welch 2007; Lee et al. 1996; Steeg et al. 1988). This scenario of a metastatic cascade seemed to fit particularly well to tumors derived from mesenchymal cells such as fibrosarcomas, as well as other tumors such as melanomas. Interestingly, these appear to use principally hematic metastasis routes. This also fit well with most cell lines, which may have undergone selection for survival in this processes in the in vitro and in vivo models used to measure invasiveness. The tumor-cell-centric view often overlooked the fact that tumors are complex tissues of numerous cell types closely interacting, the proliferating tumor cell is only one of these, and may not even be the dominant cell type within the tumor mass. Tumors are tissues that undergo constant remodeling, and as such have a chronic inflammatory component and activation of mesenchymal cell support systems resulting in an altered biology and tissue homeostasis. The contrasts with the tumor-cell-centric views first came when examining the metastatic scenario in the context of epithelial tumors, such as breast, ovarian, and intestinal tumors. One surprise was that although MMPs were clearly made within the epithelial-derived tumors, in situ hybridization showed that the cells containing high levels of mRNAs for MMPs were the stromal cells; fibroblasts and inflammatory cells within the tumors, rather than the tumor cells themselves (Autio-Harmainen et al. 1993; Soini et al. 1993). Further studies demonstrated that the tumor cells were able to induce MMP production by other cells within the tumor microenvironment, and that they could also capture the MMPs on their cell surfaces to use for invasion (Egeblad and Werb 2002). Thus the second scenario (Fig. 10.1b) is stromal dependent; these cells make the MMPs, in particular, MMP2 (primarily from fibroblasts) and MMP9 (primarily from inflammatory cells), which allow the tumor cells to acquire these on their cell surfaces and to enter the vasculature. Moreover, it has been demonstrated that breast cancer cells prompt mesenchymal cells within the tumor stroma to produce a cytokine, CCL5, which acts on cancer cells to promote their invasion and metastasis. Remarkably, these stimulated cancer cells do not acquire a stable metastatic phenotype; rather they revert to their pre-malignant state in the absence of other stimuli from the microenvironment (Karnoub et al. 2007). Interestingly, most epithelial-derived tumors show this pattern of stromal production of MMPs, and tend to metastasize via lymphatic routes. Unfortunately, the clinical trials of several MMP inhibitors did not do well in the clinic (Egeblad and Werb 2002). There were unexpected collateral effects (in particular, joint pain), and no benefit for the patients. This was likely due to several causes, including a much more complex biology than that anticipated (Egeblad and Werb 2002) and the lack of experience with combination therapy, the mode in which bevacizumab gained approval. A more detailed discussion of MMPs will be provided in other chapters. We also noted that the invasive tumor cell shares many of its properties with endothelial cells that are activated to enter into the angiogenic process (Fig. 10.2).
10 Invasion and Metastasis
219
Tumor Cells
Endothelial Cells
Proliferation Invasion n i
Proliferation rol e t on
Migration to
Basement se t Membrane Me a Target rg t Tissue u Morphogenesis orpho e i Differentiation
Chemokines Growth factors Altered matrix
VEGF, other Angiogenic factors Altered matrix
Fig. 10.2 Functional similarities in the processes of metastasis and angiogenesis
An endothelial cell in response to an angiogenic factor must undergo a series of steps that are similar to those utilized by metastatic cells (Fig. 10.3): (1) Entry into proliferation and migration toward the stimulus, both usually involving angiogenic growth factors such as VEGF and the relative tyrosine kinase receptor on the endothelial cell surface. (2) Adhesion to matrix components in the switch from a stable quiescent cell to a migratory cell, often involving the integrin avb3. (3) Activation of matrix degradation, mostly through release of MMP2. All these steps reflect those discussed for metastatic tumor cells (Figs. 10.2 and 10.3). In addition, endothelial cells undergo (4) reorganization and activation of a differentiation phase, and (5) recruitment of supporting cells (pericytes) returning to quiescence and eventual removal via apoptosis of collateral vessels in excess. Both tumor and endothelial cells can recruit fibroblasts and inflammatory cells as part of the invasive process (Figs. 10.2 and 10.3), and endothelial cells also take advantage of inflammation to promote angiogenesis (Albini et al. 2005; Balkwill and Mantovani 2001; De Visser et al. 2006; Noonan et al. 2008). Early on we found that the invasion assay we developed could be adapted to the assessment of induction of a migratory and invasive phenotype in endothelial cells (Thompson et al. 1991). This permitted the use of an in vitro system to both identify new angiogenic agents (for example, see Albini et al. 1995, 1994, 1996; Bussolino et al. 1995) and as a screen for angiogenesis inhibitors (for example, see Albini et al. 1999; Cai et al. 1999; Iurlaro et al. 1998; Valente et al. 1998). Angiogenesis also appears to be a factor linked to metastatic dissemination. In many cancer types, high vessel density is a negative prognostic factor. The effects of angiogenesis are likely to be multiple: (1) Angiogenesis facilitates tumor expansion, thus acquisition of new genetic/epigenetic events favoring invasive capacity. (2) Angiogenesis is closely linked with inflammation, improving leukocyte recruitment as an indicator
220
D.M. Noonan et al.
Fibroblasts
TK K ligands an s
MMP-2
MMP-2 MMP- 9
αβ β
αβ
Cy to
kin
es
MMP-9
FGF2
Neutrophils Macrophages
VEGF GF
Flg-1 Fl
PlGF P F
β αβ
VEGFR2
VEGFR1 E F
MMP-2 MMP- 9
MMPs
αVβ3
activated endothelial cell αVβ3
Fig. 10.3 Mechanistic similarities and interactions in tumor and endothelial cell invasion. Tumor cells produce a wide variety of autocrine and paracrine growth factors that stimulate tyrosine kinases and other cell surface receptors. Endothelial cells are activated in response to stimuli arising from tumor cells or inflammatory cells. Tumors can either produce matrix proteases, such as MMP2 and to a lesser extent, MMP9, or induce the production of these factors by surrounding fibroblasts or immune cells, respectively. Endothelial cells produce MMP2 and, to a lesser extent, MMP9, and may take advantage of the proteinases produced by inflammation as well. Activation of integrins on the tumor and endothelial cells permits directional migration toward the stimuli, resulting in breaching of cellular compartments
of an effective pro-tumorigenic “M2” inflammatory environment associated with malignancy (Albini et al. 2005; Balkwill and Mantovani 2001). (3) Angiogenesis is associated with highly increased vessel permeability, hemorrhages, and open channels that allow even minimally invasive cells’ ready access to entry into the vascular system. In some cases, tumor cells have been found to line vessels in what is known as “vascular mimicry” (Maniotis et al. 1999) and can even acquire endothelial-like functions and characteristics (Pezzolo et al. 2007). Angiogenesis and lymphangiogenesis are often associated, and similar events in lymphangiogenesis provide access to this vascular compartment as well (Alitalo et al. 2005). (4) Not only the factors associated with angiogenesis, in particular VEGF, but also hypoxia-derived signals and chemokines, released at high levels, systemically mobilize bone marrow components that may either directly participate in vessel formation within the tumors (Grunewald et al. 2006; Jin et al. 2006), or, as discussed below, directly favor formation of a metastatic niche. Counter-intuitively, however, recent studies suggest that at least in some experimental cases blocking VEGF through
10 Invasion and Metastasis
221
a nti-angiogenic therapy can promote invasion and metastasis (Loges et al. 2009), thus complicating therapeutic approaches. Another intriguing aspect is that of dormant metastases. Clinically in some tumors such as breast cancer, metastatic disease can manifest even decades after surgery and remission of the primary disease. What triggers these cells to awaken is not known. Studies have suggested that extravasation is in fact a relatively efficient process, whereas that of formation of metastatic foci is much less so (Cameron et al. 2000). Addition of an angiogenic stimuli has been found to convert dormant experimental tumors to actively growing cancer (Almog et al. 2009; Indraccolo et al. 2006; Moserle et al. 2009; Naumov et al. 2006a, b); inflammation may also be a trigger having similar effects.
Rethinking Metastasis As current theories of multistep carcinogenesis predicted that the metastatic cell was an evolution of the primary transformed cell to a genetically altered cell capable of performing the additional tasks associated with invasion and colonization, it was assumed that these cells would show a transcriptional profile distinct from that of the primary tumor. Further, it was also expected that since the hurdles along the metastatic decathlon were similar, the genetic responses would also be similar, thus metastases should show distinct profiles that were similar to each other. The use of microarrays to investigate these hypotheses in an unbiased manner rapidly created a crisis (Webb 2003). The transcriptome profiles of metastases were more similar to the primary tumor from which they were derived than to each other (Pomeroy et al. 2002; Singh et al. 2002; Van ’t Veer et al. 2002; Van De Vijver et al. 2002). One interpretation of this is that the tumors were simply “born” with metastatic potential; the transformation pathway of the cell led to acquisition of both tumorigenic potential and invasive/metastatic potential together (Webb 2003). This conclusion was in clear contrast to the established dogma of multistep carcinogenesis, and led to extensive rethinking and new investigations into if and how exactly to do tumors produce metastases (Lonneborg et al. 2009; Lonning et al. 2005; Pfeffer et al. 2003, 2009). Clinically, in some cases the available observations suggest that the “born metastatic” hypothesis could be active; many patients who present with a primary tumor often already have metastatic disease, even on occasion deriving from a relatively small tumor. On the other hand, in oncology the greatest benefit in clinical terms over recent years has been largely due to early detection through specific screening programs; clear examples are breast, prostate, colon, and melanoma. When detected early and removed, the patients have a much greater probability of disease-free survival, indicating that tumor removal indeed prevents tumor progression toward dissemination and metastasis. So what really is going on? One possible answer to the microarray dilemma was that by the time the tumor was detected clinically, it had already undergone much of the progression needed to become metastatic, thus the cells with metastatic potential could predominate in the
222
D.M. Noonan et al.
primary tumor when analyzed, therefore reflecting closely the transcriptome of the metastasis disseminated from it (Pfeffer et al. 2003, 2009). Recent investigations suggest that metastasis is likely to be much more complicated than a well-ordered decathlon. Among the key contributing factors may well be the genotype of the host itself (Hunter 2006; Hunter and Alsarraj 2009). Experimental studies using diverse murine strains have indicated that the host contributes extensively to the metastatic capacity derived from a specific mutational process (Hunter 2006; Hunter and Alsarraj 2009). This produces a theory that suggests that in one individual, a specific series of mutations may produce a primary tumor, but this tumor might lack sufficient interactions with the environment to disseminate. In another host, the same specific series of mutations instead produces a tumor that interacts with the microenvironment in a manner that favors metastatic dissemination. Potential points of interactions include local components such as fibroblasts and vascular cells, inflammation, immune skewing and immunosuppression. Other aspects may be systemic, involving the reaction of bone marrow components, dissemination of precursors, and systemic immune-suppression. In any case, the transformed primary tumors are similar in propensity for eventual metastases, while host components determine the metastatic destiny. Apart from the intense interactions within the constantly remodeling tumor tissue, there may be tumor cell subsets that show fitness for specific activities. The cancer “stem” cell, or “cancer initiating cell,” may be one of these (Gupta et al. 2009; Kakarala and Wicha 2008; Wicha et al. 2006). These cells may be either poorly differentiated cells that are able to remain in a more slowly proliferating stem cell niche which gives rise to the rapidly proliferating, partially differentiated cells that make up the majority of the tumor mass. Or they may represent the capacity of tumor cells to undergo reprogramming or “de-differentiation” toward more primitive precursor cell types. It is presumed either way that the successful metastasis must recruit these cells. One of the predictions of the multistep carcinogenesis model is that the “metastatic phenotype” might be a characteristic of only a very small subpopulation of the cells within the primary tumor and that the metastasizing cells acquired additional mutations that are not present in the primary tumor. These features would escape from detection by expression profiling (or any other genomic technique such as comparative genome hybridization on arrays) unless one analyzes each potentially unique cell population within a tumor, an essentially impossible enterprise. Recent theories suggest that meta stasis arises from cancer “stem cells” or “cancer-initiating cells” that might stay latent in the primary tumor and escape to form metastasis when local conditions induce their scattering, for instance in hypoxia. Once settled in their new location, they go back to giving rise to partially differentiated, highly replicative cells which may resemble that of the parental tumor. If we take cancer stem cells into account, the success of metastasis prediction by expression profiling of the primary tumor suggests that the primary and metastatic stem cells give rise to similar populations of partially differentiated tumor cells. The stem cell also appears to be particularly sensitive to the microenvironment. The tumor stroma is a facilitator of tumor progression, providing an environment in which stem cells can
10 Invasion and Metastasis
223
either stay quiescent or enter into asymmetric division, grow, and metastasize (Comoglio and Trusolino 2005). Another aspect is that tumors may build social networks. For example, what if one tumor cell type is not sufficient to produce a tumor? Different derivatives of tumor cells may interact with each other, and in some cases the diverse differentiation pathways taken during transformation may produce cells that do poorly alone but do well together (Bidard et al. 2008). In a simplified scenario as in Fig. 10.1c, we show two tumor cell types, one rapidly proliferating, capable of providing growth factors and a favorable environment also to the other cell type, which is specialized in invasion. The invasive cells may colonize tissues, but they would remain dormant in the absence of the proliferative cells. The proliferative cells enter the vasculature through the vessel alterations produced during the angiogenesis/ lymphangiogenesis process, where they can exit in areas preconditioned by the migratory/invasive cell subtype. Together they form a tumor colony of similar composition to the parental tumor. The ability of these two cells to meet in a distant site may be limited, potentially explaining the propensity of some tumors to produce metastatic disease from dormant micrometastases many years after removal of the primary tumor and extensive disease-free survival. Finally, recent experimental studies in vivo have suggested that tumors may send emissaries to prepare the ground for colonization (Kaplan et al. 2005; Steeg 2005). This concept has been named the “pre-metastatic niche” (Kaplan et al. 2006a, b; Wels et al. 2008), and is one area under intense investigation. It is based on the observation of mobilization of bone marrow cells by tumors that localized in the organs that will later be colonized by the tumor. These precursors could be found in the tumor target tissues very early, before the tumor cells themselves arrived, as well as with the tumor cells at later time within micrometastases. Further, the tissues colonized were specific to those targeted by the tumor cells, and cell-free supernatants could essentially reproduce a similar colonization by the premetastatic bone-marrow-derived cells. Thus the scenario becomes one of remote control (Fig. 10.1d); the tumor produces soluble factors that mobilizes bone marrow precursors and then colonizes the target tissue. These produce specific alterations that then promote tumor cell colonization. The specific alterations could include deposition of novel matrix components, degradation of matrix barriers to facilitate tumor cell entry, and production of chemoattractant/chemokine profiles that draw the tumor cell through chemotaxis. As cancer is a highly heterogeneous disease, the mechanisms leading to invasion and metastasis are likely to be as heterogeneous. If one could predict that all of the potential scenarios are clinically relevant, the task then becomes an attempt to identify which process may be occurring, and whether this can be circumvented therapeutically. Apart from early detection through screening programs, another potentially successful approach to reduction of metasatic disease could be application of prevention, in particular, angioprevention. Angioprevention is a term that refers to the tendency of chemoprevention agents to target angiogenesis (Tosetti et al. 2002), thus slowing tumor growth and progression. Bland agents with no side effects could be used in the general population before cancer appears; aspirin and
224
D.M. Noonan et al.
other COX inhibitors have shown clinical benefit in those using these agents (Cuzick et al. 2009; Harris 2009; Kraus and Arber 2009; Wang and Dubois 2009). Diet is one aspect that can be readily modified (Albini and Sporn 2007; Chambers 2009) in this population (Berrino et al. 2007; Coleman et al. 2008). Somewhat more aggressive and efficacious agents may be warranted for higher risk populations, taking into account the risk–risk balance to keep patient benefit in mind (Albini and Sporn 2007). Finally, in oncologic patients, in remission following therapy, antiangiogenesis and anti-tumor approaches may be efficacious (Albini and Sporn 2007). However, as noted above, single target anti-angiogenic agents may in fact favor metastasis (Loges et al. 2009), although the effects on dormant disease remain to be determined. One simple therapeutic approach that is often overlooked, largely because of limited economic incentive for large clinical studies, is metronomic therapy (Kerbel 2007; Kerbel and Kamen 2004). Metronomic therapy, the use of low dose, high-frequency chemotherapy to target normal cell populations (endothelial cells, inflammation) within the tumor microenvironment (Browder et al. 2000; Kerbel and Folkman 2002), is advancing in phase II trials where it seems to be effective at prolonging disease-free survival while being well tolerated (for example, see Briasoulis et al. 2009; Brizzi et al. 2009; Dellapasqua et al. 2008; GarciaSaenz et al. 2008). It is also relatively inexpensive and would act as a broad anti-angiogenic approach, potentially circumventing the pitfalls of single agent targeting. Combination of metronomic therapy with anti-inflammatory agents would also be of interest; however, substantial funding is needed for clinical trials to demonstrate clear benefit in those patients in remission following primary therapy.
Conclusions The process of metastasis remains an enigma of great clinical importance. Resolving this enigma would provide a key stride in prevention and management of cancer. Our knowledge of the mechanisms involved seems to have expanded faster than our capacity to comprehend fully the information available. However, the increasingly open-minded approach, and the realization that numerous pathways may exist and may even interact between them (some are summarized in Fig. 10.1) in reality holds great promise. The realization of the involvement of the tumor microenvironment has provided further stimuli in understanding the complex and multiple mechanisms. Novel approaches with transcriptome and proteome analyses, including microRNAs, and host genetics as well as more refined biological approaches will provide windows into how cancer cells spread, how to evaluate these clinically, and what may be the combined approaches to block the process, even before it starts. Acknowledgments These studies were funded by the AIRC (Associazione Italiana per la Ricerca sul Cancro), the Ministero della Salute, the Compagnia di San Paolo, the Ministero dell’Università e della Ricerca, and the Università degli Studi dell’Insubria.
10 Invasion and Metastasis
225
References Albini A, Benelli R (2007) The chemoinvasion assay: a method to assess tumor and endothelial cell invasion and its modulation. Nat Protocols 2:504–511 Albini A, Sporn MB (2007) The tumour microenvironment as a target for chemoprevention. Nat Rev Cancer 7:139–147 Albini A, Allavena G, Melchiori A et al. (1987a) Chemotaxis of 3t3 and sv3t3 cells to fibronectin is mediated through the cell attachment site in fibronectin and a fibronectin cell surface receptor. J Cell Biol 105:1867–1872 Albini A, Iwamoto Y, Kleinman HK et al. (1987b) A rapid in vitro assay for quantitating the invasive potential of tumor cells. Cancer Res 47:3239–3245 Albini A, Melchiori A, Santi L et al. (1991) Tumor cell invasion inhibited by timp-2. J Natl Cancer Inst 83:775–779 Albini A, Fontanini G, Masiello L et al. (1994) Angiogenic potential in vivo by kaposi sarcoma cell-free supernatants and hiv1-tat product: inhibition of ks-like lesions by timp-2. AIDS 8:1237–1244 Albini A, Barillari G, Benelli R et al. (1995) Angiogenic properties of human immunodeficiency virus type 1 tat protein. Proc Natl Acad Sci U S A 92:4838–4842 Albini A, Soldi R, Giunciuglio D et al. (1996) The angiogenesis induced by hiv-1 tat is mediated by the flk-1/kdr receptor on vascular endothelial cells. Nat Med 2:1371–1375 Albini A, Florio T, Giunciuglio D et al. (1999) Somatostatin controls kaposi’s sarcoma tumor growth through inhibition of angiogenesis. FASEB J 13:647–655 Albini A, Tosetti F, Benelli R et al. (2005) Tumor inflammatory angiogenesis and its chemoprevention. Cancer Res 65:10637–10641 Alitalo K, Tammela T, Petrova TV (2005) Lymphangiogenesis in development and human disease. Nature 438:946–953 Almog N, Ma L, Raychowdhury R et al. (2009) Transcriptional switch of dormant tumors to fastgrowing angiogenic phenotype. Cancer Res 69:836–844 Autio-Harmainen H, Karttunen T, Hurskainen T et al. (1993) Expression of 72 kilodalton type IV collagenase (gelatinase a) in benign and malignant ovarian tumors. Lab Invest 69:312–321 Balkwill F (2004) Cancer and the chemokine network. Nat Rev Cancer 4:540–550 Balkwill F, Mantovani A (2001) Inflammation and cancer: back to virchow? Lancet 357:539–545 Berrino F, De Angelis R, Sant M et al. (2007) Survival for eight major cancers and all cancers combined for european adults diagnosed in 1995–99: results of the Eurocare-4 Study. Lancet Oncol 8:773–783 Bidard FC, Pierga JY, Vincent-Salomon A et al. (2008) A “class action” against the microenvironment: do cancer cells cooperate in metastasis? Cancer Metastasis Rev 27:5–10 Briasoulis E, Pappas P, Puozzo C et al. (2009) Dose-ranging study of metronomic oral vinorelbine in patients with advanced refractory cancer. Clin Cancer Res 15:6454–6461 Brigati C, Noonan DM, Albini A et al. (2002) Tumors and inflammatory infiltrates: friends or foes? Clin Exp Metastasis 19:247–258 Brizzi MP, Berruti A, Ferrero A et al. (2009) Continuous 5-fluorouracil infusion plus long acting octreotide in advanced well-differentiated neuroendocrine carcinomas. A phase II trial of the piemonte oncology network. BMC Cancer 9:388 Browder T, Butterfield CE, Kraling BM et al. (2000) Antiangiogenic scheduling of chemotherapy improves efficacy against experimental drug-resistant cancer. Cancer Res 60:1878–1886 Bussolino F, Arese M, Montrucchio G et al. (1995) Platelet activating factor produced in vitro by kaposi’s sarcoma cells induces and sustains in vivo angiogenesis. J Clin Invest 96:940–952 Cai T, Fassina G, Morini M et al. (1999) N-acetyl cysteine inhibits endothelial cell invasion and angiogenesis. Lab Invest 79:1151–1159 Cameron MD, Schmidt EE, Kerkvliet N et al. (2000) Temporal progression of metastasis in lung: cell survival, dormancy, and location dependence of metastatic inefficiency. Cancer Res 60:2541–2546
226
D.M. Noonan et al.
Chambers AF (2009) Influence of diet on metastasis and tumor dormancy. Clin Exp Metastasis 26:61–66 Coleman MP, Quaresma M, Berrino F et al. (2008) Cancer survival in five continents: a worldwide population-based study (concord). Lancet Oncol 9:730–756 Comoglio PM, Trusolino L (2005) Cancer: the matrix is now in control. Nat Med 11:1156–1159. Coussens LM, Werb Z (2002) Inflammation and cancer. Nature 420:860–867 Crissman JD (1986) Tumor-host interactions as prognostic factors in the histologic assessment of carcinomas. Pathol Annu 21(Pt 1):29–52 Cuzick J, Otto F, Baron JA et al. (2009) Aspirin and non-steroidal anti-inflammatory drugs for cancer prevention: an international consensus statement. Lancet Oncol 10:501–507 De Visser KE, Eichten A, Coussens LM (2006) Paradoxical roles of the immune system during cancer development. Nat Rev Cancer 6:24–37 Dellapasqua S, Bertolini F, Bagnardi V et al. (2008) Metronomic cyclophosphamide and capecitabine combined with bevacizumab in advanced breast cancer. J Clin Oncol 26:4899–4905 Eccles SA, Welch DR (2007) Metastasis: recent discoveries and novel treatment strategies. Lancet 369:1742–1757 Egeblad M, Werb Z (2002) New functions for the matrix metalloproteinases in cancer progression. Nat Rev Cancer 2:161–174 Fidler I (1978) Tumor heterogeneity and the biology of cancer invasion and metastasis. Cancer Res 38:2651–2660 Fidler IJ (1990) Critical factors in the biology of human cancer metastases: 28th G. H. Clowes Memorial Award lecture. Cancer Res 50:613–618 Friedl P, Gilmour D (2009) Collective cell migration in morphogenesis, regeneration and cancer. Nat Rev Mol Cell Biol 10:445–457 Friedl P, Wolf K (2008) Tube travel: the role of proteases in individual and collective cancer cell invasion. Cancer Res 68:7247–7249 Garcia-Saenz JA, Martin M, Calles A et al. (2008) Bevacizumab in combination with metronomic chemotherapy in patients with anthracycline- and taxane-refractory breast cancer. J Chemother 20:632–639 Grunewald M, Avraham I, Dor Y et al. (2006) Vegf-induced adult neovascularization: recruitment, retention, and role of accessory cells. Cell 124:175–189 Gupta PB, Chaffer CL, Weinberg RA (2009) Cancer stem cells: mirage or reality? Nat Med 15:1010–1012 Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57–70 Harris RE (2009) Cyclooxygenase-2 (cox-2) blockade in the chemoprevention of cancers of the colon, breast, prostate, and lung. Inflammopharmacology 17:55–67 Hunter K (2006) Host genetics influence tumour metastasis. Nat Rev Cancer 6:141–146 Hunter KW, Alsarraj J (2009) Gene expression profiles and breast cancer metastasis: a genetic perspective. Clin Exp Metastasis 26:497–503 Indraccolo S, Stievano L, Minuzzo S et al. (2006) Interruption of tumor dormancy by a transient angiogenic burst within the tumor microenvironment. Proc Natl Acad Sci U S A 103:4216–4221 Iurlaro M, Benelli R, Masiello L et al. (1998) Beta interferon inhibits HIV-1 TAT induced angiogenesis; synergism with 13-cis retinoic acid. Eur J Cancer 34:570–576 Jin DK, Shido K, Kopp HG et al. (2006) Cytokine-mediated deployment of SDF-1 induces revascularization through recruitment of CXCR4+ hemangiocytes. Nat Med 12:557–567 Kakarala M, Wicha MS (2008) Implications of the cancer stem-cell hypothesis for breast cancer prevention and therapy. J Clin Oncol 26:2813–2820 Kaplan RN, Riba RD, Zacharoulis S et al. (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438:820–827 Kaplan RN, Psaila B, Lyden D (2006a) Bone marrow cells in the ‘pre-metastatic niche’: within bone and beyond. Cancer Metastasis Rev 25:521–529 Kaplan RN, Rafii S, Lyden D (2006b) Preparing the “Soil”: the premetastatic niche. Cancer Res 66:11089–11093
10 Invasion and Metastasis
227
Karnoub AE, Dash AB, Vo AP et al. (2007) Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 449:557–563 Kerbel R, Folkman J (2002) Clinical translation of angiogenesis inhibitors. Nat Rev Cancer 2:727–739 Kerbel RS (2007) Improving conventional or low dose metronomic chemotherapy with targeted antiangiogenic drugs. Cancer Res Treat 39:150–159 Kerbel RS, Kamen BA (2004) The anti-angiogenic basis of metronomic chemotherapy. Nat Rev Cancer 4:423–436 Kraus S, Arber N (2009) Inflammation and colorectal cancer. Curr Opin Pharmacol 9:405–410 Kroemer G, Pouyssegur J (2008) Tumor cell metabolism: cancer’s achilles’ heel. Cancer Cell 13:472–482 Lee JH, Miele ME, Hicks DJ et al. (1996) Kiss-1, a novel human malignant melanoma metastasissuppressor gene. J Natl Cancer Inst 88:1731–1737 Letai AG (2008) Diagnosing and exploiting cancer’s addiction to blocks in apoptosis. Nat Rev Cancer 8:121–132 Liotta LA, Thorgeirsson UP, Garbisa S (1982) Role of collagenases in tumor cell invasion. Cancer Met Rev 1:277 Loges S, Mazzone M, Hohensinner P et al. (2009) Silencing or fueling metastasis with VEGF inhibitors: antiangiogenesis revisited. Cancer Cell 15:167–170 Lonneborg A, Aaroe J, Dumeaux V et al. (2009) Found in transcription: gene expression and other novel blood biomarkers for the early detection of breast cancer. Expert Rev Anticancer Ther 9:1115–1123 Lonning PE, Sorlie T, Borresen-Dale AL (2005) Genomics in breast cancer-therapeutic implications. Nat Clin Pract Oncol 2:26–33 Maniotis AJ, Folberg R, Hess A et al. (1999) Vascular channel formation by human melanoma cells in vivo and in vitro: vasculogenic mimicry. Am J Pathol 155:739–752 Mensing H, Albini A, Krieg T et al. (1984) Enhanced chemotaxis of tumor-derived and virustransformed cells to fibronectin and fibroblast-conditioned medium. Int J Cancer 33:43–48 Moserle L, Amadori A, Indraccolo S (2009) The angiogenic switch: implications in the regulation of tumor dormancy. Curr Mol Med 9:935–941 Naumov GN, Akslen LA, Folkman J (2006a) Role of angiogenesis in human tumor dormancy: animal models of the angiogenic switch. Cell cycle 5:1779–1787 Naumov GN, Bender E, Zurakowski D et al. (2006b) A model of human tumor dormancy: an angiogenic switch from the nonangiogenic phenotype. J Natl Cancer Inst 98:316–325 Noonan DM, De Lerma Barbaro A, Vannini N et al. (2008) Inflammation, inflammatory cells and angiogenesis: decisions and indecisions. Cancer Metastasis Rev 27:31–40 Orimo A, Gupta PB, Sgroi DC et al. (2005) Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated sdf-1/cxcl12 secretion. Cell 121:335–348 Pezzolo A, Parodi F, Corrias MV et al. (2007) Tumor origin of endothelial cells in human neuroblastoma. J Clin Oncol 25:376–383 Pfeffer U, Noonan D, Albini A (2003) Re: Microarray studies challenge theories of metastasis. J Natl Cancer Inst 95:829 Pfeffer U, Romeo F, Noonan DM et al. (2009) Prediction of breast cancer metastasis by genomic profiling: where do we stand? Clin Exp Metastasis 26:547–558 Pomeroy SL, Tamayo P, Gaasenbeek M et al. (2002) Prediction of central nervous system embryonal tumour outcome based on gene expression. Nature 415:436–442 Pouyssegur J, Dayan F, Mazure NM (2006) Hypoxia signalling in cancer and approaches to enforce tumour regression. Nature 441:437–443 Simpson CD, Anyiwe K, Schimmer AD (2008) Anoikis resistance and tumor metastasis. Cancer Lett 272:177–185 Singh D, Febbo PG, Ross K et al. (2002) Gene expression correlates of clinical prostate cancer behavior. Cancer Cell 1:203–209 Singh S, Sadanandam A, Singh RK (2007) Chemokines in tumor angiogenesis and metastasis. Cancer Metastasis Rev 26:453–467
228
D.M. Noonan et al.
Soini Y, Paakko P, Autio-Harmainen H (1993) Genes of laminin b1 chain, alpha 1 (IV) chain of type IV collagen, and 72-kD type IV collagenase are mainly expressed by the stromal cells of lung carcinomas. Am J Pathol 142:1622–1630 Steeg PS (2005) Cancer biology: emissaries set up new sites. Nature 438:750–751 Steeg PS, Bevilacqua G, Kopper L et al. (1988) Evidence for a novel gene associated with low tumor metastatic potential. J Natl Cancer Inst 80:200–204 Stetler-Stevenson W, Krutzsch H, Wacher M et al. (1989) The activation of human type IV collagenase proenzyme. J Biol Chem 264:1353–1356 Tasso R, Augello A, Carida M et al. (2009) Development of sarcomas in mice implanted with mesenchymal stem cells seeded onto bioscaffolds. Carcinogenesis 30:150–157 Thompson E, Nakamura S, Shima T et al. (1991) Supernatants of acquired immunodeficiency syndrome-related kaposi’s sarcoma cells induce endothelial cell chemotaxis and invasiveness. Cancer Res 51:2670–2676 Tosetti F, Ferrari N, De Flora S et al. (2002) Angioprevention’: angiogenesis is a common and key target for cancer chemopreventive agents. FASEB J 16:2–14 Tosetti F, Noonan DM, Albini A (2009) Metabolic regulation and redox activity as mechanisms for angioprevention by dietary phytochemicals. Int J Cancer 125:1997–2003 Valente P, Fassina G, Melchori A et al. (1998) Timp-2 over-expression reduces invasion and angiogenesis and protects b16f10 melanoma cells from apoptosis. Int J Cancer 75:246–253 Van ’T Veer LJ, Dai H, Van De Vijver MJ et al. (2002) Gene expression profiling predicts clinical outcome of breast cancer. Nature 415:530–536 Van De Vijver MJ, He YD, Van’t Veer LJ et al. (2002) A gene-expression signature as a predictor of survival in breast cancer. N Engl J Med 347:1999–2009 Wang D, Dubois RN (2010) The role of cox-2 in intestinal inflammation and colorectal cancer. Oncogene 29:781–788 Webb T (2003) Microarray studies challenge theories of metastasis. J Natl Cancer Inst 95:350–351 Wels J, Kaplan RN, Rafii S et al. (2008) Migratory neighbors and distant invaders: tumor-associated niche cells. Genes Dev 22:559–574 White E, Dipaola RS (2009) The double-edged sword of autophagy modulation in cancer. Clin Cancer Res 15:5308–5316 Wicha MS, Liu S, Dontu G (2006) Cancer stem cells: an old idea – a paradigm shift. Cancer Res 66:1883–1890; discussion 1895–1886
Chapter 11
Dormancy of Disseminated Tumor Cells: Reciprocal Crosstalk with the Microenvironment Paloma Bragado, Aparna C. Ranganathan, and Julio A. Aguirre-Ghiso
Abstract Majority of cancer patients will die of metastases that develop from disseminated tumor cells (DTCs), months, years, or even decades after treatment. This pause in cancer progression suggests that, during these disease-free periods, DTCs may stop proliferating and survive in a dormant state. The mechanisms that determine whether tumor cells, after disseminating to target organs, will continue to proliferate, die or enter a protracted state of dormancy are poorly understood. Here, we review the different manifestations of dormancy and the experimental and clinical evidence supporting that the target organ microenvironment where DTCs lodge might influence the choice to enter dormancy. We also review the available animal models to study DTCs dormancy. This information is important to design strategies to maintain dormancy of DTCs or eradicate DTCs before they progress to overt metastasis. Such information would lead to anti-metastatic and/or metastasis preventive therapies which are urgently needed.
General Concepts on Tumor Cell Dormancy in the Context of Cancer Progression One of the hallmarks of cancer progression is the ability of tumor cells to move out of the primary site, invade adjacent tissues, and travel to distant sites to form discrete metastasis. Despite advances in early detection and diagnosis of cancer, as well as significant improvements in surgical techniques, general patient care and
J.A. Aguirre-Ghiso (*) Division of Hematology and Oncology, Departments of Medicine and Department of Otolaryngology, Tisch Cancer Institute, Mount Sinai School of Medicine, One Gustave L Levy Place, Box 1079, New York, NY 10029, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_11, © Springer Science+Business Media, LLC 2010
229
230
P. Bragado et al.
local and adjuvant therapies, mortality rates from cancer are still high. Particularly, almost 90% of all cancer-related deaths are due to distant metastases (Slade et al. 2005; Weigelt et al. 2005). In traditional models of metastases, metastatic cells are extremely rare, and are thought to arise only during later stages of tumor progression when the primary tumor cells are deemed “metastasis capable” having accumulated all the characteristics (genetic and epigenetic changes) required to complete the metastatic process (Fidler 2003). However, this model has been challenged by recent results from a number of laboratories that have shown that cells with metastatic potential can disseminate early from the primary tumor, even in a premalignant stage, suggesting that dissemination and progression to metastatic growth can occur in parallel to the primary tumor expansion. Accordingly, Kang et al. (2003) performed a global gene expression analysis of primary breast cancer and identified a group of genes that predicted for the development of distant metastases. These studies suggest that the genes and therefore the ability to metastasize are already present in sufficient abundance to be detected by gene arrays in primary tumors. In addition to this, recent studies have shown that oncogene activation in mouse mammary cells triggered the activation of twist, and in other studies this gene can induce dissemination of pre-invasive mammary tumor cells to lungs and bone marrow prior to the appearance of overt primary mammary tumors (Ansieau et al. 2008a, b; Husemann et al. 2008). The consequences of this new view of the metastatic cascade are that dissemination is no longer regarded as a selected mutational event and it is indeed considered an early event during cancer progression (Husemann et al. 2008). However, this does not negate that more progressed tumors will have disseminating cells and perhaps with enhanced efficiency. In fact 20–45% of patients with breast or prostate cancer will relapse years or even decades after treatment of the primary tumor, even with advanced tumors such as stage 3 breast cancer. The early dissemination theory is supported by findings from the Klein laboratory showing that DTCs can be detected in patients without metastasis and with lesions considered non-invasive (i.e., ADH, DCIS). These cells have fewer chromosomal aberrations than primary tumors or DTCs from patients with metastasis (Schmidt-Kittler et al. 2003). This suggests that in breast cancer, DTCs acquire the genomic aberrations typical of metastatic growth after having disseminated. Supporting evidence came from studies in non-invasive atypical ductal hyperplasia (ADH) and ductal carcinoma in situ (DCIS), which are considered to be ‘early’ in progression, where dissemination has already occurred (Klein 2008; Stoecklein et al. 2008). Why is this important for dormancy? These studies suggest that a part of the dormancy time is the time it takes for these DTCs to accumulate genetic alterations at distant sites and eventually give rise to metastasis. The study of the functional properties of these cells in vitro, along with their molecular analysis, can provide valuable information about the mechanisms regulating dormancy of these cells. Solakoglu et al. (2002) study chromosomal aberrations in BM-DTCs isolated from patients with breast cancer, and found that the genetic aberrations were different among different patients. This suggests that there is a high genetic heterogenicity in BM-DTCs of patients. This is supported by studies from Klein laboratory where they studied the genetic variation of disseminated
11 Dormancy of Disseminated Tumor Cells
231
tumor cells. They screened 525 bone-marrow, blood, and lymph-node samples from 474 patients with breast, prostate, and gastrointestinal cancers. What they found is that irrespective of the cancer type, there was a huge genetic difference between DTCs from MRD patients and DTC from patients with metastasis (Klein 2002; Klein et al. 2002). What is the clinical relevance of DTCs? As mentioned earlier, evidences suggest that dissemination to distant sites occurs even before primary tumor diagnosis (Husemann et al. 2008). Due to their scarcity, these disseminated tumor cells (DTCs) usually remain undetected, however, sensitive immunocytochemical techniques and molecular assays have permitted specific detection of occult disseminated tumor cells even at the single cell stage (Pantel et al. 2009; Wikman et al. 2008). These technologies allow monitoring tumor cell dissemination to the blood and bone marrow, two sites that can be easily sampled. The detection of these cells is associated with poor prognosis, and is predictive of metastases to the lung, liver, and BM (Pantel et al. 2009). Various clinical studies have given evidence that there is an association between the presence of DTC at the time of primary tumor surgery and the metastatic relapse in patients with breast, prostrate, lung, and gastrointestinal cancer among others (Alix-Panabieres et al. 2005; Bidard et al. 2008; Braun and Naume 2005). It has been proposed that these DTCs that constitute the minimal residual disease (MRD) may remain in a dormant state over the years until the conditions are propitious to resume proliferation (Aguirre-Ghiso 2007; Pantel et al. 2009; Pantel and Brakenhoff 2004). In fact, after diagnosis of the primary lesions, patients spend more time harboring residual disease than expanding lesions (metastases once overt usually grow rapidly) (Fig. 11.1). This highlights the importance of understanding the biology of residual disease, as it is an unmet opportunity to prevent the development of metastasis. Studying dormancy and residual disease is in its deepest sense the clearer avenue to the development of metastasis preventive strategies. Thus, maintaining dormancy of DTC or more simply eradicating these dormant tumor cells would represent a significant clinical breakthrough. Tumor dormancy can be broadly separated into two categories: One type is the dormancy of micrometastasis and therefore of a tumor mass and the other is the dormancy of solitary disseminated tumor cells. Detailed descriptions of these types of dormancy have been thoroughly reviewed elsewhere (Aguirre-Ghiso 2007). Here, we provide a brief updated overview.
Dormancy of Micrometastasis Angiogenic Dormancy This type of dormancy also known as tumor mass dormancy is defined by a balance between proliferation and apoptosis of micrometastasis. As a result there is no net gain in tumor mass. But this does not imply that the tumor cells are non-proliferative, in fact it is the opposite. One such mechanism, known as pre-angiogenic dormancy,
232
P. Bragado et al.
Fig. 11.1 Tumor dormancy during cancer progression. Different manifestations of dormancy. Tumor cells that have accumulated genetic and epigenetic changes that provide a growth advantage will form a primary tumor (a). These cells will invade the surrounding tissue, intravasate either in the blood or lymphatic vessels, and disseminate via circulation to distant sites (a – bottom). Tumor cells that survive dissemination lodge in the target organs (e.g., BM, liver, lungs). After treatment, the primary tumor will regress and/or be removed; however, DTCs will still be detectable in the BM of patients during long periods of time and they constitute the minimal residual disease (b, c). The crosstalk between the DTCs and the new microenvironment will determine the fate of the DCTs. They can either die from apoptosis (d – red), enter dormancy, either as solitary DTCs (e) or as a micrometastatic lesion (f ), where there is a balance between proliferation
11 Dormancy of Disseminated Tumor Cells
233
occurs because of the inability of the cells to recruit new blood vessels (Fig. 11.1). Studies from Folkman’s and other laboratories have indicated that for primary tumors to grow beyond a size of 1–2 mm, they require the recruitment of new blood vessels via angiogenesis. This mechanism is known as the “angiogenic switch” (Folkman 2002; Gimbrone et al. 1972; Holmgren et al. 1995). However, in some tumors sometimes proliferation competent tumor cells may be incapable of recruiting new blood vessels resulting in a balance between the fraction of dying cells and dividing ones. Under these circumstances, these avascular micrometastases remain viable but with no net increase in tumor mass resulting now in the entire tumor mass becoming dormant. Studies have demonstrated that such angiogenic dormancy can result from a decrease in the expression of pro-angiogenic factors such as VEGF and PDGF (platelet-derived growth factor) and/or an increase in the expression of angiogenic inhibitors such as thrombspondin, endostatin, etc. (Folkman 2002; Folkman and Hanahan 1991; Naumov et al. 2006). However such pre-angiogenic micrometastases can emerge from dormancy if they acquire the ability to become vascularized either by down-regulating angiogenesis inhibitors or by upregulating proangiogenic factors (Fig. 11.1) (Folkman 2002; Folkman and Hanahan 1991). Alternatively, this loss of dormancy can also be induced by activating or inactivating pathways that promote or suppress angiogenesis. For instance, tumor suppressors such as p53, PTEN, and Smad4 were shown to inhibit angiogenesis by regulating thrombospondin synthesis (Dameron et al. 1994a, b; Schwarte-Waldhoff et al. 2000; Su et al. 2003; Volpert et al. 1997; Wen et al. 2001). Furthermore, expression of activated oncogene c-Ha-Ras in osteosarcoma and gastric cancer cells
Fig. 11.1 (continued) (pink) and apoptosis (red) and thus no net gain in tumor mass, or they can resume proliferation and give rise to a secondary tumor. (e) Cellular dormancy represents dormancy of single disseminated tumor cells, which enter a growth arrest (G0/G1) upon arriving to a new microenvironment. The majority of DTCs studies show that these cells lack markers of proliferation such as Ki67, suggesting a deep G0–G1 arrest. Other factors that can also define prolonged quiescence and dormancy. For example, HES1, a transcription factor that appears to be required for induction of quiescence but prevents the induction of senescence and/or differentiation in several cellular systems, could favor cellular dormancy. Alternatively, the activation of genes known as metastasis suppressor genes, could contribute to dormancy. For instance, KISS1 has been shown to inhibit metastasis by inducing cellular dormancy. (g) Angiogenic dormancy occurs because proliferation competent DTCs are not able to recruit new blood vessels which leads to a balance between proliferating (pink) and dying (red) cells, and as a result the entire tumor mass becomes dormant. Angiogenic dormancy can result from a decrease in the expression of pro-angiogenic factors such as VEGF and PDGF (platelet-derived growth factor) and/or an increase in the expression of anti-angiogenic factors such as thrombospondin. However, these micrometastases can undergo an angiogenic switch and emerge from dormancy if they acquire the ability to become vascularized either by down-regulating angiogenesis inhibitors or by upregulating proangiogenic factors. (h) Immunesurveillance prevents the expansion of proliferating DTCs cells via cytotoxic T-cell-mediated cell death, thereby maintaining a constant tumor mass. In order to escape this state of dormancy, DTCs can downregulate specific tumor-associated antigens or express co-stimulatory molecules that induce apoptosis of cytotoxic CD8+ T lymphocytes (Zhang et al. 2008)
234
P. Bragado et al.
induced the loss of angiogenic dormancy in otherwise non-angiogenic cell lines (Arbiser et al. 1997; Udagawa et al. 2002). This loss of dormant mass phenotype correlated with a Ras-induced VEGF expression (Rak et al. 1995). Indirectly Ras has also been shown to stimulate angiogenesis by down-regulating angiogenesis inhibitors such as thrombospondin (Sheibani and Frazier 1996; Watnick et al. 2003) While many of the above studies were carried out on primary tumor masses, it is unclear whether angiogenic dormancy of micrometastasis occurs in patients as clinical evidence for this is weak. More importantly, if such a state of dynamic proliferation and apoptosis can be sustained by one or multiple small micrometastatic masses for years or decades has not been proven. It is tempting to speculate that given the high degree of genetic heterogeneity and the epigenetic changes that these dividing cells might experience, the angiogenic switch should occur rather quickly. Alternatively, micrometastasis can also be kept in a dormant state by an active immune response. Immunity-Driven Dormancy of Micrometastasis Another form of tumor mass dormancy is mediated by the immune system through a mechanism known as immunesurveillance or cancer editing. This form of control of dormancy is indeed microenvironmental in nature as the immune system and its influence on stromal cells can maintain tumors from expanding (see below). In this case the immune system prevents the expansion of proliferating tumor cells via cytotoxic T-cell-mediated cell death, thereby maintaining a constant tumor mass (Fig. 11.1) (Quesnel 2008). For instance, studies found that in patients with nonHodgkin’s lymphoma, residual lymphoma cells were still present in a few of the patients 3–8 years after they underwent a tumor-targeted monoclonal antibody therapy (Davis et al. 1998). Similarly, a significant number of non-proliferating residual leukemic cells were found several years after follow-up in patients with CML who had received INFg or allogenic bone marrow transplants (Chomel et al. 2000). Furthermore in several cases, cancers of donor origin have been reported in organ transplant recipients years and decades after the donor was in complete remission. One interpretation is that the tumor cells may have remained dormant in the donor and that the immunesuppression of the transplant recipients may have enabled these cells to escape from dormancy and proliferate (MacKie et al. 2003). Another evidence for the role of immune response in maintaining tumor dormancy is the observation that in colorectal cancer patients there was a strong positive correlation between immune infiltration in the tumor and the extent of tumor dissemination and clinical outcome (Galon et al. 2007). While the above clinical findings indicate a strong correlation between an immune-mediated control and tumor dormancy, the evidence is indirect and direct mechanistic evidence for this is lacking in the clinic. Current knowledge of the mechanisms of immune-mediated tumor dormancy is derived exclusively from experimental animal models. Using a mouse model of chemically induced sarcomas, Koebel et al. (2007) demonstrated that adaptive immunity maintains occult tumors containing slow cycling cells at an
11 Dormancy of Disseminated Tumor Cells
235
equilibrium with cells undergoing apoptosis. Evidence for immune-mediated tumor dormancy was supported by the finding that anti-CD4/CD8, anti-IFNg, or anti-IL-12 induced escape from dormancy (Koebel et al. 2007). In the BCL1 lymphoma mouse model, about 70% of mice develop dormant tumors in the spleen (Uhr et al. 1991; Yefenof et al. 1993a; Yefenof et al. 1993b). This occurs via an anti-idiotype immune response mediated by CD8+ T cells and IFNg in collaboration with the humoral immunity that usually induces cell cycle arrest and apoptosis (Farrar et al. 1999). Additionally, Schirmacher et al. showed that tumor-associated antigen (TAA) from residual dormant tumor cells is involved in maintaining high frequencies of long-term surviving memory CD8(+) T cells that help maintain mouse lymphoma cells at a low number in the bone marrow (Mahnke et al. 2005). In addition to its cytotoxic role in the promotion of tumor dormancy, the immune system can also maintain tumor cells in an occult or dormant state by affecting the tumor microenvironment. This occurs through anti-angiogenic mechanisms as many of the immune effectors such as IFN-g and IL-12 are potent angiogenesis inhibitors (Sidky and Borden 1987; Voest et al. 1995). Further, studies in PPAR-g knock-out mice have shown that the tumor-infiltrating neutrophils express high levels of thrombospondin that promoted tumor dormancy characterized by reduced vascularity and a balance between proliferation and apoptosis (Kaipainen et al. 2007). Establishment of tumor dormancy by immunesurveillance can not only be mediated by the direct cytotoxic and cytolytic effects of CTLs on tumor cells, but can also be promoted indirectly by the T-cell-mediated destruction of stromal cells that present antigen released from the tumor cells in the tumor microenvironment. Recent studies from the Schreiber laboratory have elegantly demonstrated that a single adoptive transfer of T cells solely resulted in the destruction of only tumor antigen-presenting stromal myeloid derived suppressor cells (MDSCs), without killing the tumor cells (Zhang et al. 2008). Such an eradication of MDSCs, which can otherwise cause CTL inhibition and promote tumor growth, resulted in a longterm equilibrium between the host stromal compartment and the tumor cells, which subsequently lead to a long-term maintenance of a residual tumor mass (Zhang et al. 2008). However, additional studies are needed to determine whether such immunesurveillance mechanisms of tumor mass dormancy are in fact operational in patients. In contrast to an efficient immune response in maintaining micrometastatic mass dormancy, solitary dormant disseminated tumor cells can persist by escaping the immune system. For instance, in the DA1-3b mouse acute myeloid leukemia model, dormant tumor cells evade the cytotoxic effects of the immune system, by overexpressing B7-H1 (also known as PD-L1 or CD274), which inhibits T-cell activation and CTL-mediated lysis (Quesnel 2006; Saudemont and Quesnel 2004). Additionally, tumor cells can escape the immune effectors and persist in a dormant state by decreasing the expression of adhesion molecules, required for interaction with CTLs that lead to tumor cell death (Mahnke et al. 2005) or by silencing tumor antigen expression either through the loss of expression or alteration in MHC molecules (Yamshchikov et al. 2005). Alternatively, dormant tumor cells can evade the immune system by resisting immune-mediated apoptosis (Fig. 11.1). Studies in the
236
P. Bragado et al.
DA1-3b model showed that methylation of SOCS1 gene resulted in progressive loss of expression and deregulation of the JAK-STAT pathway, leading to enhanced resistance to CTL-mediated apoptosis (Saudemont et al. 2007). Another mechanism of apoptosis resistance was observed in the BCL1 lymphoma dormancy model, wherein the cross-linking of membrane IgM by anti-idiotypic antibodies not only resulted in the induction of apoptosis but also in cell cycle arrest via overexpression of p21waf1, which could protect the dormant tumor cells from apoptosis (Marches et al. 1998). Thus, the interaction of dormant tumor cells with the immune system could have a bivalent effect – while on one hand an effective immune system could promote tumor dormancy, on the other, the ability of dormant DTCs to counteract the immune effectors could result in the persistence of dormant DTCs. An important question about DTCs undergoing cellular dormancy (see below) is if there is anything specific about quiescent tumor cells in their ability to regulate antigen presentation or be immunemodulatory or whether it is independent of their prolonged growth arrest.
Cellular Dormancy Unlike angiogenic or immune-mediated dormancy, cellular dormancy represents dormancy of single disseminated tumor cells, which enter a growth arrest (G0/G1) upon arriving in a new microenvironment (Fig. 11.1). The possibility that a prolonged but reversible G0–G1 arrest (quiescence) might explain DTC dormancy is due to the fact that majority of DTC studies across different cancers show that these cells lack markers of proliferation such as Ki67, suggesting a deep G0 or G1 arrest. Work by Luzzi et al. (1998) and Cameron et al. (2000) demonstrated that most of the tumor cells arrive at the secondary site survive but fail to proliferate. Such cells can remain in a non-proliferative dormant state even in the presence of a rich vascular network. However, some cells appear to survive only if in direct proximity to a blood vessel (Kienast et al. 2010). Unlike angiogenic dormancy and immunesurveillance, cellular dormancy or quiescence occur at the level of individual cells. As mentioned above such BM-DTCs are negative for proliferation markers, have been described to survive chemotherapy and hormonal therapy (Pantel et al. 2009), and can persist in the bone marrow (BM) over many years post-surgery (Braun et al. 2009). Therefore, they are thought to be the precursors of metastasis. Understanding their biology is one of the most important challenges for the future of cancer research, because these cells carry the information for the development of the disease. The reasons why a DTC that arrives at a secondary site will either die, start growing, or enter a protracted state of dormancy are unknown, but it has been proposed that the stress caused by the treatment, an unfavorable microenvironment or, the dissemination process itself may force these cells to enter a reversible growth arrest as a mechanism to survive until the conditions favor growth. An important question that remains is what factors define prolonged quiescence and dormancy in addition to a G0–G1 arrest? While much
11 Dormancy of Disseminated Tumor Cells
237
work is needed to define these variables across different cancers, we could speculate on some generalities based on the available data. A prerequisite is that the mechanism drives a long-term growth arrest that is reversible. Given that senescence, differentiation and apoptosis are for the most part irreversible states, one would speculate that dormant tumor cells might be highly resistant to induction of these terminal cell fates. HES1 is a transcription factor that appears to be required for induction of quiescence and it prevents the induction of senescence and differentiation in several cellular systems (Sang L et al., 2008). Thus, molecular mechanisms dependent on this or other transcription factors or signaling pathways that results in a similar state of quiescence might help to functionally and molecularly define cellular dormancy (see also below). Upon dissemination, tumor cells entering the new microenvironment can also enter into a state of cellular dormancy via the activation of genes known as metastasis suppressor genes (MSGs). These genes by definition suppress metastatic growth by inhibiting different stages of the metastatic process. They were originally identified by their reduced expression in metastatic cancer cells as compared to their non-metastatic counterparts (Horak et al. 2008). Of the 23 MSGs characterized, four of these genes KISS1, Kai1, MKK4/7, and Nm23-H1 have been shown to inhibit metastasis by inducing cellular dormancy. KISS1 encodes for secreted polypeptides known as kisspeptins and were shown to induce prolonged dormancy of DTCs in an experimental melanoma model (Nash et al. 2007). However, KISS1 induced dormancy in melanoma cells was found to be independent of its receptor grp54, suggesting that it is signaling via an alternate receptor. Kai1 or CD82, a tetraspanin protein was also shown to induce prolonged cellular dormancy of prostate cancer cells by binding to and interacting with the endothelial cell receptor DARC (Duffy antigen receptor for chemokines) (Bandyopadhyay et al. 2006). These studies demonstrated that interaction of Kai1 with DARC suppressed metastasis in vitro by inducing an irreversible senescence-like growth arrest that was associated with an upregulation in p21, a cell cycle inhibitor and a decrease in TBX2, an inhibitor of senescence (Prince et al. 2004). Since dormancy implies reversibility whether a reversible form of this senescence could be operational in vivo needs to be determined. The other MSGs, Nm23-H1, and MKK4/7 were also shown to induce dormancy of micrometastasis by inhibiting Ras signaling or as in the case of MKK4 by activating JNK and p38 stress pathways (Steeg et al. 2003) through yet unknown mechanisms. While strong activation of p38 pathways can induce apoptosis, more moderate levels of activation can induce growth arrest and/or survival mechanisms that favor subsequent tumor regrowth (see below). Experimental evidence indicates that Kai1 and KISS1 can promote cellular dormancy, meanwhile. MKK4/7 and Nm23-H1 were shown to induce dormancy of micrometastasis. However, this does not preclude the possibility that Kai1 and KISS1 may also suppress the growth of micrometastasis and MKK4/7 and Nm23-H1 can promote dormancy in single disseminated tumor cells. Therefore, identifying the pathways regulated by these metastasis suppressor genes could offer potential therapeutic targets to prevent metastatic growth.
238
P. Bragado et al.
Cellular Dormancy and the Microenvironment The tumor microenvironment plays a very critical role in modulating the tumor cell behavior. The fact that patients with advanced stage cancers can be free of disease for years and that they harbor DTCs suggests that these cells, despite having more genetic alterations, are still responsive to inhibitory signals present in the target organ microenvironment. Thus, such a tissue might have an instructive role in the induction of dormancy. It is also important to consider that dormancy might ensue if (1) there is a lack of appropriate signals for the tumor cells to resume growth (i.e., non-permissive microenvironment), (2) growth suppressive signals are encountered by DTCs that are responsive to these inhibitory cues, and (3) that DTCs might lack the appropriate surface receptors or their genes to integrate signals from a growth permissive microenvironment (Wikman et al. 2008). It is appreciated that the cross-talk between the surface receptors of these solitary tumor cells and the tumor microenvironment most likely dictates whether the newly seeded tumor cells will die, proliferate, or enter a state of cellular dormancy. For example, stromal fibroblasts from prostate tumors can induce tumor formation of non-transformed immortalized prostate epithelial cells when injected orthotopicaly in mice (Olumi et al. 1999). Moreover, transplantation of mammary epithelial cells bearing a weakly tumorigenic mutant-p53 from a non-irradiated to an irradiated mammary fat pad resulted in significant enhancement in tumor growth (Barcellos-Hoff and Ravani 2000). While these studies suggest that paracrine tumor–stroma interactions are critical components of tumor proliferation, they did not address its role in promoting exit from quiescence of DTCs and subsequent metastasis. Further, although these mechanisms are related to transformation (i.e., early cancer progression) it is possible that the stromal fibroblasts might influence entry or exit from dormancy. Researchers have demonstrated that reduced or aberrant signaling arising from deregulated interaction between adhesion molecules on the tumor cell and the stromal ECM could indeed promote tumor dormancy (Fig. 11.2) (White et al. 2006). For example, studies from the Bissell laboratory have shown that the inhibition of b-1 integrin and EGF signaling in transformed human breast cancer cells induced a dormancy-like state in three-dimensional cell culture in vitro and a proliferation block in vivo upon transplantation into nude mice (Weaver et al. 1997). As opposed to the disorganized aggregates formed by these transformed cells in three-dimensional cell cultures in vitro, this revertant population induced by b1integrin blockade formed well-organized acinar structures associated with a reassembled basement membrane and reestablished cell–cell adhesions and the down-regulation of cyclin D1 and upregulation of p21 expression (Weaver et al. 1997). In this study, a quasi-state of differentiation was achieved and because it was not shown to be reversible it is not clear if it represented a dormancy state. Furthermore, genetic ablation of b1 integrin in a transgenic mouse model to study the role of PyMT antigen or ERBB2 signaling did not result in cell death but rather induced a non-proliferating dormancy-like state (White et al. 2004).
Fig. 11.2 Potential mechanisms driving cellular dormancy in response to microenvironmentderived cues. Following intravasation (blood-borne or lymphatic routes), DTCs disseminate through circulation to target organs. Upon arrival at the secondary sites, a permissive microenvironment (Scenario a) in co-ordination with appropriate tumor cell surface receptors could allow metastatic cells to adapt and remodel their microenvironment to integrate growth-promoting signals and proliferate. (Scenario a) As an example (1) signals derived from fibronectin (FN) and transduced by the uPAR (metastasis associated urokinase receptor)–a5b1–integrin complex, focal adhesion kinase (FAK) and the epidermal growth factor receptor (EGFR) can result in extracellular signal-regulated kinase (ERK) activation and p38 inactivation thereby promoting metastatic proliferation. Transcriptional induction and activation of c-Jun and FoxM1 might mediate the G1 exit signals activated by ERK (124). In addition, circulating hormones present in the microenvironment (e.g., estrogen) (2) can also activate proliferative signals upon interaction and activation of hormone receptors (e.g., estrogen receptor on breast cancer cells) on the DTCs. (Scenario b) In contrast, loss of a surface uPAR, a5b1 integrin or EGFR (3) or inactivation of FAK (4) that transduce proliferative signals (e.g., from fibronectin) or interaction with a non-growth permissive ligand such as collagen-I (5) could result in stress signaling (low FAK–Ras–ERK, and high p38 activity) which in turn might lead to dormancy. This could in part be mediated by transcriptional activation of BHLHB3 and p53R213Q. This specific p53 mutation might enable growth arrest but avoid the induction of apoptosis (Adam et al. 2009). Additionally, collagen-mediated activation of DDR2 can lead to subsequent p16- and p21-mediated tumor cell growth arrest (6). However, whether the latter mechanism could lead to dormancy of DTC in vivo needs to be examined. Increased levels of TGFb in the microenvironment could have either a tumor growth promoting (7 Scenario a) or growth suppressive role (7 Scenario b) depending on the genetics of the DTCs. For instance, in DTCs that lodge in the BM microenvironment (tissue with high TGFb levels), TGFb-mediated inhibition and activation of cycD1 and p21 and p15, respectively, could lead to DTC growth arrest and subsequent cellular dormancy. Alternatively, increase mitogenic signaling via TGFb-SMAD pathway or inhibition of metastasis suppressor genes such as BHLHB3 in co-operation with specific genetic alterations (p53R175H) could lead to metastatic proliferation (Adorno et al. 2009). It is likely that other unidentified pathways are also involved in determining the DTC cell fate upon arrival at the secondary site
240
P. Bragado et al.
Additional evidence for the role of deregulated ligand-dependent signaling through b1 integrin in promoting cellular dormancy came from our laboratory, where we examined the behavior of HEp3 head and neck squamous carcinoma in vivo. In this model, the cross-talk between the fibronectin receptor a5b1 integrin and uPAR sets up a positive feedback loop that leads to the recruitment of FAK and EGFR, which resulted in a strong and persistent activation of the mitogenic MEK-ERK MAPK pathway thereby promoting rapid cell proliferation in vivo (Aguirre-Ghiso et al. 2001). In contrast, downregulation of uPAR lead to the inactivation of a5b1 and disruption of the uPA, uPAR, a5b1, FAK, and EGFR complex resulting in the deactivation of the ERK pathway and subsequent arrest of the cells in G0/G1 phase of cell cycle in vivo, triggering a cellular dormancy state that can last for months (Fig. 11.2) (Aguirre Ghiso et al. 1999; Kook et al. 1994). Such a state of dormancy could also be achieved by blocking signal transduction from the ECM and integrins via the interruption of FAK mitogenic signaling. Studies by Aguirre Ghiso (2002) showed that expression of FRNK, a dominant negative mutant of FAK in HEp3 cells resulted in inhibition of ERK activity and severely impaired their ability to proliferate in vivo without a subsequent increase in apoptosis and induction of dormancy. Later, a similar phenomenon of tumor dormancy was also observed upon mammary-specific ablation of FAK, which blocked progression of mammary hyperplasias to full carcinomas (Lahlou et al. 2007). Such FAK-deficient hyperplasias were still detectable in late-stage tumor samples resembling the induction of a cellular dormancy program (Lahlou et al. 2007). The inability of these cells to proliferate in vivo highlights the central role of FAK activation and b1-FAK signaling complex in tumor growth (Fig. 11.2). Recently, the Green laboratory has elegantly shown that MLCK and cytoskeleton dynamics are essential to promote escape from dormancy (Barkan et al. 2008). They showed that tumor cells from different cancers that exhibited either a dormant or metastatic phenotype in vivo however proliferated normally in two-dimensional culture conditions. However, upon culturing these cells in three-dimensional basement membrane cultures they displayed growth characteristics that recapitulated their in vivo dormant or proliferative phenotype. They demonstrated that the dormant cells entered a cell cycle arrest that was associated with increased p16 and p27 levels. Furthermore, the switch from dormant to proliferative state was associated with fibronectin production, signaling through b1 integrin MLC phosphorylation that lead to cytoskeleton reorganization and actin stress fiber formation (Barkan et al. 2008). A corroborative study of those presented above showed, using an experimental model of lung colony formation, that downregulation of FAK or b1 expression in mouse mammary carcinoma cells D2A1, resulted in a significant inhibition of proliferation of cancer cell colonies in the lungs obtained by direct intravenous injection (Shibue and Weinberg 2009). It became important to understand how this state of dormancy could be sustained for months and whether just the reduction of mitogenic signaling was sufficient to induce dormancy or whether other pathways had to be engaged. Further examination of the biochemical pathways downstream of the uPAR–integrin–EGFR surface complex showed that the establishment of the dormancy program in HEp3 cells
11 Dormancy of Disseminated Tumor Cells
241
required not only a powerful inhibition of ERK signaling but also the simultaneous activation of p38MAPK stress signaling pathway and robust induction of p21 (Aguirre-Ghiso et al. 2001). Activation of p38a was essential for persistent inhibition of ERK activation and induction of the expression of a number of transcription factors that lead to a prolonged and programmatic but reversible growth arrest. Activation of p38 resulted in the transcriptional downregulation of c-Jun and FOXM1, while it induced the TFs BHLHB3 (see below) and p53. Interestingly, in this model, p53 displays an R213Q mutation that appears to allow cells to switch of the apoptotic capacity of this TF while maintaining the induction of p21 (Fig. 11.2) (Pan and Haines 2000; and our unpublished data). These studies provided a novel insight into the opposing roles of both ERK and p38 signaling in the induction and maintenance of dormancy in HEp3 cells. While a low ERK to p38 activity ratio predicts for tumor dormancy, an inversion in the balance between these two opposing pathways predisposes the tumor cells to active proliferation (Aguirre-Ghiso et al. 2001). This, ratio or balance was found to be predictive of the in vivo behavior of various tumor types including prostate, breast, and fibrosarcoma indicating that it is not limited to the HEp3 model (Aguirre-Ghiso et al. 2003). Furthermore, targeted inhibition of EGFR with Gefitinib was shown to inhibit ERK and activate p38 (Sharma et al. 2006), suggesting that such a dormancy-inducing ERK/p38 ratio can also be achieved by using clinically relevant therapies. Further studies are needed to determine if the above-discussed mechanisms of cellular dormancy are in fact operational in disseminated tumor cells in patients. Nevertheless, some of these genes such as BHLHB3, FOXM1 (see below), and uPAR have important prognostic value and/or have been directly studied in DTCs. For example, expression of uPAR was found to be associated with dissemination of tumor cells to the bone marrow over time and the presence of uPAR expressing DTC in BM correlated with an unfavorable prognosis for patients with prostate and gastric cancer (Heiss et al. 1995; Thomas et al. 2009). In addition, the presence of uPAR expressing cells in the primary tumor positively correlated with the presence of DTCs in the bone marrow of breast cancer patients (Hemsen et al. 2003; Pierga et al. 2005). Therefore, targeted therapies aimed at disrupting uPAR and/or integrin signaling complexes (Chaurasia et al. 2009) in DTCs could potentially force these cells into a dormancy like program and improve patient outcome.
Stroma-Associated Factors and Tumor Cell Dormancy Collagen Matrix Signaling The above data demonstrated that if tumor cells lack appropriate surface molecules and signals transducers for interpreting the ECM, they can activate pathways that lead to prolonged dormancy. But, could ECM molecules other than fibronectin for example have an active role in inducing a state of quiescence/dormancy? Research from DeClerck’s group found that melanoma cells plated on fibrillar collagen-I
242
P. Bragado et al.
entered a dormancy-like cell cycle arrest. This was mediated by the interaction and activation of discoidin domain receptor 2 (DDR2), a specific collagen tyrosine kinase receptor, by fibrillar collagen-I and subsequent increase in p15 and p21 levels (Henriet et al. 2000; Wall et al. 2005; Wall et al. 2007). Also, signals mediated by the collagen receptor a2b1 integrin specifically activated the p38a isoform leading to the upregulation of collagen gene transcription (Ivaska et al. 1999). Moreover, induction of DDR2 by hypoxia was shown to be dependent on p38 activation (Chen et al. 2008). In agreement with these findings, we identified in our HEp3 tumor dormancy model, collagen-I and DDR2 as targets of p38SAPK-induced network of genes required for the induction of tumor cell quiescence. Therefore, although it remains to be tested it is possible that in addition to the above mechanisms, activation of p38 by collagen receptors could also activate a positive feedback loop that contributes to the quiescence as in the HEp3 dormancy model (Fig. 11.2).
Hormone Depletion and Dormancy Certain types of cancers such as breast, prostate, ovary, and endometrial cancer depend on the presence of circulating hormones in the microenvironment for their proliferation and survival. Therefore, strategies such as pharmacological or surgical depletion of endocrine hormones could promote a prolonged quiescence of tumor cells. In fact, such targeted therapies aimed at ablating primary tumors, do not completely eliminate residual tumor cells that can persist for prolonged periods of time. For example, reduction in estrogen receptor signaling can promote a dramatic regression of ER-positive tumors. In support of this, in studies done using invasive mammary carcinoma EMR-86 rat mammary carcinoma model, Wijsman et al. (1991) showed that the tumor outgrowth and metastasis of these cells in vivo was completely dependent on the presence of estrogen. Upon removal of estrogen albeit a rapid regression in the tumors, small dormant tumor remnants persisted in these animals, which could be restimulated to grow upon retreatment with the hormone (Wijsman et al. 1991). They further demonstrated that the in vivo tumor dormancy was not due to the residual tumor cells entering a growth arrest, but rather due to a balance between proliferation and apoptosis resulting in a tumor mass dormancy (Wijsman et al. 1991). The authors proposed that preexisting cancer stem cells within the growing tumor that rapidly proliferates in the presence of hormone could enter a state of dormancy upon hormone depletion. Another potentially important factor in endocrine-related dormancy is the active role of the tumor stroma in supporting growth. MMTV-induced pregnancy-dependent mammary tumors enter a prolonged state of dormancy of up to 300 days when transplanted into virgin female mice (Gattelli et al. 2004). Unlike the main mechanisms of dormancy described above, these tumor cells are dormant mainly due to a slow proliferation rate. These cells however can emerge from dormancy upon hormonal treatment. In an analogous situation, treatment of prostate cancer by androgen
11 Dormancy of Disseminated Tumor Cells
243
depletion also results in a significant decrease in tumor burden, but a few cancer cells could persist in a dormant quiescent state for extended lengths of time (Kusumi et al. 2008; Mostaghel et al. 2009). However, recurrences from these dormant breast and prostate cancers, are usually non-responsive to endocrine therapy and resistant to conventional chemotherapy. Several mechanisms have been proposed to account for these hormone-independent recurrences. These include: alterations in the expression or function of the hormone receptor (Santen et al. 2005); genetic changes such as inactivating mutations in the hormone receptor; amplification of the receptor or epigenetic changes in gene expression could also lead to such hormone independence (Iwase et al. 1998; van Agthoven et al. 1994), or development of clones that can adapt and grow in a low hormone environment (Mostaghel et al. 2009). These studies suggest that in addition to the role of adhesion molecules on the tumor cells, the nature of the ligands and soluble factors within the stroma that interacts with the tumor cells can also influence the phenotypic fate of disseminated tumor cells.
TGFb Signaling TGFb is a member of a large family of cytokines that are involved in the regulation of different cellular process such as cell proliferation, morphogenesis, migration, extracellular matrix production, angiogenesis, immune suppression, cytokine secretion, and apoptosis (see Chap. 31 in this book). TGFb is known to have a dual role in cancer. During early stages of tumorigenesis, TGFb exerts antiproliferative effects and functions as a tumor suppressor (Barcellos-Hoff and Akhurst 2009; Derynck et al. 2001). For instance, TGFb has been shown to be a potent anticancer agent that prohibits the uncontrolled proliferation of epithelial, endothelial, and hematopoietic cells (Tian and Schiemann 2009). Moreover, TGFb signaling has been shown to be required to maintain normal stem cell quiescence in the BM (Tang et al. 2007). However, at later stages during cancer progression, TGFb is considered to be pro-invasive and to induce epithelial to mesenchymal transitions (EMT) during metastatic cancer progressions (Seoane 2006; Thiery 2003) Therefore, TGFb can have a dual functional role as tumor suppressor and as metastatic promoter depending on the stage of cancer progression. But this also suggests that perhaps some tumors even having the ability to disseminate might still be negatively influenced by TGFb signaling in organs where it is abundant, such as the BM. The role of TGFb in inducing dormancy of DTCs is not well established, but since dissemination is an early event in tumorigenesis, it might be possible that DTCs still responsive to TGFb arriving at distant places like the BM or the liver remain arrested due to the high levels of TGFb on these tissues. It is well known, for example, that in breast cancer the detection of BM-DTCs is much higher (>50% of patients) (Klein 2008; Wikman et al. 2008) than the rate of metastasis in the BM.
244
P. Bragado et al.
Therefore, it has been proposed that the bone marrow microenvironment may be able to induce dormancy. BM stores a variety of growth factors including IGFs, TGFb, PDGF, and BMP (Mundy 1997; Mundy 2002; Roodman 2004). In addition, there is experimental evidence showing that TGFb is released from the bone matrix by bone osteoclasts (Guise and Chirgwin 2003). Moreover, majority of breast tumors, including their metastases, are positive for nuclear phosphorylated Smad2, which suggests they have an actively signaling TGFb pathway (Kang et al. 2005; Xie et al. 2002). In addition, Salm et al. (2005) have showed that TGFb is necessary to maintain dormancy of normal prostatic stem cells in the proximal region of ducts. Thus, TGFb is another potential signaling pathways that might participate in the induction of dormancy of DTCs in the BM. TGFb has been shown to exert its anti-proliferative activities by regulating epithelial cell and fibroblasts, which synthesize and secrete a variety of cytokines, growth factors, and extracellular matrix (ECM) proteins that mediate tissue homeostasis and inhibit cancer cell proliferation (Alexandrow and Moses 1995; Huang and Huang 2005). For example, studies show that TGFb stimulation of epithelial cells induces a G1 cell cycle arrest via the transcriptional upregulation of the cyclindependent kinase (CDK) inhibitors p21Cip1/WAF1 and p15Ink4b (Florenes et al. 1996). In addition, TGFb also causes transcriptional repression of transcription factors that regulate proliferation such as c-Myc and the inhibitors of differentiation Id1, Id2, and Id3 (Wu et al. 2009). Therefore, it is possible that TGFb induces quiescence of DTC through any of these mechanisms (Fig. 11.2). TGFb growth response can be inhibited by changes in active signaling networks or by the availability of transcriptional co-repressors or co-activators that regulate the Smad pathway. For instance, c-Ski is an important corepressor of TGF-b signaling that binds to and represses the activity of the Smad proteins. In breast and lung cancer cells, TGF-b induces Ski degradation through the ubiquitin-dependent proteasome and suppresses its ability to inhibit tumor metastasis (Deheuninck and Luo 2009; Reed et al. 2001). It has also been shown that in melanoma cells Id2 upregulation suppresses TGFb induction of p15, and this way circumvent TGFb-mediated inhibition of proliferation (Schlegel et al. 2009). The discovery that TGF-b-induced growth inhibition requires p53 activity suggests that mutational inactivation of p53 may contribute to the development of resistance to TGF-b antiproliferative effects by tumor cells (Atfi and Baron 2008; Cordenonsi et al. 2003). Recently, Adorno et al. (2009) showed that TGFb-dependent metastasis are empowered by mutantp53 and opposed by p63. They showed that TGFb collaborates with Ras and mutant-p53 to induce the assembly of a mutant-p53/p63 protein complex and inhibit p63 functions. They also have identified two candidate metastasis suppressor genes regulated by TGFb and associated with metastasis risk in a large cohort of breast cancer patients. Thus, it is possible that some of the above described regulatory mechanisms could be operational in the BM microenvironment and convert dormant DTC into proliferative active DTCs that may be even able to migrate to other tissues or grow in the bone microenvironment like in breast and prostate cancer. One of the metastasis suppressor genes regulated by TGFb described by Adorno et al. is BHLHB3, which is a transcriptional repressor involved in differentiation
11 Dormancy of Disseminated Tumor Cells
245
and a regulator of the circadian rhythm (Azmi et al. 2004; Fujimoto et al. 2007). In addition, in breast cancer, high BHLHB3 expression in primary tumors from patients correlates with good prognosis at 5-year survival (van de Vijver et al. 2002), suggesting that it influences behavior of residual disease. Work from our laboratory has recently showed that BHLHB3 is required for D-HEp3 cell quiescence in vivo and that its expression is repressed by ERK and induced by p38 (Adam et al. 2009). These results suggest that BHLHB3 is a novel negative regulator of tumor growth functionally linked to p38 signaling. Based on these results it is tempting to speculate that regulation of BHLHB3 through TGFb might limit expansion of DTCs into micrometastasis via cellular dormancy (i.e., quiescence) (Fig. 11.2). A study by Hoek et al. (2008) illustrates the dual role of TGFb. They found two different transcriptional signatures – one corresponding to a proliferative melanoma that was sensitive to TGFb inhibition and another corresponding to an invasive melanoma that was resistant to TGFb inhibition. They showed that upon transplantation in nude mice, the proliferative TGFb-sensitive cells produced growing tumors faster than the invasive variant. The authors concluded that there was a switch between both transcription signatures in vivo, probably regulated by environmental conditions (Hoek et al. 2008). Therefore, from all these studies we can conclude that there are two possible scenarios for TGFb regulation of DTCs fate. At early stages of progression, DTCs that lodge in tissues rich in TGFb like the BM, might be sensitive to its antiproliferative effect, and thus, TGFb may induce quiescence of DTC, that can be in that state for months, or even years. It is even possible that TGFB might be produced in an autocrine fashion to maintain dormancy. At a given point, these cells will become resistant to TGFb growth inhibitory effect, either because of changes in the microenvironment or DTCs newly acquired mutations, which will favor expansion.
Models to Study Dormancy Clinical dormancy has long been reported in breast, prostate, head and neck cancer, and melanoma (Fehm et al. 2008; Kell et al. 2000; Klein 2008; Morgan et al. 2007; Ossowski and Aguirre-Ghiso 2010). These dormant DTCs can be detected in the blood and bone marrow of these patients, months, years, and even decades after remission. Detection of these DTCs usually correlates with poor patient outcome (Allan et al. 2006; Braun et al. 2009; Ignatiadis et al. 2008; Janni et al. 2000). DTCs are usually found in the BM, therefore most of the studies regarding the biology of DTC are done in the BM, mainly because it is easier to obtain BM samples. However, there are other tissues where dormant DTC can also be found, for instance, studies in mice injected intravenously with melanoma and breast cancer cells, show a proportion of dormant cells in lungs and liver (Cameron et al. 2000; Logan et al. 2008). Most of the mouse models of cancer reproduce primary tumor development. Much less effort has been dedicated to model the progression of spontaneous MRD
246
P. Bragado et al.
and metastasis development despite the abundance of mouse transgenic models. Dormancy is a hidden state and thus is difficult to observe and study directly. In addition, characterization of disseminated tumor cells is very complicated because of their low abundance (one disseminated tumor cell per 106 bone marrow cells) (Braun et al. 2009). As we have discussed before, DTCs and primary tumor cells appear to be two different entities. Therefore, the models that we have classically used to study primary tumor growth cannot be used to extrapolate information about the DTCs. Because, metastasis is still the leading cause of mortality in patients and it arises from DTCs, we urgently need new models that better recapitulate MRD and DTC biology so that we can gather more information about the mechanisms that regulate their behavior. This will allow us to design new therapies specifically designed either to eradicate DTCs or to maintain them in a state of dormancy forever. The ideal model of solitary dormant tumor cells would consist of a population of cells which become dormant upon arrival to a secondary site, stay dormant for a period of time that is indicative of a prolonged inactive state (weeks to months), and then escape dormancy to grow. Importantly, it would be ideal that this occurs to all cells that disseminate. If only a subpopulation is dormant and other resume aggressive growth concomitantly in the same organ like in some of the above studies using i.v. injections, this argues that clonal heterogeneity is driving these differences. These may be informative, but if all cells that disseminate enter dormancy in a specific site it suggests more powerful microenvironmental mechanisms driving this phenotypic change independently of the genetics of those tumor cells. In order to study minimal residual disease, we need to have models that better recapitulate what happens in patients. The most commonly used metastatic models generate very rapid development of metastasis or rely on intravenous or intracardial injection and also, in the case of transgenic mice, are accompanied by exaggeratedly large primary tumor masses that shed enormous amounts of cells. Therefore, including surgery of the primary tumors that are smaller than 500 mm3, and then allowing the animals to recover and follow the residual cells might help better model MRD and dormancy. Bone marrow is the most important site to detect DTCs from epithelial tumors (Fehm et al.2008; Pantel et al. 2009; Riethdorf and Pantel 2008). DTCs are present in BM aspirate from 20 to 40% of breast and prostrate patients (Msaouel et al. 2008; Zhang et al. 2009). Spontaneous bone metastasis in animals is uncommon. For instance, most spontaneous mammary carcinomas in mice do not metastasize to bone but readily do to lungs and lymph nodes. However, this does not mean that they do not disseminate to bone and stay dormant. In fact, Husemann et al. (2008) and Husemann and Klein (2009) used HER-2 and PyMT transgenic mice, transplanted with premalignant HER-2 transgenic glands, and showed that they displayed disseminated tumor cells and micrometastasis in bone marrow and lungs. MMTV-Neu cells disseminated to the bone marrow of mice fail to expand unless they are transplanted into recipient mice that were irradiated. This suggests that the bone microenvironment may delay cancer progression and thus prolong patient survival, but also that perturbations of this microenvironment might trigger growth.
11 Dormancy of Disseminated Tumor Cells
247
Apart from breast cancer, in prostate cancer, the main cause of death is bone metastasis and BM-DTC can be found in patients (Alix-Panabieres et al. 2005; Morgan et al. 2007). In fact, in a recent publication Morgan et al. (2009) have showed that finding DTC in prostate cancer patients, who have undergone radical prostatectomy and do not show evidence of disease, predicts biochemical recurrence. BM-DTCs were detected in 72% of patients prior to radical prostatectomy. Significant effort has been taken on trying to characterize BM-DTCs. Orthotopic injection of human cancer cells into the prostate usually results in a low incidence of bone metastasis from late-stage cancers. For instance, Vesella and collaborators (2003) has developed a prostate cancer xenograft tumor cell dormancy model. They have used LuCaP 23.8 and LuCaP 35, hormone-sensitive CaP xenografts that express PSA and the wild-type androgen receptor and grew them orthotopically in SCID mice. Once the tumor volumes reached 250–500 mg, primary tumors were removed to allow micrometastases to grow. Using this procedure they were able to generate macroscopic metastases (>20 mg) in 71% (LuCaP 23.8) and 100% (LuCaP 35) of animals, respectively. Visible metastases in pelvic lymph nodes were observed in nearly half of the animals and microscopic lung metastases (1–40 cells) were observed in 80% of the animals. However, even using the prostatectomy procedure with the orthotopic implants, the authors were unable to generate bone metastasis. Due to the fact that they found that some bone marrow samples were PSA RT-PCR positive, which suggests the presence of CaP micrometastases in the bone, they propose that the lack of bone metastasis may be due to the time frame, and that if the observation time would have been longer maybe dormant BM-DTC might have developed into visible metastases. But this information also suggests that the timing to develop overt lesions is in part controlled by the target tissue microenvironment. Thus, the bone marrow microenvironment delays growth and this model can be used to study BM-DTC dormancy of CaP xenografts. In addition to this Shachaf et al. has published that inactivation of the MYC oncogene is sufficient to induce sustained regression of invasive liver cancers. In liver cells, MYC inactivation uncovers stem cell properties and triggers differentiation, but some of the cells remain in a dormant state and can become cancerous again by MYC reactivation. They showed that in MYC-induced hepatocellular carcinoma, MYC inactivation initially induces differentiation and apoptosis of most of the tumor cells. However under this circumstances, reactivation of MYC causes these differentiated tumor cells to rapidly become tumorigenic again (Shachaf and Felsher 2005; Shachaf et al. 2004 2008). These studies suggest that cancer cells retain the capacity to become biologically normal behaving cells, and that this change can be controlled by oncogene induction. Depending on the context oncogene reactivation can inhibit or activate the malignant phenotype. Based on our discussions above it is essential to include DTC as therapeutic targets. Further, molecular and functional information about these cells and the molecular mechanisms driving their behavior are necessary to design and develop more efficient therapies. Screening of BM for DTC could provide information about the therapeutic efficiency of a tested drug against DTCs. Identifying the specific
248
P. Bragado et al.
genetic and epigenetic changes, that DTC undergo during cancer progression, is essential to develop efficient therapies that target DTC, before they develop in overt metastasis. Acknowledgments This work was supported by grants from the Samuel Waxman Cancer Research Foundation Tumor Dormancy Program (to JAG), NIH/NCI (CA109182 to JAG), NIEHS (ES017146 to JAG), New York State Stem Cell Science – NYSTEM (to JAG) and NIDCR (DE020121 to AR).
References Adam AP, George A, Schewe D et al (2009) Computational identification of a p38SAPK-regulated transcription factor network required for tumor cell quiescence. Cancer Res 69: 5664–5672 Adorno M, Cordenonsi M, Montagner M et al (2009) A Mutant-p53/Smad complex opposes p63 to empower TGFbeta-induced metastasis. Cell 137:87–98 Aguirre Ghiso JA (2002) Inhibition of FAK signaling activated by urokinase receptor induces dormancy in human carcinoma cells in vivo. Oncogene 21:2513–2524 Aguirre Ghiso JA, Kovalski K, Ossowski L (1999) Tumor dormancy induced by downregulation of urokinase receptor in human carcinoma involves integrin and MAPK signaling. J Cell Biol 147:89–104 Aguirre-Ghiso JA (2007) Models, mechanisms and clinical evidence for cancer dormancy. Nat Rev Cancer 7:834–846 Aguirre-Ghiso JA, Estrada Y, Liu D et al (2003) ERK(MAPK) activity as a determinant of tumor growth and dormancy; regulation by p38(SAPK). Cancer Res 63:1684–1695 Aguirre-Ghiso JA, Liu D, Mignatti A et al (2001) Urokinase receptor and fibronectin regulate the ERK(MAPK) to p38(MAPK) activity ratios that determine carcinoma cell proliferation or dormancy in vivo. Mol Biol Cell 12:863–879 Alexandrow MG, Moses HL (1995) Transforming growth factor beta and cell cycle regulation. Cancer Res 55:1452–1457 Alix-Panabieres C, Rebillard X, Brouillet JP et al (2005) Detection of circulating prostate-specific antigen-secreting cells in prostate cancer patients. Clin Chem 51:1538–1541 Allan AL, Vantyghem SA, Tuck AB et al (2006) Tumor dormancy and cancer stem cells: implications for the biology and treatment of breast cancer metastasis. Breast Dis 26:87–98 Ansieau S, Bastid J, Doreau A et al (2008a) Induction of EMT by twist proteins as a collateral effect of tumor-promoting inactivation of premature senescence. Cancer Cell 14:79–89 Ansieau S, Hinkal G, Thomas C et al (2008b) Early origin of cancer metastases: dissemination and evolution of premalignant cells. Cell Cycle 7:3659–3663 Arbiser JL, Moses MA, Fernandez CA et al (1997) Oncogenic H-ras stimulates tumor angiogenesis by two distinct pathways. Proc Natl Acad Sci U S A 94:861–866 Atfi A, Baron R (2008) p53 brings a new twist to the Smad signaling network. Sci Signal 1:pe33 Azmi S, Ozog A, Taneja R (2004) Sharp-1/DEC2 inhibits skeletal muscle differentiation through repression of myogenic transcription factors. J Biol Chem 279:52643–52652 Bandyopadhyay S, Zhan R, Chaudhuri A et al (2006) Interaction of KAI1 on tumor cells with DARC on vascular endothelium leads to metastasis suppression. Nat Med 12:933–938 Barcellos-Hoff MH, Akhurst RJ (2009) Transforming growth factor-beta in breast cancer: too much, too late. Breast Cancer Res 11:202 Barcellos-Hoff MH, Ravani SA (2000) Irradiated mammary gland stroma promotes the expression of tumorigenic potential by unirradiated epithelial cells. Cancer Res 60:1254–1260
11 Dormancy of Disseminated Tumor Cells
249
Barkan D, Kleinman H, Simmons JL et al (2008) Inhibition of metastatic outgrowth from single dormant tumor cells by targeting the cytoskeleton. Cancer Res 68:6241–6250 Bidard FC, Vincent-Salomon A, Sigal-Zafrani B et al (2008) Time to metastatic relapse and breast cancer cells dissemination in bone marrow at metastatic relapse. Clin Exp Metastasis 25:871–875 Braun S, Auer D, Marth C (2009) The prognostic impact of bone marrow micrometastases in women with breast cancer. Cancer Invest 27:598–603 Braun S, Naume B (2005) Circulating and disseminated tumor cells. J Clin Oncol 23:1623–1626 Cameron MD, Schmidt EE, Kerkvliet N et al (2000) Temporal progression of metastasis in lung: cell survival, dormancy, and location dependence of metastatic inefficiency. Cancer Res 60:2541–2546 Chaurasia P, Mezei M, Zhou M-M et al (2009) Computer aided identification of small molecules disrupting uPAR/a5b1-integrin interaction: a new paradigm for metastasis prevention. PLoS ONE 4:e4617 Chen SC, Wang BW, Wang DL et al (2008) Hypoxia induces discoidin domain receptor-2 expression via the p38 pathway in vascular smooth muscle cells to increase their migration. Biochem Biophys Res Commun 374:662–667 Chomel JC, Brizard F, Veinstein A et al (2000) Persistence of BCR-ABL genomic rearrangement in chronic myeloid leukemia patients in complete and sustained cytogenetic remission after interferon-alpha therapy or allogeneic bone marrow transplantation. Blood 95:404–408 Cordenonsi M, Dupont S, Maretto S et al (2003) Links between tumor suppressors: p53 is required for TGF-beta gene responses by cooperating with Smads. Cell 113:301–314 Corey E, Quinn JE, Vessella RL (2003) A novel method of generating prostate cancer metastases from orthotopic implants. Prostate 56:110–114 Dameron KM, Volpert OV, Tainsky MA et al (1994a) Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1. Science 265:1582–1584 Dameron KM, Volpert OV, Tainsky MA et al (1994b) The p53 tumor suppressor gene inhibits angiogenesis by stimulating the production of thrombospondin. Cold Spring Harb Symp Quant Biol 59:483–489 Davis TA, Maloney DG, Czerwinski DK et al (1998) Anti-idiotype antibodies can induce longterm complete remissions in non-Hodgkin’s lymphoma without eradicating the malignant clone. Blood 92:1184–1190 Deheuninck J, Luo K (2009) Ski and SnoN, potent negative regulators of TGF-beta signaling. Cell Res 19:47–57 Derynck R, Akhurst RJ, Balmain A (2001) TGF-beta signaling in tumor suppression and cancer progression. Nat Genet 29:117–129 Farrar JD, Katz KH, Windsor J et al (1999) Cancer dormancy. VII. A regulatory role for CD8+ T cells and IFN-gamma in establishing and maintaining the tumor-dormant state. J Immunol 162:2842–2849 Fehm T, Muller V, Alix-Panabieres C et al (2008) Micrometastatic spread in breast cancer: detection, molecular characterization and clinical relevance. Breast Cancer Res 10(Suppl 1):S1 Fidler IJ (2003) The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nat Rev Cancer 3:453–458 Florenes VA, Bhattacharya N, Bani MR et al (1996) TGF-beta mediated G1 arrest in a human melanoma cell line lacking p15INK4B: evidence for cooperation between p21Cip1/WAF1 and p27Kip1. Oncogene 13:2447–2457 Folkman J (2002) Role of angiogenesis in tumor growth and metastasis. Semin Oncol 29:15–18 Folkman J, Hanahan D (1991) Switch to the angiogenic phenotype during tumorigenesis. Princess Takamatsu Symp 22:339–347 Fujimoto K, Hamaguchi H, Hashiba T et al (2007) Transcriptional repression by the basic helixloop-helix protein Dec2: multiple mechanisms through E-box elements. Int J Mol Med 19:925–932 Galon J, Fridman WH, Pages F (2007) The adaptive immunologic microenvironment in colorectal cancer: a novel perspective. Cancer Res 67:1883–1886
250
P. Bragado et al.
Gattelli A, Cirio MC, Quaglino A et al (2004) Progression of pregnancy-dependent mouse mammary tumors after long dormancy periods. Involvement of Wnt pathway activation. Cancer Res 64:5193–5199 Gimbrone MA Jr, Leapman SB, Cotran RS et al (1972) Tumor dormancy in vivo by prevention of neovascularization. J Exp Med 136:261–276 Guise TA, Chirgwin JM (2003) Transforming growth factor-beta in osteolytic breast cancer bone metastases. Clin Orthop Relat Res (415 Suppl):S32–S38 Heiss MM, Allgayer, H, Gruetzner KU et al (1995) Individual development and uPA-receptor expression of disseminated tumour cells in bone marrow: a reference to early systemic disease in solid cancer. Nat Med 1:1035–1039 Hemsen A, Riethdorf L, Brunner N et al (2003) Comparative evaluation of urokinase-type plasminogen activator receptor expression in primary breast carcinomas and on metastatic tumor cells. Int J Cancer 107:903–909 Henriet P, Zhong ZD, Brooks PC et al (2000) Contact with fibrillar collagen inhibits melanoma cell proliferation by up-regulating p27KIP1. Proc Natl Acad Sci U S A 97:10026–10031 Hoek KS, Eichhoff OM, Schlegel NC et al (2008) In vivo switching of human melanoma cells between proliferative and invasive states. Cancer Res 68:650–656 Holmgren L, O’Reilly MS, Folkman J (1995) Dormancy of micrometastases: balanced proliferation and apoptosis in the presence of angiogenesis suppression. Nat Med 1:149–153 Horak CE, Lee JH, Marshall JC et al (2008) The role of metastasis suppressor genes in metastatic dormancy. APMIS 116:586–601 Huang SS, Huang JS (2005) TGF-beta control of cell proliferation. J Cell Biochem 96:447–462 Husemann Y, Geigl JB, Schubert F et al (2008) Systemic spread is an early step in breast cancer. Cancer Cell 13:58–68 Husemann Y, Klein CA (2009) The analysis of metastasis in transgenic mouse models. Transgenic Res 18:1–5 Ignatiadis M, Georgoulias V, Mavroudis D (2008) Micrometastatic disease in breast cancer: clinical implications. Eur J Cancer 44:2726–2736 Ivaska J, Reunanen H, Westermarck J et al (1999) Integrin alpha2beta1 mediates isoform-specific activation of p38 and upregulation of collagen gene transcription by a mechanism involving the alpha2 cytoplasmic tail. J Cell Biol 147:401–416 Iwase H, Omoto Y, Iwata H et al (1998) Genetic and epigenetic alterations of the estrogen receptor gene and hormone independence in human breast cancer. Oncology 55(Suppl 1):11–16 Janni W, Rjosk D, Braun S (2000) Clinical relevance of occult metastatic cells in the bone marrow of patients with different stages of breast cancer. Clin Breast Cancer 1:217–225 Kaipainen A, Kieran MW, Huang S et al (2007) PPARalpha deficiency in inflammatory cells suppresses tumor growth. PLoS One 2:e260 Kang Y, He W, Tulley S et al (2005) Breast cancer bone metastasis mediated by the Smad tumor suppressor pathway. Proc Natl Acad Sci U S A 102:13909–13914 Kang Y, Siegel PM, Shu W et al (2003) A multigenic program mediating breast cancer metastasis to bone. Cancer Cell 3:537–549 Kell MR, Winter DC, O’Sullivan GC et al (2000) Biological behaviour and clinical implications of micrometastases. Br J Surg 87:1629–1639 Kienast Y, von Baumgarten L, Fuhrmann M et al (2010) Real-time imaging reveals the single steps of brain metastasis formation. Nat Med 16:116–122 Klein CA (2002) Direct molecular analysis of single disseminated cancer cells: a prerequisite for the development of adjuvant therapies? Acta Med Austriaca Suppl 59:10–13 Klein CA (2008) The direct molecular analysis of metastatic precursor cells in breast cancer: a chance for a better understanding of metastasis and for personalised medicine. Eur J Cancer 44:2721–2725 Klein CA, Blankenstein TJ, Schmidt-Kittler O et al (2002) Genetic heterogeneity of single disseminated tumour cells in minimal residual cancer. Lancet 360:683–689 Koebel CM, Vermi W, Swann JB et al (2007) Adaptive immunity maintains occult cancer in an equilibrium state. Nature 450:903–907
11 Dormancy of Disseminated Tumor Cells
251
Kook YH, Adamski J, Zelent A et al (1994) The effect of antisense inhibition of urokinase receptor in human squamous cell carcinoma on malignancy. EMBO J 13:3983–3991 Kusumi T, Koie T, Tanaka M et al (2008) Immunohistochemical detection of carcinoma in radical prostatectomy specimens following hormone therapy. Pathol Int 58:687–694 Lahlou H, Sanguin-Gendreau V, Zuo D et al (2007) Mammary epithelial-specific disruption of the focal adhesion kinase blocks mammary tumor progression. Proc Natl Acad Sci U S A 104: 20302–20307 Logan PT, Fernandes BF, Di Cesare S et al (2008) Single-cell tumor dormancy model of uveal melanoma. Clin Exp Metastasis 25:509–516 Luzzi KJ, MacDonald IC, Schmidt EE et al (1998) Multistep nature of metastatic inefficiency: dormancy of solitary cells after successful extravasation and limited survival of early micrometastases. Am J Pathol 153:865–873 MacKie RM, Reid R, Junor B (2003) Fatal melanoma transferred in a donated kidney 16 years after melanoma surgery. N Engl J Med 348:567–568 Mahnke YD, Schwendemann J, Beckhove P et al (2005) Maintenance of long-term tumour-specific T-cell memory by residual dormant tumour cells. Immunology 115:325–336 Marches R, Scheuermann RH, Uhr JW (1998) Cancer dormancy: role of cyclin-dependent kinase inhibitors in induction of cell cycle arrest mediated via membrane IgM Cancer Res 58:691–697 Morgan TM, Lange PH, Porter MP et al (2009) Disseminated tumor cells in prostate cancer patients after radical prostatectomy and without evidence of disease predicts biochemical recurrence. Clin Cancer Res 15:677–683 Morgan TM, Lange PH, Vessella RL (2007) Detection and characterization of circulating and disseminated prostate cancer cells. Front Biosci 12:3000–3009 Mostaghel EA, Montgomery B, Nelson PS (2009) Castration-resistant prostate cancer: targeting androgen metabolic pathways in recurrent disease. Urol Oncol 27:251–257 Msaouel P, Pissimissis N, Halapas A et al (2008) Mechanisms of bone metastasis in prostate cancer: clinical implications. Best Pract Res Clin Endocrinol Metab 22:341–355 Mundy GR (1997) Mechanisms of bone metastasis. Cancer 80:1546–1556 Mundy GR (2002) Metastasis to bone: causes, consequences and therapeutic opportunities. Nat Rev Cancer 2:584–593 Nash KT, Phadke PA, Navenot JM et al (2007) Requirement of KISS1 secretion for multiple organ metastasis suppression and maintenance of tumor dormancy. J Natl Cancer Inst 99:309–321 Naumov GN, Bender E, Zurakowski D et al (2006) A model of human tumor dormancy: an angiogenic switch from the nonangiogenic phenotype. J Natl Cancer Inst 98:316–325 Olumi AF, Grossfeld GD, Hayward SW et al (1999) Carcinoma-associated fibroblasts direct tumor progression of initiated human prostatic epithelium. Cancer Res 59:5002–5011 Ossowski L, Aguirre-Ghiso JA (2010) Dormancy of metastatic melanoma. Pigment Cell Melanoma Res 23:41–56. Pan Y, Haines DS (2000) Identification of a tumor-derived p53 mutant with novel transactivating selectivity. Oncogene 19:3095–3100 Pantel K, Alix-Panabieres C, Riethdorf S (2009) Cancer micrometastases. Nat Rev Clin Oncol 6:339–351 Pantel K, Brakenhoff RH (2004) Dissecting the metastatic cascade. Nat Rev Cancer 4:448–456 Pierga JY, Bonneton C, Magdelenat H et al (2005) Real-time quantitative PCR determination of urokinase-type plasminogen activator receptor (uPAR) expression of isolated micrometastatic cells from bone marrow of breast cancer patients. Int J Cancer 114:291–298 Prince S, Carreira S, Vance KW et al (2004) Tbx2 directly represses the expression of the p21(WAF1) cyclin-dependent kinase inhibitor. Cancer Res 64:1669–1674 Quesnel B (2006) Cancer vaccines and tumor dormancy: a long-term struggle between host antitumor immunity and persistent cancer cells? Expert Rev Vaccines 5:773–781 Quesnel B (2008) Tumor dormancy and immunoescape. APMIS 116:685–694 Rak J, Mitsuhashi Y, Bayko L et al (1995) Mutant ras oncogenes upregulate VEGF/VPF expression: implications for induction and inhibition of tumor angiogenesis. Cancer Res 55: 4575–4580
252
P. Bragado et al.
Reed JA, Bales E, Xu W et al (2001) Cytoplasmic localization of the oncogenic protein Ski in human cutaneous melanomas in vivo: functional implications for transforming growth factor beta signaling. Cancer Res 61:8074–8078 Riethdorf S, Pantel K (2008) Disseminated tumor cells in bone marrow and circulating tumor cells in blood of breast cancer patients: current state of detection and characterization. Pathobiology 75:140–148 Roodman GD (2004) Mechanisms of bone metastasis. N Engl J Med 350:1655–1664 Salm SN, Burger PE, Coetzee S et al (2005) TGF-{beta} maintains dormancy of prostatic stem cells in the proximal region of ducts. J Cell Biol 170:81–90 Sang L, Coller HA, Roberts JM. (2008) Control of the reversibility of cellular quiescence by the transcriptional repressor HES1. Science. 321(5892):1095–1100. Santen RJ, Song RX, Zhang Z et al (2005) Adaptive hypersensitivity to estrogen: mechanisms and clinical relevance to aromatase inhibitor therapy in breast cancer treatment. J Steroid Biochem Mol Biol 95:155–165 Saudemont A, Hamrouni A, Marchetti P et al (2007) Dormant tumor cells develop cross-resistance to apoptosis induced by CTLs or imatinib mesylate via methylation of suppressor of cytokine signaling 1. Cancer Res 67:4491–4498 Saudemont A, Quesnel B (2004) In a model of tumor dormancy, long-term persistent leukemic cells have increased B7-H1 and B7.1 expression and resist CTL-mediated lysis. Blood 104:2124–2133 Schlegel NC, Eichhoff OM, Hemmi S et al (2009) Id2 suppression of p15 counters TGF-betamediated growth inhibition of melanoma cells. Pigment Cell Melanoma Res 22:445–453 Schmidt-Kittler O, Ragg T, Daskalakis A et al (2003) From latent disseminated cells to overt metastasis: genetic analysis of systemic breast cancer progression. Proc Natl Acad Sci U S A 100:7737–7742 Schwarte-Waldhoff I, Volpert OV, Bouck NP et al (2000) Smad4/DPC4-mediated tumor suppression through suppression of angiogenesis. Proc Natl Acad Sci U S A 97:9624–9629 Seoane J (2006) Escaping from the TGFbeta anti-proliferative control. Carcinogenesis 27:2148–2156 Shachaf CM, Felsher DW (2005) Tumor dormancy and MYC inactivation: pushing cancer to the brink of normalcy. Cancer Res 65:4471–4474 Shachaf CM, Gentles AJ, Elchuri S et al (2008) Genomic and proteomic analysis reveals a threshold level of MYC required for tumor maintenance. Cancer Res 68:5132–5142 Shachaf CM, Kopelman AM, Arvanitis C et al (2004) MYC inactivation uncovers pluripotent differentiation and tumour dormancy in hepatocellular cancer. Nature 431:1112–1117 Sharma SV, Gajowniczek P, Way IP et al (2006) A common signaling cascade may underlie “addiction” to the Src, BCR-ABL, and EGF receptor oncogenes. Cancer Cell 10:425–435 Sheibani N, Frazier WA (1996) Repression of thrombospondin-1 expression, a natural inhibitor of angiogenesis, in polyoma middle T transformed NIH3T3 cells. Cancer Lett 107:45–52 Shibue T, Weinberg RA (2009) Integrin Œ£1-focal adhesion kinase signaling directs the proliferation of metastatic cancer cells disseminated in the lungs. Proc Natl Acad Sci U S A 106: 10290–10295 Sidky YA, Borden EC (1987) Inhibition of angiogenesis by interferons: effects on tumor- and lymphocyte-induced vascular responses. Cancer Res 47:5155–5161 Slade MJ, Singh A, Smith BM et al (2005) Persistence of bone marrow micrometastases in patients receiving adjuvant therapy for breast cancer: results at 4 years. Int J Cancer 114:94–100 Solakoglu O, Maierhofer C, Lahr G et al (2002) Heterogeneous proliferative potential of occult metastatic cells in bone marrow of patients with solid epithelial tumors. Proc Natl Acad Sci U S A 99:2246–2251 Steeg PS, Ouatas T, Halverson D et al (2003) Metastasis suppressor genes: basic biology and potential clinical use. Clin Breast Cancer 4:51–62 Stoecklein NH, Hosch SB, Bezler M et al (2008) Direct genetic analysis of single disseminated cancer cells for prediction of outcome and therapy selection in esophageal cancer. Cancer Cell 13:441–453
11 Dormancy of Disseminated Tumor Cells
253
Su JD, Mayo LD, Donner DB et al (2003) PTEN and phosphatidylinositol 3¢ -kinase inhibitors up-regulate p53 and block tumor-induced angiogenesis: evidence for an effect on the tumor and endothelial compartment. Cancer Res 63:3585–3592 Tang B, Yoo N, Vu M et al (2007) Transforming growth factor-beta can suppress tumorigenesis through effects on the putative cancer stem or early progenitor cell and committed progeny in a breast cancer xenograft model. Cancer Res 67:8643–8652 Thiery JP (2003) Epithelial-mesenchymal transitions in development and pathologies. Curr Opin Cell Biol 15:740–746 Thomas C, Wiesner C, Melchior SW et al (2009) Urokinase-plasminogen-activator receptor expression in disseminated tumour cells in the bone marrow and peripheral blood of patients with clinically localized prostate cancer. BJU Int 104:29–34 Tian M, Schiemann WP (2009) The TGF-beta paradox in human cancer: an update. Future Oncol 5:259–271 Udagawa T, Fernandez A, Achilles E-G et al (2002) Persistence of microscopic human cancers in mice: alterations in the angiogenic balance accompanies loss of tumor dormancy. FASEB J 16:1361–1370 Uhr JW, Tucker T, May RD et al (1991) Cancer dormancy: studies of the murine BCL1 lymphoma. Cancer Res 51:5045s–5053s van Agthoven T, van Agthoven TL, Dekker A et al (1994) Induction of estrogen independence of ZR-75–1 human breast cancer cells by epigenetic alterations. Mol Endocrinol 8: 1474–1483 van de Vijver MJ, He YD, van’t Veer LJ et al (2002) A gene-expression signature as a predictor of survival in breast cancer. N Engl J Med 347:1999–2009 Voest EE, Kenyon BM, O’Reilly MS et al (1995) Inhibition of angiogenesis in vivo by interleukin 12. J Natl Cancer Inst 87:581–586 Volpert OV, Dameron KM, Bouck N (1997) Sequential development of an angiogenic phenotype by human fibroblasts progressing to tumorigenicity. Oncogene 14:1495–1502 Wall SJ, Werner E, Werb Z et al (2005) Discoidin domain receptor 2 mediates tumor cell cycle arrest induced by fibrillar collagen. J Biol Chem 280:40187–40194 Wall SJ, Zhong ZD, DeClerck YA (2007) The cyclin-dependent kinase inhibitors p15INK4B and p21CIP1 are critical regulators of fibrillar collagen-induced tumor cell cycle arrest. J Biol Chem 282:24471–24476 Watnick RS, Cheng YN, Rangarajan A et al (2003) Ras modulates Myc activity to repress thrombospondin-1 expression and increase tumor angiogenesis. Cancer Cell 3:219–231 Weaver VM, Petersen OW, Wang F et al (1997) Reversion of the malignant phenotype of human breast cells in three-dimensional culture and in vivo by integrin blocking antibodies. J Cell Biol 137:231–245 Weigelt B, Peterse JL, van ‘t Veer LJ (2005) Breast cancer metastasis: markers and models. Nat Rev Cancer 5:591–602 Wen S, Stolarov J, Myers MP et al (2001) PTEN controls tumor-induced angiogenesis. Proc Natl Acad Sci U S A 98:4622–4627 White DE, Kurpios NA, Zuo D et al (2004) Targeted disruption of beta1-integrin in a transgenic mouse model of human breast cancer reveals an essential role in mammary tumor induction. Cancer Cell 6:159–170 White DE, Rayment JH, Muller WJ (2006) Addressing the role of cell adhesion in tumor cell dormancy. Cell Cycle 5:1756–1759 Wijsman JH, Cornelisse CJ, Keijzer R et al (1991) A prolactin-dependent, metastasising rat mammary carcinoma as a model for endocrine-related tumour dormancy. Br J Cancer 64:463–468 Wikman H, Vessella R, Pantel K (2008) Cancer micrometastasis and tumour dormancy. APMIS 116:754–770 Wu S, Hultquist A, Hydbring P et al (2009) TGF-beta enforces senescence in Myc-transformed hematopoietic tumor cells through induction of Mad1 and repression of Myc activity. Exp Cell Res 315:3099–3111
254
P. Bragado et al.
Xie W, Mertens JC, Reiss DJ et al (2002) Alterations of Smad signaling in human breast carcinoma are associated with poor outcome: a tissue microarray study. Cancer Res 62:497–505 Yamshchikov GV, Mullins DW, Chang CC et al (2005) Sequential immune escape and shifting of T cell responses in a long-term survivor of melanoma. J Immunol 174:6863–6871 Yefenof E, Picker LJ, Scheuermann RH et al (1993a) Cancer dormancy: isolation and characterization of dormant lymphoma cells. Proc Natl Acad Sci U S A 90:1829–1833 Yefenof E, Picker LJ, Scheuermann RH et al (1993b) Induction of B cell tumor dormancy by antiidiotypic antibodies. Curr Opin Immunol 5:740–744 Zhang B, Zhang Y, Bowerman NA et al (2008) Equilibrium between host and cancer caused by effector T cells killing tumor stroma. Cancer Res 68:1563–1571 Zhang Y, Ma B, Fan Q (2009) Mechanisms of breast cancer bone metastasis. Cancer Lett 292:1–7
Part III
Vasculature And Stroma
Chapter 12
Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome of Antiangiogenic Therapy: New Perspectives on an Ongoing Controversy Robert S. Kerbel, Francesco Bertolini, and Yuval Shaked Abstract Tumor angiogenesis is driven not only by proliferation of differentiated endothelial cells in sprouting capillaries from pre-existing mature vessels, but also by mobilization of bone marrow derived circulating endothelial progenitor cells (EPCs). The latter are thought to home to the tumor site and incorporate into the lumen of newly growing blood vessels. Over the past several years, a growing number of reports have challenged the hypothesis concerning the involvement of EPCs in tumor angiogenesis, and instead suggest that such cells only have a minor role, if any at all, in the formation of tumor-associated blood vessels. Consequently, these studies implicate EPCs as a minor or negligible target for cancer therapy. In this review, we discuss the arguments for and against a significant role of EPCs in tumor angiogenesis and growth, and as possible surrogate markers of angiogenesis as well as valuable therapeutic targets.
Introduction Over the last decade, various types of bone marrow derived cell (BMDCs) have emerged as key players in tumor angiogenesis, growth, and metastasis (Shaked and Voest 2009; Kaplan et al. 2007); most such cell types originate from the hematopoietic lineage (Coffelt et al. 2009; De Palma and Naldini 2006; Grunewald et al. 2006; Shojaei et al. 2007). They can contribute to angiogenesis by various mechanisms including paracrine secretion of different growth factors mediated by perivascular residing BMDCs, and by contributing to the stability of the vasculature. Research into the development of antiangiogenic drugs to treat cancer has revealed that some BMDCs may compensate for vessel-targeting antiangiogenic effects, thus allowing for continued tumor growth. These bone marrow derived Y. Shaked (*) Department of Molecular Pharmacology, Rappaport Faculty of Medicine, Technion – Israel Institute of Technology, Haifa, Israel e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_12, © Springer Science+Business Media, LLC 2010
257
258
R.S. Kerbel et al.
tumor-associated stromal cells include various myeloid, macrophage, and monocytic cell populations. They all express the pan-hematopoietic surface marker, CD45, along with various endothelial cell/angiogenesis-related surface markers. However, a subset of BMDCs which are CD45-negative has been found to promote tumor angiogenesis by acting as an alternative source of endothelial cells; these cells are termed endothelial precursor or progenitor cells (EPCs). In contrast to all other proangiogenic BMDC types, EPCs are thought to merge with the wall of a growing blood vessel, where they differentiate into mature endothelial cells, and contribute to vessel growth. The various types of BMDC promoting angiogenesis are summarized in Fig. 12.1.
Fig. 12.1 Cellular players in tumor angiogenesis. Representation of various BMDC populations which have been shown to stimulate or amplify tumor angiogenesis. The various hematopoietic (CD45+) cell types appear to have a perivascular location with respect to the tumor neovasculature, whereas the CD45-negative endothelial progenitor cells can actually incorporate into the lumen of a growing vessel and differentiate into a mature endothelial cell. TEMs tie-2 expressing monocytes, HSC hematopoietic stem cell, ECs endothelial cell, TASC tumor associated stromal cell, HPC hematopoietic progenitor cell, VLC vascular leukocytes. Mature endothelial cells detach from the basal membrane following damage to tumor blood vessel, and in the blood flow become apoptotic circulating endothelial cells (CECs). Circulating endothelial progenitors (CEPs) that might derive from hemangioblast are mobilized from the bone marrow to incorporate in the vasculature as part of physiological tissue repair. In micro-metastatic lesions CEPs may play a significant role in the angiogenic switch, which promote metastatic growth. Reproduced from Martin-Padura and Bertolini (2008), with permission from the publisher
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
259
There are a number of controversial issues regarding the origin, nature, and relevance of EPCs to angiogenesis, including tumor angiogenesis which are summarized below, along with some recent findings that may help resolve some of these controversies.
The Identification of EPCs In 1997, Isner’s group reported that post-natal vasculogenesis and angiogenesis can be driven by BMDCs circulating in peripheral blood which are home to sites of ongoing angiogenesis. The systemic process of de novo formation of capillaries and vessels from angioblasts of endothelial progenitor, previously shown to be involved in the formation of an embryonic vessel network, has been investigated in adults. Activated endothelial cells have been shown to express CD34 along with VEGFR1. In addition, CD34 is expressed by all hematopoietic stem cells but not mature hematopoietic cells. Using this distinguishing surface marker, Asahara et al. (1997) isolated CD34-negative and -positive cells and assessed their possible characteristics as endothelial cells, both morphologically and functionally. Injection of CD34+ cells labeled with the fluorescent dye Dil into ischemic hind-limb bearing mice resulted in massive homing of CD34+ but not CD34- cells to the mouse ischemic limb, and incorporation into the blood vessel wall. This study was the first to show that circulating CD34+ cells, obtained from human peripheral blood, can contribute to angiogenesis in adult tissues implicating the possibility that CD34+ cells are, in fact, EPCs. This discovery opened up new directions in angiogenesis research including tumor angiogenesis.
The Controversy Surrounding Functions of EPCs Following the first study by Asahara et al. (1997), implicating a role for EPCs in adult angiogenesis, many investigators reported results confirming the involvement of EPCs in tumor growth; various studies revealed highly contrasting results. Perhaps the most compelling evidence to support a role for EPCs in tumor angiogenesis emerged from studies using knockout mice bearing mutations in the Id family of transcriptional repressors. Id proteins are helix–loop–helix transcription factors that regulate various physiological processes including angiogenesis. These mice have been used to investigate the role of EPCs in tumor angiogenesis (Lyden et al. 1999). In vivo studies revealed that Id1-null mice exhibit decreased neovascularization as assessed by the Matrigel perfusion assay, compared to their wildtype counterparts, and this is regulated by the endogenous inhibitor of angiogenesis, thrombospondin 1 (TSP-1) (Volpert et al. 2002). It was shown that loss of Id1 in tumor endothelial cells resulted in downregulation of several proangiogenic genes, including those encoding a6 and b4 integrins, matrix metalloprotease-2, and fibroblast
260
R.S. Kerbel et al.
growth factor receptor-1 (Ruzinova et al. 2003). Benezra and colleagues showed tumors growing in mice with a defect in the Id1 and Id3 genes (Id1+/–Id3–/–) are defective in angiogenesis, and hence growth. However, if such mice are irradiated and then given a transplant of wild-type bone marrow derived cells, tumor angiogenesis and progressive growth are restored (Lyden et al. 2001). Thus this in vivo data seemingly confirm the existence of EPCs, and suggest that these cells can contribute to overall tumor growth. In contrast to the in vivo and ex vivo evidence regarding a significant role of EPCs in tumor growth, several recent studies have challenged the definition, function, and contribution of EPCs. Perhaps the major controversy is whether putative EPCs can truly incorporate into tumor blood vessels. Some investigators have raised doubts about this possibility, in part, due to the presence of other BMDCs that are found in close proximity to blood vessels, but which do not directly incorporate to the blood vessel wall. Such cells do not express VEGFR2 or CD31 surface markers like EPCs or mature endothelial cells, but express a variety of surface markers associated with angiogenesis such as Tie-2, CXCR4, and VEGFR1 (Grunewald et al. 2006; De Palma et al. 2005; Jin et al. 2006). These cells are found to colonize tumors, and contribute to their growth. Another factor which has led to this increasing controversy is the fact that some tumor models exhibit relatively high number of EPCs, whereas in others the number detected is minimal or even nonexistent: a range of 0–50% of tumor endothelium consisting of EPCs has been reported. Machein et al. (2003) reported that tumor type is one such factor as this may significantly alter vasculogenesis and relative colonization of EPCs. They reported that the majority of bone marrow cells expressing hematopoietic and/or microglia markers, did not contribute to glioma vasculature. They also showed that overexpression of VEGF in glioma cells produced highly vascularized tumors, but the number of EPCs incorporated into the tumor vasculature was not increased. In a more recent study, Machein and colleagues reported that EPCs do not contribute at all to the growth of Lewis Lung carcinomas. The study evaluated the colonization of BMDCs tagged with GFP using several imaging techniques (Wickersheim et al. 2009). Ziegelhoeffer et al. (2004) reported that in mice bearing methylcholanthrene-induced fibrosarcoma (BFS-1), there were little, if any, BMDCs found to be incorporated into the tumor vasculature. In addition to tumor type, tumor grade has also been shown to alter the percentage of EPCs. Li et al. (2004) found that Id1 and Id3 transcription factors are highly expressed in endothelial cells of poorly differentiated prostate adenocarcinomas but not in the vasculature of well-differentiated tumors. Overall, these studies suggest that neovascularization in certain tumor types and of different tumor grade may be mediated by division and migration of differentiated endothelial cells from pre-existing vessels and that EPCs may not play a significant role in these circumstances. A clinical study by Peters et al. (2005) evaluated the number of BMDCs incorporated into the tumor blood vessel wall. They studied tumors obtained from patients who had previously undergone bone marrow transplantation obtained from individuals of the opposite sex. By using fluorescent In Situ hybridization (FISH)
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
261
staining techniques and confocal laser scan microscopy the authors counted the number of endothelial cells in tumors which expressed mismatched sex chromosomes, indicating that they were derived from the transplanted donor bone marrow cells. They found that approximately 5% of tumor blood vessels consisting of bone marrow origin. These results indicate that there is a low number of EPCs incorporating into tumor blood vessels, and thus the possibility that they may function as endothelial cells contributing to the formation of a tumor vasculature. In recent studies, we have offered another possibility to explain the differences in the number of EPCs contributing to tumor growth reported by different investigators. We have shown that the genetic background of the host may play a significant role in the extent of adult vasculogenesis (Shaked et al. 2005a). Using a number of mouse strains, we evaluated the levels of EPCs circulating in the blood, and found that there are some strains, e.g., BALB/c, which exhibit high levels of EPCs in peripheral blood whereas other strains, e.g., C57Bl showed a low number of EPCs in up to a tenfold difference. We also showed that these levels strikingly correlated with the angiogenic responsiveness of the mouse strains to angiogenic stimuli measured by the corneal neovascular micropocket assay (Rohan et al. 2000) as well as in cases where mice were genetically manipulated to induce or reduce angiogenesis factors (Shaked et al. 2005a). These results indicated that in addition to tumor grade and type, the genetic background of the host may play a significant role in tumor neoangiogenesis. Only in the past few years several investigators have refuted the existence of EPCs in tumors. Purhonen et al. (2008) studied the mobilization and tumor homing of EPCs. They induced angiogenesis by using melanomas, APCmin adenomas, adenoviral VEGF delivery, and Matrigel plugs in four different genetically tagged universal or endothelial cell-specific chimeric mouse models. They failed to detect VEGFR2 cells mobilized from the bone marrow, nor incorporation directly into blood vessels. They thus concluded that EPC contribution is an extremely rare event, and perhaps might not even exist. Similarly, as noted above, Machein et al. (2003) who studied BMDCs expressing GFP, found that local VEGF induces a massive tumor infiltration of BMDCs, but no evidence of vessel wall integration of these cells. Furthermore, Kim et al. (2009) investigated acute and chronic phases of angiogenesis in wound healing assays, gel foam angiogenesis, in a parabiosis mouse model, and found that BMDCs expressing CD31 homing to sites of angiogenesis are macrophages (or monocytes) as they also express F4/80, a specific pan-macrophage marker. This approach was undertaken to avoid radiation of mice followed by a bone marrow transplant as such radiation might have affected the outcome of some of the prior studies using GFP+ bone marrow cells. These observations constitute yet another challenge to the EPC concept and the role of such cells in tumor angiogenesis, by suggesting that CD31 is expressed on angiogenic monocytes which do not differentiate into endothelial cells (Horrevoets 2009). In summary, these particular studies suggest that during tumor development and progression, new capillary formation is dependent only on “sprouting angiogenesis”, i.e., by migration and division of differential endothelial cells from pre-existing mature vessels. They clearly contradict the hypothesis of an involvement of EPCs in angiogenesis and tumor growth, and suggest that such cells might even be an
262
R.S. Kerbel et al.
experimental artifact. It should be noted, however, that several criticisms have been made regarding the study and conclusions reported by Purhonen et al. (2008). For example, APCmin mice develop only obstructive adenomas, rather than adenocarcinomas; therefore, it is an inappropriate model to study EPC incorporation in tumors (Kerbel et al. 2008). Furthermore, the study of 6-month-old VEGF-Aloaded Matrigel plugs in mice is not possible because Matrigel plugs are degraded within 2 months, particularly when VEGF-A by itself is unlikely to induce neoangiogenesis in Matrigel plugs. A recent study by Madlambayan et al. (2009) attempted to resolve this ongoing controversy by comparing a variety of models of tumor angiogenesis in a single mouse background. These included the Lewis Lung carcinoma, B16 melanoma as well as a model of retinopathy. The authors evaluated a spectrum of BMDC types including Tie2 expressing monocytes (TEMs) (De Palma et al. 2005), tumor associated macrophages (Coffelt et al. 2009), EPCs, and hemangiocytes (Jin et al. 2006). They found that the contribution of systemic BMDCs to tumor angiogenesis is dependent primarily on the expression of SDF-1 in the model/site tested. Manipulation of SDF-1 expression in such sites revealed large differences in the number of BMDCs recruited. Since the majority of EPCs express the CXCR4 chemokine receptor (Pitchford et al. 2009), the results suggest that EPCs, among other bone marrow derived CXCR4+ cells, home to active sites of angiogenesis, and promote tumor growth. This study offers a resolution for the finding of different values of BMDCs, including EPCs contributing to tumor growth in different models, as reported by some investigators.
The Controversy About the Definition of EPCs In addition to the questions raised about the possible numbers of EPCs detected in tumor vessels and their functional importance, yet another controversy which has emerged concerns the precise definition and origin of such cells. It is likely that the difficulties associated with some of the techniques used to isolate, sort, enumerate, and quantitate EPCs have contributed to this situation. In peripheral blood, EPCs are measured by using a combination of antibodies recognizing various endothelial and hematopoietic surface markers, as detected by multiparametric flow cytometry techniques (Bertolini et al. 2006). The lack of a single definitive antigen to define and detect EPC populations has contributed to the debate regarding the evaluation of EPCs. Initial studies using both in vitro quantitative and flow cytometry characterization of EPCs revealed that VEGFR2 and CD34 are the “ultimate” surface markers of EPCs; however, it has been shown that AC133 (CD133), an orphan receptor that is expressed on some hemoatopoietic progenitor cells, and lost once they differentiate, is also considered as a marker of EPCs (Peichev et al. 2000; Rafii et al. 2002). Furthermore, CD146, for example, has been evaluated as a specific endothelial cell marker (Solovey et al. 2001), and is still used to detect EPCs circulating in peripheral blood. However, other studies have recently reported that mesenchymal cells
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
263
(Elshal et al. 2005), and activated T lymphocytes (Duda et al. 2006) can also express CD146. Moreover, VE-cadherin (CD144), an endothelial-specific marker which was also thought to be expressed specifically by endothelial cells, has been reported to be expressed by some hematopoietic stem cells (Kim et al. 2005). Overall, currently, such results indicate that the most useful markers for identifying EPCs circulating in the blood are CD34, VEGFR2, and CD133, which appear to constitute a phenotypically distinct population of cells capable of differentiating into endothelial cells that can play a role in post-natal angiogenesis/vasculogenesis. To add complexity to the phenotypic characterization of EPCs (and thus to the confusion in the field) recent studies by Yoder, Ingram, and colleagues indicated that established commercially available kits used to repopulate and isolate EPCs from human peripheral blood by identifying methylcellulose endothelial cell colony forming units (CFU-ECs) do not represent truly putative EPCs (Case et al. 2007; Yoder et al. 2007). In a number of studies they identified another population of endothelial colony-forming cells (ECFCs), and tested the origin, proliferative potential, and differentiation capacity of ECFCs as well as CFU-ECs (obtained from human peripheral blood). They compared the function of CFU-ECs and ECFCs and determined that CFU-ECs are derived from the hematopoietic system, and possess myeloid progenitor cell activity, whereas ECFCs display robust proliferative potential, and form perfused vessels in vivo (Yoder et al. 2007). Overall, their results suggest that CFU-ECs, obtained from commercially available kits to isolate EPCs, are in fact not authentic EPCs, whereas ECFCs may actually represent a rare EPC population. These studies indicate that in order to improve quantitation and isolation of putative EPC populations, there is a clear need to discover a specific EPC marker which would constitute verification of EPCs in endothelial engraftment experiments in vivo. The development of monoclonal antibodies to detect endothelial cell-associated antigens has led to the extensive use of flow cytometry based assays to measure the number of EPCs circulating in peripheral blood. The use of multiparametric flow cytometry analysis on whole blood labeled simultaneously using a combination of conjugated antibodies has provided a useful tool to adequately measure rare populations in the blood. The combination of antibodies to detect EPCs in the blood usually consists of first excluding cells expressing the hematopoietic marker CD45, and then detecting cells expressing progenitor markers such as CD34, CD133, in addition to VEGFR2 as an endothelial cell marker. However, the lack of a standardized method to detect EPCs and mature endothelial cells circulating in the blood (CECs) has resulted in very different values of EPCs detected using flow cytometry techniques by different investigators, and this has caused increasing confusion about the nature of these cells, both clinically and preclinically. Since flow cytometry approaches require accurate sequential gating and are prone to operator and instrument-associated variability, a greater effort has been made recently to try and standardize the methodology for detecting EPCs and CECs in peripheral blood, especially for clinical studies (Bertolini et al. 2006; Mancuso et al. 2009). Thus, the definition and detection of EPCs requires further refinement and improvement in order to become widely accepted when used in different centers and by different operators/groups.
264
R.S. Kerbel et al.
EPCs as a Surrogate Biomarker for Antiangiogenic Therapy The identification of EPCs in peripheral blood has led to emerging data regarding the levels of EPCs in different pathologies including cardiovascular disease (Werner et al. 2005), diabetes (Cubbon et al. 2009), arthritis (Distler et al. 2009), eclampsia (Gammill et al. 2007), and cancer. With respect to cancer patients, some early studies evaluated the number of circulating endothelial cells (CECs) and their precursor subset (CEPs). It has been shown that levels of CECs, including viable CECs in cancer patients is higher than in normal healthy controls (Mancuso et al. 2001). This is thought to occur by virtue of the fact that the production of proangiogenic factors by the tumor may promote CEC and CEP mobilization. In some of our early studies, we hypothesized that certain antiangiogenic drugs may target CEPs or at least may decrease their mobilization into the bloodstream, despite the production of angiogenic stimuli by the tumor. To test this hypothesis, we took advantage of a previous study by Rohan et al. (2000) who reported that angiogenic responsiveness measured by the corneal neovascular micropocket assay using bFGF, is markedly heterogeneous in different strains of mice. We found a striking correlation between the angiogenic responsiveness and the levels of CECs and CEPs in the same mouse strain (Shaked et al. 2005a). We next asked whether CEPs are targeted by antiangiogenic drugs. We showed that mice treated with DC101, a monoclonal antibody directed to mouse VEGFR2 (Prewett et al. 1999) or ABT-510, a peptide mimetic of TSP-1 (an endogenous angiogenic inhibitor (Lawler 2002)) (Yap et al. 2005), caused a significant dose-dependent drop in levels of CEPs (defined by the phenotype CD45-,VEGFR2+,CD117+, and CD13+). We also found that the maximum decline (the “nadir”) in CEP levels correlated with the optimal biological dose (OBD) of the drug based on prior dose-response studies of anti-tumor activity (Shaked et al. 2005a). Similar results were obtained when we used another antiangiogenic treatment strategy, namely, metronomic low-dose chemotherapy (Klement et al. 2000; Hanahan et al. 2000; Shaked et al. 2005c), i.e., the administration of chemotherapy drugs using lower doses than the maximum tolerated dose (MTD), given at frequent, regular intervals with no long drug-free break periods (Klement et al. 2000). These results not only provided evidence that some antiangiogenic drugs and treatment strategies target putative CEPs, but also suggested that CEPs can be used as a surrogate pharmacodynamic biomarker to determine the OBD of antiangiogenic therapies (Schneider et al. 2005), an approach which has been further evaluated in clinical studies (Mancuso et al. 2006; Buckstein et al. 2006; Bender et al. 2008). Mancuso et al. (2006) have raised the possibility that apoptotic CECs may be used as a biomarker for antiangiogenic drug or treatment outcome. In their study of 121 breast cancer patients treated with metronomic cyclophosphamide and methrotraxate levels of CECs were evaluated. In contrast to the perception, at that time, that antiangiogenic drug therapy would reduce the levels of CECs and CEPs (Shaked et al. 2005a), the authors found that increases in CECs were associated with tumor response. When the authors distinguished between apoptotic and viable CECs, they
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
265
found that in responding patients, the majority of CECs are in fact apoptotic (Mancuso et al. 2006). Supported by preclinical models of metronomic chemotherapy, the authors showed that the elevation in apoptotic CEC levels were likely tumor derived (i.e., derived from the tumor vasculature) and not host derived, since studies in non-tumor bearing mice indicated no detectable rise in apoptotic CECs levels. Comparable results were generated in additional clinical studies. For example, in a phase I clinical study of pediatric cancer patients undergoing bevacizumab treatment (Bender et al. 2008), it was reported that the ratio between the number of apoptotic CECs and total CECs was associated with the number of bevacizumab treatment cycles, indicating patients responding to therapy (Bender et al. 2008). In another clinical study, patients with advanced breast cancer received metronomic oral capecitabine and cyclophosphamide along with bevacizumab and levels of CECs were measured during the treatment; higher baseline CECs were found to be correlated with overall response, clinical benefit, and improved progression-free survival (Dellapasqua et al. 2008). Along with this analysis, the pharmacokinetics of CECs was also evaluated, and a pattern of decreased CECs and increased levels of various angiogenic growth factors have been reported in relapsed patients (Calleri et al. 2009). Overall, the aforementioned preclinical and clinical studies suggest that CECs and perhaps also CEPs may be used as surrogate biomarkers to determining the OBD of an antiangiogenic drug as well as predictive biomarkers of clinical outcome.
Therapy-Induced EPC Mobilization and Tumor Vessel Incorporation The lack of a specific antigen to define and detect EPCs, and the use of techniques to enumerate and quantitate EPCs in peripheral blood using at least four different surface markers have become extremely challenging technical issues for investigators in the field. However, in addition to studying CECs and CEPs as possible biomarkers for treatment outcome, our efforts have also been directed to studying the functional properties and effects of EPCs in different therapeutic circumstances. The results we have obtained provide an entirely new perspective on their impact and importance to angiogenesis and cancer therapy not only using antiangiogenic drugs, but also other therapies, including chemotherapy, both conventional and metronomic. We first investigated the possible impact of CEPs on tumor angiogenesis immediately after therapy with various cytotoxic agents. The rationale was based on our initial observations that MTD chemotherapy using cyclophosphamide rapidly increases the mobilization and viability of CEPs in NOD/SCID mice bearing human lymphomas whereas, in contrast, a low-dose metronomic regimen using the same chemotherapy drug, had the very opposite effect (Bertolini et al. 2003). These results suggested that MTD chemotherapy transiently induces angiogenesis (vasculogenesis) during the follow-up drug-free break periods which are usually instituted when administrating MTD cytotoxic therapy regimens, in order to allow recovery of
266
R.S. Kerbel et al.
patients (or animals) from the toxic side effects of the treatment, e.g., myelosuppression. In contrast, the chemotherapy may induce an antiangiogenic effect during the therapy, by targeting dividing endothelial cells (Kerbel 1991). Thus this sequence of events may be likened to an “action” (antiangiogenesis) followed by a “reaction” (pro-angiogenesis). As such, the “drug holidays” might promote, at least in part, tumor cell repopulation between successive courses of MTD chemotherapy. The surviving repopulating tumor cells are obviously one of the most important causes for treatment failure or limitations in the effects of such treatments (Kim and Tannock 2005). In this regard, we have recently found that shortly after treatment with any one of a number of different cytotoxic or cytotoxic-like drugs such as microtubule inhibiting vascular disrupting agents (VDAs), and certain chemotherapy drugs, e.g., paclitaxel, the levels of CEPs rise in peripheral blood within a matter of hours, after which they subsequently home to the tumor site and incorporate into or around the lumen of tumor blood vessels (Shaked et al. 2006b; Shaked et al. 2008). In the case of VDAs, these drugs cause rapid tumor vessel ischemia and occlusion followed by massive tumor hypoxia and necrosis. However, a viable and perfused rim tumor tissue usually remains from which tumor re-growth rapidly resumes (Tozer et al. 2005; Siemann et al. 2004), similar in some respects to rapid tumor cell repopulation observed after MTD chemotherapy-induced tumor shrinkage (Kim and Tannock 2005). Since the mechanisms which underline tumor cell repopulation after cytotoxic (or cytotoxic-like) drug therapy are poorly understood, we sought to investigate whether CEPs contribute to tumor re-growth from the viable tumor rim. We found that mice treated with VDAs, e.g., combretastatin A 4 phosphate (CA4-P) or OXi-4503, a second generation (more potent derivative) of CA4-P, exhibited substantial increases in CEPs within 4h. Subsequently, these cells home to the tumor site in large numbers and promote angiogenesis which accompanies the rapid tumor growth from the viable rim. An example of the massive homing and invasion of BMDCs, including EPCs (which incorporate into the lumen of the blood vessel wall), after therapy with a VDA is presented in Fig. 12.2. We also demonstrated that when an antiangiogenic treatment (i.e., DC101) which can block CEP mobilization (Shaked et al. 2005a; Shaked et al. 2006b) was coadministered with the VDA, or when tumor bearing mice that are deficient in CEP mobilization (i.e., Id-1+/-Id-3-/- mice) (Lyden et al. 2001) are used, a minimal residual viable tumor rim was observed after OXi-4503 treatment. Importantly, this phenomenon was not restricted to VDAs, as it was also observed following administration of other (but not all) cytotoxic chemotherapeutic drug therapies. Thus we found that certain chemotherapy drugs administered near or at the MTD, e.g., paclitaxel can induce a rapid spike in CEP levels but, again, this can be largely blocked by DC101 which enhanced treatment impact. However, other chemotherapy drugs, e.g., gemcitabine, which does not induce such a CEP spike, the addition of DC101, had no additional therapeutic benefit (Bertolini et al. 2006; Shaked et al. 2008). Thus, these observations indicate that in addition to the anti-tumor activity of chemotherapy, the subsequent tumor growth that occurs may be facilitated by a rapid influx of CEPs. The administration of an antiangiogenic drug (or other treatment strategies (Shaked et al. 2005b)) concomitantly with chemotherapy that can block
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
267
Fig. 12.2 Colonization of GFP± bone marrow cells in LLC tumors grown in mice treated with OXi-4503, DC101, or the combination of the two drugs. Lewis Lung carcinoma grown in lethally irradiated mice that were previously transplanted with GFP+ bone marrow tagged cells (green) were treated with DC101, the VEGFR-2 blocking antibody, OXi-4503 or a combination of the two drugs. Three days later, tumors were removed and stained for blood vessels using the CD31 endothelial cell marker (red). Untreated mice revealed some GFP+ BMDCs that did not directly incorporated into the tumor blood vessels. In contrast, massive invasion of BMDCs to the viable tumor rim was observed in mice treated with OXi-4503, some of which were incorporated into the tumor blood vessel lumen. The combination of DC101 and OXi-4503 resulted in the absence of some BMDCs reside at the tumor viable rim. Bottom panels demonstrate the incorporation of BMDCs into the tumor vasculature (co-localization of CD31 and GFP), but also a number of other BMDCs which are adjacent to the tumor vessels and may support tumor angiogenesis in a paracrine manner. Scale bars: Upper 50 µm; bottom 20 µm. Reproduced from Shaked et al. (2006b), with permission from the publisher
tumor cell repopulation mediated by rebound angiogenesis can result in prolongation of the duration of tumor response. These results may also suggest how antiangiogenic drugs may act to enhance the efficacy of standard chemotherapy (Shaked and Kerbal 2007; Kerbal 2006). So how do these results provide an entirely new perspective regarding the ongoing controversy of the contribution of CEPs to tumor angiogenesis? The answer is that all previously published studies showing very low or non-existent levels of EPC incorporation in tumor vessels were performed on untreated tumors, which were removed at various stages of tumor growth. In contrast, there are circumstances, e.g., after acute cytotoxic therapy, where our results show that CEPs may play a significant role in tumor angiogenesis and tumor (re)growth; thus, targeting these cells at these critical time points may lead to significant therapeutic benefits. The robust and acute CEP host responses induced by drugs such as VDAs or paclitaxel may be a consequence, at least in part, of an attempt to repair damage to the tumor vasculature caused by such drugs. We should stress, however, that many other BMDC populations may also contribute to the overall host response and tumor vessel repair process following cytotoxic drug therapy, i.e., this phenomenon
268
R.S. Kerbel et al.
in all likelihood is not restricted to CEPs. It will be of great interest to assess the possibility that cytotoxic therapies, including radiation, induce the mobilization and tumor homing of other BMDC populations such as monocytes and macrophages (Brown et al. 2010). Another circumstance which implicates an important role of EPCs in tumor growth and angiogenesis even when they are present in very limited numbers comes from a study by Mittal and colleagues (Gao et al. 2008). Since the development of macroscopic metastases from micrometastases also requires angiogenesis (Woodhouse et al. 1997), Gao et al. (2008) recently studied whether EPCs are crucial regulators of the “angiogenic switch” mediating progression of metastasis. By using two models of pulmonary metastasis in mice, they demonstrated that the vessels of macrometastases contained approximately 11% EPCs. They found that this was regulated by the Id1 transcription factor (Shaked et al. 2006b), since the blockade of Id1 by short hairpin RNA (shRNA) resulted in impaired angiogenesis, and lack of formation of macrometastases. Importantly, this study not only demonstrated a role of EPCs in metastatic growth, but also indicated that even a relatively small percentage of EPC incorporation into metastatic blood vessels (~11%), can nevertheless induce a major biologic effect in tumor biology. As such, pharmacologically targeting these relatively rare cells may be therapeutically promising. The rapid and robust increase in CEP levels following cytotoxic drug therapy has been recently confirmed in clinical studies. With respect to VDAs, in a phase I clinical study patients treated with cisplatinum in combination with the tubulin binding VDA, AVE8062, exhibited increases in the number of CEPs within the first 3 days (Farace et al. 2007). In this study, CEPs were evaluated by flow cytometry and were defined by the surface markers CD45(dim), VEGFR2+, and CD34+. Similar results were reported based on a phase I clinical study of the VDA, ZD6126, conducted by Beerepoot et al. (2006). The authors measured levels of CECs (but not CEPs), and found that levels of such cells are significantly elevated within 2–8h after therapy. The authors suggested that the rise in CECs levels is probably a consequence of endothelial cell shedding from the damaged tumor vasculature in response to VDA therapy. Although the authors’ conclusion focused on the levels of CECs, no precursor marker for CEP subset was evaluated at that time, and it is plausible that some of the cells detected by flow cytometry are in fact CEPs (Shaked et al. 2006). Furthermore, an additional clinical study testing the combination of a VDA with an antiangiogenic drug has been conducted in a phase I trial involving combretastatin A 4 phosphate (CA4P) and bevacizumab (Nathan et al. 2008). Preliminary results have been presented by Nathan and colleagues (2008) who reported that the combination treatment revealed statistically significant reductions in tumor perfusion/vascular permeability which reversed after CA4P alone but which were sustained following bevacizumab treatment using DCE-MRI. In addition they observed increases in circulating CD34+ and CD133+ bone marrow progenitors following CA4P treatment as well as increases in VEGF and GCSF levels, in a similar manner to a recent preclinical study (Shaked et al. 2009). Nathan et al. (2008) concluded that CA4P in combination with bevacizumab appears safe and well tolerated with evidence of clinical activity. With respect to chemotherapy,
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
269
we reported that cancer patients treated with paclitaxel based therapy but not gemcitabine- or doxorubicin-based therapies exhibited increases in CEP levels along with SDF-1 and other proangiogenic factors (Shaked et al. 2008; Shaked et al. 2009). In addition, Voest and colleagues have also reported that MTD paclitaxel exhibit an immediate increase in CEP levels 4h after treatment was initiated, but these immediate changes did not correlate with response to therapy. However, sustained and consistent high levels of CEPs measured on day 7 (regardless of the chemotherapy drug used) were associated with an increase in tumor growth and could predict progression free and/or overall survival (Roodhart et al. 2010). Overall, these results further reinforce the concept of rapid cytotoxic drug induced host CEP mobilization, and its blockade by an antiangiogenic drug which results in increased treatment efficacy.
Conclusions The role of EPCs in post-natal angiogenesis, and specifically during tumor growth remains controversial. Although many studies have shown vessel incorporation and a functional role of EPCs in tumor biology, other studies have failed to reproduce such results. We and others have shown that in certain circumstances EPCs may play a major role in tumor angiogenesis, mainly with respect to repair of therapy-induced tumor vascular damage induced by certain drugs, e.g., VDAs and MTD chemotherapy. In such circumstances we have found that EPCs rapidly home to the tumor site and promote an active angiogenesis process. Clinically, EPCs and CECs have been investigated as possible surrogate biomarkers to predict clinical outcome in patients undergoing antiangiogenic therapies. However, such measurements require further evaluation and validation using sophisticated flow cytometry technology. Overall, a consensus regarding the nature of EPCs and their relative functional importance in tumor angiogenesis and anti-cancer therapy remains to be determined.
References Asahara T, Murohara T, Sullivan A et al (1997) Isolation of putative progenitor endothelial cells for angiogenesis. Science 275:964–967. Beerepoot LV, Radema SA, Witteveen EO et al (2006) Phase I clinical evaluation of weekly administration of the novel vascular-targeting agent, ZD6126, in patients with solid tumors. J Clin Oncol 24:1491–1498. Bender JL, Adamson PC, Reid JM et al (2008) Phase I trial and pharmacokinetic study of bevacizumab in pediatric patients with refractory solid tumors: a Children’s Oncology Group Study. J Clin Oncol 26:399–405. Bertolini F, Paul S, Mancuso P et al (2003) Maximum tolerable dose and low-dose metronomic chemotherapy have opposite effects on the mobilization and viability of circulating endothelial progenitor cells. Cancer Res 63:4342–4346.
270
R.S. Kerbel et al.
Bertolini F, Shaked Y, Mancuso P, Kerbel RS (2006) The multifaceted circulating endothelial cell in cancer: from promiscuity to surrogate marker and target identification. Nature Rev Cancer 6:835–845. Brown M et al (2010), J Clin Investig (in press). Buckstein R, Kerbel RS, Shaked Y et al (2006) High-Dose celecoxib and metronomic “low-dose” cyclophosphamide is an effective and safe therapy in patients with relapsed and refractory aggressive histology non-Hodgkin’s lymphoma. Clin Cancer Res 12:5190–5198. Calleri A, Bono A, Bagnardi V et al (2009) Predictive Potential of Angiogenic Growth Factors and Circulating Endothelial Cells in Breast Cancer Patients Receiving Metronomic Chemotherapy Plus Bevacizumab. Clin Cancer Res 15:7652–7657. Case J, Mead LE, Bessler WK et al (2007) Human CD34+AC133+VEGFR-2+ cells are not endothelial progenitor cells but distinct, primitive hematopoietic progenitors. Exp Hematol 35:1109–1118. Coffelt SB, Hughes R, Lewis CE (2009) Tumor-associated macrophages: effectors of angiogenesis and tumor progression. Biochim Biophys Acta 1796:11–18. Cubbon RM, Kahn MB, Wheatcroft SB (2009) Effects of insulin resistance on endothelial progenitor cells and vascular repair. Clin Sci (Lond)117:173–190. Dellapasqua S, Bertolini F, Bagnardi V et al (2008) Metronomic Cyclophosphamide and Capecitabine Combined With Bevacizumab in Advanced Breast Cancer. J Clin Oncol 26(30):4899-4905. De Palma M, Naldini L (2006) Role of haematopoietic cells and endothelial progenitors in tumour angiogenesis. Biochim Biophys Acta 1766:159–166. De Palma M, Venneri MA, Galli R et al (2005) Tie2 identifies a hematopoietic lineage of proangiogenic monocytes required for tumor vessel formation and a mesenchymal population of pericyte progenitors. Cancer Cell 8:211–226. Distler JH, Beyer C, Schett G, Luscher TF, Gay S, Distler O (2009) Endothelial progenitor cells: novel players in the pathogenesis of rheumatic diseases. Arthritis Rheum 60:3168–3179. Duda DG, Cohen KS, di Tomaso E et al (2006) Differential CD146 expression on circulating versus tissue endothelial cells in rectal cancer patients: implications for circulating endothelial and progenitor cells as biomarkers for antiangiogenic therapy. J Clin Oncol 24:1449–1453. Elshal MF, Khan SS, Takahashi Y, Solomon MA, McCoy JPJr (2005) CD146 (Mel-CAM), an adhesion marker of endothelial cells, is a novel marker of lymphocyte subset activation in normal peripheral blood. Blood 106:2923–2924. Farace F, Massard C, Borghi E, Bidart JM, Soria JC (2007) Vascular disrupting therapy-induced mobilization of circulating endothelial progenitor cells. Ann Oncol 18:1421–1422. Gammill HS, Lin C, Hubel CA (2007) Endothelial progenitor cells and preeclampsia. Front Biosci 12:2383–2394. Gao D, Nolan DJ, Mellick AS, Bambino K, McDonnell K, Mittal V (2008) Endothelial progenitor cells control the angiogenic switch in mouse lung metastasis. Science 319:195–198. Grunewald M, Avraham I, Dor Y et al (2006) VEGF-induced adult neovascularization: recruitment, retention, and role of accessory cells. Cell 124:175–189. Hanahan D, Bergers G, Bergsland E (2000) Less is more, regularly: metronomic dosing of cytotoxic drugs can target tumor angiogenesis in mice. J Clin Invest 105:1045–1047. Horrevoets AJ (2009) Angiogenic monocytes: another colorful blow to endothelial progenitors. Am J Pathol 174:1594–1596. Jin DK, Shido K, Kopp HG et al (2006) Cytokine-mediated deployment of SDF-1 induces revascularization through recruitment of CXCR4+ hemangiocytes. Nat Med 12:557–567. Kaplan RN, Psaila B, Lyden D (2007) Niche-to-niche migration of bone-marrow-derived cells. Trends Mol Med 13:72–81. Kerbel RS (1991) Inhibition of tumor angiogenesis as a strategy to circumvent acquired resistance to anti-cancer therapeutic agents. BioEssays 13:31–36. Kerbel RS (2006) Antiangiogenic therapy: a universal chemosensitization strategy for cancer? Science 312:1171–1175.
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
271
Kerbel RS, Benezra R, Lyden DC et al (2008) Endothelial progenitor cells are cellular hubs essential for neoangiogenesis of certain aggressive adenocarcinomas and metastatic transition but not adenomas. Proc Natl Acad Sci U S A 105:E54. Kim I, Yilmaz OH, Morrison SJ (2005) CD144 (VE-cadherin) is transiently expressed by fetal liver hematopoietic stem cells. Blood 106:903–905. Kim JJ, Tannock IF (2005) Repopulation of cancer cells during therapy: an important cause of treatment failure. Nat Rev Cancer 5:516–525. Kim SJ, Kim JS, Papadopoulos J et al (2009) Circulating monocytes expressing CD31: implications for acute and chronic angiogenesis. Am J Pathol 174:1972–1980. Klement G, Baruchel S, Rak J et al (2000) Continuous low-dose therapy with vinblastine and VEGF receptor-2 antibody induces sustained tumor regression without overt toxicity. J Clin Invest 105:R15–R24. Lawler J (2002) Thrombospondin-1 as an endogenous inhibitor of angiogenesis and tumor growth. J Cell Mol Med 6:1–12. Li H, Gerald WL, Benezra R (2004)Utilization of bone marrow-derived endothelial cell precursors in spontaneous prostate tumors varies with tumor grade. Cancer Res 64:6137–6143. Lyden D, Hattori K, Dias S et al (2001) Impaired recruitment of bone-marrow-derived endothelial and hematopoietic precursor cells blocks tumor angiogenesis and growth. Nat Med 7:1194–1201. Lyden D, Young AZ, Zagzag D et al (1999) Id1 and Id3 are required for neurogenesis, angiogenesis and vascularization of tumour xenografts. Nature 401:670–677. Machein MR, Renninger S, Lima-Hahn E, Plate KH (2003) Minor contribution of bone marrowderived endothelial progenitors to the vascularization of murine gliomas. Brain Pathol 13:582–597. Madlambayan GJ, Butler JM, Hosaka K et al (2009) Bone marrow stem and progenitor cell contribution to neovasculogenesis is dependent on model system with SDF-1 as a permissive trigger. Blood 114:4310–4319. Mancuso P, Antoniotti P, Quarna J et al (2009) Validation of a standardized method for enumerating circulating endothelial cells and progenitors: flow cytometry and molecular and ultrastructural analyses. Clin Cancer Res 15:267–273. Mancuso P, Burlini A, Pruneri G, Goldhirsch A, Martinelli G, Bertolini F (2001) Resting and activated endothelial cells are increased in the peripheral blood of cancer patients. Blood 97:3658–3661. Mancuso P, Colleoni M, Calleri A et al (2006) Circulating endothelial cell kinetics and viability predict survival in breast cancer patients receiving metronomic chemotherapy. Blood 108:452–459. Martin-Padura I, Bertolini F (2008) The circulating endothelial cell in cancer: towards marker and target identification. Curr Pharm Des 14:3780–3789. Nathan PD, Judson I, Padhani A et al (2008) A phase I study of combretastatin A4 phosphate (CA4P) and bevacizumab in subjects with advanced solid tumors. J Clin Oncol 26(15S, Abstract No. 3550), —20 May 2008. Peichev M, Naiyer AJ, Pereira D et al (2000) Expression of VEGFR-2 and AC133 by circulating human CD34(+) cells identifies a population of functional endothelial precursors. Blood 95:952–958. Peters BA, Diaz LA, Polyak K et al (2005) Contribution of bone marrow-derived endothelial cells to human tumor vasculature. Nat Med 11:261–262. Pitchford SC, Furze RC, Jones CP, Wengner AM, Rankin SM (2009) Differential mobilization of subsets of progenitor cells from the bone marrow. Cell Stem Cell 4:62–72. Prewett M, Huber J, Li Y et al (1999) Antivascular endothelial growth factor receptor (fetal liver kinase 1) monoclonal antibody inhibits tumor angiogenesis and growth of several mouse and human tumors. Cancer Res 59:5209–5218. Purhonen S, Palm J, Rossi D et al (2008) Bone marrow-derived circulating endothelial precursors do not contribute to vascular endothelium and are not needed for tumor growth. Proc Natl Acad Sci U S A 105:6620–6625.
272
R.S. Kerbel et al.
Rafii S, Lyden D, Benezra R, Hattori K, Heissig B (2002) Vascular and haematopoietic stem cells: novel targets for anti-angiogenesis therapy? Nat Rev Cancer 2:826–835. Rohan RM, Fernandez A, Udagawa T, Yuan J, D’Amato RJ (2000) Genetic heterogeneity of angiogenesis in mice. FASEB J 14:871–876. Roodhart JM, Langenberg MH, Vermaat JS et al (2010) Late release of circulating endothelial cells and endothelial progenitor cells after chemotherapy predicts response and survival in cancer patients. Neoplasia 12:87–94. Ruzinova MB, Schoer RA, Gerald W et al (2003) Effect of angiogenesis inhibition by Id loss and the contribution of bone-marrow-derived endothelial cells in spontaneous murine tumors. Cancer Cell 4:277–289. Schneider M, Tjwa M, Carmeliet P (2005) A surrogate marker to monitor angiogenesis at last. Cancer Cell 7:3–4. Shaked Y, Bertolini F, Emmenegger U, Lee CR, Kerbel RS. (2006a) On the origin and nature of elevated levels of circulating endothelial cells after treatment with a vascular disrupting agent. J Clin Oncol 24:4040–4041. Shaked Y, Bertolini F, Man S et al (2005a) Genetic heterogeneity of the vasculogenic phenotype parallels angiogenesis: implications for cellular surrogate marker analysis of antiangiogenesis. Cancer Cell 7:101–111. Shaked Y, Ciarrocchi A, Franco M et al (2006b) Therapy-induced acute recruitment of circulating endothelial progenitor cells to tumors. Science 313:1785–1787. Shaked Y, Emmenegger U, Francia G et al (2005b) Low-dose metronomic combined with intermittent bolus-dose cyclophosphamide is an effective long-term chemotherapy treatment strategy. Cancer Res 65:7045–7051. Shaked Y, Emmengger U, Man S et al (2005c) The optimal biological dose of metronomic chemotherapy regimens is associated with maximum antiangiogenic activity. Blood 106:3058–3061. Shaked Y, Henke E, Roodhart JM et al (2008) Rapid chemotherapy-induced acute endothelial progenitor cell mobilization: implications for antiangiogenic drugs as chemosensitizing agents. Cancer Cell 14:263–273. Shaked Y, Kerbel RS (2007) Antiangiogenic strategies on defense: on the possibility of blocking rebounds by the tumor vasculature after chemotherapy. Cancer Res 67:7055–7058. Shaked Y, Tang T, Woloszynek J et al (2009) Contribution of granulocyte colony-stimulating factor to the acute mobilization of endothelial precursor cells by vascular disrupting agents. Cancer Res 69:7524–7528. Shaked Y, Voest EE (2009) Bone marrow derived cells in tumor angiogenesis and growth: are they the good, the bad or the evil? Biochim Biophys Acta 1796:1–4. Shojaei F, Wu X, Zhong C et al (2007) Bv8 regulates myeloid-cell-dependent tumour angiogenesis. Nature 450:825–831. Siemann DW, Chaplin DJ, Horsman MR (2004) Vascular-targeting therapies for treatment of malignant disease. Cancer 100:2491–2499. Solovey AN, Gui L, Chang L, Enenstein J, Browne PV, Hebbel RP (2001) Identification and functional assessment of endothelial P1H12. J Lab Clin Med 138:322–331. Tozer GM, Kanthou C, Baguley BC (2005) Disrupting tumour blood vessels. Nat Rev Cancer 5:423–435. Volpert OV, Pili R, Sikder HA et al (2002) Id1 regulates angiogenesis through transcriptional repression of thrombospondin-1. Cancer Cell 2:473–483. Werner N, Kosiol S, Schiegl T et al (2005) Circulating endothelial progenitor cells and cardiovascular outcomes. N Engl J Med 353:999–1007. Wickersheim A, Kerber M, de Miguel LS, Plate KH, Machein MR (2009) Endothelial progenitor cells do not contribute to tumor endothelium in primary and metastatic tumors. Int J Cancer 125:1771–1777. Woodhouse EC, Chuaqui RF, Liotta LA (1997) General mechanisms of metastasis. Cancer 80:1529–1537.
12 Impact of Endothelial Progenitor Cells on Tumor Angiogenesis and Outcome
273
Yap R, Veliceasa D, Emmenegger U et al (2005) Metronomic low-dose chemotherapy boosts CD95-dependent antiangiogenic effect of the thrombospondin peptide ABT-510: a complementation antiangiogenic strategy. Clin Cancer Res 11:6678–6685. Yoder MC, Mead LE, Prater D et al (2007) Redefining endothelial progenitor cells via clonal analysis and hematopoietic stem/progenitor cell principals. Blood 109:1801–1809. Ziegelhoeffer T, Fernandez B, Kostin S et al (2004) Bone marrow-derived cells do not incorporate into the adult growing vasculature. Circ Res 94:230–238.
Chapter 13
Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth Pravin J. Mishra and Debabrata Banerjee
Abstract Carcinoma associated fibroblasts (CAFs) play an important role in the growth of epithelial solid tumors. The origin of these tumor or CAFs has not been conclusively established. There is experimental evidence to suggest that part of the tumor or CAFs may arise from bone marrow derived mesenchymal stromal/stem cells or MSCs. It is well known that bone marrow derived MSCs can give rise to cells of different lineages: muscle, bone, fat, and cartilage. Based on recent work from our own laboratory and that of others, we now suggest that human BM-derived MSCs exposed to tumor-conditioned medium (TCM) over a prolonged period of time can give rise to cells that assume a CAF-like phenotype. Thus, MSCs may be a source of CAFs and can be used experimentally for modeling tumor-stroma interactions. Although the importance of the dialog between cancer cells and other components of the tumor milieu has been increasingly appreciated, it is as yet unclear whether the stromal cells themselves harbor cancer promoting mutations or changes. Activated stromal cells have been shown to promote tumor growth and metastasis in experimental models and we speculate on the possibility of increased activation of bone marrow derived MSCs by higher levels of chemokines under certain physiological situations and how this may impact tumor growth.
D. Banerjee (*) Department of Medicine, The Cancer Institute of New Jersey, Robert Wood Johnson Medical School, University of Medicine and Dentistry of New Jersey, 195 Little Albany Street, New Brunswick, NJ 08901, USA and Department of Pharmacology, The Cancer Institute of New Jersey, Robert Wood Johnson Medical School, University of Medicine and Dentistry of New Jersey, 195 Little Albany Street, New Brunswick, NJ 08901, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_13, © Springer Science+Business Media, LLC 2010
275
276
P.J. Mishra and D. Banerjee
Introduction Most solid tumors are made up of two discernable compartments, tumor or neoplastic cells and the stromal platform on which the tumor cells are scattered. Besides these, there are many other cell types present in the tumor milieu including tumor associated macrophages (TAMs), lymphocytes, endothelial cells, fibroblasts, and pericytes that interact with each other either directly or via paracrine mechanisms and influence tumor growth. The origin and function of tumor stroma has been the subject of intense investigation ever since Virchow first observed that cancers were infiltrated with inflammatory and immune cells (Balkwill and Mantovani 2001). The presence of fibroblast populations within human tumors often referred to as carcinoma associated fibroblasts (CAFs) has been associated with poor outcome and an increase in metastatic potential (Tsujino et al. 2007; Yazhou et al. 2004). Experimental evidence from tumor models as well as from clinical specimens suggest that CAFs may promote tumor growth and metastasis. In this chapter we will discuss the likely source of these CAFs as well as their characteristics and hope to convince the readers that bone marrow derived mesenchymal stem/stromal cells (MSCs) may be a likely source of CAFs. We will also discuss how this can be an important model system to study tumor-stroma interaction. Lastly, we will speculate on the possibility of increased activation of bone marrow derived MSCs by higher levels of chemokines in circulation under certain physiological situations and how this may impact tumor growth.
Characteristics of CAFs Over the years, characterization of CAFs has been largely based on immunohistochemical staining and more recently by specific gene expression signatures (Allinen et al. 2004; Hu and Polyak 2008; Chang et al. 2005). These have relied upon identification and or isolation of the CAFs from tumor specimens. We and others have developed in vitro methods to generate CAFs and study them in model systems to understand the role of these CAFs in tumor growth and metastasis as well as to better define paracrine factors that participate in the tumor stroma dialog (Mishra et al. 2008; Spaeth et al. 2009). Advantages of an in vitro system include better control over cell type, identification and characterization of CAFs and generation of large number of cells in culture to carry out studies in various model systems including xenograft studies. In order to generate CAFs (also referred to as tumor associated fibroblasts or TAFs) in vitro, long term exposure to tumor cell conditioned medium has been the method of choice. Spaeth et al. (2009) have defined the tumor or CAFs involved in ovarian cancer growth (as well as breast and prostate) by expression of four qualifying factors: (1) fibroblast markers fibroblast-specific protein (FSP)
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
277
and fibroblast activating protein (FAP); (2) genes associated with an increase of tumor aggression, including stromelysin-1 (SL-1), thrombospondin-1 (Tsp-1), and tenascin-C (Tn-C); (3) myofibroblast/provascularizing potential including desmin, a-smooth muscle actin (a-SMA), and vascular endothelial growth factor (VEGF); and lastly, (4) growth factors, transforming growth factor-beta (TGF-b), HGF/ scatter factor (SF), basic fibroblast growth factor (bFGF), and epidermal growth factor (EGF). Bone marrow derived MSCs have been suggested to play a major role in tissue regeneration and repair based on their multilineage differentiation potential (Pittenger et al. 1999; Picinich et al. 2007) as well as the propensity of these cells to migrate to areas of injury, inflammation and tumor (Fox et al. 2007). Studeny et al. (2002) originally demonstrated the capacity of MSCs, also known as multipotent stromal cells, to home to tumors and participate in tumor stroma formation, suggesting MSCs as a potential source of stroma. The observed tendency of MSCs to migrate to sites of injury, inflammation and tumors is attributed to common chemoattractants or factors present in the wound and tumor milieu leading to the concept that tumors are unhealed wounds (Dvorak 1986). Our own work on role of MSCs in wound healing has revealed remarkable similarities in the process of wound healing and tumor growth with some subtle differences such as in production of cytokines by MSCs in the wound bed or in the presence of keratinocytes, the principal epithelial cells in skin (Mishra et al. unpublished observations). The ability of MSCs to migrate to tumor sites has been exploited to deliver therapeutic genes to tumor sites using MSCs as carriers. (Studeny et al. 2002, 2004; Nakamizo et al. 2005; Hung et al. 2005).
Bone Marrow Derived MSCs as Source of CAFs Accumulating evidence suggests that CAFs or TAFs play an important role in the growth of epithelial solid tumors. It has been demonstrated that a significant fraction of the stroma in some breast cancers consists of fibroblasts (Bissell and Radisky 2001). More recent studies have revealed that CAFs from breast cancer specimens promote tumor cell growth as compared to fibroblasts obtained from non-neoplastic locations. In addition to tumor growth, tumor stroma has also been shown to be involved in other important processes such as angiogenesis and metastasis. Orimo and colleagues (2005) defined several important characteristics of breast CAFs including promotion of breast carcinoma cell growth, promotion of angiogenesis, and expression of myofibroblast traits. Expression of the chemokine stromal-derived factor 1 (SDF-1) has also been shown to be of paramount importance in the interaction between tumor cells and stromal fibroblasts (Orimo et al. 2005). Given the importance of tumor stroma in promoting tumor growth, it becomes necessary to identify the likely source(s) of cells that make up the tumor stroma, particularly the carcinoma associated fibroblasts or CAFs.
278
P.J. Mishra and D. Banerjee
It has been suggested that the likely source (s) of tumor stroma include (1) the locally resident tissue stem cells that are recruited to the tumor site; (2) epithelial to mesenchymal transition or EMT; (3) fibroblast recruitment into the tumor stroma; and (4) recruitment of bone marrow-derived cells from the circulation (Udagawa et al. 2006; Koyama et al. 2008; Jodele et al. 2005). We and others have demonstrated that bone marrow derived MSCs, like other bone marrow-resident cells, have the capacity to differentiate within the tumor microenvironment into myofibroblastic-like cells commonly referred to as; myofibroblasts, tumor-associated (myo)fibroblasts (TAF), CAF, fibrocytes or pericytes (Ogawa et al. 2006; Mishra et al. 2008; Spaeth et al. 2009) which have been shown to play an important role in tumor formation, growth and metastasis. Although the cell type of origin of myofibroblasts has not been conclusively established it has been shown that they may be bone marrow derived (Direkze et al. 2004). In a recent study we demonstrated that human bone marrow-derived mesenchymal stem cells (hMSCs) when exposed to tumor-conditioned medium (TCM) over a prolonged period of time assume a CAF-like phenotype (Mishra et al. 2008) and this was confirmed by Spaeth et al. (2009). More importantly, these cells exhibit functional properties of CAFs including sustained expression of SDF-1 and the ability to promote tumor cell growth both in vitro and in an in vivo co-implantation model. These cells also express myofibroblast markers including a-SMA and FSP. Gene expression profiling revealed similarities between TCM exposed hMSCs and CAFs. This suggests that hMSCs are a source of CAFs and can be used in modeling tumor-stroma interactions (Mishra et al. 2008; Karnoub et al. 2007). Further evidence for a bone marrow source of CAFs comes from studies using a gastric cancer mouse model (Gan mice) in which prostaglandin E2 (PG E2) and Wnt signaling were simultaneously activated in the gastric mucosa (Guo et al. 2008). Since PGE2 and Wnt pathways both play a role in human gastric tumorigenesis, the Gan mouse model may display important aspects of the molecular etiology of human gastric cancer. Bone marrow transplantation (BMT) experiments revealed that subsets of gastric myofibroblasts were derived from bone marrow (Guo et al. 2008). In a recent report, Worthley et al. (2009) have shown for the first time that human gastrointestinal neoplasia associated myofibroblasts can develop from bone marrow derived cells following allogeneic BMT based on identification of SMA expressing, Y chromosome positive CD45 negative cells in the tumor milieu of female recipients of BMT who went on to develop solid tumors following allogeneic male bone marrow/stem cells transplants. This clinical study supports the previous in vitro as well as in vivo animal model studies that suggested bone marrow origin of CAFs (Ogawa et al. 2006). Although literature supports the bone marrow origin of CAFs, alternate sources including resident fibroblasts and adipose tissue cells are also sources of CAFs. In a recent report Zhang et al. (2009) demonstrate that white adipose tissue derived stromal cells are also recruited to tumor stroma and promote cancer progression in animal models of human cancers including Kaposi’s sarcoma, prostate and breast cancer models.
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
279
Alterations in Tumor Associated Stromal Cells Although the importance of the dialog between cancer cells and other components of the tumor milieu has been increasingly appreciated, the question of whether the stromal cells themselves harbor cancer promoting mutations or changes is just beginning to be addressed. For example, Patocs et al. (2007) hypothesized that mutational inactivation of the tumor-suppressor gene TP53 and genomic alterations in stromal cells of a tumor’s microenvironment may be correlated to clinical outcome. The authors, in an analysis of somatic and stromal cell mutations, demonstrated that stroma-specific loss of heterozygosity or allelic imbalance was associated with somatic p53 mutations and regional lymph-node metastases in sporadic breast cancer but not in hereditary breast cancer. A previous study in head and neck squamous cell carcinoma (SCC) by the same investigators looked at whether the apparently “normal” stroma surrounding the tumor epithelium can acquire genomic alterations and contribute to cancer initiation and or progression. Tumor-associated stroma of head and neck SCC from smokers was found to have a high degree of genomic alterations. The results indicated that stroma-specific genetic alterations may have been associated with smoking-related head and neck SCC genesis (Weber et al. 2007). Whether the identified mutations correlate with stromal changes or whether this is a reflection of methodological artifacts remains unresolved and larger studies on tumor associated stromal cells from a variety of tumor specimens may be helpful in settling this contentious issue (Eng et al. 2009; Campbell et al. 2009). It has become clear that the initiation and progression of carcinomas depend not only on alterations in tumor epithelial cells, but also on changes in their microenvironment. To study changes in stromal cells within the tumor milieu Hasawi et al. (2008) undertook to characterize CAFs and their tumor counterpart fibroblasts (TCFs) at the cellular and molecular level in a small subset of breast cancer patients using normal breast fibroblasts (NBFs) from plastic surgery as a control. The results suggested that the p53/p21 response to gamma-radiation was attenuated in 70% of CAFs, whereas it was normal in all the TCF and NBF cells. These results indicate that alterations in the p53 pathway can occur in breast CAFs and their corresponding adjacent counterparts, further pointing to the important role that stroma may play in breast carcinogenesis and treatment. Alterations in the tumor suppressor p53 have been reported in tumor-associated stromal cells, however, the consequence of these alterations are not well understood. Dudley et al. (2008) have investigated p53 status and response to p53-activating drugs using tumor-associated stromal cells from A375 melanoma and PC3 prostate carcinoma xenografts, as well as from a spontaneous prostate tumor model (TRAMP). Unlike normal stromal cells, tumor associated stromal cells failed to arrest in G2 after etoposide treatment, failed to upregulate p53-inducible genes, and failed to undergo apoptosis after treatment with vincristine. Tumor-associated stromal cells were also found to be less sensitive to p53-activating drugs. Knockdown of p53 in normal stromal cells produced similar results strongly supporting the
280
P.J. Mishra and D. Banerjee
contention that there was loss of p53 response in tumor associated stromal cells. Investigations into the role of tumor associated stromal cells in supporting aerobic glycolysis (also called the Warburg effect) has revealed that activation of highly conserved mammalian uncoupling proteins through interaction with the tumor milieu may facilitate the Warburg effect in the absence of permanent respiratory impairment (Samudio et al. 2008). Additionally, tumor associated stromal cells have been reported to cooperate with tumor cells by taking up the lactate produced by tumor cells and secreted into the tumor microenvironment. The stromal cells take up the secreted lactate via monocarboxylate transporters (MCT1 and MCT2) and after conversion to pyruvate secrete it back into the extracellular milieu where it can be used by local cell constituents for oxidative phosphorylation (Samudio et al. 2008; Koukourakis et al. 2006). Our own observations indicate that exposure of bone marrow derived MSCs to 15 mM lactate for periods as short as 5 min results in upregulation of MCT1 protein as determined by western blotting (Rattigan and Banerjee 2009, unpublished data). Although little is known regarding how changes in stromal gene expression affect epithelial tumor progression, it is becoming clear that cancer is influenced by signals emanating from tumor stroma. Finak et al. (2008) studied gene expression profiles of tumor stroma from 53 primary breast tumors obtained by laser capture microdissection and were able to generate a novel stroma-derived prognostic predictor (SDPP) that correlated with disease outcome independently of standard clinical prognostic factors. The SDPP gene expression signature when applied to several previously published tumor-derived expression data sets was able to identify poor-outcome individuals from amongst multiple clinical subtypes, including lymph node-negative tumors. The SDDP signature appears to be an improved prognostic predictor compared to previously published methods, particularly for HER2-positive breast tumors. Genes showing strong prognostic tendencies included those associated with differential immune responses and angiogenic and hypoxic responses underscoring the importance of stromal biology in tumor progression (Kroemer and Pouyssegur 2008). The plasticity of both the epithelial tumor cells and bone marrow derived MSCs and its impact on tumor biology remain a subject of intense investigation. On the one hand tumor cells are known to undergo epithelial mesenchymal transition (Salomon and Thiery 2003; Radisky 2005) while on the other hand, MSCs are also capable of mesenchymal epithelial transition (Chaffer et al. 2006) thus adding to the complexity of cell types in the microenvironment.
Implications of MSCs as a Source of CAFs: A Model to Study Tumor Stroma Interactions In vitro generation of CAFs following exposure to TCM as well as the admixed injection of tumor cells and bone marrow derived MSCs hold great promise as systems where tumor stroma interaction can be manipulated and studied. Additionally, this may also provide a cell culture method for generating one of the important cell types of the tumor stroma, the activated myofibroblasts.
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
281
A better understanding of the dialog between various bone marrow derived cell types and tumor cells within the tumor milieu will be important for improved tumor therapy that takes into account influence of the tumor milieu on survival and growth of tumor cells. We have initiated studies to model the tumor-stroma interaction in vitro by culturing TCM exposed BMDMSCs (representing CAFs), reporter gene marked tumor cells (to report growth of tumor cells in the cell mix) and other components of the tumor milieu. It is now becoming clearer that CAFs and other cells in the tumor milieu actively participate in altering the growth and drug response of tumors (Iwamoto et al. 2007; Mantovani et al. 2006). For example, MSCs, which have very high levels of asparaginase expression, can protect leukemic cells from asparaginese cytotoxicity by providing increased concentrations of asparagine in the vicinity of leukemic cells. This system can be exploited to study the contribution of freshly harvested TAFs from dissected tumors on growth of a similar type of tumor cell. By establishing a panel of reporter gene marked tumor cells representing a variety of tumor types and subtypes, it may be possible to determine influence of TAFs on tumor growth. Specific reporters can also be generated for pathways that may be activated in different tumors, for example one can study whether androgen independent prostate cancer cells are still influenced by the tumor stroma. The relative ease with which the reconstituted tumor in its microenvironment can be transplanted as a xenograft may also permit improved drug studies in vivo. An important aspect of the reconstituted tumor microenvironment is the ability to evaluate chemopreventive measures in vitro and in vivo. By targeted pretreatment of critical components of the reconstituted tumor microenvironment, it may be possible to rank chemopreventive agents by potency as well as cell type specificity. Further applications of the reconstituted tumor milieu system include study of gap junctions and other direct communication means between tumor cells and other cell types. This type of experimental system may provide a better simulation of an in vivo solid tumor and may be a more realistic model for investigation of tumor biology and experimental therapeutics.
Activation of BMD MSCs and Growth of Tumors Speculation on Role of Chemokines on Activation of Circulating MSCs and Effect on Tumor Growth in African American Individuals with Breast Cancer Caucasian American women have higher age-adjusted breast cancer incidence rates compared with African American women (143 per 100,000 versus 119 per 100,000; Ries et al. 2002). However, African American women have higher age-adjusted mortality rates from the disease. Breast cancer mortality rates among younger African American women are approximately twice that of younger Caucasian American women (Amend et al. 2006; Albain et al. 2009). African American breast
282
P.J. Mishra and D. Banerjee
cancer patients also more frequently develop basal like subset of breast cancer (triple negative; ER-, PR- and Her2 low; greater stromal involvement), which has a poorer prognosis (van de Vijver et al. 2002; Sorlie et al. 2001). Gene expression studies reveal that pathways related to tumor angiogenesis and chemotaxis may be upregulated in tumors from African American women as compared with women of Caucasian origin. Increased microvessel density and increased TAMs in tumors from African American patients have been reported (Martin et al. 2009) suggesting involvement of the tumor microenvironment. A majority of African Americans are negative (~68% as compared to 3.4% of Caucasian Americans) for Duffy antigen receptor for chemokines (DARC) and it has been suggested that absence of this non-signaling receptor can increase migration of several types of blood cells from circulation to tissues (Tournamille et al. 1995; Fukuma et al. 2003; Lentsch 2002; Pruenster et al. 2009) supporting the idea of increased migration of activated MSCs to the tumor milieu. A mutation in the erythroid specific promoter GATA-1 binding site in the promoter of the gene encoding DARC results in lack of expression of DARC on erythrocytes. The erythroid DARC clears ELR chemokines such as CXCL8 from circulation while the endothelial DARC may facilitate transport of chemokines from tissue to vascular lumen for presentation (Lee et al. 2003). Overexpression of DARC in breast cancer cells correlates with a tumor and metastasis suppressor phenotype underscoring the importance of the DARC-negative phenotype in tumor progression. Wang et al. (2006) have shown a strong negative correlation between DARC expression and lymph node metastasis, ER status and microvessel density suggesting that low DARC expression correlated with aggressive disease in an analysis of 75 breast tumors. We have re-analyzed data compiled previously from two separate gene expression profiling studies and discovered that lower relative expression of DARC correlates with basal subtypes (unpublished data; analysis carried out in collaboration with Dr. S. Ganesan at CINJ from data published by Alexe et al. 2007; Wang et al. 2005). Our analysis suggests that lower relative expression of DARC correlates with basal subtypes of breast cancer. The basal subtypes, basal A1 and basal A2 had the lowest expression as compared with the luminal type as well as normal breast tissue. African American breast cancer patients more frequently develop basal subtype of breast cancer. Our analysis did not separate the population by ethnicity supporting an independent association between DARC negativity and the basal subtype. Lack of DARC expression in African American individuals correlates with elevated levels of free chemokines, such as CXCL8, and leads to pathological conditions as shown below. Lack of DARC Expression, Circulating Chemokines and Pathological Conditions It has been shown that DARC negative individuals have higher levels of CXCL8 in supernatants of red cell concentrates than DARC positive individuals suggesting that plasma levels of CXCL8 are higher in the former (Wadhwa et al. 2000).
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
283
Moreover, high levels of ELR chemokines such as CXCL8 are found in renal biopsies from patients with HIV associated renal neuropathies seen predominantly in African American children with HIV. It is postulated that individuals who lack the DARC sink for chemokines would have higher levels of circulating chemokines available for endothelial cells to bind and deliver to renal tissues and increase risk of renal injury (Liu et al. 1999). DARC Expression and Cancer in African American Men DARC has been found to act as a multispecific receptor for chemokines of both the C–C and C–X–C families. Using a transgenic model of prostate cancer with DARC-deficient mice, Shen et al. (2006) demonstrated that lack of DARC expression contributed to enhanced prostate tumor growth. Their studies showed that DARC functions to clear angiogenic CXC chemokines (ELR chemokines, the key one being CXCL8) from the prostate tumor microcirculation and that lack of DARC expression in 70% of African Americans may be a contributing factor to the increased mortality to prostate cancer in this population but not to tumor initiation. Although DARC and elevated chemokine levels were implicated, no mechanistic insight was provided by these studies. We have developed the idea further and have generated the novel hypothesis that increased circulating chemokines results in greater number of activated MSCs leading to aggressive tumor growth in DARC negative African American breast cancer patients. Recent data from Mayr et al. (2007) demonstrates that individuals of African origin have higher levels of CXCL8 than individuals of Caucasian origin. In the analysis from Mayr et al. (2007) 41 Caucasians were of European origin, whereas the remaining 31 individuals of sub-Saharan African origin. CXCL-8 levels averaged 5.7 pg/mL (CI 4.4–7.0) in Africans and 3.3 pg/mL (CI 2.7–3.8) in Caucasians (P< 0.0008*). These levels are consistent with previously determined values in Caucasians. This shows that the lack of erythroid DARC expression correlates with elevated CXCL8 level in plasma.
Activation of Bone Marrow-Derived MSCs and Metastasis Karnoub et al. (2007) demonstrated that bone-marrow-derived human MSCs increased metastasis potency of MDAMB231 (otherwise weakly metastatic human breast carcinoma cells), when this cell mixture was introduced subcutaneously in nude mice. The MSCs were stimulated by the breast cancer cells to secrete CCL5, which acted in a paracrine fashion on the cancer cells to enhance metastasis. Thus, MSCs as part of the tumor microenvironment promoted metastatic spread by eliciting reversible changes in the phenotype of cancer cells. The influence of MSCs on tumor growth and metastasis was recently described by Shinagawa et al. (2009) who examined the role of human MSCs in the
284
P.J. Mishra and D. Banerjee
Experimental model systems for Support Tumor growth studying tumor stroma interactions
Bone Marrow
CAFs Injected with cancer cell Therapeutic Potential /Regenerative medicine CAFs
TCM-Activated
MSCs
Differentiation
In DARC negative individuals Increased levels of circulating bioactive molecules such as CXCL8
Aggressive tumor growth ?
Fig. 13.1 Mesenchymal stem cells and tumor growth. TCM tumor conditioned medium, CAF carcinoma associated fibroblasts
tumor stroma using an orthotopic nude mice model of KM12SM colon cancer. Systemically injected MSCs migrated to the stroma of colon tumors, both orthotopically implanted as well as metastatic liver tumors. Orthotopically implanted KM12SM cells mixed with MSCs resulted in larger tumors than KM12SM cells alone suggesting that MSCs promoted tumor growth. Survival rate was significantly lower in the group of animals receiving admixed MSCs, and liver metastasis was seen only in this group. Tumors resulting from admixed cells had a significantly higher proliferating cell nuclear antigen labeling index, a significantly greater proportion of microvessels and a significantly lower apoptotic index. Splenic injection of KM12SM cells admixed with MSCs resulted in greater number of liver metastases than when KM12SM cells were injected alone. MSCs incorporated into the stroma of primary and metastatic tumors and expressed markers of carcinoma-associated fibroblasts (CAFs) such as a-SMA and platelet-derived growth factor receptor-b. In in vitro experiments, KM12SM cells attracted MSCs and MSCs stimulated migration and invasion of tumor cells. The results of studies carried out by Shinagawa et al. (2009) suggest that MSCs migrate and differentiate to CAFs in tumor stroma, and promote growth and metastasis of human colon cancer in a nude mouse model.
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
285
Conclusion Solid tumors are often surrounded by a variety of cell types including myofibroblasts also known as tumor TAFs or CAFs, tumor associated macrophages (TAMs), lymphocytes, and pericytes. We have discussed how BMD MSCs can be activated in vitro to generate myofibroblasts that resemble CAFs and the advantages of this system to model tumor stroma interaction. Bone marrow derived MSCs are suggested to be a source of tumor or CAFs and influence tumor growth and metastasis. There is emerging clinical data to suggest that in some instances tumor associated myofibroblasts may indeed be of bone marrow origin. We have also taken the liberty to speculate on the possibility that activation of MSCs derived from bone marrow under certain physiological conditions may lead to aggressive tumor growth such as seen in African American individuals with breast and prostate cancers (Fig. 13.1).
References Albain KS, Unger JM, Crowley JJ, Coltman CA Jr, Hershman DL (2009) Racial disparities in cancer survival among randomized clinical trials patients of the Southwest Oncology Group. J Natl Cancer Inst 101(14):984–992 Alexe G, Dalgin GS, Scanfeld D et al (2007) High expression of lymphocyte-associated genes in node-negative HER2+ breast cancers correlates with lower recurrence rates. Can Research 67:10669–10676 Allinen M, Beroukhim R et al (2004) Molecular characterization of the tumor microenvironment in breast cancer. Cancer Cell 6(1):17–32 Amend K, Hicks D, Ambrosone, CB (2006) Breast cancer in African-American women: differences in tumor biology from European American women. Cancer Res 66:8327–8330 Balkwill F, Mantovani A (2001) Inflammation and cancer: back to Virchow? Lancet 357:539–545 Bissell MJ, Radisky D. (2001) Putting tumors in context. Nature Rev Cancer 1:46–54 Campbell I, Polyak K, Haviv I (2009) Clonal mutations in the cancer associated fibroblasts: the case against genetic coevolution. Cancer Res 69:6765–6769 Chaffer CL, Brennan JP, Slavin JL et al (2006) Mesenchymal-to-epithelial transition facilitates bladder cancer metastasis: role of fibroblast growth factor receptor-2. Cancer Res 66:11271–11278 Chang HY, Nuyten DS, Sneddon JB et al (2005) Robustness scalability and integration of a wound response gene expression signature in predicting breast cancer survival. Proc Natl Acad Sci U S A 102(10): 3738–3743 Direkze NC, Hodivala-Dilke K, Jeffery R et al (2004) Bone marrow contribution to tumor-associated myofibroblasts and fibroblasts. Cancer Res 64:8492–8495 Dudley AC, Shih SC, Cliffe AR et al (2008) Attenuated p53 activation in tumor-associated stromal cells accompanies decreased sensitivity to etoposide and vincristine. Br J Cancer 99:118–125 Dvorak HF (1986) Tumors: wounds that do not heal. Similarities between tumor stroma generation and wound healing. N Engl J Med 315:1650–1659 Eng C, Leone G, Orloff MS, Ostrowski MC (2009) Genomic alterations in tumor stroma. Cancer Res 69:6759–6764 Finak G, Bertos N, Pepin F et al (2008) Stromal gene expression predicts clinical outcome in breast cancer. Nat Med 14:518–527
286
P.J. Mishra and D. Banerjee
Fox J M, Chamberlain G, Ashton BA, Middleton J (2007) Recent advances into the understanding of mesenchymal stem cell trafficking. Br J Haematol 137:491–502 Fukuma N, Akimitsu N, Hamamoto H, Kusuhara H, Sugiyama Y, Sekimizu K (2003) A role of the Duffy antigen for the maintenance of plasma chemokine concentrations. Biochem Biophys Res Commun 303:137–139 Guo X, Oshima H, Kitmura T et al (2008) Stromal fibroblasts activated by tumor cells promote angiogenesis in mouse gastric cancer. J Biol Chem 283:19864–19871 Hawsawi NM, Ghebeh H, Hendrayani SF et al (2008) Breast carcinoma-associated fibroblasts and their counterparts display neoplastic-specific changes. Cancer Res 68:2717–2725 Hu M, Polyak K (2008) Molecular characterisation of the tumour microenvironment in breast cancer. Eur J Cancer 44(18):2760–2765 Hung SC, Deng WP, Yang WK et al (2005) Mesenchymal stem cell targeting of microscopic tumors and tumor stroma development monitored by noninvasive in vivo positron emission tomography imaging. Clin Cancer Res 11:7749–7756 Iwamoto S, Mihara K, Dowining JR, Pui C-H, Campana D (2007) Mesenchymal cells regulate the response of acute lymphoblastic leukemia to asparaginase. J Clin Invest 117:1049–1057 Jodele S, Chantrain CF, Blavier L, Lutzko C, Crooks GM et al (2005) The contribution of bone marrow-derived cells to the tumor vasculature in neuroblastoma is matrix metalloproteinase-9 dependent. Cancer Res 65:3200–3208 Karnoub AE, Dash AB, Vo AP et al (2007) Mesenchymal stem cells within tumor stroma promote breast cancer metastasis. Nature 449:557–563 Koukourakis MI, Giatromanolaki A, Harris AL et al (2006) Comparison of metabolic pathways between cancer cells and stromal cells in colorectal carcinomas: a metabolic survival role for tumor-associated stroma. Cancer Res 66:632–637 Koyama H, Kobayashi N, Harada M, Takeoka M, Kawai Y et al (2008) Significance of tumor-associated stroma in promotion of intratumoral lymphangiogenesis: pivotal role of a hyaluronan-rich tumor microenvironment. Am J Pathol 172:179–193 Kroemer G, Pouyssegur J (2008) Tumor cell metabolism: cancer’s Achilles’ heel. Cancer Cell 13:472–482 Lee JS, Frevert CW, Wurfel MM et al (2003) Duffy antigen facilitates movement of chemokine across the endothelium in vitro and promotes neutrophil transmigration in vitro and in vivo. J Immunol 170(10):5244–5251 Lentsch AB (2002) The Duffy antigen/receptor for chemokines (DARC) and prostate cancer. A role as clear as black and white? The FASEB Journal 16:1093–1095 Liu XH, Hadley TJ, Xu L, Peiper SC, Ray PE (1999) Up-regulation of Duffy antigen receptor expression in children with renal disease. Kidney Int 55(4):1491–1500 Mantovani A, Schioppa T, Porta C, Allavena P, Sica A (2006) Role of tumor-associated macrophages in tumor progression and invasion. Cancer Metastasis Rev 25:315–322 Martin DN, Boersma BJ et al (2009). Differences in the tumor microenvironment between African-American and European-American breast cancer patients. PloS One 4:1–14 Mayr FB, Spiel AO, Leitner JM, Firbas C, Kliegel T, Jilma B (2007) Ethnic differences in plasma levels of interleukin-8 (IL-8) and granulocyte colony stimulating factor (G-CSF). Transl Res 149(1):10–14 Mishra PJ, Mishra PJ, Humeniuk R et al (2008) Carcinoma-associated fibroblast-like differentiation of human mesenchymal stem cells. Cancer Res 68:4331–4339 Nakamizo A, Marini F, Amano T et al (2005) Human bone marrow-derived mesenchymal stem cells in the treatment of gliomas. Cancer Res 65:3307–3318 Ogawa M, LaRue AC, Drake CJ (2006) Hematopoietic origin of fibroblasts/myofibroblasts: its pathophysiologic implications. Blood 108:2893–2896 Orimo A, Gupta PB, Sgroi DC et al (2005) Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell 121:335–348
13 Bone Marrow Derived Mesenchymal Stem/Stromal Cells and Tumor Growth
287
Patocs A, Zhang L, Xu Y et al (2007) Breast-cancer stromal cells with TP53 mutations and nodal metastases. N Engl J Med 357:2543–2551 Picinich SC, Mishra PJ, Mishra PJ, Glod J, Banerjee D (2007) The therapeutic potential of mesenchymal stem cells. Expert Opin Biol Ther 7:965–973 Pittenger MF, Mackay AM, Beck SC et al (1999) Multilineage potential of adult human mesenchymal stem cells. Science 284:143–147 Pruenster M, Mudde L, Bombosi P et al (2009) The Duffy antigen receptor for chemokines transports chemokines and supports their promigratory activity. Nat Immunol 10:101–108 Radisky DC (2005) Epithelial-mesenchymal transition. J Cell Sci 118:4325–4326 Ries L, Eisner M, Kosary M (2002) SEER Cancer Statistics Review, 1973–1999. National Cancer Institute, Bethesda, MD Salomon AV, Thiery JP. (2003) Host microenvironment in breast cancer development Epithelial– mesenchymal transition in breast cancer development. Breast Cancer Res 5:101–106 Samudio I, Fiegl M, McQueen T et al (2008) The Warburg effect in leukemia-stroma cocultures is mediated by mitochondrial uncoupling associated with uncoupling protein-2 activation. Cancer Res 68:5198–5205 Shen H, Schuster R, Stringer KF, Waltz SE, Lentsch AB (2006) The Duffy antigen/receptor for chemokines (DARC) regulates prostate tumor growth. FASEB J 20:59–64 Shinagawa K, Kiatadai Y, Ohara E et al (2009) Mesenchymal stem cells migrate to tumor stroma and enhance tumor growth in an orthotopic nude mouse model of colon cancer. Abstract 3156. p314 Proc of 100th Annual Meeting AACR, Denver,CO, 18–22 April 2009 Sorlie T, Perou CM, Tibshirani R et al (2001) Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proc Natl Acad Sci U S A 98: 10869–10874 Spaeth EL, Dembinski JL, Kate Sasser A et al (2009) Mesenchymals stem cell transition to tumor associated fibroblasts contributes to fibrovascular network expansion and tumor progression. PloS One 4:1–11 Studeny M, Marini FC, Champlin RE et al (2002) Bone marrow-derived mesenchymal stem cells as vehicles for interferon-beta delivery into tumors. Cancer Res 62:3603–3608 Studeny M, Marini FC, Dembinski JL et al (2004) Mesenchymal stem cells: potential precursors for tumor stroma and targeted-delivery vehicles for anticancer agents. J Natl Cancer Inst 96:1593–1603 Tournamille C, Colin Y, Cartron JP, Le Van Kim C (1995) Disruption of a GATA motif in the Duffy gene promoter abolishes erythroid gene expression in Duffy-negative individuals. Nat Genet 10:224–228 Tsujino T, Seshimo I, Yamamoto H, Ngan CY, Ezumi K et al (2007) Stromal myofibroblasts predict disease recurrence for colorectal cancer. Clin Cancer Res 13:2082–2090 Udagawa T, Puder M, Wood M, Schaefer BC, D’Amato RJ (2006) Analysis of tumor-associated stromal cells using SCID GFP transgenic mice: contribution of local and bone marrow-derived host cells. FASEB J 20:95–102 van de Vijver MJ, He YD, van’t Veer LJ et al (2002) Gene expression profiling predicts clinical outcome of breast cancer. N Engl J Med 347(25):1999–2009 Wadhwa M, Seghatchian MJ, Dilger P, Contreras M, Thorpe R (2000) Cytokine accumulation in stored red cell concentrates: effect of buffy-coat removal and leucoreduction. Transfus Sci 23(1):7–16 Wang Y, Klijn JG, Zhang Y et al (2005) Gene-expression profiles to predict distant metastasis of lymph-node-negative primary breast cancer. Lancet 365:671–679 Wang J, Ou ZL, Hou YF, Luo JM, Shen ZZ, Ding J, Shao ZM (2006) Enhanced expression of Duffy antigen receptor for chemokines by breast cancer cells attenuates growth and metastasis potential. Oncogene 25:7201–7211 Weber F, Xu Y, Zhang L, Patocs A, Shen L, Platzer P, Eng C (2007) Microenvironmental genomic alterations and clinicopathological behavior in head and neck squamous cell carcinoma. JAMA 297:187–195
288
P.J. Mishra and D. Banerjee
Worthley DL, Ruszkiewicz A, Davies R et al (2009) Human Gastrointestinal neoplasia associated myofibroblasts can develop from bone marrow derived cells following allogeneic stem cell transplantation. Stem Cells 27:1463–1468 Yazhou C, Wenlv S, Weidong Z, Licun W (2004) Clinicopathological significance of stromal myofibroblasts in invasive ductal carcinoma of the breast. Tumor Biol 25:290–295 Zhang Y, Daquinag A, Traktuev DO et al (2009) White adipose tissue cells are recruited by experimental tumors and promote cancer progression in mouse models. Cancer Res 69:5259–5265
Chapter 14
Integrin Signaling in Lymphangiogenesis Barbara Garmy-Susini
Abstract The lymphatic vasculature is essential for the maintenance of normal fluid balance and immune response. The lymphatic vessels are also crucially involved in pathological processes such as lymphedema and tumor metastasis. In the adult, lymphangiogenesis is only involved in pathological conditions. Induction of tumoral lymphangiogenesis allows tumor cells to enter the lymphatics to induce metastasis to the lymph nodes and distant loci. Dynamic interactions between lymphatic endothelial cells (LEC) and components of their surrounding extracellular matrix are necessary for the invasion, migration, and survival of LEC during lymphangiogenesis. A major part of these interactions are mediated by integrins, a cell matrix receptors family that initiate intracellular signaling cascades in response to binding to specific extracellular matrix molecules. Understanding the molecular events that define this subset of invasive LEC will facilitate the development of new treatment strategies. Therefore, understanding the role of integrin in tumor lymphangiogenesis remains crucial. Here, we review the role of integrins as major regulator of lymphangiogenesis.
Introduction Lymphatic system collects the extravasated fluid to maintain normal tissue fluid balance. In addition, lymphatic vessels absorb and transport fat released by enterocytes and represent an important part of immune surveillance by carrying immune cells and antigens. Furthermore to these physiologic tasks, lymphatic system participates to pathological conditions such as lymphedema, inflammatory diseases, and tumor metastasis. Many studies have demonstrated the existence of proliferative peri- and intratumoral lymphatic vessels (Skobe et al. 2001; Alitalo et al. 2005). Additionally, tumoral lymphangiogenesis correlates with an increase of
B. Garmy-Susini (*) Unité mixte Inserm U858, Institut de Médecine Moléculaire de Rangueil, IFR 150, 1, Avenue Jean Poulhès, BP 84225 31432, Toulouse Cedex 4, France e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_14, © Springer Science+Business Media, LLC 2010
289
290
B. Garmy-Susini
metastases (Williams et al. 2003; Mumprecht and Detmar 2009) and detection of lymphangiogenic growth factors is associated with poor prognosis in many human tumors (Stacker et al. 2004; Renyi-Vamos et al. 2005; Zeng et al. 2005). Development and remodeling of lymphatic system requires complex interactions between lymphatic endothelial cells (LECs) and extarcellular matrix (ECM). The molecular mechanisms leading to LEC–ECM interactions remain crucial for the understanding of lymphatic development by contributing to cell movement, proliferation, and survival. Integrins represent the most important family of receptors mediating cell adhesion to ECM. A large body of evidences has demonstrated that integrins are not only adhesion receptors, but influenc e the biological activity of several other molecular systems within the cell. Their role in angiogenesis has also been clearly established. Here, we reviewed emerging results highlighting new roles of integrins in lymphangiogenesis.
Lyphangiogenesis Lymphatic Vasculature Lymphangiogenesis, the outgrowth of novel lymphatic vessels, plays a central role in maintaining the interstitial fluid balance via transporting extravasated tissue fluid, macromolecules and cells back into the blood circulation. Lymphatic vessels also play a crucial role in promoting the immune functions by controlling cellular and antigen trafficking. Initial lymphatics combine to form larger vessels called precollectors and collectors, which in turn lead to four major groups of lymph nodes in the axillary and inguinal regions. Present in the skin and in most internal organs, the lymphatic vasculature is composed of vessels with distinct morphological features. Similar to blood capillaries, lymphatic capillaries consist of a single layer of thin-walled, nonfenestrated LECs, but they are not covered by pericytes or smooth muscle cells, and have an absent or poorly developed basement membrane. In addition, they lack tight junctions and adherens junctions, which allow easy access for fluid, macromolecules, and cells into the vessel lumen (Leak 1976). Endothelial cells of lymphatic capillaries are oak leaf shaped and are interconnected by specialized discontinuous button-like junctions, whereas collecting lymphatic vessels downstream have continuous zipperlike junctions found also in blood vessels (Baluk et al. 2007). Overlapping endothelial cell–cell contacts (also called primary valves) in initial lymphatic vessels prevent fluid escaping back into the interstitial space (Trzewik et al. 2001; Schmid-Schonbein 2003). The recent discovery of specific markers and growth factors for lymphatic endothelium and the establishment of genetic mouse models with impairment of lymphatic function have provided novel insights into the molecular control of the lymphatic system in physiology and in embryonic development. Based upon these findings, novel therapeutic strategies are currently being developed that aim at inhibiting or promoting the formation and function of lymphatic vessels in disease.
14 Integrin Signaling in Lymphangiogenesis
291
Lymphatic Makers Research into the role of the lymphatic system in cancer metastasis has been hampered by the lack of specific markers that distinguish lymphatic vessels from blood vessels, and also by the lack of known lymphatic-specific growth factors (Alitalo et al. 2005). However, novel lymphatic-specific markers have been recently identified such as lymphatic vascular endothelial-cell hyaluronan receptor-1 (LYVE-1) (Prevo et al. 2001) a new homologue of the CD44 glycoprotein, which is a lymphspecific receptor for hyaluronan; Prox1, a homeobox transcription factor that induces lymphatic lineage-specific differentiation and is essential for the embryonic development of the lymphatic system (Wigle and Oliver 1999; Hong et al. 2002; Petrova et al. 2002) and Podoplanin, a transmembrane glycoprotein molecule (Wicki et al. 2006) expressed in the cardinal vein and in Prox1-positive lymphatic progenitor cells in the embryos and restricted to lymphatic endothelium during later development. VEGFR-3 (also known as FLT-4) has largely been described as an major marker of lymphatics (Karkkainen et al. 2000; Kilic et al. 2007) as its expression in the adult becomes restricted to the lymphatic endothelium (Makinen et al. 2001; Bridenbaugh 2005; Breslin et al. 2007). However, recent finding have shown that VEGFR-3 is also upregulated on vascular endothelial cells in angiogenic sprouts and is present on vessels in tumors and wounds (Petrova et al. 2008; Tammela et al. 2008).
Induction of Lymphangiogenesis The spread of cancer cells from the primary tumor to distant organs usually first occurs via the sentinel lymph node. Tumors can induce lymphangiogenesis in both primary tumor and lymph node via release of the lymphangiogenic growth factors vascular endothelial growth factor VEGF-C or VEGF-D, leading to increased rates of metastasis to the draining sentinel lymph nodes and beyond (Mandriota et al. 2001; Skobe et al. 2001; Stacker et al. 2001) (Fig. 14.1). Several mediators of lymphangiogenesis have been identified. Fibroblast growth factor FGF-2 promotes lymphatic vessel growth in the mouse cornea (Kubo et al. 2002; Chang et al. 2004) and also stimulates proliferation and migration of LECs by binding to the receptor FGFR-3, which is upregulated by the transcription factor Prox1 in lymphatic endothelium (Shin et al. 2006). Hepatocyte growth factor (HGF; also known as scatter factor) also induces proliferation, migration, and tube formation of LECs and increases lymphangiogenesis in vivo (Kajiya et al. 2005). Despite the growing number of novel potential lymphangiogenic factors, the VEGF family represent the most extensive lymphangiogenic stimuli in the majority of human and experimental cancers, in particular VEGF-C and VEGF-D. VEGF-C promotes lymphangiogenesis by activating VEGF receptor (VEGFR)-2 and
292
B. Garmy-Susini
Fig. 14.1 Mechanisms regulating tumor lymphangiogenesis. (1) Tumor cells near preexisting blood vessels secrete growth factors and chemokines such as VEGF-A and FGF-2 that stimulate quiescent vascular endothelium to enter the cell cycle. (2)Tumors also secrete factors such as VEGF-C, VEGF-D, VEGF-A, FGF-2 and HGF that stimulate the growth of new lymphatic vessels in the peritumoral space. (3) These growth factors activate or upregulate expression of integrins such as avb3, avb5, a4b1, and a5b1 on blood vessels and a1b1, a2b1, a4b1, and a9b1 on lymphatic vessels. (4) These integrins then promote lymphatic endothelial cell migration, proliferation and survival, resulting in the creation of new lymphatic vessels. (5) Tumor-derived VEGF-C also promotes new lymphatic vessel growth in sentinel lymph nodes. (6) Whereas the new blood vessels induce tumor growth providing oxygen and nutriments, new lymphatic vessels represent an easier way for tumor cells to metastasize due to their vascular wall more permeable. (7) Tumor cells form metastases to proximal draining lymph nodes and then to distant organs such as lung, liver or bone
VEGFR-3 on LECs (Makinen et al. 2001). VEGF-C-deficient mice fail to develop a functional lymphatic system (Karkkainen et al. 2004), and transgenic expression of soluble VEGFR-3 results in pronounced lymphedema (Makinen et al. 2001). Recently, VEGF-A was identified as a strong lymphangiogenic mediator. Adenoviral delivery of murine VEGF-A to the skin of mice strongly promotes lymphatic vessel growth, and transgenic mice that overexpress murine VEGF-A, specifically in the skin, show increased lymphangiogenesis during wound healing and inflammation (Nagy et al. 2002; Hong et al. 2004; Hirakawa et al. 2005). Moreover both VEGF-C- and VEGF-A-overexpressing primary tumors induce lymphangiogenesis in sentinel lymph nodes before metastasizing to these tissues (Hirakawa et al. 2007) indicates that primary tumors can prepare their future metastatic site in advance of their arrival, partly by producing lymphangiogenic factors that mediate their transport. Thus, tumors might actively modify their future distant loci to make it more suitable for their further metastatic spread. This hypothesis is further supported by recent findings that attraction of bone marrow-derived hematopoietic progenitor
14 Integrin Signaling in Lymphangiogenesis
293
cells to premetastatic sites was related to tumor-specific upregulation of fibronectin in resident fibroblasts (Kaplan et al. 2005). It will be important to determine the exact mechanisms by which lymphangiogenesis induces tumor-cell migration, via the lymphatic vasculature, to the lymph nodes and other tissues. Systemic blockade of VEGFR-3 has been recently shown to block tumor metastasis not only to lymph nodes, but also to lungs in experimental models of breast cancer (Krishnan et al. 2003; Roberts et al. 2006). Treatment with a VEGF-C inhibitor could potentially prevent the systemic spread of an early or even advanced-stage malignancy, by inhibiting lymphangiogenesis in tumors and lymph nodes. Recent reports have shown that the overexpression of VEGF-C or VEGF-D induces tumor lymphangiogenesis and promotes lymphatic metastasis in mouse tumor models (Stacker and Achen 2008). However, few clinical studies have investigated the association between the expression of VEGF-C and VEGF-D, and lymphangiogenesis and lymphatic metastasis. VEGF-D increases lymphatic vessel growth and lymphatic metastasis (Stacker et al. 2001). Recent studies revealed that VEGF-C is more significantly correlated with lymph node metastasis than VEGF-D. These findings suggest that VEGF-D is less important in lymphatic metastasis than VEGF-C, but is still necessary for metastasis (Sugiura et al. 2009).
Lymphangiogenesis and Pathology In adult organisms, lymphangiogenesis takes place only in certain pathological conditions. Abnormal function of the lymphatics is implicated in the diseases, such as lymphedema, inflammation, immune diseases, and tumor metastasis. Lymphedema is a disorder of the lymphatic vascular system characterized by impaired lymphatic return and swelling of the extremities. When the lymphatic system has been damaged during surgery or radiation treatment, its capacity to absorb excess water and cells from the interstitial space is reduced. If the transport capacity of the lymphatic system is reduced enough so that it cannot handle this increase in lymphatic load, an insufficiency of the lymphatic system may occur. Lymphedema can be an unfortunate side effect of cancer treatment. It is a chronic condition that, if ignored, can lead to disfigurement, immobilization, and severe infections. Without treatment, the swelling may continue to increase. Inflammation is thought to contribute to the development and progression of various cancers, including lung (Ardies 2003), breast (Van der Auwera et al. 2004), gastrointestinal (Jaiswal et al. 2001; Biarc et al. 2004; Brower 2005), ovarian (Altinoz and Korkmaz 2004), prostate (Wang et al. 2004), skin (Hussein and Ahmed 2005), and liver cancers (Bartsch and Nair 2004). Inflammatory breast cancer exhibits increased angiogenesis and lymphangiogenesis and has a higher metastatic potential than noninflammatory breast cancer (Angelo and Kurzrock 2007). Blocking lymphangiogenesis in chronic inflammatory diseases may become an important means of ameliorating the severity of some of these pathologies.
294
B. Garmy-Susini
The extent of lymph node metastasis is a major determinant for the staging and the prognosis of most human malignancies. Although the clinical significance of lymph node involvement is well documented, molecular mechanisms that promote tumor spread into the lymphatic or blood vascular systems and widespread dissemination are not well understood. Recent studies have shown a large body of evidence that newly visualized lymphatics facilitates formation of metastases. High tumor interstitial fluid pressure is thought to promote tumor cell entry into lymphatic vessels that have lower fluid pressure (Jain 1994; Padera et al. 2002). Intratumoral lymphatic vessel growth often correlates with metastasis of human melanoma, breast, or head and neck cancers (Maula et al. 2003; Choi et al. 2005; Dadras et al. 2005), where tumor cells can be observed within lymphatic vessels, demonstrating that lymphatic vessel growth is important for tumor spread.
Integrins Many studies have implicated a number of endothelial cell integrins in the regulation of cell growth, survival and migration during angiogenesis; nevertheless, their role in lymphangiogenesis remains unclear. The role of cell adhesion molecules and extracellular matrix proteins (ECM) in various pathological processes including angiogenesis, thrombosis, apoptosis, cell migration, and proliferation are well documented. These processes can lead to both acute and chronic disease states such as ocular diseases, metastasis, myocardial infarction, stroke, osteoporosis, a wide range of inflammatory diseases, vascular remodeling, and neurodegenerative disorders. As several integrin-targeted therapeutic agents are in clinical trials for cancer therapy, future clinical studies will likely determine whether integrin inhibitors will be best used against select tumors, such as those in which tumor cells themselves express the targeted integrin. As integrins are clearly a family of critical and fundamental regulators of angiogenesis and lymphangiogenesis, the future of integrin antagonists in cancer therapy is promising.
Integrin Expression and Function Integrin induce several physiological processes including cell activation, migration, proliferation, differentiation, and many other processes require direct contact between cells or ECM proteins. Cell–cell and cell–matrix interactions are mediated through several different families of CAM including the selectins, the integrins, the cadherins, and the immunoglobulins. The integrin family is an extensive group of structurally related receptors for ECM proteins and immunoglobulin superfamily molecules. Integrins are heterodimeric membrane glycoproteins comprised of noncovalently associated a and b subunits that promote cell attachment and migration on the surrounding extracellular matrix. Eighteen a and eight b subunits can
14 Integrin Signaling in Lymphangiogenesis
295
Fig. 14.2 Integrin family of adhesion receptors. Each integrin receptor heterodimer binds a specific set of endogenous ligands, which may include ligands in the ECM, soluble ligands, and ligands on other cell surfaces. Integrins are divalent cation-dependent heterodimeric membrane glycoproteins comprised of noncovalently associated a and b subunits. Each integrin subunit consists of an extracellular domain, a single transmembrane region, and a short (approximately 30–40 amino acids) cytoplasmic region. Upon ligand binding, a series of intracellular signaling events is initiated. These pathways are associated with enhanced cell proliferation, migration, and survival. The role of integrins lymphangiogenesis is at an early stage of understanding. Nevertheless, recent studies have shown an important role integrins a1b1, a2b1, a4b1, and a9b1 in promoting lymphatic endothelial cell migration and survival (green)
associate to form 24 unique integrin heterodimers (Fig. 14.2). Each integrin subunit consists of an extracellular domain, a single transmembrane region, and a short (approximately 30–40 amino acids) cytoplasmic region (Hynes 2002). The main ligands for integrins in the extracellular space are extracellular matrix proteins, such as laminin and collagen, as well as cellular counter-receptors. Integrins are linked to the cytoskeleton through their cytoplasmic domains. Integrins modulate the cytoskeleton via various submembrane adaptor proteins and kinases (Zamir and Geiger 2001). They transduce signals across the plasma membrane in both directions. Most of the time, they are expressed on the cell surface in an inactive state. The activation of integrins leads to conformational changes in the extracellular domain allowing ligand binding. Integrin binding to its ligands requires its activation by inside-out signals. Conversely, integrin ligation triggers outside–in signals that regulate different aspects of cell behavior, including cell survival, control of transcription, cell proliferation, cell motility, and cytoskeletal organization (Hynes 2002). Integrin bidirectional signals are thought to be transmitted via protein–protein interactions but how the integrin cytoplasmic tail orchestrates the transmission of bidirectional signals is not well understood. Specific binding of the cytoskeletal protein to integrin b subunit cytoplasmic tails leads to the conformational rearrangements of integrin extracellular domains that increase their affinity (Tadokoro et al. 2003). Studies support a model in which intracellular protein interactions with the integrin b
296
B. Garmy-Susini
subunit tail play key roles for both inside–out and outside–in signal transduction (Arias-Salgado et al. 2003; Tadokoro et al. 2003; Arias-Salgado et al. 2005). Outside–in signal: During migration, cells are constantly making and breaking integrin contacts, activating intra-cellular signaling pathways. These signals are associated with growth factors signals and regulate cell invasion into their microenvironment (Hood and Cheresh 2002). They are mediated by intracellular signaling molecules such as FAK, SHC, and RHO family of small GTPases. Inside–out signal: Integrins not only send signals to the cell in response to the extracellular environment, but also respond to intracellular stimuli to modify the way in which they interact with extracellular matrix proteins (Hood and Cheresh 2002). This process modulates the affinity and the avidity of integrins for their ligands.
Role of Integrins in Promoting Endothelial Cells Migration, Proliferation, and Survival During tumorigenesis, a switch of integrin expression can be observed, in as much as growth-promoting and growth-attenuating integrins are up- and downregulated, respectively. ECM-ligand binding to an integrin initiates signals, which eradiating from the integrins are transmitted via different interconnecting pathways and elicit various cell functions, such as morphological changes, adhesion, migration, and gene activation. Any of these functions takes part in the metastatic cascade of tumor progression, such as epithelial-to-mesenchymal transition of carcinoma cells, tumor cell contact with the basement membrane, invasion into neighboring tissues as well as production and activation of ECM-degrading MMPs. Besides their direct involvement in tumor progression as cell surface molecules on tumor cells, integrins on endothelial surrounding a tumor can also determine various cancer characteristics, such as tumor-induced neoangiogenesis, lymphangiogenesis, and immune resistance. Therefore, integrins are relevant pharmacological targets in tumor biology.
Ligand Specificity of Integrins Each unique a-chain combines with a b-chain to form heterodimers with unique selectivity for ECM, cell surface molecules, and plasma proteins (Loftus et al. 1994; Plow et al. 2000; van der Flier and Sonnenberg 2001). Integrins binds to their ligand to a divalent-cation dependent fashion (Smith et al. 1994; Mould et al. 1995). While some integrins selectively recognize primarily a single ECM protein ligand (e.g., integrin a5b1 recognizes primarily fibronectin), others can bind several ligands (e.g., integrin avb3 binds vitronectin, fibronectin, collagen, and other matrix proteins).
14 Integrin Signaling in Lymphangiogenesis
297
Many integrins, including avb3, a5b1, aIIbb3, avb6, and a3b1, recognize the tripeptide Arg-Gly-Asp (RGD) in their ligands (Plow et al. 2000). Sequences flanking the RGD peptide are also important for integrin selectivity (Haas and Plow 1994). Other integrins recognize alternative short peptide sequences; for example, integrin a4b1 recognizes Glu-Ile-Leu-Asp-Val (EILDV) and Arg-Glu-Asp-Val (REDV) in the alternatively spliced fibronectin domain known as IIICS (Komoriya et al. 1991). Some integrins, such as a4b1, can also bind cell surface receptors, such as Vascular Cell Adhesion Molecule-1 (VCAM-1), to promote cell–cell adhesion (Jin and Varner 2004) in addition to alternatively spliced CS-1 fibronectin. Spurred by the recent success to generate pharmaceutical mimetics of RGDdependent integrins and by the integrin’s easy accessibility on the cell surface, the hope is rising that also RGD-independent integrins, such as the collagen- and laminin-binding integrins, can be pharmacologically manipulated to fight integrindependent functions of cancer cells, which are necessary and at least partially specific for their proliferation and progression.
Integrin Signaling Integrins are essential for cell migration and invasion. They mediate adhesion to extracellular matrix and regulate intracellular signaling pathways that control cytoskeletal organization. These pathways involve phosphorylation of focal adhesion kinase (FAK), recruitment of adaptator proteins, activation of small GTPases, and downstream effector molecules (Fig. 14.3). Integrin ligation promotes integrin clustering and subsequent integrin-mediated intracellular signal transduction. Unlike growth factor receptors, integrins have no intrinsic enzymatic or kinase activities, but activate complex signaling pathways by coclustering with kinases and adaptor proteins in focal adhesion complexes. A number of signaling pathways are activated by integrins and many of these are found within focal adhesion complexes. Focal adhesion complexes are comprised of integrins, protein kinases – such as FAK and Src – adaptor proteins such as Shc, signaling intermediates such as Rho family GTPases, actin-binding cytoskeletal proteins such as talin, a-actinin, paxillin, tensin, and vinculin and other signaling proteins. Integrin signaling promotes cell migration, proliferation, and survival. Loss of integrin ligation inhibits these events and unligated integrins can actively initiate apoptosis, even without loss of cell attachment. This form of death is stress response- and death-receptor-independent, but caspase 8-dependent, and has been called “integrin mediated death” (Stupack et al. 2006). Fak FAK is a cytoplasmic protein kinase that colocalizes with integrins at structures called focal adhesion. Integrins binding of ECM ligands induces integrin clustering
298
B. Garmy-Susini
Fig. 14.3 Each integrin subunit consists of an extracellular domain, a single transmembrane region, and a short (approximately 30–40 amino acids) cytoplasmic cytoplasmic tails and no intrinsic enzymatic or kinase activities. Upon ligand binding, a series of intracellular signaling events is initiated. These pathways are associated with enhanced cell proliferation, migration, and survival. To integrate signals and activate intracellular signaling pathways, integrins cocluster with serine, threonine and tyrosine kinases, phosphatases, and adaptor proteins in focal adhesions. Focal adhesion complexes are comprised of integrins, protein kinases such as focal adhesion kinase (FAK), Src and many other kinases, adaptor proteins such as Shc, signaling intermediates such as PI-3-kinase, Rho and Rac GTPases and actin-binding cytoskeletal proteins such as talin, a-actinin, paxillin, tensin, and vinculin. Integrin signaling promotes cell migration by providing traction along the extracellular matrix and by promoting actin remodeling through the Rho family small GTPases. This actin remodeling leads to cytoplasmic flow in the direction of cell migration and cell body retraction at the trailing end of the cells. Individual components of integrin-mediated signaling cascades, such as FAK, Shc and Raf, play essential roles in angiogenesis. For example, FAK is a mediator of signal transduction by integrins and growth factor receptors in endothelial cells. In addition, Shc, an important adaptor protein that potentiates MAP kinase pathway signaling is activated by both integrins and growth factor receptors and plays critical roles in early vascular development. Like Shc, Raf-1 is an integral component of the MAP kinase signaling pathway. This signaling intermediate is activated by integrins and is critical for vascular morphogenesis. Thus, integrin mediated signaling likely plays important roles in vascular development in the adulthood
and FAK activation (Schlaepfer et al. 1999). Numerous studies have linked FAKmediated signaling pathways to cancer and a variety of other biological and disease processes. Integrin-mediated cell adhesion is the major upstream activator of FAK, and increased activation and tyrosine phosphorylation of FAK have been observed in essentially all adherent cells analyzed so far. Once it is activated, FAK undergoes
14 Integrin Signaling in Lymphangiogenesis
299
autophosphorylated at Y397, which creates a binding site for Src via the SH2 domain (Schaller and Parsons 1994; Xing et al. 1994; Cary et al. 1999). FAK also binds p85 subunit of PI3K through autophosphorylated Y397 (Reiske et al. 2000). Phospholipid production stimulated by FAK association and activation of PI3K can activate Akt kinase, which inhibits apoptosis by regulating various cell death machinery proteins (Luo et al. 2003; Hennessy et al. 2005). Several FAK signaling pathways have been shown to play an important role in cell proliferation. The formation of FAK/Src complex allows Src to phosphorylate Y925 on FAK to mediate its interaction with Grb2 leading to the activation of RAS-Erk signaling pathway (Schlaepfer et al. 1994). FAK interaction with another adaptor molecule SHC also contributes to the activation of this pathway (Schlaepfer and Hunter 1996). Numerous studies have shown that autophosphorylation of FAK at Y397 and its association with Src at the site is essential for FAK’s ability to promote cell migration, as mutation of Y397 to F both disrupted FAK association with Src and its stimulation of cell migration in many cell types including FAK-/- cells (Cary et al. 1996; Cary et al. 1998; Owen et al. 1999). Inhibition of cell spreading by blocking FAK functions with FRNK (C-terminal FAK fragment functioning as a dominant negative FAK) could be partially rescued by overexpression of Src (Richardson and Parsons 1996). Taken together, these studies demonstrate the crucial role of FAK in cell migration, an essential process during lymphangiogenesis. Shc The adapter protein Shc is a prototype adapter protein that has been quite useful in the understanding of the function of adapter proteins in cellular signaling. The importance of Shc in vivo has been demonstrated by the knockout of the shcA gene, which results in embryonic lethality at day 11.5 (Lai and Pawson 2000) due to defects in blood vessels formation and cardiovascular development (Lai and Pawson 2000). Although a role for Shc in activation of the Ras/MAPK pathway and its role in mitogenic signaling has been better described (Bonfini et al. 1996), what other roles it may play in signaling and why this protein is so often recruited/targeted for phosphorylation by many different types of receptors is not clear. SHC is recruited to activated tyrosine kinases in response to ligation of integrins a1b1, a6b4, a5b1, and avb3 (Wary et al. 1996, 1998). Studies from various laboratories in different model systems have shown an important role for Shc in leading to Ras activation. Shc is tyrosine-phosphorylated by receptor activation and it subsequently interacts with Grb2. Grb2, binds to Ras guanine nucleotide exchange factor, Sos. The Shc/Grb2/Sos complex gets localized to the membrane through the interaction of Shc with the phosphorylated receptor. In the case of the integrin family of receptors and G-protein-coupled receptors, the precise mechanism by which the Shc/Grb2/Sos complex is localized to the membrane is not clear. Sos has been found preferentially in complexes that also contain Shc (Pronk et al. 1994). Shc may also influence the extent of Ras activation. A number of studies using
300
B. Garmy-Susini
dominant negative Shc proteins and mice lacking Shc expression have definitively established a role for Shc in MAPK activation (Salcini et al. 1994; Gotoh et al. 1997; Pratt et al. 1999; Lai and Pawson 2000). Together, studies identify Shc as a critical regulator of integrin signaling pathways. Rho Family of Small GTPases Integrin engagement and subsequent clustering of these receptors in focal adhesions leads to the generation of intracellular macromolecular complexes (Miyamoto et al. 1995). Numerous proteins present at the cytoplasmic face of focal adhesions are considered to be structural components of focal adhesions, including cytoskeletal proteins such as vinculin and talin (Jockusch and Rudiger 1996). In addition, numerous “signaling” proteins with enzymatic activity (e.g., kinases and GTPases) are also components of focal adhesions. The Rho family of GTPases appears poised to contribute to these integrin-mediated signals, in particular, signals that control cytoskeletal organization involved in changes in cell morphology. Rho family members such as Cdc42, Rac1, and RhoA are part of the Ras superfamily of proteins that cycle between active and an active state. Activated RhoA is capable of stimulating microfilament bundling in serum-starved cells that are already adherent (Ridley and Hall 1992), similar to the response of cells to plating on an ECM-coated surface. Rho is also essential for the formation of focal complexes (Hotchin and Hall 1995). The Rho family member Rac controls growth factor-stimulated membrane ruffling and formation of lamellipodia (Ridley and Hall 1992). Finally, Cdc42 activation triggers the extension of filopodia (Kozma et al. 1995; Nobes and Hall 1995). Studies such as these have defined how soluble extracellular factors induce the assembly of focal adhesions and stress fibers in serum-starved adherent Swiss 3T3 fibroblasts through activation of the Rho family of GTPases (Ridley 1996; Tapon and Hall 1997). Rho family GTPases are therefore necessary to induce the signaling and cellular responses that are required to reorganize the actin cytoskeleton into an invasive and migratory phenotype. Talin Talin is a major cytoskeletal protein that colocalizes with and binds to integrins, and to actin and actin-binding proteins such as vinculin (Calderwood et al. 1999, 2002; Tadokoro et al. 2003; Rose et al. 2007). Talin interacts directly with the b-chain cytoplasmic domain of integrins. Knockdown of talin expression in CHO (Chinesehamster ovary) cells inhibits the activation of both b1 and b3 integrins without altering integrin expression, and this cannot be compensated for by the expression of activating molecules such as activated R-Ras or the CD98 heavy chain (Tadokoro et al. 2003). Talin colocalizes with activated integrins, and overexpression of talin’s N-terminus activates integrins (Calderwood et al. 1999). Talin represents therefore a potential key regulator of integrins signaling during lymphangiogenesis.
14 Integrin Signaling in Lymphangiogenesis
301
Vinculin Vinculin is a cytoskeletal protein associated with the cytoplasmic part of matrix adherens-type junctions. Vinculin is one of several interacting proteins involved in anchoring actin to the membrane (Weller et al. 1990). Vinculin also plays a pivotal role in cell adhesion and migration by providing the link between the actin cytoskeleton and the transmembrane receptors, integrin and cadherin. Additionaly, modulation of b1-integrin association with vinculin in human coronary artery endothelial cells (HCAEC) alter endothelial wound closure under shear stress (Albuquerque and Flozak 2003) demonstrating its role in integrin-mediated endothelial migration.
Paxillin Paxillin connect the actin cytoskeleton to the extracellular matrix within focal adhesions. Studies have shown that mutant of paxillin formed focal adhesions and exhibited limited movement associated with cell migration and wound-healing assays (Huang et al. 2003). Paxillin has also been shown to bind specifically the cytoplasmic tail of alpha4 chain to induce cell migration (Liu et al. 2000; Liu and Schnellmann 2003). This a4 binding protein suggests a potential role of the integrin alpha chain in regulating activation and signaling pathways.
Intergrins and Lymphangiogenesis The role of integrins in angiogenesis has been previously indicated in tumor angiogenesis (Hynes et al. 2002), but their role in the remodeling of lymphatic vessels is at an early stage of understanding.
a 9b1 Recent studies have shown that integrin a9b1 is expressed on quiescent LEC. Integrin a9b1 binds tenascin and oxteopontin (Kanayama et al. 2009) and is required for development of the fully functional lymphatic system because mice deficient in a9b1 integrin die 6–12 days after birth due to chylothorax, an accumulation of lymph in the pleural cavity (Huang et al. 2000). Integrin a9b1 participates to the lymphatic valves morphogenesis as it has been recently demonstrated an upregulation of integrin a9b1 on LEC and deposition of its ligand fibronectin in the extracellular matrix during embryonic development (Bazigou et al. 2009). Integrin a9b1 plays a role in growth factor mediated lymphangiogenesis as Prox-1, a LEC selective transcription factor coordinately upregulates integrin a9b1 and VEGFR-3
302
B. Garmy-Susini
expression and endothelial cell motility in vivo (Mishima et al. 2007). Furthermore, studies have shown that this integrin promotes VEGF-C- and -D-stimulated cell migration by directly binding these growth factors (Vlahakis et al. 2005). Importantly, antagonism of a9b1 suppresses VEGF-C induced motility. Taken together, these studies indicate that a9b1 plays unique yet critical roles in lymphangiogenesis.
a 1b1 and a2b1 Other studies have shown that integrins a1b1 and a2b1 are expressed on lymphatic endothelium in healing wounds in response to VEGF-A. Inhibition of these integrins blocked lymphangiogenesis in these wounds (Hong et al. 2004). These integrins bind laminin and collagen and also play roles in regulating angiogenesis. Studies in cultured LECs revealed that VEGF-A induced expression of the alpha1 and alpha2 integrins, which promoted their in vitro tube formation and their haptotactic migration toward type I collagen. It has also been shown that systemic blockade of the alpha1 and alpha2 integrins inhibits VEGF-A-driven lymphangiogenesis in vivo (Hong et al. 2004).
a 5b1 Integrin a5b1 is receptor for fibronectin. The role of integrins a5b1 in tumor angiogenesis has been clearly demonstrated (Yang et al. 1993; Francis et al. 2002) its their role in tumor lymphangiogenesis remains unclear. Integrins a5b1 and av are expressed by a subpopulation of lymphatic vessels in the inflamed cornea and small molecule antagonists of this integrin inhibited inflammatory lymphangiogenesis (Dietrich et al. 2007). Based on these studies, the role of integrin a5b1 during lymphangiogenesis seems to be restricted to inflammatory diseases. Recently, integrin a5b1 has been shown to be upregulated on lymphatic sprouts in a mice model of airways inflammation (Okazaki et al. 2009) suggesting a role in lymphatic growth. The av integrins appear to play little or no role in lymphangiogenesis and integrin a5b1 appears to play no role in tumor lymphangiogenesis due to the lack of expression on lymphangiogenic LEC during tumor progression (Garmy-Susini et al. 2007; Avraamides et al. 2008; Garmy-Susini and Varner 2008).
a 4b1 Integrin a4b1 not only allows cells to interact with the ECM after binding a spliced variant of fibronectin (CS-1 fibronectin), but also participates to cell–cell adhesion after binding a cell-surface molecule VCAM-1, a member of the immunoglobulin
14 Integrin Signaling in Lymphangiogenesis
303
superfamily. Integrin a4b1 promotes angiogenesis after binding its ligand VCAM-1 on mural cells during blood vessels formation (Garmy-Susini et al. 2005). Integrin a4b1 regulate embryonic development as loss of its gene causes embryonic lethality by E11.5-E12.5 due to vascular defects (Kwee et al. 1995; Yang et al. 1995; Sengbusch et al. 2002). Integrin a4b1 is also highly expressed on tumor lymphatic endothelium (Garmy-Susini et al. 2007) and antagonists of this integrin can block lymphangiogenesis and tumor metastasis.
Conclusion Most of blood and lymphatic vessels remain quiescent in the adulthood and are activated during pathological conditions such as wound healing, inflammatory disease, or tumor development. Several integrins appear to play important roles in regulating lymphangiogenesis. Their profile during lymphangiogenesis is distinct from those regulating angiogenesis. Integrin expressed during lymphangiogenesis promote LEC migration, proliferation and survival, providing new ways for tumor cells to escape and metastasize to distant loci. They represent potentially relevant targets for antilymphangiogenic therapy. Many studies exhibit the evidence that drug therapies have to be considered in connection with the cell microenvironment that controls lymphangiogenesis. Targeting integrins for antilymphangiogenic therapy remains crucial, as they are expressed on activated vessels but not on quiescent vessels and therefore accessible by drugs only during lymphangiogenesis. Antagonists of these integrins may be useful in preventing tumor metastasis by blocking lymphangiogenesis. Moreover, agents targeting integrins receptors are now in clinical development for treating solid tumors. Due to the lack of side effect associated with integrins inhibitors the challenge for the future will be to optimize their use in combination with other treatment to fully exploit their therapeutic potential in lymphangiogenesis inhibition. However, the molecular events underlying the regulatory function of integrin receptors are not well understood and targeting of integrin function and signaling may be an alternative strategy to extracellular integrin antagonists for the therapeutic inhibition of tumor lymphangiogenesis.
References Albuquerque ML, Flozak AS (2003) Lamellipodial motility in wounded endothelial cells exposed to physiologic flow is associated with different patterns of beta1-integrin and vinculin localization. J Cell Physiol 195, 50–60. Alitalo K, Tammela T, Petrova TV (2005) Lymphangiogenesis in development and human disease. Nature 438, 946–953. Altinoz MA, Korkmaz R (2004) NF-kappaB, macrophage migration inhibitory factor and cyclooxygenase-inhibitions as likely mechanisms behind the acetaminophen- and NSAIDprevention of the ovarian cancer. Neoplasma 51, 239–247.
304
B. Garmy-Susini
Angelo LS, Kurzrock R (2007) Vascular endothelial growth factor and its relationship to inflammatory mediators. Clin Cancer Res 13, 2825–2830. Ardies CM (2003) Inflammation as cause for scar cancers of the lung. Integr Cancer Ther 2, 238–246. Arias-Salgado EG, Lizano S, Sarkar S, Brugge JS, Ginsberg MH, Shattil SJ (2003) Src kinase activation by direct interaction with the integrin beta cytoplasmic domain. Proc Natl Acad Sci U S A 100, 13298–13302. Arias-Salgado EG, Lizano S, Shattil SJ, Ginsberg MH (2005) Specification of the direction of adhesive signaling by the integrin beta cytoplasmic domain. J Biol Chem 280, 29699–29707. Avraamides CJ, Garmy-Susini B, Varner JA (2008) Integrins in angiogenesis and lymphangiogenesis. Nat Rev Cancer 8, 604–617. Baluk P, Fuxe J, Hashizume H, Romano T, Lashnits E, Butz S, Vestweber D, Corada M, Molendini C, Dejana E, McDonald DM (2007) Functionally specialized junctions between endothelial cells of lymphatic vessels. J Exp Med 204, 2349–2362. Bartsch H, Nair J (2004) Oxidative stress and lipid peroxidation-derived DNA-lesions in inflammation driven carcinogenesis. Cancer Detect Prev 28, 385–391. Bazigou E, Xie S, Chen C, Weston A, Miura N, Sorokin L, Adams R, Muro AF, Sheppard D, Makinen T (2009) Integrin-alpha9 is required for fibronectin matrix assembly during lymphatic valve morphogenesis. Dev Cell 17, 175–186. Biarc J, Nguyen IS, Pini A, Gosse F, Richert S, Thierse D, Van Dorsselaer A, Leize-Wagner E, Raul F, Klein JP, Scholler-Guinard M (2004) Carcinogenic properties of proteins with proinflammatory activity from Streptococcus infantarius (formerly S.bovis). Carcinogenesis 25, 1477–1484. Bonfini L, Migliaccio E, Pelicci G, Lanfrancone L, Pelicci PG (1996) Not all Shc’s roads lead to Ras. Trends Biochem Sci 21, 257–261. Breslin JW, Gaudreault N, Watson KD, Reynoso R, Yuan SY, Wu MH (2007) Vascular endothelial growth factor-C stimulates the lymphatic pump by a VEGF receptor-3-dependent mechanism. Am J Physiol Heart Circ Physiol 293, H709–718. Bridenbaugh E (2005) Literature watch. Complete and specific inhibition of adult lymphatic regeneration by a novel VEGFR-3 neutralizing antibody. Lymphat Res Biol 3, 87–88. Brower V (2005) Researchers attempting to define role of cytokines in cancer risk. J Natl Cancer Inst 97, 1175–1177. Calderwood DA, Yan B, de Pereda JM, Alvarez BG, Fujioka Y, Liddington RC, Ginsberg MH (2002) The phosphotyrosine binding-like domain of talin activates integrins. J Biol Chem 277, 21749–21758. Calderwood DA, Zent R, Grant R, Rees DJ, Hynes RO, Ginsberg MH (1999) The Talin head domain binds to integrin beta subunit cytoplasmic tails and regulates integrin activation. J Biol Chem 274, 28071–28074. Cary LA, Chang JF, Guan JL (1996) Stimulation of cell migration by overexpression of focal adhesion kinase and its association with Src and Fyn. J Cell Sci 109 (Pt 7), 1787–1794. Cary LA, Han DC, Guan JL (1999) Integrin-mediated signal transduction pathways. Histol Histopathol 14, 1001–1009. Cary LA, Han DC, Polte TR, Hanks SK, Guan JL (1998) Identification of p130Cas as a mediator of focal adhesion kinase-promoted cell migration. J Cell Biol 140, 211–221. Chang LK, Garcia-Cardena G, Farnebo F, Fannon M, Chen EJ, Butterfield C, Moses MA, Mulligan RC, Folkman J, Kaipainen A (2004) Dose-dependent response of FGF-2 for lymphangiogenesis. Proc Natl Acad Sci U S A 101, 11658–11663. Choi WW, Lewis MM, Lawson D, Yin-Goen Q, Birdsong GG, Cotsonis GA, Cohen C, Young AN (2005) Angiogenic and lymphangiogenic microvessel density in breast carcinoma: correlation with clinicopathologic parameters and VEGF-family gene expression. Mod Pathol 18, 143–152. Dadras SS, Lange-Asschenfeldt B, Velasco P, Nguyen L, Vora A, Muzikansky A, Jahnke K, Hauschild A, Hirakawa S, Mihm MC, Detmar M (2005) Tumor lymphangiogenesis predicts melanoma metastasis to sentinel lymph nodes. Mod Pathol 18, 1232–1242.
14 Integrin Signaling in Lymphangiogenesis
305
Dietrich T, Onderka J, Bock F, Kruse FE, Vossmeyer D, Stragies R, Zahn G, Cursiefen C (2007) Inhibition of inflammatory lymphangiogenesis by integrin alpha5 blockade. Am J Pathol 171, 361–372. Francis SE, Goh KL, Hodivala-Dilke K, Bader BL, Stark M, Davidson D, Hynes RO (2002) Central roles of alpha5beta1 integrin and fibronectin in vascular development in mouse embryos and embryoid bodies. Arterioscler Thromb Vasc Biol 22, 927–933. Garmy-Susini B, Jin H, Zhu Y, Sung RJ, Hwang R, Varner J (2005) Integrin alpha4beta1-VCAM1-mediated adhesion between endothelial and mural cells is required for blood vessel maturation. J Clin Invest 115, 1542–1551. Garmy-Susini B, Makale M, Fuster M, Varner JA (2007) Methods to study lymphatic vessel integrins. Methods Enzymol 426, 415–438. Garmy-Susini B, Varner JA (2008) Roles of integrins in tumor angiogenesis and lymphangiogenesis. Lymphat Res Biol 6, 155–163. Gotoh N, Toyoda M, Shibuya M (1997) Tyrosine phosphorylation sites at amino acids 239 and 240 of Shc are involved in epidermal growth factor-induced mitogenic signaling that is distinct from Ras/mitogen-activated protein kinase activation. Mol Cell Biol 17, 1824–1831. Haas TA, Plow EF (1994) Integrin-ligand interactions: a year in review. Curr Opin Cell Biol 6, 656–662. Hennessy BT, Smith DL, Ram PT, Lu Y, Mills GB (2005) Exploiting the PI3K/AKT pathway for cancer drug discovery. Nat Rev Drug Discov 4, 988–1004. Hirakawa S, Brown LF, Kodama S, Paavonen K, Alitalo K, Detmar M (2007) VEGF-C-induced lymphangiogenesis in sentinel lymph nodes promotes tumor metastasis to distant sites. Blood 109, 1010–1017. Hirakawa S, Kodama S, Kunstfeld R, Kajiya K, Brown LF, Detmar M (2005) VEGF-A induces tumor and sentinel lymph node lymphangiogenesis and promotes lymphatic metastasis. J Exp Med 201, 1089–1099. Hong YK, Harvey N, Noh YH, Schacht V, Hirakawa S, Detmar M, Oliver G (2002) Prox1 is a master control gene in the program specifying lymphatic endothelial cell fate. Dev Dyn 225, 351–357. Hong YK, Lange-Asschenfeldt B, Velasco P, Hirakawa S, Kunstfeld R, Brown LF, Bohlen P, Senger DR, Detmar M (2004) VEGF-A promotes tissue repair-associated lymphatic vessel formation via VEGFR-2 and the alpha1beta1 and alpha2beta1 integrins. Faseb J 18, 1111–1113. Hood JD, Cheresh DA (2002) Role of integrins in cell invasion and migration. Nat Rev Cancer 2, 91–100. Hotchin NA, Hall A (1995) The assembly of integrin adhesion complexes requires both extracellular matrix and intracellular rho/rac GTPases. J Cell Biol 131, 1857–1865. Huang C, Rajfur Z, Borchers C, Schaller MD, Jacobson K (2003) JNK phosphorylates paxillin and regulates cell migration. Nature 424, 219–223. Huang XZ, Wu JF, Ferrando R, Lee JH, Wang YL, Farese RV, Jr., Sheppard D (2000) Fatal bilateral chylothorax in mice lacking the integrin alpha9beta1. Mol Cell Biol 20, 5208–5215. Hussein MR, Ahmed RA (2005) Analysis of the mononuclear inflammatory cell infiltrate in the non-tumorigenic, pre-tumorigenic and tumorigenic keratinocytic hyperproliferative lesions of the skin. Cancer Biol Ther 4, 819–821. Hynes RO (2002) Integrins: bidirectional, allosteric signaling machines. Cell 110, 673–687. Hynes RO, Lively JC, McCarty JH, Taverna D, Francis SE, Hodivala-Dilke K, Xiao Q (2002) The diverse roles of integrins and their ligands in angiogenesis. Cold Spring Harb Symp Quant Biol 67, 143–153. Jain RK (1994) Barriers to drug delivery in solid tumors. Sci Am 271, 58–65. Jaiswal M, LaRusso NF, Gores GJ (2001) Nitric oxide in gastrointestinal epithelial cell carcinogenesis: linking inflammation to oncogenesis. Am J Physiol Gastrointest Liver Physiol 281, G626–634. Jin H, Varner J (2004) Integrins: roles in cancer development and as treatment targets. Br J Cancer 90, 561–565.
306
B. Garmy-Susini
Jockusch BM, Rudiger M (1996) Crosstalk between cell adhesion molecules: vinculin as a paradigm for regulation by conformation. Trends Cell Biol 6, 311–315. Kajiya K, Hirakawa S, Ma B, Drinnenberg I, Detmar M (2005) Hepatocyte growth factor promotes lymphatic vessel formation and function. Embo J 24, 2885–2895. Kanayama M, Kurotaki D, Morimoto J, Asano T, Matsui Y, Nakayama Y, Saito Y, Ito K, Kimura C, Iwasaki N, et al. (2009) Alpha9 integrin and its ligands constitute critical joint microenvironments for development of autoimmune arthritis. J Immunol 182, 8015–8025. Kaplan RN, Riba RD, Zacharoulis S, Bramley AH, Vincent L, Costa C, MacDonald DD, Jin DK, Shido K, Kerns SA, et al. (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438, 820–827. Karkkainen MJ, Ferrell RE, Lawrence EC, Kimak MA, Levinson KL, McTigue MA, Alitalo K, Finegold DN (2000) Missense mutations interfere with VEGFR-3 signalling in primary lymphoedema. Nat Genet 25, 153–159. Karkkainen MJ, Haiko P, Sainio K, Partanen J, Taipale J, Petrova TV, Jeltsch M, Jackson DG, Talikka M, Rauvala H, et al. (2004) Vascular endothelial growth factor C is required for sprouting of the first lymphatic vessels from embryonic veins. Nat Immunol 5, 74–80. Kilic N, Oliveira-Ferrer L, Neshat-Vahid S, Irmak S, Obst-Pernberg K, Wurmbach JH, Loges S, Kilic E, Weil J, Lauke H, et al. (2007) Lymphatic reprogramming of microvascular endothelial cells by CEA-related cell adhesion molecule-1 via interaction with VEGFR-3 and Prox1. Blood 110, 4223–4233. Komoriya A, Green LJ, Mervic M, Yamada SS, Yamada KM, Humphries MJ (1991) The minimal essential sequence for a major cell type-specific adhesion site (CS1) within the alternatively spliced type III connecting segment domain of fibronectin is leucine-aspartic acid-valine. J Biol Chem 266, 15075–15079. Kozma R, Ahmed S, Best A, Lim L (1995) The Ras-related protein Cdc42Hs and bradykinin promote formation of peripheral actin microspikes and filopodia in Swiss 3T3 fibroblasts. Mol Cell Biol 15, 1942–1952. Krishnan J, Kirkin V, Steffen A, Hegen M, Weih D, Tomarev S, Wilting J, Sleeman JP (2003) Differential in vivo and in vitro expression of vascular endothelial growth factor (VEGF)-C and VEGF-D in tumors and its relationship to lymphatic metastasis in immunocompetent rats. Cancer Res 63, 713–722. Kubo H, Cao R, Brakenhielm E, Makinen T, Cao Y, Alitalo K (2002) Blockade of vascular endothelial growth factor receptor-3 signaling inhibits fibroblast growth factor-2-induced lymphangiogenesis in mouse cornea. Proc Natl Acad Sci U S A 99, 8868–8873. Kwee L, Baldwin HS, Shen HM, Stewart CL, Buck C, Buck CA, Labow MA (1995) Defective development of the embryonic and extraembryonic circulatory systems in vascular cell adhesion molecule (VCAM-1) deficient mice. Development 121, 489–503. Lai KM, Pawson T (2000) The ShcA phosphotyrosine docking protein sensitizes cardiovascular signaling in the mouse embryo. Genes Dev 14, 1132–1145. Leak LV (1976) The structure of lymphatic capillaries in lymph formation. Fed Proc 35, 1863–1871. Liu S, Rose DM, Han J, Ginsberg MH (2000) Alpha4 integrins in cardiovascular development and diseases. Trends Cardiovasc Med 10, 253–257. Liu X, Schnellmann RG (2003) Calpain mediates progressive plasma membrane permeability and proteolysis of cytoskeleton-associated paxillin, talin, and vinculin during renal cell death. J Pharmacol Exp Ther 304, 63–70. Loftus JC, Smith JW, Ginsberg MH (1994) Integrin-mediated cell adhesion: the extracellular face. J Biol Chem 269, 25235–25238. Luo J, Manning BD, Cantley LC (2003) Targeting the PI3K-Akt pathway in human cancer: rationale and promise. Cancer Cell 4, 257–262. Makinen T, Veikkola T, Mustjoki S, Karpanen T, Catimel B, Nice EC, Wise L, Mercer A, Kowalski H, Kerjaschki D, et al. (2001) Isolated lymphatic endothelial cells transduce growth, survival and migratory signals via the VEGF-C/D receptor VEGFR-3. Embo J 20, 4762–4773.
14 Integrin Signaling in Lymphangiogenesis
307
Mandriota SJ, Jussila L, Jeltsch M, Compagni A, Baetens D, Prevo R, Banerji S, Huarte J, Montesano R, Jackson DG, et al. (2001) Vascular endothelial growth factor-C-mediated lymphangiogenesis promotes tumour metastasis. Embo J 20, 672–682. Maula SM, Luukkaa M, Grenman R, Jackson D, Jalkanen S, Ristamaki R (2003) Intratumoral lymphatics are essential for the metastatic spread and prognosis in squamous cell carcinomas of the head and neck region. Cancer Res 63, 1920–1926. Mishima K, Watabe T, Saito A, Yoshimatsu Y, Imaizumi N, Masui S, Hirashima M, Morisada T, Oike Y, Araie M, et al. (2007) Prox1 induces lymphatic endothelial differentiation via integrin alpha9 and other signaling cascades. Mol Biol Cell 18, 1421–1429. Miyamoto S, Teramoto H, Coso OA, Gutkind JS, Burbelo PD, Akiyama SK, Yamada KM (1995) Integrin function: molecular hierarchies of cytoskeletal and signaling molecules. J Cell Biol 131, 791–805. Mould AP, Akiyama SK, Humphries MJ (1995) Regulation of integrin alpha 5 beta 1-fibronectin interactions by divalent cations. Evidence for distinct classes of binding sites for Mn2+, Mg2+, and Ca2+. J Biol Chem 270, 26270–26277. Mumprecht V, Detmar M (2009) Lymphangiogenesis and cancer metastasis. J Cell Mol Med 13, 1405–1416. Nagy JA, Vasile E, Feng D, Sundberg C, Brown LF, Manseau EJ, Dvorak AM, Dvorak HF (2002) VEGF-A induces angiogenesis, arteriogenesis, lymphangiogenesis, and vascular malformations. Cold Spring Harb Symp Quant Biol 67, 227–237. Nobes CD, Hall A (1995) Rho, rac, and cdc42 GTPases regulate the assembly of multimolecular focal complexes associated with actin stress fibers, lamellipodia, and filopodia. Cell 81, 53–62. Okazaki T, Ni A, Ayeni OA, Baluk P, Yao LC, Vossmeyer D, Zischinsky G, Zahn G, Knolle J, Christner C, McDonald DM (2009) alpha5beta1 Integrin blockade inhibits lymphangiogenesis in airway inflammation. Am J Pathol 174, 2378–2387. Owen JD, Ruest PJ, Fry DW, Hanks SK (1999) Induced focal adhesion kinase (FAK) expression in FAK-null cells enhances cell spreading and migration requiring both auto- and activation loop phosphorylation sites and inhibits adhesion-dependent tyrosine phosphorylation of Pyk2. Mol Cell Biol 19, 4806–4818. Padera TP, Kadambi A, di Tomaso E, Carreira CM, Brown EB, Boucher Y, Choi NC, Mathisen D, Wain J, Mark EJ, et al. (2002) Lymphatic metastasis in the absence of functional intratumor lymphatics. Science 296, 1883–1886. Petrova TV, Bono P, Holnthoner W, Chesnes J, Pytowski B, Sihto H, Laakkonen P, Heikkila P, Joensuu H, Alitalo K (2008) VEGFR-3 expression is restricted to blood and lymphatic vessels in solid tumors. Cancer Cell 13, 554–556. Petrova TV, Makinen T, Makela TP, Saarela J, Virtanen I, Ferrell RE, Finegold DN, Kerjaschki D, Yla-Herttuala S, Alitalo K (2002) Lymphatic endothelial reprogramming of vascular endothelial cells by the Prox-1 homeobox transcription factor. Embo J 21, 4593–4599. Plow EF, Haas TA, Zhang L, Loftus J, Smith JW (2000) Ligand binding to integrins. J Biol Chem 275, 21785–21788. Pratt JC, van den Brink MR, Igras VE, Walk SF, Ravichandran KS, Burakoff SJ (1999) Requirement for Shc in TCR-mediated activation of a T cell hybridoma. J Immunol 163, 2586–2591. Prevo R, Banerji S, Ferguson DJ, Clasper S, Jackson DG (2001) Mouse LYVE-1 is an endocytic receptor for hyaluronan in lymphatic endothelium. J Biol Chem 276, 19420–19430. Pronk GJ, de Vries-Smits AM, Buday L, Downward J, Maassen JA, Medema RH, Bos JL (1994) Involvement of Shc in insulin- and epidermal growth factor-induced activation of p21ras. Mol Cell Biol 14, 1575–1581. Reiske HR, Zhao J, Han DC, Cooper LA, Guan JL (2000) Analysis of FAK-associated signaling pathways in the regulation of cell cycle progression. FEBS Lett 486, 275–280. Renyi-Vamos F, Tovari J, Fillinger J, Timar J, Paku S, Kenessey I, Ostoros G, Agocs L, Soltesz I, Dome B (2005) Lymphangiogenesis correlates with lymph node metastasis, prognosis, and angiogenic phenotype in human non-small cell lung cancer. Clin Cancer Res 11, 7344–7353.
308
B. Garmy-Susini
Richardson A, Parsons T (1996) A mechanism for regulation of the adhesion-associated proteintyrosine kinase pp125FAK. Nature 380, 538–540. Ridley AJ (1996) Rho: theme and variations. Curr Biol 6, 1256–1264. Ridley AJ, Hall A (1992) The small GTP-binding protein rho regulates the assembly of focal adhesions and actin stress fibers in response to growth factors. Cell 70, 389–399. Roberts N, Kloos B, Cassella M, Podgrabinska S, Persaud K, Wu Y, Pytowski B, Skobe M (2006) Inhibition of VEGFR-3 activation with the antagonistic antibody more potently suppresses lymph node and distant metastases than inactivation of VEGFR-2. Cancer Res 66, 2650–2657. Rose DM, Alon R, Ginsberg MH (2007) Integrin modulation and signaling in leukocyte adhesion and migration. Immunol Rev 218, 126–134. Salcini AE, McGlade J, Pelicci G, Nicoletti I, Pawson T, Pelicci PG (1994) Formation of Shc-Grb2 complexes is necessary to induce neoplastic transformation by overexpression of Shc proteins. Oncogene 9, 2827–2836. Schaller MD, Parsons JT (1994) Focal adhesion kinase and associated proteins. Curr Opin Cell Biol 6, 705–710. Schlaepfer DD, Hanks SK, Hunter T, van der Geer P (1994) Integrin-mediated signal transduction linked to Ras pathway by GRB2 binding to focal adhesion kinase. Nature 372, 786–791. Schlaepfer DD, Hauck CR, Sieg DJ (1999) Signaling through focal adhesion kinase. Prog Biophys Mol Biol 71, 435–478. Schlaepfer DD, Hunter T (1996) Evidence for in vivo phosphorylation of the Grb2 SH2-domain binding site on focal adhesion kinase by Src-family protein-tyrosine kinases. Mol Cell Biol 16, 5623–5633. Schmid-Schonbein GW (2003) The second valve system in lymphatics. Lymphat Res Biol 1, 25–29; discussion 29–31. Sengbusch JK, He W, Pinco KA, Yang JT (2002) Dual functions of [alpha]4[beta]1 integrin in epicardial development: initial migration and long-term attachment. J Cell Biol 157, 873–882. Shin JW, Min M, Larrieu-Lahargue F, Canron X, Kunstfeld R, Nguyen L, Henderson JE, Bikfalvi A, Detmar M, Hong YK (2006) Prox1 promotes lineage-specific expression of fibroblast growth factor (FGF) receptor-3 in lymphatic endothelium: a role for FGF signaling in lymphangiogenesis. Mol Biol Cell 17, 576–584. Skobe M, Hawighorst T, Jackson DG, Prevo R, Janes L, Velasco P, Riccardi L, Alitalo K, Claffey K, Detmar M (2001) Induction of tumor lymphangiogenesis by VEGF-C promotes breast cancer metastasis. Nat Med 7, 192–198. Smith JW, Piotrowicz RS, Mathis D (1994) A mechanism for divalent cation regulation of beta 3-integrins. J Biol Chem 269, 960–967. Stacker SA, Achen MG (2008) From anti-angiogenesis to anti-lymphangiogenesis: emerging trends in cancer therapy. Lymphat Res Biol 6, 165–172. Stacker SA, Caesar C, Baldwin ME, Thornton GE, Williams RA, Prevo R, Jackson DG, Nishikawa S, Kubo H, Achen MG (2001) VEGF-D promotes the metastatic spread of tumor cells via the lymphatics. Nat Med 7, 186–191. Stacker SA, Williams RA, Achen MG (2004) Lymphangiogenic growth factors as markers of tumor metastasis. Apmis 112, 539–549. Stupack DG, Teitz T, Potter MD, Mikolon D, Houghton PJ, Kidd VJ, Lahti JM, Cheresh DA (2006) Potentiation of neuroblastoma metastasis by loss of caspase-8. Nature 439, 95–99. Sugiura T, Inoue Y, Matsuki R, Ishii K, Takahashi M, Abe M, Shirasuna K (2009) VEGF-C and VEGF-D expression is correlated with lymphatic vessel density and lymph node metastasis in oral squamous cell carcinoma: Implications for use as a prognostic marker. Int J Oncol 34, 673–680. Tadokoro S, Shattil SJ, Eto K, Tai V, Liddington RC, de Pereda JM, Ginsberg MH, Calderwood DA (2003) Talin binding to integrin beta tails: a final common step in integrin activation. Science 302, 103–106.
14 Integrin Signaling in Lymphangiogenesis
309
Tammela T, Zarkada G, Wallgard E, Murtomaki A, Suchting S, Wirzenius M, Waltari M, Hellstrom M, Schomber T, Peltonen R, et al. (2008) Blocking VEGFR-3 suppresses angiogenic sprouting and vascular network formation. Nature 454, 656–660. Tapon N, Hall A (1997) Rho, Rac and Cdc42 GTPases regulate the organization of the actin cytoskeleton. Curr Opin Cell Biol 9, 86–92. Trzewik J, Mallipattu SK, Artmann GM, Delano FA, Schmid-Schonbein GW (2001) Evidence for a second valve system in lymphatics: endothelial microvalves. Faseb J 15, 1711–1717. Van der Auwera I, Van Laere SJ, Van den Eynden GG, Benoy I, van Dam P, Colpaert CG, Fox SB, Turley H, Harris AL, Van Marck EA, et al. (2004) Increased angiogenesis and lymphangiogenesis in inflammatory versus noninflammatory breast cancer by real-time reverse transcriptase-PCR gene expression quantification. Clin Cancer Res 10, 7965–7971. van der Flier A, Sonnenberg A (2001) Function and interactions of integrins. Cell Tissue Res 305, 285–298. Vlahakis NE, Young BA, Atakilit A, Sheppard D (2005) The lymphangiogenic vascular endothelial growth factors VEGF-C and -D are ligands for the integrin alpha9beta1. J Biol Chem 280, 4544–4552. Wang W, Bergh A, Damber JE (2004) Chronic inflammation in benign prostate hyperplasia is associated with focal upregulation of cyclooxygenase-2, Bcl-2, and cell proliferation in the glandular epithelium. Prostate 61, 60–72. Wary KK, Mainiero F, Isakoff SJ, Marcantonio EE, Giancotti FG (1996) The adaptor protein Shc couples a class of integrins to the control of cell cycle progression. Cell 87, 733–743. Wary KK, Mariotti A, Zurzolo C, Giancotti FG (1998) A requirement for caveolin-1 and associated kinase Fyn in integrin signaling and anchorage-dependent cell growth. Cell 94, 625–634. Weller PA, Ogryzko EP, Corben EB, Zhidkova NI, Patel B, Price GJ, Spurr NK, Koteliansky VE, Critchley DR (1990) Complete sequence of human vinculin and assignment of the gene to chromosome 10. Proc Natl Acad Sci U S A 87, 5667–5671. Wicki A, Lehembre F, Wick N, Hantusch B, Kerjaschki D, Christofori G (2006) Tumor invasion in the absence of epithelial-mesenchymal transition: podoplanin-mediated remodeling of the actin cytoskeleton. Cancer Cell 9, 261–272. Wigle JT, Oliver G (1999) Prox1 function is required for the development of the murine lymphatic system. Cell 98, 769–778. Williams CS, Leek RD, Robson AM, Banerji S, Prevo R, Harris AL, Jackson DG (2003) Absence of lymphangiogenesis and intratumoural lymph vessels in human metastatic breast cancer. J Pathol 200, 195–206. Xing Z, Chen HC, Nowlen JK, Taylor SJ, Shalloway D, Guan JL (1994) Direct interaction of v-Src with the focal adhesion kinase mediated by the Src SH2 domain. Mol Biol Cell 5, 413–421. Yang JT, Rayburn H, Hynes RO (1993) Embryonic mesodermal defects in alpha 5 integrin-deficient mice. Development 119, 1093–1105. Yang JT, Rayburn H, Hynes RO (1995) Cell adhesion events mediated by alpha 4 integrins are essential in placental and cardiac development. Development 121, 549–560. Zamir E, Geiger B (2001) Molecular complexity and dynamics of cell-matrix adhesions. J Cell Sci 114, 3583–3590. Zeng Y, Opeskin K, Horvath LG, Sutherland RL, Williams ED (2005) Lymphatic vessel density and lymph node metastasis in prostate cancer. Prostate 65, 222–230.
Chapter 15
Role of Pericytes in Resistance to Antiangiogenic Therapy Koji Matsuo, Chunhua Lu, Mian M.K. Shazad, Robert L. Coleman, and Anil K. Sood
Abstract Angiogenesis plays an important role in the progressive growth of primary tumor and metastasis, and targeting tumor angiogenesis as a therapeutic strategy is showing promise. While these approaches have shown improved survival for some cancer patients, most eventually develop progressive disease due to resistance to antiangiogenic therapy. Recent evidence suggests a functional role for pericytes in acquired resistance to antiangiogenesis agents. Pericytes play an important role in stabilizing blood vessels in the microvasculature regulated by the PDGF ligand (PDGF-BB) and receptor (PDGFR-b) homeostasis, and serve as a local source of survival factors for endothelial cells. Therefore, dual targeting of pericytes (PDGF axis blockers) and endothelial cells (VEGF pathway blockers) may be more efficacious than targeting either cell type alone.
Introduction Oxygen is vitally important to the survival of most organisms. Due to limitations in oxygen diffusion, mammalian cells require blood supply within 100–200 µm of its cell lining (Carmeliet and Jain 2000). As a result, it has been estimated that for a tumor to grow beyond 1mm in size, angiogenesis is required (Naumov et al. 2006). Recruitment of new blood vessels can occur by many pathways including vasculogenesis and angiogenesis (Carmeliet and Jain 2000; Semenza 2003). Pericytes and endothelial cells are
A.K. Sood (*) Department of Gynecologic Oncology, University of Texas M. D. Anderson Cancer Center, Houston, TX, USA and Center for RNA Interference and Non-Coding RNA, University of Texas M. D. Anderson Cancer Center, Houston, TX, USA and Department of Cancer Biology, University of Texas M. D. Anderson Cancer Center, Houston, TX, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_15, © Springer Science+Business Media, LLC 2010
311
312
K. Matsuo et al.
two critical components of blood vessel formation, and the physical and chemical interactions between the two cells are indispensable for normal endothelial function (Allt and Lawrenson 2001; Cuevas et al. 1984; Gerhardt and Betsholtz 2003). For cancer, angiogenesis plays an important role in progressive growth of primary tumor and metastasis. Therefore, targeting tumor angiogenesis may provide an efficient strategy to block tumor growth. Previous studies targeting angiogenesis have widely focused on tumor endothelial cells and their progenitors. While these approaches have shown improved survival for some cancer patients, most eventually develop progressive disease due to resistance to antiangiogenic therapy (Kerbel 2008). Several mechanisms for this acquired antiangiogenic resistance have been hypothesized, one of which is the functional role of pericytes in the tumor microenvironment (Bergers and Hanahan 2008; Kerbel 2008). In this chapter, we will focus on physiology of pericytes in tumor vasculature, targeting pericytes as a therapeutic modality, role of pericytes in resistance to antiangiogenesis therapy, and strategies to overcome pericyte-associated antiangiogenic resistance.
Biology, Physiology, and Pathology of Pericytes Despite decades of research, the origin of pericytes is still not fully understood. It is currently believed that pericytes develop from various types of progenitors depending on their anatomic location in the body. For example, epicardial, mesenchymal, and neural crest cells are believed to be a source for pericytes in the cardiac coronary vasculature, dorsal aorta, and cardiac outflow tract, respectively (Bergwerff et al. 1998; Drake et al. 1998; Vrancken Peeters et al. 1999). Pericytes play an important role in stabilizing blood vessels in the microvasculature (Nehls and Drenckhahn 1993; Sims 1986). A hallmark of pericyte function is their ability to provide vascular stability by depositing matrix or releasing factors that can promote endothelial cell differentiation or quiescence (Armulik et al. 2005). For example, it was demonstrated that pericytes release angiopoietin-1 (Ang-1) that binds to epidermal growth factor homology domains-2 (Tie2) receptor (von Tell et al. 2006). Tie2 receptor expression recently has been identified in mesenchymal cells that are present in the stroma; implicating a repository for tumor vessel pericytes (De Palma et al. 2005). Pericyte homeostasis in normal biology is regulated in significant part by signaling through the PDGF ligand and receptor system (Fig. 15.1) (Lindahl et al. 1999; Pietras and Hanahan 2005). PDGF is a potent mitogen for pericytes and fibroblasts and is composed of A, B, C, and D polypeptide chains and forms homodimers PDGF-AA, BB, CC, and DD and heterodimer PDGF-AB (Andrae et al. 2008). Its biological activities are linked to two tyrosine kinase receptors, PDGFR-a and PDGFR-b (Heldin et al. 1998; Kelly et al. 1991). PDGFR-a binds to PDGF-AA, BB, AB, and CC, whereas PDGFR-b interacts with BB and DD (Betsholtz et al. 2001). Previous studies have shown that up to 90% reduction in pericyte coverage in mice is compatible with postnatal survival (Enge et al. 2002), whereas loss of >95% of pericytes is lethal (Enge et al. 2002; Lindahl et al. 1997), suggesting that a rather
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
313
Fig. 15.1 PDGF pathways in tumor biology. PDGF ligands and receptor pathway is suggested to be involved in the recruitment of pericytes and stabilization of blood vessels
low threshold density of pericytes is required for basal function of microvasculature. Pericyte deficiency, seen in knockout mice lacking platelet-derived growth factor-B (PDGF-B) and its receptor, PDGF receptor-beta (PDGFR-b), results in various changes in microvasculature, including endothelial hyperplasia, vessel dilation, tortuosity, leakage, and rupture, leading to widespread and lethal microhemorrhage and edema in late gestation (Hellstrom et al. 1996b, 2001; Lindahl et al. 1997). Studies of implanted tumors have shown that pericytes initially accumulate at the interface of tumor and host tissue and later around new blood vessels exhibiting close contacts with endothelial cells. Maturation of the tumor-associated vasculature is accompanied quantitatively by a reduced pericyte volume and qualitatively by morphological changes wherein pericytes become flattened and elongated (Verhoeven and Buyssens 1988).
Pericytes and Tumor Angiogenesis Pericytes are tightly attached to the endothelial cells in normal vasculature; however, it is becoming apparent that pericyte coverage in tumor blood vessels is quite distinct. Studies with insulinoma, breast, and lung carcinoma models have demonstrated that pericytes were present on most blood vessels, but 30–50% of the endothelial surface had no pericyte coverage (Morikawa et al. 2002). The extent of pericyte coverage in different tumors ranges from expansive (Schlingemann et al. 1990; Wesseling et al. 1995) to little or none (Benjamin et al. 1999; Eberhard et al. 2000; Johnson and Bruce 1997). Some of these differences may be explained by
314
K. Matsuo et al.
Fig. 15.2 Pericyte coverage on the microvasculature. Dual immunofluorescence staining for CD31 (red) and desmin (green) was performed to visualize endothelial cells and pericytes, respectively. (a) In the normal vasculature, pericytes cover vessels extensively and tightly. (b) In the tumor vasculature, pericytes are morphologically abnormal and tortuous in shape, attach to endothelial cells loosely, and paradoxically extend cytoplasmic processes away from the vessel wall
differences in pericyte marker expression among tumors. Pericytes in tumor vasculature are morphologically abnormal and tortuous in shape, attach to endothelial cells loosely, and paradoxically extend cytoplasmic processes away from the vessel wall (Fig. 15.2) (Lu et al. 2008). Several markers, including a-SMA, NG2, and desmin, have been found to be useful for detecting pericytes (Schlingemann et al. 1990, 1991). Pericytes also play an important role in regulating local blood flow, phagocytosis, modulation of new blood vessel growth and protecting endothelial cells from apoptosis (Hirschi and D’Amore 1996; Reinmuth et al. 2001). It has been demonstrated that pericytes can serve as a local source of vascular endothelial growth factor (VEGF), which is a survival factor for endothelial cells (Brown et al. 2001; Fukumura et al. 1998; Lu et al. 2008; Reinmuth et al. 2001) (Fig. 15.2). Tumor vessels lacking pericytes appear to be more dependent on VEGF for their survival than are vessels invested by pericytes (Benjamin et al. 1999). It is possible that decreased pericyte coverage may make the vascular endothelium vulnerable to VEGF blockage. Tyrosine kinase inhibitors affecting multiple receptor tyrosine kinases may exert their antitumor activity in part by reducing pericyte density in tumor vessels, thereby sensitizing them to inhibition of endothelial receptor tyrosine kinases (Lindblom et al. 1999, 2003; Shaheen et al. 2001). Variable recruitment of pericytes into tumor-associated vasculature may be explained by differences in expression of the PDGF ligands. Recent studies suggest that PDGF-B expression stimulates pericyte recruitment in gliomas (Guo et al. 2003). Both PDGF-AA and BB ligands are expressed in most ovarian cancer samples (Apte et al. 2004). PDGF-B produced by tumor endothelial cells has been shown to be required for the recruitment of adequate number of pericytes as well as proper integration of pericytes in the vascular wall (Abramsson et al. 2003). This may be further enhanced by PDGF-B production directly from tumor cells (Abramsson et al. 2003). Genetic ablation of PDGF-B in endothelial cells leads to
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
315
impaired recruitment of pericytes, resulting in glomerular, cardiac and placental abnormalities using Cre-lox techniques (Bjarnegard et al. 2004). The close interrelationship between these control mechanisms suggests a short-range paracrine system. Using tyrosine kinase inhibitors, a functional role for PDGFR-b has also been implicated in pericyte recruitment in mouse insulinomas (Bergers et al. 2003). Interestingly, the tyrosine kinase inhibitors had no effect on normal tissues with regard to pericyte detachment, but they disrupted the association of pericytes from endothelial cells in tumors (Bergers et al. 2003). These results suggest that there may be differential sensitivity of pericytes in developing immature blood vessels compared to the mature vasculature with regard to dependence on PDGF.
Pericytes and Resistance to Antiangiogenic Therapy One of the first lines of evidence regarding the role of pericytes in protecting the tumor endothelium came from observations following treatment with anti-VEGF therapy. It was noted that anti-VEGF treatment with receptor tyrosine kinase inhibitors in RIP-Tag mice resulted in significant reductions in tumor vasculature, however, the antiangiogenic therapy failed to affect the basement membrane and pericytes (Mancuso et al. 2006).
VEGF Pathway and Pericytes in Tumor Angiogenesis VEGF affects not only endothelial cell proliferation but also cell survival (Carmeliet et al. 1996; Senger et al. 1983). Among several strategies for platinum refractory ovarian cancer, it was concluded that antiangiogenesis approaches targeting VEGF pathway were among the most efficacious (Burger et al. 2007). However, it is possible that anti-VEGF therapy selectively “prunes” blood vessels that have little or no pericyte coverage (Bergers and Hanahan 2008). Upon stimulation of pericytes (e.g., PDGF-BB), local production of factors such as VEGF may serve as a survival factor for the remaining endothelial cells (Lu et al. 2008; Reinmuth et al. 2001). For example, antiangiogenesis therapy with a VEGF receptor inhibitor (AEE788) significantly reduced microvessel density in orthotopic mouse models of ovarian cancer, but the percentage of pericyte-coated vessels was significantly increased, suggesting that pericytes may protect these vessels (Lu et al. 2007).
Resistance to Antiangiogenic Therapy The mechanisms of resistance to antiangiogenic compounds are not well understood, but may overlap to some degree with those factors responsible for resistance to cytotoxic agents (Broxterman and Georgopapadakou 2007; Shahzad et al. 2009).
316
K. Matsuo et al.
For instance, similar to cytotoxic drugs, antiangiogenic agents may face resistance due to decreased bioavailability in the tumors or by upregulation of factors that are involved in apoptosis inhibition. Antiangiogenic agents that affect endothelial cells may not be effectively targeting other components of the tumor vasculature. There are currently many antiangiogenic agents in clinical development for ovarian cancer. While antiangiogenesis treatment strategies are showing promise in clinical trials for patients with ovarian and other cancers (prolonging progressionfree and/or overall survival), most patients eventually develop progression resulting in eventual death (Burger et al. 2007; Jain et al. 2006; Spannuth et al. 2008). Emerging evidence from preclinical and clinical data suggest at least two modes of resistance to antiangiogenic therapy (Bergers and Hanahan 2008). The first mode is adaptive (evasive) resistance and the second is intrinsic (preexisting) nonresponsiveness (Bergers and Hanahan 2008). Possible mechanisms associated with adaptive resistance to anti-VEGF therapy include: (1) activation and/or upregulation of alternative proangiogenic pathways in the tumor; (2) activation and enhancement of invasion and metastasis to provide access to normal tissue vasculature without neovascularization; (3) recruitment of bone marrow-derived proangiogenic cells that differentiate into pericytes; and (4) increased pericyte coverage of tumor microvasculature serving to support survival functions for tumor endothelial cells (Bergers and Hanahan 2008). Antiangiogenic approaches are highly effective in tumor vessels lacking adequate pericyte coverage (Benjamin et al. 1999). It was found that antiangiogenic therapy can effectively reduce the overall vessel density and increase endothelial cell apoptosis in pericyte-negative vessels, but the fraction of pericyte-positive vessels significantly increases after such therapy (McCarty et al. 2004). For example, a VEGFR inhibitor was effective against endothelial cells, affecting early stage angiogenic vasculature, but not against the more mature vasculature in larger tumors (Bergers et al. 2003). The vessels in the tumors intrinsically resistant to antiangiogenic therapy showed slim and tightly covered blood vessels with pericytes (Jain and Booth 2003; Mancuso et al. 2006). These blood vessels are quite distinct from the untreated tumor vasculature, which is often characterized by pericytes that are loosely attached to endothelial cells (Jain 2005; Jain and Booth 2003). On the basis of such data, it is thought that endothelial cells may induce pericyte recruitment to protect themselves from the effects of antiangiogenic therapy. While disruption of the connections between pericytes and endothelial cells may be an attractive approach to increase the antiangiogenic effects on endothelial cells, there is an emerging concern of increased risk of metastasis associated with facilitation of tumor cell entry into the circulation through the disrupted vasculature (Xian et al. 2006). Recruitment of pericytes for external coating and reinforcement of immature blood vessels is essential in the process of normal vascular development (Allt and Lawrenson 2001; Cuevas et al. 1984; Gerhardt and Betsholtz 2003). Antiangiogenic agents may inhibit other kinases as well, and whether their effects on pericytes are mediated by kinases other than PDGFR-b cannot be ruled out. In addition, PDGF-B and PDGFR-b have other roles beside pericyte recruitment,
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
317
including autocrine growth stimulation, formation of tumor stroma, and control of interstitial tumor pressure that may also affect tumor growth (Forsberg et al. 1993; Heldin and Westermark 1999; Pietras et al. 2001; Westermark and Heldin 1991). Other growth factors that may affect pericyte function include endothelin-1 (produced by endothelial cells) (Egginton et al. 1996), interleukin-2 (may increase localization of pericytes to endothelial cell junctions) (Sims et al. 1994), and transforming growth factor-b (produced by pericytes and may inhibit endothelial cell proliferation) (Wakui et al. 1997). However, the precise mechanism underlying recruitment of pericytes to endothelial cells is not well understood. PDGF signaling is suggested as the key factor for pericyte recruitment due to its critical role in pericyte development and function (Hellstrom et al. 1999a, 2001; Lindahl et al. 1997). Tumor hypoxia due to the vascular regression following antiangiogenic treatment appears to induce recruitment of various bone marrow-derived cells to the tumor microenvironment (Du et al. 2008). In the presence of inducible HIF1-a, populations of CD45+ myeloid cells containing Tie2, VEGFR1, CD11b, and F4/80 subpopulations, as well as endothelial and pericyte progenitor cells are released into the circulation where they promote neovascularization in glioblastoma (Du et al. 2008). In absence of HIF1-a, fewer bone marrow-derived cells are recruited to the tumors, which severely impair tumor growth. These data suggest paradoxical induction of tumor angiogenesis via bone marrow-derived vessel progenitor cells after antiangiogenic therapy.
Targeting Pericytes for Antivascular Strategies Signaling through PDGF ligand and receptor system significantly regulates pericyte homeostasis (Lindblom et al. 2003; Pietras and Hanahan 2005). On the basis of the functional and biological roles of the PDGD-BB/PDGFR-b axis on pericyte function, inhibitors of this pathway may be effective in targeting pericytes. Several available clinical or preclinical approaches to inhibiting PDGF-B and PDGFR-b pathway include: (1) tyrosine kinase inhibitors of PDGF receptor, such as imatinib mesylate (Lu et al. 2007 2008), SU6668 (Erber et al. 2004); (2) PDGF-B aptamer, a modified DNA-based aptamer to PDGF-B chain that blocks binding of PDGF-B to its cell-surface receptor (Fredriksson et al. 2002); (3) PDGF Trap targeting PDGF-BB (Lu et al. 2008); and (4) RNA interference. Treatment of RIP-Tag tumors with AX102, an aptamer blocking PDGF-BB, showed loss of pericytes with subsequent loss of endothelial cells; thereby emphasizing the importance of pericytes in tumor associated vasculature (Sennino et al. 2007). However, not surprisingly, monotherapy with these agents has been ineffective and may actually accelerate tumor growth by making the tumor vasculature more immature (Jayson et al. 2005; Spannuth et al. 2008). In contrast, combination of PDGF-BB/ PDGFR-b inhibitors with anti-VEGF or cytotoxic agents may enhance efficacy of
318
K. Matsuo et al.
antiangiogenic therapy. For example, several studies with various tumor models recently have shown that dual targeting of endothelial cells with agents such as AEE788 or SU5416, and pericytes with STI571 or SU6668, is more efficacious than targeting either cell type alone, even in established or drug-resistant tumors (Bergers et al. 2003; Lu et al. 2007). Furthermore, triple combination therapy with a cytotoxic agent, AEE788, and STI571 was effective in regressing bulky tumors in orthotopic mouse models of ovarian cancer using in vivo bioluminescence imaging (Lu et al. 2007). Other dual targeting approaches that showed efficacy in inhibiting in vitro tumor growth include combination of VEGF Trap and PDGF Trap targeting PDGF-BB (Lu et al. 2008), and bevacizumab with AX102. We and others have shown that the PDGF-BB aptamer AX102 can significantly reduce pericyte coverage, and also enhance efficacy of anti-VEGF therapy. It is possible that the efficacy of some of the new tyrosine kinase inhibitors such as SU11248 and SU14813 may be related to targeting both endothelial cells (VEGFR) and pericytes (PDGFR) (Hartmann et al. 2009; Hopkins et al. 2008; Patyna et al. 2006). Similarly, dual targeting of endothelial cells and pericytes has been reported to cause regression of pancreatic tumors in RIP1 Tag 2 mice (Bergers et al. 2003). Thus, for bulky tumors, which are often seen in recurrent disease, these approaches may have clinical relevance. In addition to PDGF-B/PDGFR-b pathway, other pathways such as Ang1 and a receptor tyrosine kinase with immunoglobulin and Tie2 pathway, sphingosine 1-phosphate (S1P) and endothelial differentiation gene-1 (Edg-1), TGF-b1 and activin-like kinase receptor (Alk5) may also offer additional opportunities for therapeutic targeting of pericytes (Erber et al. 2004). Neuropilin is a receptor of semaphorin that was initially implicated in the development of nervous system and in axon guidance (Capparuccia and Tamagnone 2009). However, recent studies have found that neuropilin is expressed in pericytes associated with angiogenesis. It is also suggested that blocking neuropilin-1 function inhibits vascular remodeling, rendering vessels more susceptible to anti-VEGF therapy (Pan et al. 2007). Neuropilin-2 is a high-affinity kinase-deficient receptor for VEGF and semaphorin, and expressed in certain types cancer (Gray et al. 2008). Dual targeting of neuropilin in tumor and pericytes may offer novel approaches for antivascular therapy.
Conclusions In summary, pericytes appear to be emerging as an important target for antivascular approaches. Pericytes not only function to stabilize the vasculature, but may play an important role in conferring resistance to anti-VEGF therapy. Therefore, dual endothelial and pericyte targeting holds promise for more efficacious cancer treatment strategies.
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
319
Fig. 15.3 Interactions between tumor, endothelial cells, and pericytes. Cancer and endothelial cells produce PDGF-BB. PDGF-BB promotes VEGF production by pericytes. VEGF provides local survival signals for mature vasculature (adapted and modified, Lu et al. 2008)
Acknowledgments The authors declare that there is no conflict of interest. KM is supported by the GCF/OCRF Ann Schreiber Ovarian Cancer Research grant and an award from the Meyer and Ida Gordon Foundation #2. MMS is supported by the GCF-Molly Cade ovarian cancer research grant and the NIH/NICHD Baylor WRHR scholarship grant (HD050128). Portions of this work were supported by NIH grants (CA 110793 and 109298), the Ovarian Cancer Research Fund, Inc. (Program Project Development Grant), U. T. M. D. Anderson Cancer Center SPORE (P50CA083639 and P50CA098258), the Marcus Foundation, the Entertainment Industry Foundation, the Blanton-Davis Ovarian Cancer Research Program, and the Betty Anne Asche Murray Distinguished Professorship. We thank Ms. Alison E. Schroeer for assistance with graphic design for Figs. 15.1 and 15.3.
320
K. Matsuo et al.
References Abramsson A, Lindblom P, Betsholtz C (2003) Endothelial and nonendothelial sources of PDGF-B regulate pericyte recruitment and influence vascular pattern formation in tumors. J Clin Invest 112:1142–1151. Allt G, Lawrenson JG (2001) Pericytes: cell biology and pathology. Cells Tissues Organs 169:1–11. Andrae J, Gallini R, Betsholtz C (2008) Role of platelet-derived growth factors in physiology and medicine. Genes Dev 22:1276–1312. Apte SM, Bucana CD, Killion JJ, Gershenson DM, Fidler IJ (2004) Expression of platelet-derived growth factor and activated receptor in clinical specimens of epithelial ovarian cancer and ovarian carcinoma cell lines. Gynecol Oncol 93:78–86. Armulik A, Abramsson A, Betsholtz C (2005) Endothelial/pericyte interactions. Circ Res 97:512–523. Benjamin LE, Golijanin D, Itin A, Pode D, Keshet E (1999) Selective ablation of immature blood vessels in established human tumors follows vascular endothelial growth factor withdrawal. J Clin Invest 103:159–165. Bergers G, Hanahan D (2008) Modes of resistance to anti-angiogenic therapy. Nat Rev Cancer 8:592–603. Bergers G, Song S, Meyer-Morse N, Bergsland E, Hanahan D (2003) Benefits of targeting both pericytes and endothelial cells in the tumor vasculature with kinase inhibitors. J Clin Invest 111:1287–1295. Bergwerff M, Verberne ME, DeRuiter MC, Poelmann RE, Gittenberger-de Groot AC (1998) Neural crest cell contribution to the developing circulatory system: implications for vascular morphology? Circ Res 82:221–231. Betsholtz C, Karlsson L, Lindahl P (2001) Developmental roles of platelet-derived growth factors. Bioessays 23:494–507. Bjarnegard M, Enge M, Norlin J, Gustafsdottir S, Fredriksson S, Abramsson A, Takemoto M, Gustafsson E, Fassler R, Betsholtz C (2004) Endothelium-specific ablation of PDGFB leads to pericyte loss and glomerular, cardiac and placental abnormalities. Development 131: 1847–1857. Brown EB, Campbell RB, Tsuzuki Y, Xu L, Carmeliet P, Fukumura D, Jain RK (2001) In vivo measurement of gene expression, angiogenesis and physiological function in tumors using multiphoton laser scanning microscopy. Nat Med 7:864–868. Broxterman HJ, Georgopapadakou NH (2007) Anticancer therapeutics: a surge of new developments increasingly target tumor and stroma. Drug Resist Updat 10:182–193. Burger RA, Sill MW, Monk BJ, Greer BE, Sorosky JI (2007) Phase II trial of bevacizumab in persistent or recurrent epithelial ovarian cancer or primary peritoneal cancer: a Gynecologic Oncology Group Study. J Clin Oncol 25:5165–5171. Capparuccia L, Tamagnone L (2009) Semaphorin signaling in cancer cells and in cells of the tumor microenvironment – two sides of a coin. J Cell Sci 122:1723–1736. Carmeliet P, Ferreira V, Breier G, Pollefeyt S, Kieckens L, Gertsenstein M, Fahrig M, Vandenhoeck A, Harpal K, Eberhardt C, Declercq C, Pawling J, Moons L, Collen D, Risau W, Nagy A (1996) Abnormal blood vessel development and lethality in embryos lacking a single VEGF allele. Nature 380:435–439. Carmeliet P, Jain RK (2000) Angiogenesis in cancer and other diseases. Nature 407:249–257. Cuevas P, Gutierrez-Diaz JA, Reimers D, Dujovny M, Diaz FG, Ausman JI (1984) Pericyte endothelial gap junctions in human cerebral capillaries. Anat Embryol (Berl) 170:155–159. De Palma M, Venneri MA, Galli R, Sergi Sergi L, Politi LS, Sampaolesi M, Naldini L (2005) Tie2 identifies a hematopoietic lineage of proangiogenic monocytes required for tumor vessel formation and a mesenchymal population of pericyte progenitors. Cancer Cell 8:211–226. Drake CJ, Hungerford JE, Little CD (1998) Morphogenesis of the first blood vessels. Ann N Y Acad Sci 857:155–179.
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
321
Du R, Lu KV, Petritsch C, Liu P, Ganss R, Passegue E, Song H, Vandenberg S, Johnson RS, Werb Z, Bergers G (2008) HIF1alpha induces the recruitment of bone marrow-derived vascular modulatory cells to regulate tumor angiogenesis and invasion. Cancer Cell 13:206–220. Eberhard A, Kahlert S, Goede V, Hemmerlein B, Plate KH, Augustin HG (2000) Heterogeneity of angiogenesis and blood vessel maturation in human tumors: implications for antiangiogenic tumor therapies. Cancer Res 60:1388–1393. Egginton S, Hudlicka O, Brown MD, Graciotti L, Granata AL (1996) In vivo pericyte-endothelial cell interaction during angiogenesis in adult cardiac and skeletal muscle. Microvasc Res 51:213–228. Enge M, Bjarnegard M, Gerhardt H, Gustafsson E, Kalen M, Asker N, Hammes HP, Shani M, Fassler R, Betsholtz C (2002) Endothelium-specific platelet-derived growth factor-B ablation mimics diabetic retinopathy. EMBO J 21:4307–4316. Erber R, Thurnher A, Katsen AD, Groth G, Kerger H, Hammes HP, Menger MD, Ullrich A, Vajkoczy P (2004) Combined inhibition of VEGF and PDGF signaling enforces tumor vessel regression by interfering with pericyte-mediated endothelial cell survival mechanisms. FASEB J 18:338–340. Forsberg K, Valyi-Nagy I, Heldin CH, Herlyn M, Westermark B (1993) Platelet-derived growth factor (PDGF) in oncogenesis: development of a vascular connective tissue stroma in xenotransplanted human melanoma producing PDGF-BB. Proc Natl Acad Sci U S A 90:393–397. Fredriksson S, Gullberg M, Jarvius J, Olsson C, Pietras K, Gustafsdottir SM, Ostman A Landegren U (2002) Protein detection using proximity-dependent DNA ligation assays. Nat Biotechnol 20:473–477. Fukumura D, Xavier R, Sugiura T, Chen Y, Park EC, Lu N, Selig M, Nielsen G, Taksir T, Jain RK, Seed B (1998) Tumor induction of VEGF promoter activity in stromal cells. Cell 94:715–725. Gerhardt H, Betsholtz C (2003) Endothelial-pericyte interactions in angiogenesis. Cell Tissue Res 314:15–23. Gray MJ, Van Buren G, Dallas NA, Xia L, Wang X, Yang AD, Somcio RJ, Lin YG, Lim S, Fan F, Mangala LS, Arumugam T, Logsdon CD, Lopez-Berestein G, Sood AK, Ellis LM (2008) Therapeutic targeting of neuropilin-2 on colorectal carcinoma cells implanted in the murine liver. J Natl Cancer Inst 100:109–120. Guo P, Hu B, Gu W, Xu L, Wang D, Huang HJ, Cavenee WK, Cheng SY (2003) Platelet-derived growth factor-B enhances glioma angiogenesis by stimulating vascular endothelial growth factor expression in tumor endothelia and by promoting pericyte recruitment. Am J Pathol 162:1083–1093. Hartmann JT, Haap M, Kopp HG, Lipp HP (2009) Tyrosine kinase inhibitors – A review on pharmacology, metabolism and side effects. Curr Drug Metab 10:470–481. Heldin CH, Ostman A, Ronnstrand L (1998) Signal transduction via platelet-derived growth factor receptors. Biochim Biophys Acta 1378:F79–F113. Heldin CH, Westermark B (1999) Mechanism of action and in vivo role of platelet-derived growth factor. Physiol Rev 79:1283–1316. Hellstrom AC, Tegerstedt G, Silfversward C, Pettersson F (1999a) Malignant mixed mullerian tumors of the ovary: histopathologic and clinical review of 36 cases. Int J Gynecol Cancer 9:312–316. Hellstrom M, Gerhardt H, Kalen M, Li X, Eriksson U, Wolburg H, Betsholtz C (2001) Lack of pericytes leads to endothelial hyperplasia and abnormal vascular morphogenesis. J Cell Biol 153:543–553. Hellstrom M, Kalen M, Lindahl P, Abramsson A, Betsholtz C (1999b) Role of PDGF-B and PDGFR-beta in recruitment of vascular smooth muscle cells and pericytes during embryonic blood vessel formation in the mouse. Development 3047–3055. Hirschi KK, D’Amore PA (1996) Pericytes in the microvasculature. Cardiovasc Res 32:687–698. Hopkins TG, Marples M, Stark D (2008) Sunitinib in the management of gastrointestinal stromal tumours (GISTs). Eur J Surg Oncol 34:844–850.
322
K. Matsuo et al.
Jain RK (2005) Normalization of tumor vasculature: an emerging concept in antiangiogenic therapy. Science 307:58–62. Jain RK, Booth MF (2003) What brings pericytes to tumor vessels? J Clin Invest 112:1134–1136. Jain RK, Duda DG, Clark JW, Loeffler JS (2006) Lessons from phase III clinical trials on antiVEGF therapy for cancer. Nat Clin Pract Oncol 3:24–40. Jayson GC, Parker GJ, Mullamitha S, Valle JW, Saunders M, Broughton L, Lawrance J, Carrington B, Roberts C, Issa B, Buckley DL, Cheung S, Davies K, Watson Y, Zinkewich-Peotti K, Rolfe L, Jackson A (2005) Blockade of platelet-derived growth factor receptor-beta by CDP860, a humanized, PEGylated di-Fab’, leads to fluid accumulation and is associated with increased tumor vascularized volume. J Clin Oncol 23:973–981. Johnson JP, Bruce JN (1997) Angiogenesis in human gliomas: prognostic and therapeutic implications. EXS 79:29–46. Kelly JD, Haldeman BA, Grant FJ, Murray MJ, Seifert RA, Bowen-Pope DF, Cooper JA, Kazlauskas A (1991) Platelet-derived growth factor (PDGF) stimulates PDGF receptor subunit dimerization and intersubunit trans-phosphorylation. J Biol Chem 266:8987–8992. Kerbel RS (2008) Tumor angiogenesis. N Engl J Med 358:2039–2049. Lindahl P, Bostrom H, Karlsson L, Hellstrom M, Kalen M, Betsholtz C (1999) Role of plateletderived growth factors in angiogenesis and alveogenesis. Curr Top Pathol 93:27–33. Lindahl P, Johansson BR, Leveen P, Betsholtz C (1997) Pericyte loss and microaneurysm formation in PDGF-B-deficient mice. Science 277:242–245. Lindblom P, Gerhardt H, Liebner S, Abramsson A, Enge M, Hellstrom M, Backstrom G, Fredriksson S, Landegren U, Nystrom HC, Bergstrom G, Dejana E, Ostman A, Lindahl P, Betsholtz C (2003) Endothelial PDGF-B retention is required for proper investment of pericytes in the microvessel wall. Genes Dev 17:1835–1840. Lindblom P, Valdemarsson S, Westerdahl J, Tennvall J, Bergenfelz A (1999) Hyperthyroidism after surgery for primary hyperparathyroidism. Langenbecks Arch Surg 384:568–575. Lu C, Kamat AA, Lin YG, Merritt WM, Landen CN, Kim TJ, Spannuth W, Arumugam T, Han LY, Jennings NB, Logsdon C, Jaffe RB, Coleman RL, Sood AK (2007) Dual targeting of endothelial cells and pericytes in antivascular therapy for ovarian carcinoma. Clin Cancer Res 13:4209–4217. Lu C, Thaker PH, Lin YG, Spannuth W, Landen CN, Merritt WM, Jennings NB, Langley RR, Gershenson DM, Yancopoulos GD, Ellis LM, Jaffe RB, Coleman RL, Sood AK (2008) Impact of vessel maturation on antiangiogenic therapy in ovarian cancer. Am J Obstet Gynecol 198:477 e471–e479; discussion 477 e479–e410. Mancuso MR, Davis R, Norberg SM, O’Brien S, Sennino B, Nakahara T, Yao VJ, Inai T, Brooks P, Freimark B, Shalinsky DR, Hu-Lowe DD, McDonald DM (2006) Rapid vascular regrowth in tumors after reversal of VEGF inhibition. J Clin Invest 116:2610–2621. McCarty MF, Wey J, Stoeltzing O, Liu W, Fan F, Bucana C, Mansfield PF, Ryan AJ, Ellis LM (2004) ZD6474, a vascular endothelial growth factor receptor tyrosine kinase inhibitor with additional activity against epidermal growth factor receptor tyrosine kinase, inhibits orthotopic growth and angiogenesis of gastric cancer. Mol Cancer Ther 3:1041–1048. Morikawa S, Baluk P, Kaidoh T, Haskell A, Jain RK, McDonald DM (2002) Abnormalities in pericytes on blood vessels and endothelial sprouts in tumors. Am J Pathol 160:985–1000. Naumov GN, Akslen LA, Folkman J (2006) Role of angiogenesis in human tumor dormancy: animal models of the angiogenic switch. Cell Cycle 5:1779–1787. Nehls V, Drenckhahn D (1993) The versatility of microvascular pericytes: from mesenchyme to smooth muscle? Histochemistry 99:1–12. Pan Q, Chanthery Y, Liang WC, Stawicki S, Mak J, Rathore N, Tong RK, Kowalski J, Yee SF, Pacheco G, Ross S, Cheng Z, Le Couter J, Plowman G, Peale F, Koch AW, Wu Y, Bagri A, Tessier-Lavigne M Watts RJ (2007) Blocking neuropilin-1 function has an additive effect with anti-VEGF to inhibit tumor growth. Cancer Cell 11:53–67. Patyna S, Laird AD, Mendel DB, O’Farrell M, Liang A,C, Guan H, Vojkovsky T, Vasile S, Wang X, Chen J, Grazzini M, Yang CY, Haznedar JO, Sukbuntherng J, Zhong WZ, Cherrington JM, Hu-Lowe D (2006) SU14813: a novel multiple receptor tyrosine kinase inhibitor with potent antiangiogenic and antitumor activity. Mol Cancer Ther 5:1774–1782.
15 Role of Pericytes in Resistance to Antiangiogenic Therapy
323
Pietras K, Hanahan D (2005) A multitargeted, metronomic, and maximum-tolerated dose “chemoswitch” regimen is antiangiogenic, producing objective responses and survival benefit in a mouse model of cancer. J Clin Oncol 23:939–952. Pietras K, Ostman A, Sjoquist M, Buchdunger E, Reed RK, Heldin CH, Rubin K (2001) Inhibition of platelet-derived growth factor receptors reduces interstitial hypertension and increases transcapillary transport in tumors. Cancer Res 61:2929–2934. Reinmuth N, Liu W, Jung YD, Ahmad SA, Shaheen RM, Fan F, Bucana CD, McMahon G, Gallick GE, Ellis LM (2001) Induction of VEGF in perivascular cells defines a potential paracrine mechanism for endothelial cell survival. FASEB J 15:1239–1241. Schlingemann RO, Rietveld FJ, de Waal RM, Ferrone S, Ruiter DJ (1990) Expression of the high molecular weight melanoma-associated antigen by pericytes during angiogenesis in tumors and in healing wounds. Am J Pathol 136:1393–1405. Schlingemann RO, Rietveld FJ, Kwaspen F, van de Kerkhof PC, de Waal RM, Ruiter DJ (1991) Differential expression of markers for endothelial cells, pericytes, and basal lamina in the microvasculature of tumors and granulation tissue. Am J Pathol 138:1335–1347. Semenza GL (2003) Angiogenesis in ischemic and neoplastic disorders. Annu Rev Med 54:17–28. Senger DR, Galli SJ, Dvorak AM, Perruzzi CA, Harvey VS, Dvorak HF (1983) Tumor cells secrete a vascular permeability factor that promotes accumulation of ascites fluid. Science 219:983–985. Sennino B, Falcon BL, McCauley D, Le T, McCauley T, Kurz JC, Haskell A, Epstein DM, McDonald DM (2007) Sequential loss of tumor vessel pericytes and endothelial cells after inhibition of platelet-derived growth factor B by selective aptamer AX102. Cancer Res 67:7358–7367. Shaheen RM, Tseng WW, Davis DW, Liu W, Reinmuth N, Vellagas R, Wieczorek AA, Ogura Y, McConkey DJ, Drazan KE, Bucana CD, McMahon G, Ellis LM (2001) Tyrosine kinase inhibition of multiple angiogenic growth factor receptors improves survival in mice bearing colon cancer liver metastases by inhibition of endothelial cell survival mechanisms. Cancer Res 61:1464–1468. Shahzad MM, Lopez-Berestein G, Sood AK (2009) Novel strategies for reversing platinum resistance. Drug Resist Updat 12:148–152. Sims DE (1986) The pericyte – a review. Tissue Cell 18:153–174. Sims DE, Miller FN, Horne MM, Edwards MJ (1994) Interleukin-2 alters the positions of capillary and venule pericytes in rat cremaster muscle. J Submicrosc Cytol Pathol 26:507–513. Spannuth WA, Sood AK, Coleman RL (2008) Angiogenesis as a strategic target for ovarian cancer therapy. Nat Clin Pract Oncol 5:194–204. Verhoeven D, Buyssens N (1988) Desmin-positive stellate cells associated with angiogenesis in a tumour and non-tumour system. Virchows Arch B Cell Pathol Incl Mol Pathol 54:263–272. von Tell D, Armulik A, Betsholtz C (2006) Pericytes and vascular stability. Exp Cell Res 312:623–629. Vrancken Peeters MP, Gittenberger-de Groot AC, Mentink MM, Poelmann RE (1999) Smooth muscle cells and fibroblasts of the coronary arteries derive from epithelial-mesenchymal transformation of the epicardium. Anat Embryol (Berl) 199:367–378. Wakui S, Furusato M, Muto T, Ohshige H, Takahashi H, Ushigome S (1997) Transforming growth factor-beta and urokinase plasminogen activator presents at endothelial cell-pericyte interdigitation in human granulation tissue. Microvasc Res 54:262–269. Wesseling P, Schlingemann RO, Rietveld FJ, Link M, Burger PC, Ruiter DJ (1995) Early and extensive contribution of pericytes/vascular smooth muscle cells to microvascular proliferation in glioblastoma multiforme: an immuno-light and immuno-electron microscopic study. J Neuropathol Exp Neurol 54:304–310. Westermark B, Heldin CH (1991) Platelet-derived growth factor in autocrine transformation. Cancer Res 51:5087–5092. Xian X, Hakansson J, Stahlberg A, Lindblom P, Betsholtz C, Gerhardt H, Semb H (2006) Pericytes limit tumor cell metastasis. J Clin Invest 116:642–651.
Chapter 16
Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas Urszula M. Polanska, Kieran T. Mellody, and Akira Orimo
Abstract Carcinomas are complex tissues comprised of neoplastic cells and a non-cancerous compartment referred to as the “stroma.” The stroma consists of an extracellular matrix (ECM) and a variety of mesenchymal cells, notably including fibroblasts, myofibroblasts, endothelial cells, pericytes and leukocytes. During tumourigenesis, the tumour-associated stroma is continuously exposed to substantial paracrine signals released by nearby carcinoma cells, and is often populated by considerable numbers of myofibroblasts. These cells are a hallmark of “activated fibroblasts” that are commonly observed in injured and fibrotic tissue. Importantly, their presence in large numbers within the stroma of human carcinomas is related to high-grade malignancies and poor prognoses in patients. Tumour-derived stromal myofibroblast-rich cell populations propagated in vitro stably maintain their myofibroblastic state, as well as an ability to significantly promote tumour growth in xenograft models. Differentiation of stromal cells into myofibroblasts within the tumour corroborates the evolution of the normal stroma towards a tumour-promoting stroma. However, the selective pressures responsible for instigating the generation of an altered stroma, and the molecular alterations that stably maintain the unique tumour-promoting myofibroblastic phenotype remain unclear. This chapter highlights the biological role of the tumour-associated stroma, with a particular focus on myofibroblasts and their ability to promote tumour progression through their interactions with carcinoma cells.
Introduction The epithelium is the tissue composed of highly organised cells that line the cavities and surfaces of structures in organs. The epithelium in the mammary gland, for example, consists of an outer layer of myoepithelial cells and an inner layer of A. Orimo () CR-UK Stromal-Tumour Interaction Group, Paterson Institute for Cancer Research, The University of Manchester, Wilmslow Road, Manchester M20 4BX, UK e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_16, © Springer Science+Business Media, LLC 2010
325
326
U.M. Polanska et al.
Fig. 16.1 Histology shows differences between stroma of the normal mammary gland and of a breast carcinoma. Paraffin sections prepared from an invasive breast cancer region (c, f ) and a non-tumour region (b, e) dissected from the breast tissue of a patient, and the normal breast region (a, d) prepared from a healthy donor, were stained with hematoxylin and eosin (H&E) (a–c) or immunostained with an anti-a-smooth muscle actin (a-SMA) antibody (d–f). Large numbers of myofibroblasts (indicated by arrows) in the cancer region (f) and myoepithelial cells (indicated by an arrow in d and e) surrounding luminal epithelial cells are shown. Scale bar, 75 mm [from Orimo et al. (2005)]
luminal epithelial cells attached to a basement membrane (Fig. 16.1a–e). The epithelium is surrounded by an adjacent stroma that consists of mesenchymal cells and a specialised ECM that helps maintain tissue integrity and homeostasis (Fig. 16.1a, b). Not only is the stroma important for cell differentiation and proliferation during tissue development, but it also plays a major role in wound healing, fibrosis and tumourigenesis. The malignant transformation of epithelial cells depends on genetic and/or epigenetic alterations that accumulate within certain oncogenes and/or tumour suppressor genes [reviewed in Hanahan and Weinberg (2000) and Vogelstein and Kinzler (2004)]. These perturbations alone, however, are not sufficient for the carcinoma to fully develop into a high-grade malignancy [for reviews, see DePinho (2000), Bissell and Radisky (2001), Coussens and Werb (2002) and Mueller and Fusenig (2004)]. Malignant epithelial cells require additional support from the surrounding stroma that evolves alongside and promotes tumourigenic progression [reviewed in Rowley (1998), Cunha et al. (2003), Bhowmick et al. (2004b), Lorusso and Ruegg (2008) and Polyak et al. (2009)]. The presence of large numbers of stromal cells and a dense ECM is characteristic of the stroma found in many types of carcinomas, including breast cancer, and is often referred to as a “desmoplastic” or “reactive” stroma (Fig. 16.1c, f). The desmoplastic stroma is believed to play a major role in promoting the growth and progression of the carcinoma [reviewed in Ronnov-Jessen et al. (1996), Bissell and Radisky (2001), Elenbaas and Weinberg (2001), Desmouliere et al. (2004), Kalluri and Zeisberg (2006), De Wever et al. (2008) and Shimoda et al. (2009)].
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
327
Fibroblasts and myofibroblasts, collectively designated carcinoma-associated fibroblasts (CAFs), are one of the most significant cell types found within the desmoplastic stroma. CAFs interact with cancer cells and collaborate with other components of the stroma through their production and secretion of various growth factors, cytokines and chemokines. During tumourigenesis, signalling molecules secreted by these cells effectively mediate neoangiogenesis, carcinoma cell proliferation, survival, motility and invasion [reviewed in Ronnov-Jessen et al. (1996), De Wever and Mareel (2003), Kalluri and Zeisberg (2006), Orimo and Weinberg (2006) and Ostman and Augsten (2009)].
Myofibroblasts Involved in Tissue Fibrosis Share Characteristics with Tumour-Associated Myofibroblasts Myofibroblasts exhibit a highly contractile phenotype through their expression of a-smooth muscle actin (a-SMA), which assembles into a robust network of stress fibers. These cells share characteristics with smooth muscle cells and pericytes [Desmouliere et al. 1993; reviewed in Serini and Gabbiani (1999)]. Myofibroblasts are commonly found in injured tissues (Gabbiani et al. 1971), and play essential roles in regulating the healing response by facilitating contraction of the wound. They also secrete a number of soluble growth factors, pro-inflammatory cytokines, chemokines and matrix metalloproteinases (MMPs), which act together in concert to induce angiogenesis and further facilitate tissue repair [reviewed in Powell et al. (1999) and Ronnov-Jessen et al. (1996)]. Myofibroblasts are activated and proliferate in injured tissues to promote wound healing. These cells are cleared from the tissue once the healing process is complete. Myofibroblasts are also found in a number of pathological fibrotic diseases including rheumatoid arthritis, Dupuytren’s disease, idiopathic pulmonary fibrosis, scleroderma, pulmonary hypertension and arteriosclerosis [Kim et al. 2009; reviewed in Varga (2002), Leask (2006), Varga and Abraham (2007) and Ihn (2008)]. In contrast to the normal healing response, myofibroblasts in fibrotic tissues continue to persist in an activated state. Their presence perpetuates the inflammatory process resulting in an improper tissue repair response that disrupts organ and tissue function. Several studies have identified crucial roles for the pleiotropic cytokine TGF-b in promoting fibrosis [for review, see Varga and Abraham (2007)]. TGF-b promotes mesenchymal cell proliferation, migration, differentiation, adhesion, survival, cytokine induction and ECM protein synthesis [reviewed in Derynck et al. (2001), Bierie and Moses (2006) and Massague (2008)]. It also induces and activates many profibrotic gene signalling pathways in stromal cells such as PDGF signalling that promotes the proliferation and survival of myofibroblasts during fibrogenesis [reviewed in Bonner (2004)]. This cytokine also induces expression level of endothelin-1 (ET-1) which signals through the ETA and ETB receptors. Binding of ET1 to the ETA and ETB receptors activates the JNK pathway that increases further expression of endogenous ET-1 (Shi-Wen et al. 2006). The resulting autocrine
328
U.M. Polanska et al.
signalling loop contributes to the persistence of the myofibroblastic phenotype in myofibroblasts derived from scleroderma (Shi-Wen et al. 2006). In addition, TGF-b induces Jagged1-Notch autocrine signalling that is involved in myofibroblast differentiation during fibrosis (Liu et al. 2009). Importantly, myofibroblasts extracted from pathological fibrotic tissues, frequently show the establishment of a TGF-b-Smad autocrine signalling loop that is stably maintained during propagation of these cells in vitro [Kim et al. 2009; reviewed in Varga (2002)]. Myofibroblasts, derived from fibrotic tissues, secrete high levels of TGF-b and show upregulated expression levels of TGF-b receptors. This results in a constitutive activation of the TGF-b-Smad signalling pathway in an autocrine fashion that helps to maintain the myofibroblastic phenotype. Inhibition of this pathway not only substantially abates the myofibroblastic phenotype, as demonstrated using both in vitro and in vivo models of fibrosis, but is also associated with decreased ECM deposition and suppression of the inflammatory cytokine response [reviewed in Varga and Abraham (2007) and Varga and Pasche (2009)]. However, the mechanism(s) by which TGF-b autocrine signalling is initially triggered and programmed in these cells remain unclear. One possibility is that sustained exposure of stromal cells to chronic inflammation, and other selective pressures, results in altered gene expression and epigenetic modifications that generate the myofibroblastic phenotype. The tissue remodelling processes involved in wound healing and fibrosis physiologically resemble those occurring within the tumour-associated stroma [reviewed in Dvorak (1986) and Schafer and Werner (2008)]. An important study highlighted the ability of activated fibroblasts, derived from injured tissue, to actively promote tumourigenesis (Hu et al. 2009). Such fibroblasts, extracted from an inflamed rheumatoid arthritis patient, were co-injected with carcinoma cells into a recipient mouse. These activated fibroblasts increased levels of cyclooxygenase-2 (COX-2) expression in the carcinoma cells, an enzyme that is involved in inflammationassociated tumourigenesis and which boosts tumour progression. The carcinoma exploits the host’s wound healing response in order to promote tumourigenesis [reviewed in Bissell and Radisky (2001), Furlow (2005), Kalluri and Zeisberg (2006) and Orimo and Weinberg (2006)]. However, it is not known if molecular signalling, such as TGF-b autocrine signalling, which plays a central role in myofibrogenesis during fibrosis, also mediates the tumour-promoting myofibroblastic phenotype observed in tumour-associated myofibroblasts.
Carcinoma-Associated Fibroblast Characterised as Tumour-Promoting Myofibroblasts The presence of myofibroblasts in large numbers in the stroma of human tumours (Fig. 16.1c, f) is often associated with high-grade malignancies and poor prognoses in patients (Cardone et al. 1997; Maeshima et al. 2002; Tsujino et al. 2007; Kellermann et al. 2008). In order to examine the functional role of tumour-associated myofibroblasts during tumourigenesis, CAFs have been isolated from various different
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
329
types of human carcinomas including carcinomas of the breast (Allinen et al. 2004; Orimo et al. 2005; Lebret et al. 2007; Casey et al. 2008; Mercier et al. 2008), prostate (Olumi et al. 1999; Ao et al. 2007), ovary (Yang et al. 2006), pancreas (Hwang et al. 2008), skin (Sneddon et al. 2006) and esophagus (Zhang et al. 2009). Our own observations indicate that CAFs extracted from a human breast tumour mass exhibit a different morphology in culture compared to “counterpart fibroblasts” extracted from a noncancerous region of the breast in the same individual or “normal fibroblasts” isolated from a healthy donor (Fig. 16.2a). CAFs appear more slender and spindly and upon reaching confluency, they often lose their contact inhibition of cell growth, continuing to proliferate and stack on top of one another (Fig. 16.2a). These observations are consistent with the findings of an earlier pioneer study that characterised CAFs isolated from human prostate carcinomas (Olumi et al. 1999). Primary cultured fibroblasts, extracted from human normal breast tissues, are a heterogeneous cell population that contains a small number of cells exhibiting a myofibroblastic characteristic, as determined by a-SMA expression. However, CAFs contain a noticeably larger proportion of myofibroblasts compared to those present within counterpart and normal fibroblasts (Fig. 16.2b) (Ronnov-Jessen et al. 1992; Orimo et al. 2005). Moreover, CAFs in comparison with control fibroblasts, when embedded within a collagen gel, display a more contractile phenotype that is indicative of the activated fibroblastic trait in myofibroblasts (Fig. 16.2c). They are also able to retain the activated phenotype during propagation in vitro for at least ten population doublings (PDs), without further ongoing interaction with carcinoma cells (Fig. 16.2c). Taken together, these observations demonstrate that CAFs include increased proportions of myofibroblasts that stably maintain an activated fibroblast phenotype. Genetic analyses of CAFs have previously been performed to determine if these cells are themselves neoplastic. Karyotype analysis to date has not detected any chromosomal abnormalities in CAFs (Olumi et al. 1999; Orimo et al. 2005). Consistently, genome-wide genetic analyses have also failed to detect any significant loss of heterozygosity (LOH) in these cells (Allinen et al. 2004; Qiu et al. 2008; Walter et al. 2008); (discussed in “Carcinoma-Associated Fibroblast Characterised as Tumour-Promoting Myofibroblasts” section). CAFs exhibit no anchorage-independent growth ability in vitro and no tumourigenicity in vivo when injected alone, without carcinoma cells, into a recipient mouse (Fig. 16.3a–c). They also undergo cellular senescence in vitro after 15 PDs (Orimo et al. 2005). Collectively, these findings strongly suggest that CAFs are intrinsically non-tumourigenic. CAFs, however, can promote tumour growth through their interactions with carcinoma cells. When injected along with human breast carcinoma MCF-7-ras cells subcutaneously into an immunodeficient mouse, CAFs show an ability to more significantly increase the rate and size of tumour growth compared to control counterpart and normal fibroblasts (Orimo et al. 2005). Importantly, CAFs retain such a tumour-promoting phenotype until, at least, ten PDs in the absence of continuous interaction with carcinoma cells (Orimo et al. 2005). Several other independent studies have also confirmed the ability of CAFs extracted from a number of different human carcinomas to promote tumourigenesis
330
a
U.M. Polanska et al. i
Normal fibroblasts
iii
CAFs
b
Normal fibroblasts
iii
Counterpart fibroblasts
iv
i
α-SMA Ab
ii
Counterpart fibroblasts
iv
ii
CAFs
c
α-SMA Ab
α-SMA Ab
Confluent
d
CAFs
CAFs
Control IgG
180 170 160
Normal fibroblasts-2
Counterpart fibroblasts
Area of collagen gels (mm2)
Normal fibroblasts-1
150 140 130 120 110 100 90 80
Normal fibroblast-1 Normal fibroblast-2 Counterpart fibroblasts CAFs (5 PDs) CAFs (10 PDs)
70 60
CAFs
50
0
10
20
30
40
hr
50
60
70
80
Fig. 16.2 CAFs include a larger proportion of myofibroblasts. (a) Normal fibroblasts (i) were extracted from a reduction mammoplasty performed on a healthy donor. Counterpart fibroblasts (ii) were also extracted from a noncancerous region of the breast at least 2 cm from the outer tumour margin. In addition, CAFs (iii, iv) were extracted from an invasive ductal breast carcinoma region in the same individual. CAFs (iv) show a more spindle-like morphology when they reach near confluency compared with normal (i) and counterpart fibroblasts (ii) (scale bar 50 mm). (b) An increased proportion of a-SMA-positive myofibroblasts are observed in CAFs. Immunofluorescence of normal fibroblasts (i), counterpart fibroblasts (ii) or CAFs (iii) was performed using an antia-SMA antibody. CAFs (iv) were stained with control IgG as a negative control (scale bar: 50 mm); a–d from Orimo et al. (2005). (c) Equal numbers of various fibroblasts were incubated in a collagen gel for 72 h. CAFs exhibit a greater ability to contract collagen gels compared to normal or counterpart fibroblasts. The graph shows the rate of contraction of the collagen gels for each group of fibroblasts tested. Error bars indicate standard deviation of the mean [from Orimo et al. (2005)]
(Olumi et al. 1999; Yang et al. 2006; Hwang et al. 2008; Hu et al. 2009). Immunohistochemical analysis using a human-specific anti-vimentin antibody, which detects human mesenchymal cells but not MCF-7-ras carcinoma cells or the mouse host cells, shows that either the injected CAFs or control fibroblasts are present in considerable numbers within the advanced tumour xenografts (Fig. 16.3d) (Orimo et al. 2005). Taken together, these studies suggest that CAFs promote
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
331
Fig. 16.3 CAFs themselves are not neoplastic but are capable of promoting tumour growth through their interactions with apposed carcinoma cells. (a) Various human mammary fibroblasts, including CAFs, show no anchorage-independent growth activity when propagated in vitro. Tissue sections from mice inoculated with CAFs alone without cancer cells were stained with hematoxylin and eosin (b), and with a human-specific anti-vimentin antibody and hematoxylin (c). (d) CAFs survive alongside carcinoma cells within the advanced tumour xenografts. Immunohistochemical analysis of a paraffin section prepared from a 62-day-old tumour with antibodies against human vimentin (brown) and GFP (pink), allowing specific detection of human vimentin-expressing CAFs and GFP-labelled MCF-7-ras carcinoma cells, respectively. Nuclear staining was performed using hemotoxylin (blue) [from Orimo et al. (2005)]. Scale bars: 100 mm (a) and 50 mm (b–d)
tumour growth via their ongoing interactions with the apposed carcinoma cells throughout tumour progression. Moreover, these fibroblasts are able to retain their stable tumour-promoting properties in a cell-autonomous fashion.
Somatic Genetic and Epigenetic Alterations in Tumour-Associated Stroma The CAFs are non-neoplastic cells capable of promoting tumourigenesis through their interactions with carcinoma cells. However, there is much debate as to whether tumour-associated stromal cells harbour somatic genetic alterations [reviewed in
332
U.M. Polanska et al.
Weinberg (2008b), Haviv et al. (2009) and Polyak et al. (2009)]. Furthermore, the molecular mechanism(s) responsible for maintaining the stable tumour-promoting CAF phenotype also remain unclear. A number of studies support the presence of somatic genetic alterations, including LOH and mutations, in stromal regions microdissected from various human carcinomas, including those of the breast (Moinfar et al. 2000; Wernert et al. 2001; Kurose et al. 2002; Patocs et al. 2007), ovary (Tuhkanen et al. 2004), colon (Wernert et al. 2001), bladder (Paterson et al. 2003) and head and neck (Weber et al. 2007). Indeed, 25.6% in the hereditary group and 19.4% in the sporadic group of breast cancer patients show somatic p53 mutations in microdissected tumour-associated stromal regions (Patocs et al. 2007). In addition, there is a higher risk associated with regional lymph node metastases in these patients. In prostatic cell-specific oncogene-driven transgenic mice with a p53+/- background, the developing tumour puts selective pressure on the tumour-associated stroma. This results in frequent loss of the wild type p53 allele within the stroma (Hill et al. 2005). Carcinoma cell-derived paracrine signals may act upon the adjacent stroma to give rise to p53-negative highly proliferative mesenchymal cells. However, whilst alterations in the p53 gene may confer upon stromal cells their tumour-promoting properties, it is unclear whether such alterations do give rise to the myofibroblast-rich desmoplastic stroma often seen in human carcinomas. Furthermore, patients with Li-Fraumeni syndrome are prone to developing soft tissue and bone sarcomas due to inherent p53 alterations, as do p53 knockout transgenic mice. However, it is not known why breast cancer patients, who may lose p53 function within the tumourassociated stroma, very rarely develop tumours of mesenchymal origin (Bolton and Sieunarine 1990; Tokudome et al. 2005). Most of the studies that detected genetic alterations in the tumour stromal compartments were performed using paraffin-fixed tumour tissues. Conversely, genome-wide genetic analyses, including single nucleotide polymorphism and comparative genomic hybridisation arrays, performed using freshly frozen tissues derived from human breast carcinomas, fail to detect any significant genetic alterations in cancer-associated stroma (Qiu et al. 2008). This was highlighted in a recent study that reported only one chromosomal LOH event among 35 samples of cancer-associated stromal regions dissected from different breast and ovarian tumours [Qiu et al. 2008; reviewed in Haviv et al. (2009)]. Similar analyses performed using CAFs have also failed to detect any significant genetic alterations, as described earlier (Allinen et al. 2004; Qiu et al. 2008; Walter et al. 2008). Therefore, detection of genetic alterations in some studies, whilst not in others, may be in part due to the different methods used to acquire and prepare tissues for analysis, the use of paraffin-fixed tissues instead of frozen tissues or the different types of genomic analyses performed [commented in Weinberg (2008b) and reviewed in Haviv et al. (2009)]. Future independent analyses are therefore required to clarify these issues. The presence of altered profiles of epigenetic modifications, such as DNA methylation within the genome of CAFs, has also been reported (Hu et al. 2005; Fiegl et al. 2006; Hanson et al. 2006; Jiang et al. 2008). In addition, CAFs, unlike
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
333
normal fibroblasts, are able to confer epigenetic modifications upon apposed human breast epithelial cells. This was demonstrated in an important study in which CAFs were co-cultured for three weeks with human breast epithelial cells in an assay that allowed cell-to-cell contacts to form (Lin et al. 2008). CAFs activated the AKT signalling pathway in the epithelial cells, which hypermethylated the promoter region for the tumour suppressor gene cystatin M. Future studies are required to determine if the CAF-induced epigenetic modifications are stably reprogrammed within the epithelial cells, and if this, in turn, promotes tumourigenesis in vivo. Understanding what epigenetic modifications in CAFs mediate their phenotype, and how these cells confer epigenetic alterations upon neighbouring epithelial cells, will further highlight the molecular mechanisms that underlie their tumour-promoting properties.
Heterogeneous Cellular Origins of Carcinoma-Associated Myofibroblasts Human dermal fibroblasts are heterogeneous in nature and often display locationspecific gene expression signatures (Chang et al. 2002). Similarly, tumour-associated stromal fibroblasts are also considered to be a heterogeneous cell population (Sugimoto et al. 2006; Anderberg and Pietras 2009). Although the expression of a-SMA defines the myofibroblastic phenotype, to date, no specific protein markers have been identified that determine the cellular origins of myofibroblasts found in the stroma of tumour. These cells are therefore characterised by expression of a number of mesenchymal protein markers that include tenascin-C, fibroblast specific protein-1, fibroblast activation protein, PDGF receptor-a/b, NG2 chondroitin sulfate proteoglycan, vimentin and prolyl 4-hydroxylase [Sugimoto et al. 2006; reviewed in Kalluri and Zeisberg (2006), De Wever et al. (2008) and Anderberg and Pietras (2009)]. Various cell types are thought to act as a source for the emergence of tumourpromoting CAF myofibroblasts (Fig. 16.4). Resident fibroblasts are likely a major source of myofibroblasts [reviewed in Powell et al. (1999) and Hinz (2007)], whilst other mesenchymal cell types including endothelial cells, pericytes, smooth muscle cells and preadipocytes are also capable of converting into myofibroblasts [Zeisberg et al. 2007; reviewed in Kalluri and Zeisberg (2006) and De Wever et al. (2008)]. Bone marrow-derived progenitors, such as fibrocytes and mesenchymal stem cells (MSCs), are also reported to differentiate into myofibroblasts within the tumour. It has been shown, for example, that bone marrow-derived cells contribute to the generation of nearly 40% of myofibroblasts present in the tumour-associated stroma of 28-day-old pancreatic tumour xenografts (Ishii et al. 2003), and of approximately 25% in spontaneous pancreatic tumours developing for 16–18 weeks in another transgenic mouse model (Direkze et al. 2004). It is also possible that a small number of myofibroblasts present within the normal fibroblast populations are clonally expanded in response to selective pressure imposed by the tumour
334
U.M. Polanska et al.
Endothelial cells
Myofibroblasts
Fibroblasts
Fibroblasts with genetic alterations (e.g., p53)
?
Pericytes
Differentiation Selection Differentiation/ Selection
Myofibroblasts BMDCs/MSCs
blood vessels
Tumour progression
Fibroblasts cancer cells
Carcinoma in situ Invasive carcinoma Fig. 16.4 Schematic representation of cellular origins of CAF myofibroblasts during tumour progression; CAFs are proposed to be derived from various cell types, such as pre-existing fibroblasts, pericytes or endothelial cells. Bone marrow-derived progenitor cells (BMDCs) including MSCs and fibrocytes also differentiate into myofibroblasts. A small population of residual myofibroblasts or mesenchymal cells that carry genetic alterations may be clonally selected during tumour progression. The genetically altered mesenchymal cells may or may not differentiate into myofibroblasts
microenvironment. In addition, the stromal cells that have acquired epigenetic or genetic alterations may differentiate into myofibroblasts and these cells are then also likely to undergo clonal expansion. The tumour stroma continues to remodel itself during tumour progression and actively recruits various cell types into the tumour mass where they act as different sources for myofibroblasts. It is not known, however, if different cells of origins in the myofibroblast populations in the stroma of tumour exhibit different tumourpromoting properties. Once generated, myofibroblasts maintain their ability to substantially promote tumour growth and progression in many aspects via their interactions with carcinoma cells and other host stromal cells.
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
335
Normal Stroma-Derived Tumour-Suppressive Signalling and Tumour Stroma-Derived Tumour-Promoting Signalling TGF-b signalling functions as a potent tumour suppressor in epithelial cells during the early stages of tumourigenesis. However, the role of this signalling pathway within the stroma is unclear. A study using a transgenic mouse model, in which TGF-b type II receptor (TbRII) gene is deleted specifically in stromal fibroblasts, has demonstrated the inherent tumour-suppressing effects of stromal TGF-b signalling on epithelial cells (Bhowmick et al. 2004a). The TbRII-deficient fibroblasts secrete high levels of hepatocyte growth factor (HGF) that causes spontaneous malignant transformation of c-Met receptor-expressing forestomach epithelial cells and tumourigenesis within the mice. More recently, the tumour-suppressive role of Notch signalling within the stroma has been demonstrated in a transgenic mouse model. The chimeric deletion of the Notch 1 gene, within the epidermis of a conditional Notch 1flox/flox mouse, causes a wound-like microenvironment, as characterised by the infiltration of myofibroblasts and leukocytes into the affected tissue (Demehri et al. 2009). These cells secrete elevated levels of TGF-b, keratinocyte growth factor and stromal cell-derived factor 1 (SDF-1, also called CXCL12). This altered microenvironment, rich in growth factors and cytokines, promotes chemical carcinogeninduced skin tumourigenesis of Notch 1-expressing keratinocytes in a non-cell autonomous manner. The inherent tumour-suppressive role of stromal Notch signalling is therefore important in maintaining tissue integrity and homeostasis within the skin. In addition to TGF-b and Notch signalling, phosphatase and tensin homologue (PTEN) signalling within the stroma plays an important function in inhibiting tumourigenesis. Myofibroblasts in lung tissues from idiopathic pulmonary fibrosis patients show decreased levels of PTEN expression (White et al. 2006). Inhibition of PTEN also increases numbers of myofibroblasts observed in bleomycin-induced fibrotic lung (White et al. 2006). Furthermore, suppression of PTEN expression and inhibition of its activity stimulate differentiation of cultured fibroblasts into myofibroblasts. Indeed, the loss of PTEN expression in mammary stromal fibroblasts within a transgenic mouse model accelerates the initiation and progression of ErbB2-driven breast epithelial cancers (Trimboli et al. 2009). The ErbB2 tumours grow more rigorously in the presence of PTEN-deficient stromal fibroblasts and show a highly desmoplastic stromal reaction, as determined by an increase in the numbers of infiltrating macrophages. However, it is not known if the PTENdeficient stromal fibroblasts per se are a source of tumour-associated myofibroblasts in the stroma. In addition, a gene expression profile obtained from PTEN-null mammary stromal fibroblasts indicates upregulation of genes predominantly involved in ECM remodelling, inflammation and fibrosis. These observations further highlight the inherent tumour-suppressive role of the normal stroma, and specifically stromal PTEN, TGF-b and Notch signalling pathways, involved in suppressing tumourigenesis.
336
U.M. Polanska et al.
Through its interactions with carcinoma cells, the initially tumour-suppressing stroma, however, overtime acquires an ability to promote tumourigenesis. SDF-1 is an important proangiogenic chemokine that is known to actively regulate tumourigenesis [reviewed in Burger and Kipps (2006), Nagasawa (2006), Ratajczak et al. (2006), Kryczek et al. (2007) and Petit et al. (2007)]. Unlike fibroblasts found within the non-cancerous stroma, myofibroblasts present within the tumour stroma actively secrete SDF-1 (Allinen et al. 2004; Orimo et al. 2005; Ao et al. 2007; Tait et al. 2007). This chemokine, upon binding to its cognate receptor CXCR4 constitutively expressed on the cell surface of carcinoma cells, stimulates carcinoma cell proliferation (Hall and Korach 2003; Allinen et al. 2004; Orimo et al. 2005). In the later stages of tumour progression, TGF-b signalling in epithelial cancer cells also promotes tumourigenesis. A recent study showed that elevated secretion of TGF-b by CAFs (San Francisco et al. 2004; Rosenthal et al. 2004; Ao et al. 2007), upregulates the expression levels of CXCR4 in prostatic epithelial cells. This in turn boosts stromal SDF-1-stimulated tumourigenesis in vivo (Ao et al. 2007). TGF-b within the stroma therefore also plays a permissive role in promoting some aspects of tumourigenesis. The Hedgehog (Hh) signaling is a key regulator of embryonic development but has also been implicated in the development of many different type of cancers [Karhadkar et al. 2004; reviewed in Taipale and Beachy (2001)]. The activity of the transmembrane receptor Smoothened (Smo) is constitutively suppressed by the unbound Patched 1 (PTCH1) receptor, the cognate receptor for Hh ligands. However, subsequent binding of Hh ligands to PTCH1 reverses this effect, and thus activates the Smo signalling pathway. Recent studies have demonstrated that ligand-dependent activation of the Hh pathway in the tumour-associated stroma plays an important role in influencing tumourigenesis [reviewed in Theunissen and de Sauvage (2009)]. For example, the human pancreatic carcinoma cell-secreted Sonic hedgehog (Shh) ligand has been shown to induce a desmoplastic stromal reaction within the tumour (Bailey et al. 2008). Exposure of cultured pancreatic stellate cells to Shh also facilitated their differentiation into myofibroblasts (Bailey et al. 2008). Moreover, the expression levels of both PTCH1 and Gli1, a downstream target of the Smo signalling pathways, are elevated in the stromal region of human pancreatic carcinomas (Tian et al. 2009). Hh ligands, however, are expressed within the epithelial carcinoma compartment. Importantly, inhibition of Smo protein activity on the tumour-associated stromal cells using a potent, small molecular antagonist of the Hh pathway attenuated growth of human colon adenocarcinoma cells in a tumour xenograft model (Yauch et al. 2008). These findings demonstrate that the Hh ligands secreted by carcinoma cells instigate tumour-promoting Hh signalling in nearby stromal cells in a paracrine fashion, through binding of the PTCH1 receptor expressed by stromal cells. However, the contribution of soluble paracrine factors, released by the tumourpromoting stroma in response to the activation of Hh signalling, to tumourigenesis remains unknown. A recent study highlighted an ability of Hh signalling within the tumour-associated stroma to affect drug delivery in cancer (Olive et al. 2009). Systemic administration
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
337
of a Smo inhibitor into a murine model of pancreatic cancer, which expresses endogenous mutant K-ras and p53 alleles, inhibits Hh signalling. This alters the composition of the desmoplastic stroma and attenuates proliferation of stromal myofibroblasts. The authors also claim that the associated increase of the tumour vascular density facilitates efficient delivery of the chemotherapeutic agent into the tumour. As another possible mechanism, the increase in drug delivery may be due to decreased interstitial fluid pressure that is normally instigated by the tumourassociated stroma with high Hh signalling. This may prevent transport of the drug from the vasculature into the tumour tissue [for reviews, see Jain (2001) and Pietras et al. (2003)]. Stromal fibroblasts interact with carcinoma epithelial cells within the tumour, yet this relationship between the fibroblasts and carcinoma cells is reciprocal. It is likely that carcinoma cells initially secrete cell signalling molecules (e.g., TGF-b), which can induce the conversion of fibroblasts and/or their progenitors into myofibroblasts. In vitro experiments have demonstrated that stromal fibroblasts exposed to medium conditioned by carcinoma cells can undergo differentiation into myofibroblasts (Guo et al. 2008; Kellermann et al. 2008; Noma et al. 2008). Importantly, the ability of stromal fibroblasts to promote tumourigenesis and tumour progression is substantially associated with myofibroblast differentiation [De Wever et al. 2004; Orimo et al. 2005; reviewed in Desmouliere et al. (2004) and De Wever et al. (2008)]. Carcinoma cells can induce myofibroblast differentiation of the apposed stromal cells and the resulting myofibroblasts can in turn reciprocate to promote growth and progression of carcinoma cells. The development of the malignant tumour and promotion towards advanced stages of carcinomas require the successful co-evolution of both cancer cells and stromal cells [reviewed in De Wever et al. (2008), Weinberg (2008b), Polyak et al. (2009) and Shimoda et al. (2009)].
Tumour-Associated Stroma Promotes Neoangiogenesis The CAF-promoted neoangiogenesis is mediated by a number of proangiogenic cytokines and chemokines including SDF-1, vascular endothelial growth factor-A (VEGF-A) and basic fibroblast growth factor 2 (FGF2) (Fig. 16.5a) (Orimo et al. 2005; Guo et al. 2008; Noma et al. 2008; Pietras et al. 2008). The elevated levels of VEGF promoter activity have been detected in stromal fibroblasts present in tumours developing within transgenic mice expressing VEGF promoter-driven GFP (Fukumura et al. 1998). Other studies have shown that inhibition of stroma cell-derived VEGF significantly decreases the growth rate of fibrosarcoma cells (Dong et al. 2004). In addition, TGF-b-primed myofibroblasts secrete increased levels of VEGF that stimulates vascular formation by endothelial cells co-cultured in a 3D assay (Noma et al. 2008). These studies further support the pivotal roles of stromal VEGF in stimulating neoangiogenesis. A recent study has reported that tumour-associated stromal fibroblasts not only stimulate tumour angiogenesis but also mediate resistance to anti-angiogenic therapy.
338
U.M. Polanska et al.
Fibroblast-enriched cell fractions extracted from the mouse lymphomas, which were resistant to anti-VEGFA antibody treatment, upregulated PDGF-C expression by 200-fold compared to fibroblasts isolated from non-refractory tumours (Crawford et al. 2009). This increase in stromal PDGF-C expression in the therapy-resistant tumours was attributed to an increase in neoangiogenesis occurring during the course of anti-VEGF treatment. Similarly, in a cervical squamous cell carcinoma
a TGFβ PDGF Shh
HGF SDF-1 IGF-2 Gremlin-1 SFRP-1
proliferation
FGF2 VEGF-A SDF-1 Endothelial cells/EPCs TGFβ SDF-1 Tenascin C and W HGF MMPs
angiogenesis
SFRP1
invasion
survival
Fig. 16.5 Tumour-promoting effects exerted by CAFs. (a) Tumour-promoting paracrine signalling exerted by tumour-associated myofibroblasts: HGF (Tokunou et al. 2001; De Wever et al. 2004), SDF-1 (Orimo et al. 2005; Daly et al. 2008), IGF-2 (insulin-like growth factors 2) (Zhu et al. 2007), Gremlin-1 (Sneddon et al. 2006), SFRP1 (secreted frizzled-related protein) (Joesting et al. 2005), FGF2 (fibroblasts growth factor 2) (Pietras et al. 2008), VEGF-A (vascular endothelial growth factor A) (Fukumura et al. 1998), TGF-b (Weinberg 2007), MMPs, tenascin C (De Wever et al. 2004) and tenascin W (Degen et al. 2007). Carcinoma-derived signalling molecules causing myofibroblast differentiation and proliferation: TGF-b (Desmouliere et al. 1993; Ronnov-Jessen et al. 1996; Serini and Gabbiani 1999; Hinz et al. 2001), PDGF (Bostrom et al. 1996; Jester et al. 2002; Bonner 2004), Sonic hedgehog (Shh) (Bailey et al. 2008; Olive et al. 2009). (b) CAFs stimulate neoangiogenesis; Human breast MCF-7-ras tumours that developed in the presence of CAFs are highly angiogenic. Sections prepared from MCF-7-ras tumour xenografts containing various fibroblasts were stained with anti-CD31 antibody to identify endothelial cells (iv, v, vi) or by Masson’s trichrome (i, ii, iii), which stains collagen in blue, red blood cells in red and cancer cells in purple. Tumours, developed in the presence of CAFs, are rich in collagen (iii) and have a highly vascularised stroma (iii, vi). Scale bar: 100 mm; (c) Human breast MCF-7-ras tumours developed in the presence of CAFs contain increased numbers of Sca1+CD31+ endothelial progenitor cells (EPCs) compared to tumours derived from MCF-7-ras admixed with control fibroblasts. The advanced tumours were dissociated into a single cell suspension and stained with anti-Sca1 and anti-CD31 antibodies; p<0.05 [from Orimo et al. (2005)]
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
b
+Normal fibroblasts
i
339
+CAFs
+Counterpart fibroblasts
ii
iii
v
vi
capillary vessels
Masson’s trichrome
iv
CD31 Ab
c
0.9
n=8
Sca1+/CD31+ EPCs in tumour (%)
0.8 0.7 0.6 0.5 0.4 0.3 n=7
0.2 0.1 0
Fig. 16.5 (continued)
capillary vessels
*
n=6 n=6
340
U.M. Polanska et al.
transgenic mouse model, treatment with the PDGF receptor kinase inhibitor also resulted in suppression of FGF2 and FGF7 expression in tumour stromal fibroblasts, which attenuated neoangiogenesis and tumour growth (Pietras et al. 2008). The increased level of SDF-1 secreted by CAFs also substantially promotes neoangiogenesis as it actively recruits circulating endothelial progenitor cells (EPCs) into the primary tumour (Orimo et al. 2005) (Fig. 16.5c). These cells are able to differentiate into vascular endothelial cells that form tumour-associated capillaries. Also, varieties of leukocytes including neutrophils, macrophages, CD11b+Gr-1+myeloid-derived suppressor cells (MDSCs) and lymphocytes are often recruited into the tumour-associated stroma. Experimental data derived from both in vitro and in vivo studies demonstrate crucial roles for these tumourassociated leukocytes in boosting neoangiogenesis and tumour progression [DeNardo et al. 2009; reviewed in Condeelis and Pollard (2006), de Visser et al. (2006) and Mantovani et al. (2008)]. Importantly, the tumour microenvironment polarises tumour-associated macrophages towards either a pro-tumour or an anti-tumour phenotype (Sica et al. 2008). Recent literature also indicates that the immunosuppressive cytokine TGF-b, abundantly produced within the tumour microenvironment, plays an essential role in switching the polarity of tumourassociated neutrophils towards a pro-tumourigenic phenotype (Fridlender et al. 2009). These studies highlight an important role for the tumour microenvironment in promoting tumour progression by regulating the polarity of tumour-associated leukocytes.
Roles of Tumour-Associated Stroma in Promoting Cancer Cell Invasion and Metastasis Invasion and metastasis are responsible for 90% of cancer-associated mortality, and the majority of cancer cells at the time of death are often found in metastases rather than in the primary tumour (Weinberg 2007). The tumour invasion-metastasis cascade is a complex multi-step process that includes localised invasion of carcinoma cells, entrance into the systemic circulation, survival during transportation, extravasation, the establishment of micrometastases in distal tissues and colonisation to form macroscopic metastases [reviewed by Fidler (2003)]. It has long been assumed that dissemination of metastatic carcinoma cells depends largely on their cell-autonomous effects due to epigenetic and/or genetic alterations that accumulate within these malignant cells. However, emerging evidence now proposes a different schema in which metastatic spread is determined relatively early on in tumour progression and is not totally dependant on the acquisition of additional genetic alterations within carcinoma cells [reviewed in Weinberg (2008a) and Klein (2009)]. Two distinct populations of normal mammary epithelial cells, originating from breast tissue within the same individual, upon receiving the same set of genetic aberrations, give rise to different types of carcinoma cells with distinct metastatic traits (Ince et al. 2007). This suggests that the differentiation
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
341
program in the parental cell of origin, from which carcinoma cells are derived, is also actively present in the derivative malignant cells. This differentiation program appears to strongly influence the development of their metastatic propensity (Gupta et al. 2005; Ince et al. 2007). The tumour microenvironment likely serves as another important determinant that encourages carcinoma cells in the primary tumour to become motile and invasive, and to disseminate into distant organs. Interaction of carcinoma cells with the tumour-associated stroma facilitates the invasion-metastasis cascade (Karnoub et al. 2007). Tumour-associated stromal fibroblasts play a significant role in regulating such migratory and invasive behaviours in carcinoma cells. This is supported by evidence indicating that stromal myofibroblasts are frequently present at the invasive front of human carcinomas [Nakayama et al. 1998; Sivridis et al. 2005; reviewed in De Wever and Mareel (2003)]. In addition, it has been shown that CAF myofibroblasts, extracted from human carcinomas, increase the migratory and invasive propensity of the cancer cells co-cultured alongside them in collagen gels [De Wever et al. 2004; Gaggioli et al. 2007; Jodele et al. 2006; Daly et al. 2008; reviewed in Desmouliere et al. (2004)]. Tenascin C, HGF and SDF-1, which are secreted by CAFs, also play a role in mediating CAF-stimulated invasion of carcinoma cells (De Wever et al. 2004; Daly et al. 2008) (Fig. 16.5a). A number of additional signalling molecules are involved in initiating and regulating the metastatic spread of cancer cells to distant organs [reviewed in Siegel and Massague (2003) and Nguyen et al. (2009)] (Fig. 16.5a). TGF-b signalling is a key molecular pathway involved in this process. Human cancer cells, interacting with the stroma at the leading edge of the invasive front, often undergo epithelial to mesenchymal transition (EMT), as determined by the downregulation of epithelial markers and upregulation of mesenchymal markers [reviewed in Thiery and Sleeman (2006), Weinberg (2007) and Thiery et al. (2009)]. Therefore EMT, induced in cancer cells by signals released from the tumour-associated stroma (e.g., TGF-b), may be an early cellular event that promotes the invasion and metastatic dissemination of cancer cells. Cancer cells that undergo EMT lose their cell-to-cell junctions and adopt a more migratory and invasive phenotype. The reverse process, mesenchymal to epithelial transition, allows carcinoma cells that have undergone metastasis to anchor themselves to basement membranes, form cell-to-cell contacts and colonise distal organs and tissues [reviewed in Brabletz et al. (2005), Lee et al. (2006), Thiery and Sleeman (2006), Hugo et al. (2007) and Kalluri and Weinberg (2009)]. A recent study, using in vivo imaging to monitor activation of TGF-b-Smad signalling in a rat mammary carcinoma xenograft model, found that TGF-b signalling is activated in carcinoma cells rapidly moving alone, but not slowly moving as an aggregate (Giampieri et al. 2009). The tumour-associated stroma, which is rich in TGF-b, may induce activation of the TGF-b signalling in carcinoma cells. Activation of this signalling pathway is likely to increase their motility along with their ability to invade basement membranes and intravasate into blood vessels. In contrast, carcinoma cells that metastasise into distal organs, where the stroma is not rich in TGF-b, lose activation of the TGF-b signalling pathway. These cells therefore
342
U.M. Polanska et al.
lose their migratory phenotype and are able to proliferate and colonise these tissues. Furthermore, primary tumour microenvironment-derived TGF-b primes breast carcinoma cells to elevate expression levels of angiopoietin-like 4, which disassembles the vascular endothelial cell-to-cell junctions and facilitates extravasation into the lung (Padua et al. 2008). Taken together, these observations reinforce the view that the stromal microenvironment in the primary tumour has the ability to influence tumour invasion and metastasis. The concept of a pre-metastatic niche in which non-cancer cells initiate metastasis of carcinoma cells into distal specific organs has recently been proposed (Kaplan et al. 2005). Bone marrow-derived hematopoietic progenitor cells, positive for vascular endothelial growth factor receptor-1 create pre-metastatic niches in a distal organ prior to the arrival of carcinoma cells. The pre-metastatic niche supports recruitment of disseminated carcinoma cells and facilitates the further formation of micro- and macro-metastases [Hiratsuka et al. 2002; reviewed in Psaila and Lyden (2009)]. Tumour cell-derived paracrine soluble factors such as VEGF-A, placental growth factor, inflammatory S100 chemokines and the extracellular enzyme lysyl oxidase, play key roles in generating the pre-metastatic niche (Kaplan et al. 2005; Hiratsuka et al. 2008; Erler et al. 2009). Some of these factors are also substantially produced by the stroma within the primary tumour. Future studies are therefore likely to help clarify the role of the tumour-associated stroma in generating the pre-metastatic niche. During colonisation of tissues and organs distal to the site of the primary tumour, disseminated metastatic cells establish micro- and macro-metastases. In order to do so, cancer cells actively remodel the surrounding stroma, changing its inherent tumour-suppressing properties. The malignant cells thereby create a hospitable tissue microenvironment that supports their proliferation and the manifestation of metastatic lesions. Interestingly, the actively remodelled stroma often shares a histopathological resemblance with stroma present in the primary tumour.
Conclusions/Perspectives The importance of the tumour-associated stroma in regulating tumour growth and progression is becoming evermore apparent. During tumourigenesis, the stroma evolves in response to signals released by carcinoma cells. Myofibroblast differentiation during tumourigenesis corroborates the evolution of the normal stroma towards a tumour-promoting stroma. CAF myofibroblasts, frequently found within the stroma of many different types of carcinomas, contribute enormously to tumourigenesis. However, many aspects of the mesenchymal–epithelial interactions in the tumour have still to be elucidated. Delineating these complex paracrine and autocrine signalling pathways within the tumour microenvironment will be necessary to fully understand the role of the stroma in instigating and assisting the progression of cancer. This research will
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
343
also hopefully allow us to understand the mechanisms that initiate the differentiation of cells into myofibroblasts, and the factors involved in the maintenance of their tumour-promoting phenotype. Importantly, recent research highlights noticeable roles for the tumour-associated stroma in promoting invasion and metastasis. Understanding better the function of myofibroblasts during wound healing and tissue fibrosis may provide additional insights into the broad spectrum of stromal signalling that progresses cancer cell growth, invasion and metastasis. Drug resistance often occurs in carcinoma cells that are genetically unstable and thus accumulate adaptive mutations during the course of chemoprevention study. Novel therapies are being developed that target the tumour-promoting signalling in stromal cells that possess a more stable genome compared to carcinoma cells. Whilst the debate continues as to whether the tumour-associated stroma is genetically altered and may develop resistance to chemotherapies, it still remains an attractive therapeutic target in the fight against cancer. Disrupting the interactions that occur between the stroma and carcinoma cells using novel therapies, alone or in combination with conventional treatments that target the carcinoma cells directly, may result in greater efficacy. Drug design may focus on preventing progenitor cells, fibroblasts or other precursory cell types from differentiating into myofibroblasts. Understanding the complex mesenchymal–epithelial interactions will further our understanding of the nature of human cancer. Such research is challenging but holds great promise, particularly with the advent of new pharmacological agents designed to specifically target these stromal–tumour interactions. Acknowledgements We would like to extend our gratitude to Dr. Robert A. Weinberg, Whitehead Institute for Biomedical Research, for his ongoing guidance and excellent supervision. We also thank Dr. Radoslaw Polanski for the critical reading of this chapter and Cancer research UK for the funding.
References Allinen M, Beroukhim R, Cai L et al (2004) Molecular characterization of the tumor microenvironment in breast cancer. Cancer Cell 6:17–32. Anderberg C, Pietras K (2009) On the origin of cancer-associated fibroblasts. Cell Cycle 8:1461–1462. Ao M, Franco OE, Park D et al (2007) Cross-talk between paracrine-acting cytokine and chemokine pathways promotes malignancy in benign human prostatic epithelium. Cancer Res 67:4244–4253. Bailey JM, Swanson BJ, Hamada T et al (2008) Sonic hedgehog promotes desmoplasia in pancreatic cancer. Clin Cancer Res 14:5995–6004. Bhowmick NA, Chytil A, Plieth D et al (2004a) TGF-beta signaling in fibroblasts modulates the oncogenic potential of adjacent epithelia. Science 303:848–851. Bhowmick NA, Neilson EG, Moses HL (2004b) Stromal fibroblasts in cancer initiation and progression. Nature 432:332–337. Bierie B, Moses HL (2006) Tumour microenvironment: TGFbeta: the molecular Jekyll and Hyde of cancer. Nat Rev Cancer 6:506–520.
344
U.M. Polanska et al.
Bissell MJ, Radisky D (2001) Putting tumours in context. Nat Rev Cancer 1:46–54. Bolton B, Sieunarine K (1990) Carcinosarcoma: a rare tumour of the breast. Aust N Z J Surg 60:917–919. Bonner JC (2004) Regulation of PDGF and its receptors in fibrotic diseases. Cytokine Growth Factor Rev 15:255–273. Bostrom H, Willetts K, Pekny M et al (1996) PDGF-A signaling is a critical event in lung alveolar myofibroblast development and alveogenesis. Cell 85:863–873. Brabletz T, Hlubek F, Spaderna S et al (2005) Invasion and metastasis in colorectal cancer: epithelial-mesenchymal transition, mesenchymal-epithelial transition, stem cells and betacatenin. Cells Tissues Organs 179:56–65. Burger JA, Kipps TJ (2006) CXCR4: a key receptor in the crosstalk between tumor cells and their microenvironment. Blood 107:1761–1767. Cardone A, Tolino A, Zarcone R et al (1997) Prognostic value of desmoplastic reaction and lymphocytic infiltration in the management of breast cancer. Panminerva Med 39:174–177. Casey TM, Eneman J, Crocker A et al (2008) Cancer associated fibroblasts stimulated by transforming growth factor beta1 (TGF-beta 1) increase invasion rate of tumor cells: a population study. Breast Cancer Res Treat 110:39–49. Chang HY, Chi JT, Dudoit S et al (2002) Diversity, topographic differentiation, and positional memory in human fibroblasts. Proc Natl Acad Sci USA 99:12877–12882. Condeelis J, Pollard JW (2006) Macrophages: obligate partners for tumor cell migration, invasion, and metastasis. Cell 124:263–266. Coussens LM, Werb Z (2002) Inflammation and cancer. Nature 420:860–867. Crawford Y, Kasman I, Yu L et al (2009) PDGF-C mediates the angiogenic and tumorigenic properties of fibroblasts associated with tumors refractory to anti-VEGF treatment. Cancer Cell 15:21–34. Cunha GR, Hayward SW, Wang YZ et al (2003) Role of the stromal microenvironment in carcinogenesis of the prostate. Int J Cancer 107:1–10. Daly AJ, McIlreavey L, Irwin CR (2008) Regulation of HGF and SDF-1 expression by oral fibroblasts—implications for invasion of oral cancer. Oral Oncol 44:646–651. Degen M, Brellier F, Kain R et al (2007) Tenascin-W is a novel marker for activated tumor stroma in low-grade human breast cancer and influences cell behavior. Cancer Res 67:9169–9179. Demehri S, Turkoz A, Kopan R (2009) Epidermal Notch1 loss promotes skin tumorigenesis by impacting the stromal microenvironment. Cancer Cell 16:55–66. DeNardo DG, Barreto JB, Andreu P et al (2009) CD4(+) T cells regulate pulmonary metastasis of mammary carcinomas by enhancing protumor properties of macrophages. Cancer Cell 16: 91–102. DePinho RA (2000) The age of cancer. Nature 408:248–254. Derynck R, Akhurst RJ, Balmain A (2001) TGF-beta signaling in tumor suppression and cancer progression. Nat Genet 29:117–129. Desmouliere A, Geinoz A, Gabbiani F et al (1993) Transforming growth factor-beta 1 induces alpha-smooth muscle actin expression in granulation tissue myofibroblasts and in quiescent and growing cultured fibroblasts. J Cell Biol 122:103–111. Desmouliere A, Guyot C, Gabbiani G (2004) The stroma reaction myofibroblast: a key player in the control of tumor cell behavior. Int J Dev Biol 48:509–517. de Visser KE, Eichten A, Coussens LM (2006) Paradoxical roles of the immune system during cancer development. Nat Rev Cancer 6:24–37. De Wever O, Mareel M (2003) Role of tissue stroma in cancer cell invasion. J Pathol 200:429–447. De Wever O, Nguyen QD, Van Hoorde L et al (2004) Tenascin-C and SF/HGF produced by myofibroblasts in vitro provide convergent proinvasive signals to human colon cancer cells through RhoA and Rac. FASEB J 18:1016–1018. De Wever O, Demetter P, Mareel M et al (2008) Stromal myofibroblasts are drivers of invasive cancer growth. Int J Cancer 123:2229–2238. Direkze NC, Hodivala-Dilke K, Jeffery R et al (2004) Bone marrow contribution to tumor-associated myofibroblasts and fibroblasts. Cancer Res 64:8492–8495.
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
345
Dong J, Grunstein J, Tejada M et al (2004) VEGF-null cells require PDGFR alpha signalingmediated stromal fibroblast recruitment for tumorigenesis. EMBO J 23:2800–2810. Dvorak HF (1986) Tumors: wounds that do not heal. Similarities between tumor stroma generation and wound healing. N Engl J Med 315:1650–1659. Elenbaas B, Weinberg RA (2001) Heterotypic signaling between epithelial tumor cells and fibroblasts in carcinoma formation. Exp Cell Res 264:169–184. Erler JT, Bennewith KL, Cox TR et al (2009) Hypoxia-induced lysyl oxidase is a critical mediator of bone marrow cell recruitment to form the premetastatic niche. Cancer Cell 15:35–44. Fidler IJ (2003) The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nat Rev Cancer 3:453–458. Fiegl H, Millinger S, Goebel G et al (2006) Breast cancer DNA methylation profiles in cancer cells and tumor stroma: association with HER-2/neu status in primary breast cancer. Cancer Res 66:29–33. Fridlender ZG, Sun J, Kim S et al (2009) Polarization of tumor-associated neutrophil phenotype by TGF-beta: “N1” versus “N2” TAN. Cancer Cell 16:183–194. Fukumura D, Xavier R, Sugiura T et al (1998) Tumor induction of VEGF promoter activity in stromal cells. Cell 94:715–725. Furlow B (2005) Wound healing processes and tumour growth. Lancet Oncol 6:363. Gabbiani G, Ryan GB, Majne G (1971) Presence of modified fibroblasts in granulation tissue and their possible role in wound contraction. Experientia 27:549–550. Gaggioli C, Hooper S, Hidalgo-Carcedo C et al (2007) Fibroblast-led collective invasion of carcinoma cells with differing roles for RhoGTPases in leading and following cells. Nat Cell Biol 9:1392–1400. Giampieri S, Manning C, Hooper S et al (2009) Localized and reversible TGFbeta signalling switches breast cancer cells from cohesive to single cell motility. Nat Cell Biol 11:1287–1296. Guo X, Oshima H, Kitmura T et al (2008) Stromal fibroblasts activated by tumor cells promote angiogenesis in mouse gastric cancer. J Biol Chem 283:19864–19871. Gupta PB, Kuperwasser C, Brunet JP et al (2005) The melanocyte differentiation program predisposes to metastasis after neoplastic transformation. Nat Genet 37:1047–1054. Hall JM, Korach KS (2003) Stromal cell-derived factor 1, a novel target of estrogen receptor action, mediates the mitogenic effects of estradiol in ovarian and breast cancer cells. Mol Endocrinol 17:792–803. Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57–70. Hanson JA, Gillespie JW, Grover A et al (2006) Gene promoter methylation in prostate tumorassociated stromal cells. J Natl Cancer Inst 98:255–261. Haviv I, Polyak K, Qiu W et al (2009) Origin of carcinoma associated fibroblasts. Cell Cycle 8:589–595. Hill R, Song Y, Cardiff RD et al (2005) Selective evolution of stromal mesenchyme with p53 loss in response to epithelial tumorigenesis. Cell 123:1001–1011. Hinz B (2007) Formation and function of the myofibroblast during tissue repair. J Invest Dermatol 127:526–537. Hinz B, Mastrangelo D, Iselin CE et al (2001) Mechanical tension controls granulation tissue contractile activity and myofibroblast differentiation. Am J Pathol 159:1009–1020. Hiratsuka S, Nakamura K, Iwai S et al (2002) MMP9 induction by vascular endothelial growth factor receptor-1 is involved in lung-specific metastasis. Cancer Cell 2:289–300. Hiratsuka S, Watanabe A, Sakurai Y et al (2008) The S100A8-serum amyloid A3-TLR4 paracrine cascade establishes a pre-metastatic phase. Nat Cell Biol 10:1349–1355. Hu M, Yao J, Cai L et al (2005) Distinct epigenetic changes in the stromal cells of breast cancers. Nat Genet 37:899–905. Hu M, Peluffo G, Chen H et al (2009) Role of COX-2 in epithelial-stromal cell interactions and progression of ductal carcinoma in situ of the breast. Proc Natl Acad Sci USA 106:3372–3377. Hugo H, Ackland ML, Blick T et al (2007) Epithelial-mesenchymal and mesenchymal-epithelial transitions in carcinoma progression. J Cell Physiol 213:374–383.
346
U.M. Polanska et al.
Hwang RF, Moore T, Arumugam T et al (2008) Cancer-associated stromal fibroblasts promote pancreatic tumor progression. Cancer Res 68:918–926. Ihn H (2008) Autocrine TGF-beta signaling in the pathogenesis of systemic sclerosis. J Dermatol Sci 49:103–113. Ince TA, Richardson AL, Bell GW et al (2007) Transformation of different human breast epithelial cell types leads to distinct tumor phenotypes. Cancer Cell 12:160–170. Ishii G, Sangai T, Oda T et al (2003) Bone-marrow-derived myofibroblasts contribute to the cancer-induced stromal reaction. Biochem Biophys Res Commun 309:232–240. Jain RK (2001) Delivery of molecular and cellular medicine to solid tumors. Adv Drug Deliv Rev 46:149–168. Jester JV, Huang J, Petroll WM et al (2002) TGFbeta induced myofibroblast differentiation of rabbit keratocytes requires synergistic TGFbeta, PDGF and integrin signaling. Exp Eye Res 75:645–657. Jiang L, Gonda TA, Gamble MV et al (2008) Global hypomethylation of genomic DNA in cancerassociated myofibroblasts. Cancer Res 68:9900–9908. Jodele S, Blavier L, Yoon JM et al (2006) Modifying the soil to affect the seed: role of stromalderived matrix metalloproteinases in cancer progression. Cancer Metastasis Rev 25:35–43. Joesting MS, Perrin S, Elenbaas B et al (2005) Identification of SFRP1 as a candidate mediator of stromal-to-epithelial signaling in prostate cancer. Cancer Res 65:10423–10430. Kalluri R, Weinberg RA (2009) The basics of epithelial-mesenchymal transition. J Clin Invest 119:1420–1428. Kalluri R, Zeisberg M (2006) Fibroblasts in cancer. Nat Rev Cancer 6:392–401. Kaplan RN, Riba RD, Zacharoulis S et al (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438:820–827. Karhadkar SS, Bova GS, Abdallah N et al (2004) Hedgehog signalling in prostate regeneration, neoplasia and metastasis. Nature 431:707–712. Karnoub AE, Dash AB, Vo AP et al (2007) Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 449:557–563. Kellermann MG, Sobral LM, da Silva SD et al (2008) Mutual paracrine effects of oral squamous cell carcinoma cells and normal oral fibroblasts: induction of fibroblast to myofibroblast transdifferentiation and modulation of tumor cell proliferation. Oral Oncol 44:509–517. Kim KK, Wei Y, Szekeres C et al (2009) Epithelial cell alpha3beta1 integrin links beta-catenin and Smad signaling to promote myofibroblast formation and pulmonary fibrosis. J Clin Invest 119:213–224. Klein CA (2009) Parallel progression of primary tumours and metastases. Nat Rev Cancer 9:302–312. Kryczek I, Wei S, Keller E et al (2007) Stroma-derived factor (SDF-1/CXCL12) and human tumor pathogenesis. Am J Physiol Cell Physiol 292:C987–C995. Kurose K, Gilley K, Matsumoto S et al (2002) Frequent somatic mutations in PTEN and TP53 are mutually exclusive in the stroma of breast carcinomas. Nat Genet 32:355–357. Leask A (2006) Scar wars: is TGFbeta the phantom menace in scleroderma? Arthritis Res Ther 8:213–219. Lebret SC, Newgreen DF, Thompson EW et al (2007) Induction of epithelial to mesenchymal transition in PMC42-LA human breast carcinoma cells by carcinoma-associated fibroblast secreted factors. Breast Cancer Res 9:R19. Lee JM, Dedhar S, Kalluri R et al (2006) The epithelial-mesenchymal transition: new insights in signaling, development, and disease. J Cell Biol 172:973–981. Lin HJ, Zuo T, Lin CH et al (2008) Breast cancer-associated fibroblasts confer AKT1-mediated epigenetic silencing of Cystatin M in epithelial cells. Cancer Res 68:10257–10266. Liu T, Hu B, Choi YY et al (2009) Notch1 signaling in FIZZ1 induction of myofibroblast differentiation. Am J Pathol 174:1745–1755. Lorusso G, Ruegg C (2008) The tumor microenvironment and its contribution to tumor evolution toward metastasis. Histochem Cell Biol 130:1091–1103.
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
347
Maeshima AM, Niki T, Maeshima A et al (2002) Modified scar grade: a prognostic indicator in small peripheral lung adenocarcinoma. Cancer 95:2546–2554. Mantovani A, Allavena P, Sica A et al (2008) Cancer-related inflammation. Nature 454: 436–444. Massague J (2008) TGFbeta in Cancer. Cell 134:215–230. Mercier I, Casimiro MC, Wang C et al (2008) Human breast cancer-associated fibroblasts (CAFs) show caveolin-1 downregulation and RB tumor suppressor functional inactivation: Implications for the response to hormonal therapy. Cancer Biol Ther 7:1212–1225. Moinfar F, Man YG, Arnould L et al (2000) Concurrent and independent genetic alterations in the stromal and epithelial cells of mammary carcinoma: implications for tumorigenesis. Cancer Res 60:2562–2566. Mueller MM, Fusenig NE (2004) Friends or foes —bipolar effects of the tumour stroma in cancer. Nat Rev Cancer 4:839–849. Nagasawa T (2006) Microenvironmental niches in the bone marrow required for B-cell development. Nat Rev Immunol 6:107–116. Nakayama H, Enzan H, Miyazaki E et al (1998) The role of myofibroblasts at the tumor border of invasive colorectal adenocarcinomas. Jpn J Clin Oncol 28:615–620. Nguyen DX, Bos PD, Massague J (2009) Metastasis: from dissemination to organ-specific colonization. Nat Rev Cancer 9:274–284. Noma K, Smalley KS, Lioni M et al (2008) The essential role of fibroblasts in esophageal squamous cell carcinoma-induced angiogenesis. Gastroenterology 134:1981–1993. Olive KP, Jacobetz MA, Davidson CJ et al (2009) Inhibition of hedgehog signaling enhances delivery of chemotherapy in a mouse model of pancreatic cancer. Science 324: 1457–1461. Olumi AF, Grossfeld GD, Hayward SW et al (1999) Carcinoma-associated fibroblasts direct tumor progression of initiated human prostatic epithelium. Cancer Res 59:5002–5011. Orimo A, Weinberg RA (2006) Stromal fibroblasts in cancer: a novel tumor-promoting cell type. Cell Cycle 5:1597–1601. Orimo A, Gupta PB, Sgroi DC et al (2005) Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell 121:335–348. Ostman A, Augsten M (2009) Cancer-associated fibroblasts and tumor growth—bystanders turning into key players. Curr Opin Genet Dev 19:67–73. Padua D, Zhang XH, Wang Q et al (2008) TGFbeta primes breast tumors for lung metastasis seeding through angiopoietin-like 4. Cell 133:66–77. Paterson RF, Ulbright TM, MacLennan GT et al (2003) Molecular genetic alterations in the lasercapture-microdissected stroma adjacent to bladder carcinoma. Cancer 98:1830–1836. Patocs A, Zhang L, Xu Y et al (2007) Breast-cancer stromal cells with TP53 mutations and nodal metastases. N Engl J Med 357:2543–2551. Petit I, Jin D, Rafii S (2007) The SDF-1-CXCR4 signaling pathway: a molecular hub modulating neo-angiogenesis. Trends Immunol 28:299–307. Pietras K, Sjöblom T, Rubin K et al (2003) PDGF receptors as cancer drug targets. Cancer Cell 3:439–443. Pietras K, Pahler J, Bergers G et al (2008) Functions of paracrine PDGF signaling in the proangiogenic tumor stroma revealed by pharmacological targeting. PLoS Med 5:e19. Polyak K, Haviv I, Campbell IG (2009) Co-evolution of tumor cells and their microenvironment. Trends Genet 25:30–38. Powell DW, Mifflin RC, Valentich JD et al (1999) Myofibroblasts. I. Paracrine cells important in health and disease. Am J Physiol 277:C1–C9. Psaila B, Lyden D (2009) The metastatic niche: adapting the foreign soil. Nat Rev Cancer 9:285–293. Qiu W, Hu M, Sridhar A et al (2008) No evidence of clonal somatic genetic alterations in cancerassociated fibroblasts from human breast and ovarian carcinomas. Nat Genet 40:650–655.
348
U.M. Polanska et al.
Ratajczak MZ, Zuba-Surma E, Kucia M et al (2006) The pleiotropic effects of the SDF-1-CXCR4 axis in organogenesis, regeneration and tumorigenesis. Leukemia 20:1915–1924. Ronnov-Jessen L, Van Deurs B, Nielsen M et al (1992) Identification, paracrine generation, and possible function of human breast carcinoma myofibroblasts in culture. In Vitro Cell Dev Biol 28A:273–283. Ronnov-Jessen L, Petersen OW, Bissell MJ (1996) Cellular changes involved in conversion of normal to malignant breast: importance of the stromal reaction. Physiol Rev 76:69–125. Rosenthal E, McCrory A, Talbert M et al (2004) Elevated expression of TGF-beta1 in head and neck cancer-associated fibroblasts. Mol Carcinog 40:116–121. Rowley DR (1998) What might a stromal response mean to prostate cancer progression? Cancer Metastasis Rev 17:411–419. San Francisco IF, DeWolf WC, Peehl DM et al (2004) Expression of transforming growth factorbeta 1 and growth in soft agar differentiate prostate carcinoma-associated fibroblasts from normal prostate fibroblasts. Int J Cancer 112:213–218. Schafer M, Werner S (2008) Cancer as an overhealing wound: an old hypothesis revisited. Nat Rev Mol Cell Biol 9:628–638. Serini G, Gabbiani G (1999) Mechanisms of myofibroblast activity and phenotypic modulation. Exp Cell Res 250:273–283. Shimoda M, Mellody KT, Orimo A (2010) Carcinoma-associated fibroblasts are a rate-limiting determinant for tumour progression. Semin Cell Dev Biol 21:19–25. Shi-Wen X, Rodriguez-Pascual F, Lamas S et al (2006) Constitutive ALK5-independent c-Jun N-terminal kinase activation contributes to endothelin-1 overexpression in pulmonary fibrosis: evidence of an autocrine endothelin loop operating through the endothelin A and B receptors. Mol Cell Biol 26:5518–5527. Sica A, Larghi P, Mancino A et al (2008) Macrophage polarization in tumour progression. Semin Cancer Biol 18:349–355. Siegel PM, Massague J (2003) Cytostatic and apoptotic actions of TGF-beta in homeostasis and cancer. Nat Rev Cancer 3:807–821. Sivridis E, Giatromanolaki A, Koukourakis MI (2005) Proliferating fibroblasts at the invading tumour edge of colorectal adenocarcinomas are associated with endogenous markers of hypoxia, acidity, and oxidative stress. J Clin Pathol 58:1033–1038. Sneddon JB, Zhen HH, Montgomery K et al (2006) Bone morphogenetic protein antagonist gremlin 1 is widely expressed by cancer-associated stromal cells and can promote tumor cell proliferation. Proc Natl Acad Sci USA 103:14842–14847. Sugimoto H, Mundel TM, Kieran MW et al (2006) Identification of fibroblast heterogeneity in the tumor microenvironment. Cancer Biol Ther 5:1640–1646. Taipale J, Beachy PA (2001) The Hedgehog and Wnt signalling pathways in cancer. Nature 411:349–354. Tait LR, Pauley RJ, Santner SJ et al (2007) Dynamic stromal-epithelial interactions during progression of MCF10DCIS.com xenografts. Int J Cancer 120:2127–2134. Theunissen JW, de Sauvage FJ (2009) Paracrine Hedgehog signaling in cancer. Cancer Res 69:6007–6010. Thiery JP, Sleeman JP (2006) Complex networks orchestrate epithelial-mesenchymal transitions. Nat Rev Mol Cell Biol 7:131–142. Thiery JP, Acloque H, Huang RJ et al (2009) Epithelial-mesenchymal transitions in development and disease. Cell 139:871–890. Tian H, Callahan CA, DuPree KJ et al (2009) Hedgehog signaling is restricted to the stromal compartment during pancreatic carcinogenesis. Proc Natl Acad Sci USA 106:4254–4259. Tokudome N, Sakamoto G, Sakai T et al (2005) A case of carcinosarcoma of the breast. Breast Cancer 12:149–153. Tokunou M, Niki T, Eguchi K et al (2001) c-MET expression in myofibroblasts: role in autocrine activation and prognostic significance in lung adenocarcinoma. Am J Pathol 158:1451–1463. Trimboli AJ, Cantemir-Stone CZ, Li F et al (2009) Pten in stromal fibroblasts suppresses mammary epithelial tumours. Nature 461:1084–1091.
16 Tumour-Promoting Stromal Myofibroblasts in Human Carcinomas
349
Tsujino T, Seshimo I, Yamamoto H et al (2007) Stromal myofibroblasts predict disease recurrence for colorectal cancer. Clin Cancer Res 13:2082–2090. Tuhkanen H, Anttila M, Kosma VM et al (2004) Genetic alterations in the peritumoral stromal cells of malignant and borderline epithelial ovarian tumors as indicated by allelic imbalance on chromosome 3p. Int J Cancer 109:247–252. Varga J (2002) Scleroderma and Smads: dysfunctional Smad family dynamics culminating in fibrosis. Arthritis Rheum 46:1703–1713. Varga J, Abraham D (2007) Systemic sclerosis: a prototypic multisystem fibrotic disorder. J Clin Invest 117:557–567. Varga J, Pasche B (2009) Transforming growth factor beta as a therapeutic target in systemic sclerosis. Nat Rev Rheumatol 5:200–206. Vogelstein B, Kinzler KW (2004) Cancer genes and the pathways they control. Nat Med 10:789–799. Walter K, Omura N, Hong SM et al (2008) Pancreatic cancer associated fibroblasts display normal allelotypes. Cancer Biol Ther 7:882–888. Weber F, Xu Y, Zhang L et al (2007) Microenvironmental genomic alterations and clinicopathological behavior in head and neck squamous cell carcinoma. JAMA 297:187–195. Weinberg RA (2007) The biology of cancer. Garland Science, New York, pp 34, 584–585, 605–614, 652. Weinberg RA (2008a) Mechanisms of malignant progression. Carcinogenesis 29:1092–1095. Weinberg RA (2008b) Coevolution in the tumor microenvironment. Nat Genet 40:494–495. Wernert N, Locherbach C, Wellmann A et al (2001) Presence of genetic alterations in microdissected stroma of human colon and breast cancers. Anticancer Res 21:2259–2264. White ES, Atrasz RG, Hu B et al (2006) Negative regulation of myofibroblast differentiation by PTEN (phosphatase and tensin homolog deleted on chromosome 10). Am J Respir Crit Care Med 173:112–121. Yang G, Rosen DG, Zhang Z et al (2006) The chemokine growth-regulated oncogene 1 (Gro-1) links RAS signaling to the senescence of stromal fibroblasts and ovarian tumorigenesis. Proc Natl Acad Sci USA 103:16472–16477. Yauch RL, Gould SE, Scales SJ et al (2008) A paracrine requirement for hedgehog signalling in cancer. Nature 455:406–410. Zeisberg EM, Potenta S, Xie L et al (2007) Discovery of endothelial to mesenchymal transition as a source for carcinoma-associated fibroblasts. Cancer Res 67:10123–10128. Zhang C, Fu L, Fu J et al (2009) Fibroblast growth factor receptor 2-positive fibroblasts provide a suitable microenvironment for tumor development and progression in esophageal carcinoma. Clin Cancer Res 15:4017–4027. Zhu CQ, Popova SN, Brown ER et al (2007) Integrin alpha 11 regulates IGF2 expression in fibroblasts to enhance tumorigenicity of human non-small-cell lung cancer cells. Proc Natl Acad Sci USA 104:11754–11759.
Part IV
Immune-Mediated Cells
Chapter 17
Mast Cells and Tumor Microenvironment Theoharis C. Theoharides, Konstantinos-Dionysios Alysandratos, Asimenia Angelidou, and Bodi Zhang
Abstract Increasing evidence indicates that a unique immune cell, the mast cell, accumulates in the stroma surrounding certain tumors, especially mammary and pancreatic adenocarcinoma, as well as melanoma. Many molecules secreted by mast cells could benefit the tumor in at least four ways: (1) angiogenin, heparin and vascular endothelial growth factor (VEGF), which induce neovascularization; (2) proteases that disrupt the surrounding matrix and facilitate metastases; (3) growth factors such as, epidermal growth factor (EGF), nerve growth factor (NGF), platelet derived growth factor (PDGF) and stem cell factor (SCF); (4) histamine, IL-10 and transforming growth factor-b (TGF-b), which are immunosuppressant, along with activation of certain dendritic cells that induce immunologic anergy. These actions could only occur through the unique ability of mast cells to release certain mediators selectively without degranulation. Blocking such release of pro-tumor mediators may constitute a novel therapeutic approach. Abbreviations BBB CRH CT CTMC
blood-brain-barrier cortocotropin-releasing hormone tryptase and chymase mast cells connective tissue mast cells
T.C. Theoharides (*) Molecular Immunopharmacology and Drug Development Laboratory, Department of Pharmacology and Experimental Therapeutics, Tufts University School of Medicine and Tufts Medical Center, 136 Harrison Avenue, Boston, MA 02111, USA and Department of Biochemistry, Tufts University School of Medicine and Tufts Medical Center, Boston, MA, USA and Department of Internal Medicine, Tufts University School of Medicine and Tufts Medical Center, Boston, MA, USA e-mail:
[email protected]
R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_17, © Springer Science+Business Media, LLC 2010
353
354
DMBA EGF hCBMCs HDC HMC-1 IFN-a IFN-g MDSCs MMC MMP-9 NGF NMU NO NSCLC NT PAR-1 and -2 PDAC PDGF RBL-1 SCF SCLC SP T mast cells TAMs TGF-b TNF TRAIL TSLP VEGF VPF
T.C. Theoharides et al.
7, 12-dimethylbenz(a)anthracene epidermal growth factor human umbilical cord blood-derived cultured mast cells histidine decarboxylase human leukemic mast cells interferon-a interferon-g myeloid-derived stem cells mucosal mast cells metalloproteinase-9 nerve growth factor nitrosomethylurea nitric oxide non-small cell lung carcinomas neurotensin protease-activated receptors pancreatic ductal adenocarcinoma platelet-derived growth factor rat basophil leukemia cells stem cell factor small cell lung carcinomas substance P tryptase mast cells tumor-associated macrophages transforming growth factor-b tumor necrosis factor TNF-related apoptosis-related ligand thymic stromal lymphopoietin vascular endothelial growth factor vascular permeability factor
Introduction Despite substantial resources invested in basic cancer research, mortality rates for the most frequent forms of cancer have not decreased significantly. Metastasis facilitated by stromal proteolytic enzymes (Almholt and Johnsen 2003) and chemokines (Murphy 2001) remains the chief cause of morbidity and mortality. The stroma surrounding the tumor is increasingly acquiring importance for its growth and dissemination, with infiltrating inflammatory cells actually contributing to cancer proliferation (Mantovani et al. 2002). For instance, tumor-associated macrophages (TAMs) and tumor-associated fibroblasts can be beneficial to tumor angiogenesis and growth (Silzle et al. 2003; Yu and Rak 2003) through secretion of vascular endothelial growth factor (VEGF) (Barbera-Guillem et al. 2002) and platelet-derived growth factor (PDGF) (Kataki et al. 2002). In one model of
17 Mast Cells and Tumor Microenvironment
355
subcutaneous melanoma, both angiogenesis and growth rate correlated with tumor infiltration by macrophages that expressed angiotensin II type 1 receptor and VEGF (Egami et al. 2003). TAMs were also significantly correlated with squamous cell carcinoma invasion (Li et al. 2002). Mast cells are important in allergic and late phase reactions, but also in inflammation and T-cell mediated immune responses (Mekori and Metcalfe 2000; Redegeld and Nijkamp 2003; Pedotti et al. 2003). Yet, mast cells appear to be recruited by tumors and accumulate in the stroma (Theoharides 1988) (Fig. 17.1). Mast cells could be helpful to the tumor, but only if secretion of beneficial molecules could occur selectively without degranulation (Theoharides et al. 2007). In fact, the tumor stroma microenvironment could alter the phenotypic behavior of mast cells. For instance, acidity created by rapid cancer cell proliferation inhibits mast cell degranulation, but enhances IL-4 production (Frossi et al. 2003). Nitric oxide (NO) generated by new vessel growth inhibits mast cell degranulation (Coleman 2002), as do oxidized polyamines secreted by the tumor (Vliagoftis et al. 1992). Mast cells can promote tumor development by: (a) disturbing the normal stroma-epithelium
Fig. 17.1 Schematic representation of the possible role of increased number of mast cells in the stroma of certain tumors. Mast cells could be recruited by tumor-derived chemoattractants such as adrenomedullin, MCP-1, RANTES and SCF, to selectively secrete molecules beneficial to the tumor; these could include growth factors, histamine which is mitogenic (H1) and immunosuppressant (H2), neovascularization agents such as heparin, VEGF and IL-8, as well as proteases that could permit new blood vessel formation and metastases
356
T.C. Theoharides et al.
communication as was shown for matrix degradation at sites of tumor invasion in rat mammary adenocarcinoma, (b) facilitating tumor angiogenesis, (c) releasing growth factors (Conti et al. 2007), and (d) inducing a state of immunosuppression. The tumor enhancing effect of mast cells has been shown repeatedly with the use of W/Wv mast cell deficient mice, which developed fewer lung metastases to subcutaneous B16-BL6 tumors (Starkey et al. 1988), and in which mice premalignant angiogenesis of squamous epithelial carcinogenesis was blocked (Coussens et al. 1999). The development of 1, 2-dimethylhydrazine-induced intestinal tumors was slowed by 60% in W/Wv mice (Wedemeyer and Galli 2005). There was also reduced microvessel formation and tumor size in W/Wv mice injected with MB49 murine bladder carcinoma (Dethelfsen et al. 1994).
Mast Cell Biology Mast cells derive from a specific bone marrow progenitor cell, they migrate into tissues where they mature depending upon microenvironmental conditions, and they participate in allergic reactions, as well as innate and acquired immunity (Mekori and Metcalfe 2000; Galli et al. 2005a, b). Mast cells are located perivascularly close to neurons and could have a critical role in neuroinflammatory diseases (Theoharides and Cochrane 2004), as well as in stress-induced brain metastases (Theoharides et al. 2008). Mast cells vary considerably in their cytokine and proteolytic enzyme content: connective tissue mast cells (CTMC) contain tryptase and chymase (CT mast cells), while mucosal mast cells (MMC) contain only tryptase (T mast cells). However, the phenotypic expression of mast cells is not fixed, since MMC can develop into CTMC given the appropriate microenvironmental conditions (Galli et al. 2005a). Moreover, addition of IL-5 to human umbilical cord blood-derived cultured mast cells (hCBMCs) augmented IgE-induced production of distinct cytokines, such as tumor necrosis factor (TNF) and MIP-1a, but without histamine (Ochi et al. 2000). In addition to IgE and antigen, the main trigger in allergic reactions, anaphylatoxins (C3a, C5a), cytokines (IL-1, IL-33), hormones (CRH) and neuropeptides can stimulate mast cell activation; the latter include endorphins, substance P (SP), neurotensin (NT), and nerve growth factor (NGF) leading to secretion of numerous biologically active mediators, but through different pathways (Theoharides and Kalogeromitros 2006). In addition, thymic stromal lymphopoietin (TSLP) released from epithelial cells in response to infection, trauma, and inflammation activates mast cells in the absence of IgE, but in the presence of IL-1, to release IL-5 and IL-13 (Allakhverdi et al. 2007; Al-Shami et al. 2005). Mast cells can secrete either the content of individual granules (Theoharides and Douglas 1978) or distinct mediators selectively (Theoharides et al. 1982), possibly through regulation by specific phosphoproteins (Sieghart et al. 1978; Theoharides et al. 1980). Vascular permeability/vascular endothelial cell growth factor (VPF/VEGF) can be secreted from bone marrow-derived mouse mast cells
17 Mast Cells and Tumor Microenvironment
357
(Boesiger et al. 1998). In view of the fact that acute stress increased tumor size and decreased survival (Sklar and Anisman 1979; Antoni et al. 2006), we investigated if cortocotropin-releasing hormone (CRH), secreted under stress, could induce VEGF release from hCBMCs. We reported that CRH induced selective VEGF release without histamine (Cao et al. 2005). We further showed that IL-1 could induce selective secretion of IL-6 from hCBMCs without degranulation through a unique vesicular shuttle (Kandere-Grzybowska et al. 2003). IL-1 can further stimulate secretion of VEGF (Salven et al. 2002), thus promoting angiogenesis (Salven et al. 2002) and lung carcinoma growth (Saijo et al. 2002). Stem cell factor (SCF) can also induce selective release of IL-6 without histamine and without degranulation (Gagari et al. 1997). This process has been termed “differential release”, “intragranular activation” or “piecemeal degranulation” (Letourneau et al. 1996). Moreover, in certain diseases such as scleroderma and interstitial cystitis (Theoharides et al. 1995), mast cells could be almost totally depleted of their granule content, without classic degranulation, rendering them undetectable by light microscopy (“phantom mast cells”) (Claman et al. 1986).
Mast Cells Could Be Beneficial to the Tumor Mast cells could accumulate at sites of tumor growth in response to numerous chemoattractants (Table 17.1) such as RANTES or MCP-1 (Conti et al. 1997), and are associated with poor prognosis (Molin et al. 2002). In addition, SCF at low doses mediates chemotaxis of mast cells, while a higher dose is necessary for release of mediators (Huang et al. 2008), such as metalloproteinase-9 (MMP-9) (Huang et al. 2008). Adrenomedullin can stimulate histamine release from rat peritoneal mast cells (Yoshida et al. 2001), but can also be released from human cultured A549 lung carcinoma cells and stimulate human leukemic mast cells (HMC-1) (Zudaire et al. 2006). Moreover, adrenomedullin is also produced by HMC-1 cells, it augments growth of lung cancer cells (Zudaire et al. 2006), and adrenomedullin-producing mast cells were shown to infiltrate human lung cancers (Zudaire et al. 2006). Mast cells have been consistently implicated in tumor angiogenesis (Crivellato et al. 2008), along with other myeloid cells (Murdoch et al. 2008). Mast cell-deficient W/Wv mice exhibited decreased rate of tumor angiogenesis (Starkey et al. 1988). Mast cells could facilitate tumor angiogenesis through heparin-like molecules that would also permit metastases through their anti-clotting effects (Fig. 17.1). Mast cells
Table 17.1 Tumor-derived mast cell chemoattractants/ triggers
• Adrenomedullin • MCP-1 • RANTES • SCF
358
T.C. Theoharides et al.
also generate and secrete IL-8 which has been shown to be an angiogenesis factor, as well as a tumor cell chemotactic factor and tumor mitogen (Waugh and Wilson 2008). Mast cells secrete a number of growth factors, such as EGF, PDGF, NGF, and SCF (Galli et al. 2005b). Moreover, VPF/VEGF is secreted from mouse bone marrowderived and human cultured mast cells (Boesiger et al. 1998), as well as from HMC-1 cells (Grutzkau et al. 1998) (Table 17.2). Mast cells are rich in metalloproteinases, such as MMP-9, that can facilitate tumor invasiveness (Almholt and Johnsen 2003). Such enzymes can disturb the normal stroma-epithelium communication, as was shown for matrix degradation at sites of tumor invasion in rat mammary adenocarcinoma (Dabbous et al. 1986). Mast cells and stress could also disrupt the blood-brain-barrier (BBB) and promote brain metastases (Theoharides et al. 2008). Acute stress can activate mast cells and increase BBB permeability that is mast cell dependent (Esposito et al. 2002a). These findings are important in view of the fact that acute stress increases metastases in breast and other tumors (Sklar and Anisman 1979; Antoni et al. 2006), and over 30% of breast cancer patients develop brain metastases with poor associated prognosis (Schouten et al. 2002). In fact, a number of cancers express CRH receptors (Reubi et al. 2003), prompting the suggestion that CRH may affect tumor cell behavior. For instance, a human breast cancer line MCF7 expresses CRH mRNA and secrets immunoreactive CRH (Graziani et al. 2006b), prompting the possibility of autocrine or paracrine effects. Another aspect of tumor microenvironment is immunosuppression. Histamine induces tumor cell proliferation through H1 receptors, while suppressing the immune system through H2 and possibly H4 receptors (Tiligada et al. 2009; Gutzmer et al. 2005). It was also shown that the histamine content of human breast Table 17.2 Mast cell mediators relevant to tumor microenvironment Mediators Main pathophysiologic effects Histamine Vasodilation, mitogenesis, immunosuppression Enzymes Chymase Tissue damage Metalloproteinases Tissue disruption, metastases Tryptase Tissue damage, metastases Growth factors Tumor growth EGF, GM-CSF, NGF, PDGF, SCF Angiogenic factors Angiogenesis, neovascularization Angiogenin Heparin VEGF Cytokines IL-8 Chemoattractant, tumor mitogen EGF epidermal growth factor, GM-CSF granulocyte monocyte-colony stimulating factor, NGF nerve growth factor, PDGF platelet-derived growth factor, SCF stem cell factor, VEGF vascular endothelial cell growth factor
17 Mast Cells and Tumor Microenvironment
359
cancer tissue was much higher than adjacent normal tissue and sufficient to act as a local immunosuppressant (Reynolds et al. 1998). The H2 receptor antagonist, famotidine, given preoperatively enhanced tumor infiltrating lymphocytes and increased metastatic lymph node reactive changes in breast cancer in humans (Parshad et al. 2002). Mast cells can promote immunosuppression by also secreting the immunosuppressants transforming growth factor-b (TGF-b) and IL-10 (Conti et al. 2003). SCF-mediated mast cell infiltration of tumors further enhances immunosuppression (Huang et al. 2008). Mast cells may further contribute to tumor anergy by promoting the development/recruitment of “tolerogenic host antigenpresenting cells” (Wasiuk et al. 2009) and Treg cells (Wasiuk et al. 2009). Mast cells may also influence migration and function of dendritic cells through distinct prostaglandins (PGE2 and PGD2) (Wasiuk et al. 2009). In addition, mast cells could indirectly down-regulate anti-tumor immunity by influencing the functions of immune suppressive myeloid-derived stem cells (MDSCs). MDSCs also secrete VEGF (Marx 2008), but this action requires MMP-9 (Yang et al. 2004), which is produced by mast cells. Mast cell-derived IL-1b induces MDSCs in mice with transplanted mammary carcinoma or fibrosarcoma (Bunt et al. 2006).
Breast Cancer Disruption of the normal flow of information between stroma and parenchyma could permit neoplastic progression (Barcellos-Hoff 1998). Stromal matrix metalloproteinases, rather than the target cell, were shown to promote mammary tumorigenesis (Sternlicht et al. 1999), while irradiated mammary gland stroma promoted carcinogenesis of unirradiated epithelial cells (Barcellos-Hoff and Ravani 2000). Mammary carcinogenesis in Wistar/Furth rats occurs when only the stroma of the mammary gland (fat pad) is exposed to the carcinogen nitrosomethylurea (NMU) (Maffini et al. 2004). The earliest effects of carcinogen administration in mammary gland carcinogenesis are manifested in the stroma with infiltration of inflammatory cells including mast cells (Maffini et al. 2004). The number of mast cells was significantly increased in malignant as compared to benign, lesions in human breast biopsies (Kashiwase et al. 2004). Moreover, the number of mast cells was greater in scirrhous than papillotubular carcinoma (Kashiwase et al. 2004). The histamine content of human breast cancer tissue, an index of mast cell presence, was much higher than adjacent normal tissue (Reynolds et al. 1998). Recent papers also confirmed high number of mast cells in human mammary adenocarcinoma (Rajput et al. 2007; Ribatti et al. 2007). Some of these papers suggested that the presence of mast cells may indicate a favorable prognosis (Rajput et al. 2007; Dabiri et al. 2004); however, in these instances mast cells were “identified” by staining for c-kit (Rajput et al. 2007; Dabiri et al. 2004), which is not specific, as cancer cells also express c-kit (Charpin et al. 2009). Moreover, reduction of c-kit expression was associated with malignant transformation of breast epithelium in human breast cancer (Polat 2007), and in carcinogen-induced rat mammary carcinoma (Maffini
360
T.C. Theoharides et al.
et al. 2008). Consequently, tumor c-kit expression and not the presence of c-kit positive mast cells, appears to be associated with favorable outcomes. An increased number of mast cells was reported in cis-hydroxyproline-induced mammary tumors in Buffalo rats (Strum et al. 1981). Similar findings were obtained in 7, 12-dimethylbenz(a)anthracene (DMBA)-induced mammary adenocarcinoma in which mast cells accumulated, but appeared to be intact and resistant to the mast cell degranulator compound 48/80 (Andersson et al. 1976). The mast cell inhibitor disodium cromoglycate increased blood clotting and hypoxia in invasive murine breast cancer (Samoszuk and Corwin 2003). The location of mast cells in relation to tumor cells may also be important. Lymph nodes may behave differently (Munn and Mellor 2006) upon cancer cell entrapment. Tryptase-positive mast cells correlated with angiogenesis and presence of micrometastases in sentinel lymph nodes from 80 patients with breast cancer (Ribatti et al. 2007). In contrast, intermammary lymph node enlargement with mast cell infiltration was considered to be a positive prognostic sign (Quan et al. 2002).
Melanoma and Basal Cell Carcinoma Mast cells accumulate especially around invasive melanoma (Reed et al. 1996; Dvorak et al. 1980), and their numbers correlate with increased neovascularization, mast cell overexpression of VEGF, tumor aggressiveness, and poor prognosis (Ch’ng et al. 2006). Tumor vascularity and tryptase-positive mast cells correlated with poor melanoma prognosis (Ribatti et al. 2003). Moreover, SCF splice variants were detected in melanoma (Welker et al. 2000), and could present new forms of mast cell growth factors related to melanoma growth. Mast cells have also been repeatedly noted to accumulate around basal cell carcinoma lesions and are thought to contribute to cancer growth by inducing immunosuppression (Grimbaldeston et al. 2000). Increased dermal mast cell numbers are associated with higher risk of developing basal cell carcinoma in humans possibly through UVB-induced immunosuppression (Grimbaldeston et al. 2002). Mast cells can also mediate TNF-a dependent dendritic cell migration and consequently increase skin tumor antigen presentation, but in a manner that does not elicit an immune response (Munn and Mellor 2006; Hart et al. 2002).
Pancreatic Cancer Pancreatic ductal adenocarcinoma (PDAC) is the 4th cause of cancer-related deaths in the USA, with a prognosis of less than 6.0 months and 5-year survival of less than 5% (Welsch et al. 2007; Hezel et al. 2006). PDAC escapes early detection and resists treatment (Tuveson and Hingorani 2005). Even though PDAC had been
17 Mast Cells and Tumor Microenvironment
361
Fig. 17.2 Light photomicrograph of ductal pancreatic adenocarcinoma with a number of infiltrating intact mast cells (solid arrows) stained with acidified toluidine blue, magnification = 20×
thought not to be particularly vascular, it was shown to secrete VEGF (Luo et al. 2001). Adrenomedullin, discussed earlier as a mast cell trigger, is expressed by pancreatic cancer cells and increases their growth and survival (Nakamura et al. 2006; Keleg et al. 2007). Increased numbers of mast cells have also been noted in pancreatic cancer (Esposito et al. 2002b, 2004). Moreover, a recent paper reported that mast cells were an absolute requirement for angiogenesis and cancer development in Myc-activated mouse pancreatic tumors (Soucek et al. 2007). Mast cells are actually localized among ductal adenocarcinoma cells, and they appear morphologically intact (Fig. 17.2), suggesting that pro-tumor molecules may be released without degranulation.
Lung Cancer The relevance of lung mast cells to lung cancer is intriguing, given that mast cells are well known to be increased in asthma and other inflammatory lung diseases (Krishnaswamy et al. 2007). In fact, lung mast cells are also increased in the epithelium of smokers (Lamb and Lumsden 1982), especially in the small airways (Battaglia et al. 2007). Gene profiling of disaggregated human lung mast cells from patients with interstitial diseases showed increased expression of matrix metalloproteinases (Edwards et al. 2005). The number of mast cells was increased in bronchoalveolar lavage of patients with bronchial carcinoma (Walls et al. 2007). Also, increased mast cell density using tryptase and surface CD34 immunocytochemistry was significantly correlated with tumor progression, angiogenesis and poor prognosis in human adenocarcinomas (Takanami et al. 2000). High counts of
362
T.C. Theoharides et al.
chymase-positive mast cells also correlated with worse prognosis in bronchoalveolar carcinoma (Nagata et al. 2003) and in lung adenocarcinoma (Ibaraki et al. 2005). This direct correlation between increased numbers of mast cells and lung cancer was apparently independent of tumor angiogenesis, as measured by the presence of endothelial cells stained with anti-human factor VIII antibody (Tomita et al. 2000). Nevertheless, increased mast cell density correlated with increased VEGF expression and poor prognosis in 33/53 cases of non-small cell lung carcinomas (NSCLC) (Imada et al. 2000). Histidine decarboxylase (HDC) immunoreactivity, an index of mast cell presence/activation, could distinguish 18/23 cases of small cell lung carcinomas (SCLC), but only 6/12 cases of NSCLC (Matsuki et al. 2003), suggesting that higher number of mast cells were infiltrating the SCLC, known to be more aggressive.
Mast Cells Could be Detrimental to the Tumor Even though mast cells could be detrimental to tumor growth, they apparently cannot secrete such mediators as they may be inhibited from degranulation by tumorderived blockers, such as oxidized polyamines (Vliagoftis et al. 1992), or chondroitin sulfate, which inhibits mast cell activation (Theoharides et al. 2000). It would be fascinating if one could inhibit mast cells from secreting pro-tumor mediators, but promote secretion of anti-tumor molecules. For instance, tryptase stimulates protease-activated receptors (PAR-1 and -2), also activated by thrombin and trypsin respectively, and induces widespread inflammation (D’Andrea et al. 2001). IL-4 binds to IL-4 receptors expressed by human breast carcinoma cells, and leads to apoptosis (Gooch et al. 1998). TNF-a could also induce tumor cell death (Gordon and Galli 1990). Histamine inhibited human primary melanoma cell proliferation presumably by acting through H1 receptors, an action enhanced by IL-6 (LazarMolnar et al. 2002). Heparan sulfate proteoglycans could block binding of heparin to the cell surface and prevent neovascularization (Fannon et al. 2003). For instance protamine, which binds to heparin and neutralizes its anticoagulant properties, induced selective thrombosis of blood vessels within mammary adenocarcinoma (Su et al. 2001). On the other hand, cancer cell-associated chondroitin sulfate accumulates in mammary gland tumors and in metastatic lesions (Hinrichs et al. 1999); in fact, tumor cells metastasize through binding of their surface chondroitin sulfate to the interstitial matrix (Kokenyesi 2001). Interferon-a (IFN-a) or mast cell-derived interferon-g (IFN-g) may enhance TNF-related apoptosis-related ligand (TRAIL) gene expression and translation, leading to apoptosis of tumor cells in an autocrine and paracrine manner (Abadie et al. 2004; Wang et al. 2004). Moreover, treatment of melanoma patients with IFN-a increased TRAIL levels in serum (Tecchio et al. 2004). Receptor binding of TRAIL, in turn, activates an number of down steam events leading to regulation of the inflammatory response (Collison et al. 2009). In addition to its cytotoxic role,
17 Mast Cells and Tumor Microenvironment
363
TRAIL can inhibit angiogenesis indirectly by facilitating apoptosis of endothelial cells (Li et al. 2003). Certain studies indicate that CRH may inhibit growth of endometrial (Graziani et al. 2002) and breast cancer cells (Graziani et al. 2006b) in culture. Other studies suggest that endometrial cancer not expressing CRHR-1 may be associated with a more aggressive phenotype in humans (Graziani et al. 2006a). Mast cells (Kempuraj et al. 2004) and other inflammatory cells (Karalis et al. 1997) can synthesize and release CRH. It would, therefore, be interesting if mast cells could release CRH, but at the same time be prevented from CRH inducing VEGF release in an autocrine fashion.
Conclusion Mounting evidence indicates that mast cells accumulate in tumor stroma and can promote tumor growth and metastases. Mast cells may, therefore, serve as a new target for the adjuvant treatment of tumors (Groot et al. 2009), such as mammary adenocarcinoma or pancreatic cancer (Theoharides 2008), through the selective inhibition of tumor-promoting molecules. A possible therapeutic approach could involve the use of select flavonoids. Flavonoids are naturally occurring polyphenolic compounds present in green plants and seeds with anti-oxidant, anti-inflammatory and cancer-inhibiting properties (Middleton et al. 2000). A multiethnic epidemiological study identified an inverse relationship between flavonoid intake and pancreatic cancer (Nothlings et al. 2007). An inverse relationship has also been reported between intake of flavonoids, such as quercetin, and risk of lung cancer in general (Le Marchand et al. 2000), as well as in male smokers in particular (Hirvonen et al. 2001). The flavonoids quercetin, luteolin and epigallocatechin also decrease proliferation of pancreatic carcinoma cells in culture (Lee et al. 2002; Shankar et al. 2008; Takada et al. 2002), but the mechanism of this action is not known. We showed that these flavonoids inhibit proliferation and secretion from rat basophil leukemia (RBL-1) cells (Alexandrakis et al. 1999) and HMC-1 cells (Alexandrakis et al. 2003), as well as pro-inflammatory cytokine release from normal hCBMCs (Kempuraj et al. 2005). In fact, recent reviews have re-emphasized the potential use of select flavonoids, such as epigallocatechin, quercetin and curcumin in cancer treatment (Saif et al. 2009; Sogno et al. 2009; Jagtap et al. 2009). Future studies should investigate any unique ability of select flavonoids with anti-cancer properties to also inhibit secretion of pro-cancer molecules, while permitting secretion of anti-cancer mediators from mast cells. Acknowledgments Work discussed was supported in part by a “Concept Award” No. BC024430 from the United States Department of Defense to (TCT), and funds from Theta Biomedical Consulting and Development Co., Inc. (Brookline, MA). The possible therapeutic role of inhibiting mast cell-derived molecules beneficial to tumor growth is covered by patent application US 10/811,838 submitted by TCT and assigned to Theta, Inc.
364
T.C. Theoharides et al.
Many thanks are due to Dr. Eva Karamitropoulou (Attikon General Hospital, Athens Medical School, Athens, Greece) for the pancreatic cancer photo, and to Ms. Jessica Christian for her word processing skills.
References Abadie A, Besancon F, Wietzerbin J (2004) Type I interferon and TNFalpha cooperate with type II interferon for TRAIL induction and triggering of apoptosis in SK-N-MC EWING tumor cells. Oncogene 23:4911–4920 Alexandrakis M, Singh L, Boucher W, Letourneau R, Theofilopoulos P et al (1999) Differential effect of flavonoids on inhibition of secretion and accumulation of secretory granules in rat basophilic leukemia cells. Int J Immunopharmacol 21:379–390 Alexandrakis MG, Letourneau R, Kempuraj D, Kandere K, Huang M et al (2003) Flavones inhibit proliferation and increase mediator content in human leukemic mast cells (HMC-1). Eur J Haematol 71:448–454 Allakhverdi Z, Comeau MR, Jessup HK, yoon BP, Breuer A et al (2007) Thymic stromal lymphopoietin is released by human epithelial cell in response to microbes, trauma, or inflammation and potently activates mast cells. J Exp Med 19:253–258 Almholt K, Johnsen M (2003) Stromal cell involvement in cancer. Cancer Res 162:31–42 Al-Shami A, Spolski R, Kelly J, Keane-Myers A, Leonard WJ (2005) A role for TSLP in the development of inflammation in an asthma model. J Exp Med 202:829–839 Andersson A-C, Henningsson S, Lundell L, Rosengren E, Sundler F (1976) Diamines and polyamines in DMBA-induced breast carcinoma containing mast cells resistant to compound 48/80. Agents Actions 6/5:577–583 Antoni MH, Lutgendorf SK, Cole SW, Dhabhar FS, Sephton SE et al (2006) The influence of bio-behavioural factors on tumour biology: pathways and mechanisms. Nat Rev Cancer 6:240–248 Barbera-Guillem E, Nyhus JK, Wolford CC, Friece CR, Sampsel JW (2002) Vascular endothelial growth factor secretion by tumor-infiltrating macrophages essentially supports tumor angiogenesis, and IgG immune complexes potentiate the process. Cancer Res 62:7042–7049 Barcellos-Hoff MH (1998) The potential influence of radiation-induced microenvironments in neoplastic progression. J Mammary Gland Biol Neoplasia 3:165–175 Barcellos-Hoff MH, Ravani SA (2000) Irradiated mammary gland stroma promotes the expression of tumorigenic potential by unirradiated epithelial cells. Cancer Res 60:1254–1260 Battaglia S, Mauad T, van Schadewijik AM, Vignola AM, Rabe KF et al (2007) Differential distribution of inflammatory cells in large and small airways in smokers. J Clin Pathol 60: 907–911 Boesiger J, Tsai M, Maurer M, Yamaguchi M, Brown LF et al (1998) Mast cells can secrete vascular permeability factor/vascular endothelial cell growth factor and exhibit enhanced release after immunoglobulin E-dependent upregulation of Fce receptor I expression. J Exp Med 188:1135–1145 Bunt SK, Sinha P, Clements VK, Leips J, Ostrand-Rosenberg S (2006) Inflammation induces myeloid-derived suppressor cells that facilitate tumor progression. J Immunol 176:284–290 Cao J, Papadopoulou N, Kempuraj D, Boucher WS, Sugimoto K et al (2005) Human mast cells express corticotropin-releasing hormone (CRH) receptors and CRH leads to selective secretion of vascular endothelial growth factor. J Immunol 174:7665–7675 Charpin C, Giusiano S, Charfi S, Secq V, Carpentier S et al (2009) Quantitative immunohistochemical expression of c Kit in breast carcinomas is predictive of patients’ outcome. Br J Cancer 101:48–54
17 Mast Cells and Tumor Microenvironment
365
Ch’ng S, Waallis RA, Yuan L, Davis PF, Tan ST (2006) Mast cells and cutaneous malignancies. Mod Pathol 19:149–159 Claman HN, Choi KL, Sujansky W, Vatter AE (1986) Mast cell “disappearance” in chronic murine graft-vs-host disease (GVHD)-ultrastructural demonstration of “phantom mast cells”. J Immunol 137:2009–2013 Coleman JW (2002) Nitric oxide: a regulator of mast cell activation and mast cell-mediated inflammation. Clin Exp Immunol 129:4–10 Collison A, Foster PS, Mattes J (2009) The emerging role of TRAIL as key regulator of inflammatory responses. Clin Exp Pharmacol Physiol 36:1049–1053 Conti P, Pang X, Boucher W, Letourneau R, Reale M et al (1997) Impact of Rantes and MCP-1 chemokines on in vivo basophilic mast cell recruitment in rat skin injection model and their role in modifying the protein and mRNA levels for histidine decarboxylase. Blood 89: 4120–4127 Conti P, Kempuraj D, Kandere K, Gioacchino MD, Barbacane RC et al (2003) IL-10, an inflammatory/inhibitory cytokine, but not always. Immunol Lett 86:123–129 Conti P, Castellani ML, Kempuraj D, Salini V, Vecchiet J et al (2007) Role of mast cells in tumor growth. Ann Clin Lab Sci 37:315–322 Coussens LM, Raymond WW, Nergers G, Laig-Webster M, Behrendtsen O et al (1999) Inflammatory mast cells up-regulate angiogenesis during squamous epithelial carcinogenesis. Genes Dev 13:1382–1397 Crivellato E, Nico B, Ribatti D (2008) Mast cells and tumour angiogenesis: new insight from experimental carcinogenesis. Cancer Lett 269:1–6 D’Andrea MR, Derian CK, Santulli RJ, Andrade-Gordon P (2001) Differential expression of protease-activated receptors-1 and -2 in stromal fibroblasts of normal, benign, and malignant human tissues. Am J Pathol 158:2031–2041 Dabbous MK, Walker R, Haney L, Carter LM, Nicolson GL et al (1986) Mast cells and matrix degradation at sites of tumor invasion in rat mammary adenocarinoma. Br J Cancer 54:459–465 Dabiri S, Huntsman D, Makretsov N, Cheang M, Gilks B et al (2004) The presence of stromal mast cells identifies a subset of invasive breast cancers with a favorable prognosis. Mod Pathol 17:690–695 Dethelfsen SM, Matsumura N, Zetter BR (1994) Mast cell accumulation at sites of murine tumor implantation: implications for angiogenesis and tumor metastasis. Invasion Metastasis 14:395–408 Dvorak AM, Mihm MCJ, Osage JE, Dvorak HF (1980) Melanoma. An ultrastructural study of the host inflammatory and vascular responses. J Invest Dermatol 75:388–393 Edwards ST, Cruz AC, Donnelly S, Dazin PF, Schulman ES et al (2005) c-Kit immunophenotyping and metalloproteinase expression profiles of mast cells in intestinal lung diseases. J Pathol 206:279–290 Egami K, Murohara T, Shimada T, Sasaki K, Shintani S et al (2003) Role of host angiotensin II type 1 receptor in tumor angiogenesis and growth. J Clin Invest 112:67–75 Esposito P, Chandler N, Kandere-Grzybowska K, Basu S, Jacobson S et al (2002a) Corticotropinreleasing hormone (CRH) and brain mast cells regulate blood-brain-barrier permeability induced by acute stress. J Pharmacol Exp Ther 303:1061–1066 Esposito I, Kleeff J, Bischoff SC, Fischer L, Collecchi P et al (2002b) The stem cell factor-c-kit system and mast cells in human pancreatic cancer. Lab Invest 82:1481–1492 Esposito I, Menicagli M, Funel N, Bergmann F, Boggi U et al (2004) Inflammatory cells contribute to the generation of an angiogenic phenotype in pancreatic ductal adenocarcinoma. J Clin Pathol 57:630–636 Fannon M, Forsten-Williams K, Dowd CJ, Freedman DA, Folkman J et al (2003) Binding inhibition of angiogenic factors by heparan sulfate proteoglycans in aqueous humor: potential mechanism for maintenance of an avascular environment. FASEB J 17:902–904
366
T.C. Theoharides et al.
Frossi B, De Carli M, Daniel KC, Rivera J, Pucillo C (2003) Oxidative stress stimulates IL-4 and IL-6 production in mast cells by an APE/Ref-1-dependent pathway. Eur J Immunol 33:2168–2177 Gagari E, Tsai M, Lantz CS, Fox LG, Galli SJ (1997) Differential release of mast cell interleukin-6 via c-kit. Blood 89:2654–2663 Galli SJ, Kalesnikoff J, Grimbaldeston MA, Piliponsky AM, Williams CM et al (2005a) Mast cells as “tunable” effector and immunoregulatory cells: recent advances. Annu Rev Immunol 23:749–786 Galli SJ, Nakae S, Tsai M (2005b) Mast cells in the development of adaptive immune responses. Nat Immunol 6:135–142 Gooch JL, Lee AV, Yee D (1998) Interleukin 4 inhibits growth and induces apoptosis in human breast cancer cells. Cancer Res 58:4199–4205 Gordon JR, Galli SJ (1990) Mast cells as a source of both preformed and immunologically inducible TNF-a/cachectin. Nature 346:274–276 Graziani G, Tentori L, Portarena I, Barbarino M, Tringali G et al (2002) CRH inhibits cell growth of human endometrial adenocarcinoma cells via CRH-receptor 1-mediated activation of cAMP-PKA pathway. Endocrinology 143:807–813 Graziani G, Ferrandina G, Pozzoli G, Vergati M, Muzi A et al (2006a) Corticotropin-releasing hormone receptor-1 in human endometrial cancer. Oncol Rep 15:375–379 Graziani G, Tentori L, Muzi A, Vergati M, Tringali G et al (2006b) Evidence that corticotropinreleasing hormone inhibits cell growth of human breast cancer cells via the activation of CRHR1 receptor subtype. Mol Cell Endocrinol 264:44–49 Grimbaldeston MA, Skov L, Baadsgaard O, Skov BG, Marshman G et al (2000) Communications: high dermal mast cell prevalence is a predisposing factor for basal cell carcinoma in humans. J Invest Dermatol 115:317–320 Grimbaldeston MA, Skov L, Finlay-Jones JJ, Hart PH (2002) Increased dermal mast cell prevalence and susceptibility to development of basal cell carcinoma in humans. Methods 28:90–96 Groot KT, Abudukelimu A, Redegeld FA (2009) Mast cells as target in cancer therapy. Curr Pharm Des 15:1868–1878 Grutzkau A, Kruger-Krasagakes S, Baumeister H, Schwarz C, Kogel H et al (1998) Synthesis, storage and release of vascular endothelial growth factor/vascular permeability factor (VEGF/ VPF) by human mast cells: Implications for the biological significance of VEGF206. Mol Biol Cell 9:875–884 Gutzmer R, Diestel C, Mommert S, Kother B, Stark H et al (2005) Histamine H4 receptor stimulation suppresses IL-12p70 production and mediates chemotaxis in human monocyte-derived dendritic cells. J Immunol 174:5224–5232 Hart PH, Townley SL, Grimbaldeston MA, Khalil Z, Finlay-Jones JJ (2002) Mast cells, neuropeptides, histamine, and prostaglandins in UV-induced systemic immunosuppression. Methods 28:79–89 Hezel AF, Kimmelman AC, Stanger BZ, Bardeesy N, Depinho RA (2006) Genetics and biology of pancreatic ductal adenocarcinoma. Genes Dev 20:1218–1249 Hinrichs U, Rutteman GR, Nederbragt H (1999) Stromal accumulation of chondroitin sulphate in mammary tumours of dogs. Br J Cancer 80:1359–1365 Hirvonen T, Tornwali ME, Pietinen P, Albanes D, Vitamo J (2001) Flavonol and flavone intake and the risk of cancer in male smokers (Finland). Cancer Causes Control 12:789–796 Huang B, Lei Z, Zhang GM, Li D, Song C et al (2008) SCF-mediated mast cell infiltration and activation exacerbate the inflammation and immunosuppression in tumor microenvironment. Blood 112:1269–1279 Ibaraki T, Muramatsu M, Takai S, Jin D, Maruyama H et al (2005) The relationship of tryptase and chymase positive mast cells to angiogenesis in stage 1 non-small cell lung cancer. Eur J Cardiothorac Surg 28:621 Imada A, Shijubo N, Kojima H, Abe S (2000) Mast cells correlate with angiogenesis and poor outcome in stage I lung adenocarcinoma. Eur Respir J 15:1087–1093
17 Mast Cells and Tumor Microenvironment
367
Jagtap S, Meganathan K, Wagh V, Winkler J, Hescheler J et al (2009) Chemoprotective mechanism of the natural compounds, epigallocatechin-3-O-gallate, quercetin and curcumin against cancer and cardiovascular diseases. Curr Med Chem 16:1451–1462 Kandere-Grzybowska K, Letourneau R, Kempuraj D, Donelan J, Poplawski S et al (2003) IL-1 induces vesicular secretion of IL-6 without degranulation from human mast cells. J Immunol 171:4830–4836 Karalis K, Louis JM, Bae D, Hilderbrand H, Majzoub JA (1997) CRH and the immune system. J Neuroimmunol 72:131–136 Kashiwase Y, Morioka J, Inamura H, Yoshizawa Y, Usui R et al (2004) Quantitative analysis of mast cells in benign and malignant breast lesions. Immunohistochemical study on formalinfixed, paraffin-embedded tissues. Int Arch Allergy Immunol 134:199–205 Kataki A, Scheid P, Piet M, Marie B, Martinet N et al (2002) Tumor infiltrating lymphocytes and macrophages have a potential dual role in lung cancer by supporting both host-defense and tumor progression. J Lab Clin Med 140:320–328 Keleg S, Kayed H, Jiang X, Penzel R, Giese T et al (2007) Adrenomedullin is induced by hypoxia and enhances pancreatic cancer cell invasion. Int J Cancer 121:21–32 Kempuraj D, Papadopoulou NG, Lytinas M, Huang M, Kandere-Grzybowska K et al (2004) Corticotropin-releasing hormone and its structurally related urocortin are synthesized and secreted by human mast cells. Endocrinology 145:43–48 Kempuraj D, Madhappan B, Christodoulou S, Boucher W, Cao J et al (2005) Flavonols inhibit proinflammatory mediator release, intracellular calcium ion levels and protein kinase C theta phosphorylation in human mast cells. Br J Pharmacol 145:934–944 Kokenyesi R (2001) Ovarian carcinoma cells synthesize both chondroitin sulfate and heparan sulfate cell surface proteoglycans that mediate cell adhesion to interstitial matrix. J Cell Biochem 83:259–270 Krishnaswamy G, Ajitawi O, Chi DS (2007) The human mast cells. Methods Mol Biol 315:13–34 Lamb D, Lumsden A (1982) Intra-epithelial mast cells in human airway epithelium: evidence for smoking-induced changes in their frequency. Thorax 37:334–342 Lazar-Molnar E, Hegyesi H, Pallinger E, Kovacs P, Toth S et al (2002) Inhibition of human primary melanoma cell proliferation by histamine is enhanced by interleukin-6. Eur J Clin Invest 32:743–749 Le Marchand L, Murphy SP, Hankin JH, Wilkens LR, Kolonel LN (2000) Intake of flavonoids and lung cancer. J Natl Cancer Inst 92:154–160 Lee LT, Huang YT, Hwang JJ, Lee PP, Ke FC et al (2002) Blockade of the epidermal growth factor receptor tyrosine kinase activity by quercetin and luteolin leads to growth inhibition and apoptosis of pancreatic tumor cells. Anticancer Res 22:1615–1627 Letourneau R, Pang X, Sant GR, Theoharides TC (1996) Intragranular activation of bladder mast cells and their association with nerve processes in interstitial cystitis. Br J Urol 77:41–54 Li C, Shintani S, Terakado N, Nakashiro K, Hamakawa H (2002) Infiltration of tumor-associated macrophages in human oral squamous cell carcinoma. Oncol Rep 9:1219–1223 Li JH, Kirkiles-Smith NC, McNiff JM, Pober JS (2003) TRAIL induces apoptosis and inflammatory gene expression in human endothelial cells. J Immunol 171:1526–1533 Luo J, Guo P, Matsuda K, Truong N, Lee A et al (2001) Pancreatic cancer cell-derived vascular endothelial growth factor is biologically active in vitro and enhances tumorigenicity in vivo. Int J Cancer 92:361–369 Maffini MV, Soto AM, Calabro JM, Ucol AA, Sonnenschein C (2004) The stroma as a crucial target in rat mammary gland carcinogenesis. J Cell Sci 117:1495–1502 Maffini MV, Soto AM, Sonnenschein C, Papadopoulos N, Theoharides TC (2008) Lack of c-kit receptor promotes mammary tumors in N-nitrosomethylurea-treated Ws/Ws rats. Cancer Cell Int 8:5 Mantovani A, Sozzani S, Locati M, Allavena P, Sica A (2002) Macrophage polarization: tumorassociated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol 23:549–555
368
T.C. Theoharides et al.
Marx J (2008) Cancer immunology. Cancer’s bulwark against immune attack: MDS cells. Science 319:154–156 Matsuki Y, Tanimoto A, Hamada T, Sasaguri Y (2003) Histidine decarboxylase expression as a new sensitive and specific marker for small cell lung carcinoma. Mod Pathol 16:72–78 Mekori YA, Metcalfe DD (2000) Mast cells in innate immunity. Immunol Rev 173:131–140 Middleton E Jr, Kandaswami C, Theoharides TC (2000) The effects of plant flavonoids on mammalian cells: implications for inflammation, heart disease and cancer. Pharmacol Rev 52:673–751 Molin D, Edstrom A, Glimelius I, Glimelius B, Nilsson G et al (2002) Mast cell infiltration correlates with poor prognosis in Hodgkin’s lymphoma. Br J Haematol 119:122–124 Munn DH, Mellor AL (2006) The tumor-draining lymph node as an immune-privileged site. Immunol Rev 213:146–158 Murdoch C, Muthana M, Coffelt SB, Lewis CE (2008) The role of myeloid cells in the promotion of tumour angiogenesis. Nat Rev Cancer 8:618–631 Murphy PM (2001) Chemokines and the molecular basis of cancer metastasis. N Engl J Med 345:833–835 Nagata M, Shijubo N, Waalls AF, Ichimiya s, Abe S et al (2003) Chymase positive mast cells in small sized adenocarcinoma of the lung. Virchows Arch 443:565–573 Nakamura M, Han B, Nunobiki O, Kakudo K (2006) Adrenomedullin: a tumor progression factor via angiogenic control. Curr Cancer Drug Targets 6:635–643 Nothlings U, Murphy SP, Wilkens LR, Henderson BE, Kolonel LN (2007) Flavonols and pancreatic cancer risk: the multiethnic cohort study. Am J Epidemiol 166:924–931 Ochi H, DeJesus NH, Hsieh FH, Austen KF, Boyce JA (2000) IL-4 and -5 prime human mast cells for different profiles of IgE-dependent cytokine production. Proc Natl Acad Sci U S A 97:10509–10513 Parshad R, Kapoor S, Gupta SD, Kumar A, Chattopadhyaya TK (2002) Does famotidine enhance tumor infiltrating lymphocytes in breast cancer? Results of a randomized prospective pilot study. Acta Oncol 41:362–365 Pedotti R, De Voss JJ, Steinman L, Galli SJ (2003) Involvement of both ‘allergic’ and ‘autoimmune’ mechanisms in EAE, MS and other autoimmune diseases. Trends Immunol 24:479–484 Polat A (2007) c-KIT expression in columnar cell lesions of the breast accompanied by benign and malignant breast diseases. Pathol Res Pract 203:765–769 Quan C, Wang HL, Lu SJ (2002) Resistance to mammary carcinogenesis in Copenhagen rats: potential roles of vascular endothelial growth factor and mast cells. Cancer Lett 186:165–175 Rajput AB, Turbin DA, Cheang MCU, Voduc DK, Leung S et al (2007) Stromal mast cells in invasive breast cancer are a marker of favourable prognosis: a study of 4,444 cases. Breast Cancer Res 107:249–257 Redegeld FA, Nijkamp FP (2003) Immunoglobulin free light chains and mast cells: pivotal role in T-cell-mediated immune reactions? Trends Immunol 24:181–185 Reed JA, McNutt NS, Bogdany JK, Albino AP (1996) Expression of the mast cell growth factor interleukin-3 in melanocytic lesions correlates with an increased number of mast cells in the perilesional stroma: implications for melanoma progression. J Cutan Pathol 23:495–505 Reubi JC, Waser B, Vale W, Rivier J (2003) Expression of CRF1 and CRF2 receptors in human cancers. J Clin Endocrinol Metab 88:3312–3320 Reynolds JL, Akhter JA, Magarey CJ, Schwartz P, Adams WJ et al (1998) Histamine in human breast cancer. Br J Surg 85:538–541 Ribatti D, Ennas MG, Vacca A, Ferreli F, Nico B et al (2003) Tumor vascularity and tryptasepositive mast cells correlate with a poor prognosis in melanoma. Eur J Clin Invest 33: 420–425 Ribatti D, Finato N, Crivellato E, Guidolin D, Longo V et al (2007) Angiogenesis and mast cells in human breast cancer sentinel lymph nodes with and without micrometastases. Histopathology 51:837–842
17 Mast Cells and Tumor Microenvironment
369
Saif MW, Tytler E, Lansigan F, Brown DM, Husband AJ (2009) Flavonoids, phenoxodiol, and a novel agent, triphendiol, for the treatment of pancreaticobiliary cancers. Expert Opin Investig Drugs 18:469–479 Saijo Y, Tanaka M, Miki M, Usui K, Suzuki T et al (2002) Proinflammatory cytokine IL-1 beta promotes tumor growth of Lewis lung carcinoma by induction of angiogenic factors: in vivo analysis of tumor-stromal interaction. J Immunol 169:469–475 Salven P, Hattori K, Heissig B, Rafii S (2002) Interleukin-1alpha promotes angiogenesis in vivo via VEGFR-2 pathway by inducing inflammatory cell VEGF synthesis and secretion. FASEB J 16:1471–1473 Samoszuk M, Corwin MA (2003) Mast cell inhibitor cromolyn increases blood clotting and hypoxia in murine breast cancer. Int J Cancer 107:159–163 Schouten LJ, Rutten J, Huveneers HA, Twijnstra A (2002) Incidence of brain metastases in a cohort of patients with carcinoma of the breast, colon, kidney, and lung and melanoma. Cancer 94:2698–2705 Shankar S, Ganapathy S, Hingorani SR, Srivastava RK (2008) EGCG inhibits growth, invasion, angiogenesis and metastasis of pancreatic cancer. Front Biosci 13:440–452 Sieghart W, Theoharides TC, Alper SL, Douglas WW, Greengard P (1978) Calcium-dependent protein phosphorylation during secretion by exocytosis in the mast cell. Nature 275:329–331 Silzle T, Kreutz M, Dobler MA, Brockhoff G, Knuechel R et al (2003) Tumor-associated fibroblasts recruit blood monocytes into tumor tissue. Eur J Immunol 33:1311–1320 Sklar LS, Anisman H (1979) Stress and coping factors influence tumor growth. Science 205:513–515 Sogno I, Vannini N, Lorusso G, Cammarota R, Noonan DM et al (2009) Anti-angiogenic activity of a novel class of chemopreventive compounds: oleanic acid terpenoids. Recent Results Cancer Res 181:209–212 Soucek L, Lawlor ER, Soto D, Shchors K, Swigart LB et al (2007) Mast cells are required for angiogenesis and macroscopic expansion of Myc-induced pancreatic islet tumors. Nat Med 13:1211–1218 Starkey JR, Crowle PK, Taubenberger S (1988) Mast-cell-deficient W/Wv mice exhibit a decreased rate of tumor angiogenesis. Int J Cancer 42:48–52 Sternlicht MD, Lochter A, Sympson CJ, Huey B, Rougier JP et al (1999) The stromal proteinase MMP3/Stromelysin-1 promotes mammary carcinogenesis. Cell 98:134–146 Strum JM, Lewko WM, Kidwell WR (1981) Structural alterations within N-nitrosomethylureainduced mammary tumors after in vivo treatment with cis-hydroxyproline. Lab Invest 45:347–354 Su MY, Najafi A, Nalcioglu O (2001) Selective thrombosis of tumor blood vessels in mammary adenocarcinoma implants in rats. Am J Pathol 159:245–251 Takada M, Nakamura Y, Koizumi T, Toyama H, Kamigaki T et al (2002) Suppression of human pancreatic carcinoma cell growth and invasion by epigallocatechin-3-gallate. Pancreas 25:45–48 Takanami I, Takeuchi K, Naruke M (2000) Mast cell density is associated with angiogenesis and poor prognosis in pulmonary adenocarcinoma. Cancer 88:2686–2692 Tecchio C, Huber V, Scapini P, Calzetti F, Margotto D et al (2004) IFNalpha-stimulated neutrophils and monocytes release a soluble form of TNF-related apoptosis-inducing ligand (TRAIL/ Apo-2 ligand) displaying apoptotic activity on leukemic cells. Blood 103:3837–3844 Theoharides TC (1988) Neuroimmunology of tumor growth: the role of mast cells. Int J Immunopath Pharmacol 1:89–98 Theoharides TC (2008) Mast cells and pancreatic cancer. N Engl J Med 358:1860–1861 Theoharides TC, Cochrane DE (2004) Critical role of mast cells in inflammatory diseases and the effect of acute stress. J Neuroimmunol 146:1–12 Theoharides TC, Douglas WW (1978) Secretion in mast cells induced by calcium entrapped within phospholipid vesicles. Science 201:1143–1145 Theoharides TC, Kalogeromitros D (2006) The critical role of mast cell in allergy and inflammation. Ann N Y Acad Sci 1088:78–99
370
T.C. Theoharides et al.
Theoharides TC, Sieghart W, Greengard P, Douglas WW (1980) Antiallergic drug cromolyn may inhibit histamine secretion by regulating phosphorylation of a mast cell protein. Science 207:80–82 Theoharides TC, Bondy PK, Tsakalos ND, Askenase PW (1982) Differential release of serotonin and histamine from mast cells. Nature 297:229–231 Theoharides TC, Sant GR, El-Mansoury M, Letourneau RJ, Ucci AA Jr et al (1995) Activation of bladder mast cells in interstitial cystitis: a light and electron microscopic study. J Urol 153:629–636 Theoharides TC, Patra P, Boucher W, Letourneau R, Kempuraj D et al (2000) Chondroitin sulfate inhibits connective tissue mast cells. Br J Pharmacol 131:1039–1049 Theoharides TC, Kempuraj D, Tagen M, Conti P, Kalogeromitros D (2007) Differential release of mast cell mediators and the pathogenesis of inflammation. Immunol Rev 217:65–78 Theoharides TC, Rozniecki JJ, Sahagian G, Kempuraj D, Kalogeromitros D et al (2008) Impact of stress and mast cells on brain metastases. J Neuroimmunol 205:1–7 Tiligada E, Zampeli E, Sander K, Stark H (2009) Histamine H3 and H4 receptors as novel drug targets. Expert Opin Investig Drugs 18:1519–1531 Tomita M, Matsuzaki Y, Onitsuka T (2000) Effect of mast cells on tumor angiogenesis in lung cancer. Ann Thorac Surg 69:1686–1690 Tuveson DA, Hingorani SR (2005) Ductal pancreatic cancer in humans and mice. Cold Spring Harb Symp Quant Biol 70:65–72 Vliagoftis H, Boucher WS, Mak LL, Theoharides TC (1992) Inhibition of mast cell secretion by oxidation products of natural polyamines. Biochem Pharmacol 43:2237–2245 Walls AF, Roberts JA, Godfrey RC, Holgate ST (2007) Histochemical heterogeneity of human mast cells: disease-related differences in mast cell subsets recovered by bronchoalveolar lavaage. Int Arch Allergy Appl Immunol 92:233–241 Wang SH, Mezosi E, Wolf JM, Cao Z, Utsugi S et al (2004) IFNgamma sensitization to TRAIL-induced apoptosis in human thyroid carcinoma cells by upregulating Bak expression. Oncogene 23:928–935 Wasiuk A, de V V, Hartmann K, Roers A, Noelle RJ (2009) Mast cells as regulators of adaptive immunity to tumours. Clin Exp Immunol 155:140–146 Waugh DJ, Wilson C (2008) The interleukin-8 pathway in cancer. Clin Cancer Res 14:6735–6741 Wedemeyer J, Galli SJ (2005) Decreased susceptibility of mast cell-deficient kit(W)Kit(W-v) mice to the development of 1,2-dimethyldrazine-induced intestinal tumors. Lab Invest 85:388–396 Welker P, Schadendorf D, Artuc M, Grabbe J, Henz BM (2000) Expression of SCF splice variants in human melanocytes and melanoma cell lines: potential prognostic implications. Br J Cancer 82:1453–1458 Welsch T, Kleeff J, Friess H (2007) Molecular pathogenesis of pancreatic cancer: advances and challenges. Curr Mol Med 7:504–521 Yang L, DeBusk LM, Fukuda K, Fingleton B, Green-Jarvis B et al (2004) Expansion of myeloid immune suppressor Gr+CD11b+ cells in tumor-bearing host directly promotes tumor angiogenesis. Cancer Cell 6:409–421 Yoshida M, Yoshida H, Kitaichi K, Hiramatsu K, Kimura T et al (2001) Adrenomedullin and proadrenomedullin N-terminal 20 peptide induce histamine release from rat peritoneal mast cells. Regul Pept 101:163–168 Yu JL, Rak JW (2003) Host microenvironment in breast cancer development: inflammatory and immune cells in tumour angiogenesis and arteriogenesis. Breast Cancer Res 5:83–88 Zudaire E, Martinez A, Garayoa M, Pio R, Kaur G et al (2006) Adrenomedullin is a cross-talk molecule that regulates tumor and mast cell function during human carcinogenesis. Am J Pathol 168:280–291
Chapter 18
Macrophages in the Tumor Microenvironment Monica Escorcio-Correia and Thorsten Hagemann
Abstract Solid tumors consist of neoplastic cells, non-malignant stromal cells and migratory haematopoietic cells. Complex interactions between the cell types in this microenvironment regulate tumor growth, progression, metastasis and angiogenesis. There is also strong evidence that this microenvironment is inflammatory and that activation of the innate immune system plays a role in the progression of cancer. One such inflammatory cell that has the potential to promote cancer progression is the macrophage. There is a growing body of pre-clinical and clinical evidence associating abundance of tumor-associated macrophages (TAM) with poor prognosis. According to Condeelis and Pollard, TAM are obligate partners for malignant cell migration, invasion and metastases in many different cancers. These conclusions are based not only on association studies, but also on experiments that show ablation of macrophage function, or their infiltration into experimental tumors, inhibits growth and metastases. The monocyte/macrophage lineage is composed of innate immune cells that can have a variety of functions. Monocytes are produced in the bone marrow and circulate in the blood before being recruited into various tissues and differentiating into tissue-resident macrophages, prototypical phagocytes that play an important role in tissue homeostasis and pathogen clearance. Their ability to produce immunostimulatory cytokines such as interleukin-12 (IL-12) means that they can also play an important role in the activation of other immune cells, for example natural killer (NK) cells and T lymphocytes. In addition, they are also professional antigen presenting cells and can therefore participate in the mounting of the adaptive immune response (Gordon and Taylor 2005). Similarly to what is observed in the context of T cell responses, which can be polarized to a particular phenotype such as Th1 or Th2, macrophages can also be polarized depending on the context in which they are activated. In fact, an M1 and T. Hagemann (*) Barts and The London School of Medicine and Dentistry, Institute of Cancer, Charterhouse Square, London, EC1M 6BQ, UK e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_18, © Springer Science+Business Media, LLC 2010
371
372
M. Escorcio-Correia and T. Hagemann
M2 nomenclature that parallels, to a certain extent, with Th1 and Th2 has also been used for macrophages. The M1 phenotype is observed when macrophages are classically activated by stimuli such as lipolysaccharide (LPS) and interferon gamma (IFN-g). M1 cells are characterized by an enhanced capacity for antigen presentation and cell cytotoxicity, thanks to the increased expression of surface molecules such as MHCII, CD80, CD86, as well as reactive oxygen species and pro-inflammatory cytokines. On the other hand, M2 macrophages are typically activated in the presence of cytokines such as IL-4 and IL-13, usually produced during immune responses to parasites and allergens as part of a Th2 response. M2 macrophages can be characterized by their enhanced expression of mannose and scavenger receptors and arginase (Solinas et al. 2009; Gordon 2003). However, macrophages are plastic cells and similarly to what has been observed for T cells, it can be difficult to assign them to an absolute category as any cell population is likely to be heterogeneous and have characteristics of more than one subtype (Stout et al. 2009). For example, M1 macrophages are usually characterized by the expression inducible nitric oxide synthase (iNOS) which is required for the metabolism of arginine into nitric oxide (NO), an important mediator in macrophage cytotoxic function (Sica and Bronte 2007). On the other hand, M2 macrophages have higher expression of arginase, which allows them to metabolize arginine into urea and l-ornithine (Sica and Bronte 2007). However, some tumor-associated macrophages (TAMs) simultaneously express high levels of both iNOS and arginase (Tsai et al. 2007; Kusmartsev and Gabrilovich 2005). Tumors are composed not only of malignant cells and can include a variety of cell types, with macrophages sometimes representing a significant proportion of these cells, up to 40% of non-malignant cells in certain types of cancer (Pollard 2004). The macrophage population in tumors is composed not only of resident tissue macrophages but also includes differentiated monocytes that are recruited to the tumor from the blood, attracted by chemokines such as CCL2 (MCP-1), which can be highly expressed in tumors. Other chemoattractants involved in macrophage recruitment that are found to be expressed in the tumor microenvironment include CCL3 (MIP-1a), CCL4 (MIP-1b), CCL5 (RANTES), CXCL12 (SDF-1) and vascular endothelial growth factor (VEGF) (Pollard 2004; Bingle et al. 2002). In human patients, the abundance of TAMs has been correlated with poor prognosis in certain types of malignancy and several studies have shown that macrophages can contribute to different aspects of tumor growth and progression (Pollard 2004; Bingle et al. 2002). In accordance, the high expression of macrophage growth factors and chemokines such as macrophage-colony stimulating factor (M-CSF) and CCL2, have also been correlated with poor prognosis (Pollard 2004).
Pro-tumor Aspects The role of macrophages in tumor growth has been studied in both experimental and clinical studies. We will discuss evidence that supports the involvement of macrophages in tumor growth and progression, including immune suppression and promotion of angiogenesis and metastasis.
18 Macrophages in the Tumor Microenvironment
373
Immunosuppressive Phenotype Malignant cells have been shown to express unique antigens that are capable of triggering immune responses and potent cytotoxicity from T cells (Dunn et al. 2002). In addition, studies with transgenic animals that are deficient for specific immune cell subsets, including NK cells, NKT and both ab and gd T cells, have shown that the ablation of these cells leads to increased susceptibility to tumors (Dunn et al. 2002). However, despite this evidence supporting the potential of the immune system to participate in tumor clearance, anti-tumor immune responses often seem to be insufficient in preventing tumor growth. One of the reasons for this is the presence of several tumor-derived factors such as IL-10, VEGF and transforming growth factor beta (TGF-b) that render the tumor microenvironment immunosuppressive (Kim et al. 2006; Lewis and Pollard 2006). In fact, it has been shown that there are functional differences between tumor specific cytotoxic T cells depending on whether they are circulating through peripheral lymphoid organs or have infiltrated into tumor sites. Specifically, it has been shown that in melanoma CD8+ T cells collected from unaffected lymph nodes are capable of potent cytotoxicity in vitro towards their cognate antigen (Zippelius et al. 2004). In contrast, CD8+ T cells with the same antigen specificity and from the same individual, but directly associated with tumor sites are functionally tolerant in vitro (Zippelius et al. 2004). In many tumors, in addition to the malignant cells themselves, two other types of immunosuppressive cells are commonly found: Foxp3+ T regulatory cells (Tregs) and TAMs. The abundance of both these cells types in tumor lesions has been correlated with poor prognosis in human patients and their suppression of anti-tumor immunity has also been demonstrated (Pollard 2004; Khazaie and von Boehmer 2006; Zou 2005). One of the main roles of Tregs in the promotion of tumor growth seems to be the suppression of cytolytic activity by antigen-specific CD8+ T cells, with TGF-b having been identified as an important mediator of suppression (Khazaie and von Boehmer 2006). In addition, Tregs have also been shown to interact with innate immune cells and to play a role in the modulation of the tumor cytokine microenvironment. One study by Taams and colleagues showed that Tregs are capable of skewing macrophages into alternative activation (M2), which has been correlated with a tumor-promoting phenotype (Tiemessen et al. 2007). Following co-culture of human macrophages with Tregs (CD4+CD25+CD127low) or effector T cells (CD4+CD25CD127+), LPS was added to the cultures to induce macrophage activation. It was observed that in the presence of Tregs there was significant upregulation of alternative activation surface markers such as CD206 and CD163 on the macrophages when compared to cultures with effector T cells or macrophages alone (Tiemessen et al. 2007). Moreover, Tregs also prevented the production of pro-inflammatory cytokines, including tumor necrosis factor alpha (TNF-a), IL-6, IL-1, IL-8, macrophage inhibitory protein-1 alpha (MIP-1a) and macrophage chemoattractant protein-1 (MCP-1), likely to result from a defect in NF-kB activation, which the authors also observed (Tiemessen et al. 2007). In tumors, this interaction between TAMs and Tregs is perpetuated as the macrophages are able to recruit CD4+CD25+ T regulatory cells, for example via the production of IL-10 (Zhou et al. 2009).
374
M. Escorcio-Correia and T. Hagemann
In the tumor microenvironment, malignant cells are also capable of polarizing resident macrophages and recruited monocytes to a tumor-promoting M2-skewed phenotype. This polarization prevents the classical (M1) macrophage activation that would enable these cells to mediate tumor cytotoxicity and effectively present antigens to promote an anti-tumor adaptive immune response. This means that, in addition to other effector immune cells such as cytotoxic T cells, macrophages themselves are also suppressed by the tumor microenvironment. Moreover, their acquired TAM phenotype means that these macrophages are not only unable to participate in an anti-tumor immune response, but become suppressor cells themselves. TAMs have been shown to express a gene profile that shares characteristics with that of (alternatively activated) M2 macrophages (Biswas et al. 2006; Saccani et al. 2006; Lepique et al. 2009; Mantovani et al. 2002). Common patterns of gene expression between these two macrophage subsets include high levels of IL-10, scavenger receptors and mannose receptor, as well as low IL-12. One marked difference between TAMs and M2 macrophages is however, the high expression in TAMs of the interferon-inducible chemokines CCL5, CXCL9, CXCL10 and CXCL16 (Biswas et al. 2006). In any case, and true to their plastic nature, macrophages might have a different phenotype depending on the tumor region in which they are found (Stout et al. 2009; Biswas et al. 2008). One example is the modulation of TAM function by hypoxia in tumors, which can induce a particular macrophage phenotype that promotes angiogenesis (Murdoch et al. 2005). There are also studies that have correlated an increase in infiltrating macrophages in certain human tumors with improved prognosis, depending on their location within the tumor, which indicates that these macrophages might maintain an anti-tumor phenotype (Forssell et al. 2007; Kawai et al. 2008; Ohno et al. 2004). Interestingly, another study showed that following chemotherapy a high TAM infiltrate also correlated with increased patient survival (Taskinen et al. 2007). In support of their immunosuppressive phenotype, IL-10 and TGF-b are highly expressed by TAMs and these cytokines, usually produced by Th2 cells and Tregs, can act as potent inhibitors of Th1 cytokines such as IFN-g. The production of prostaglandins and arginase-1 by TAMs has also been implicated in the suppression of T cell anti-tumor immunity (Balkwill and Mantovani 2001). Several studies further indicate that TAMs can thwart immune responses to tumors by promoting the polarization of T cells towards a regulatory rather than a cytotoxic phenotype characterized by increased expression of IL-10 and defective induction of IFN-g (Lepique et al. 2009; Loercher et al. 1999). Other aspects of macrophage-mediated immune suppression in tumors include their poor antigen presenting capacity and their contribution to the differentiation of suppressive dendritic cells (DCs) (Mantovani et al. 2002). A study by Curiel et al. (2003) identified a population of dendritic cells that express B7-H1 on their surface and act to downregulate T cell anti-tumor responses. The authors showed that TAMs, but not normal macrophages, were able to induce B7-H1 expression on DCs via IL-10 and these cells could no longer effectively stimulate T cell proliferation (Curiel et al. 2003).
18 Macrophages in the Tumor Microenvironment
375
Macrophages and Inflammation Another important aspect of the pro-tumor effect of TAMs is their contribution to persistent inflammation in the tumor microenvironment. There are several lines of evidence that indicate a potential link between cancer and inflammation. Malignancies often arise at sites of chronic inflammation and epidemiological studies have shown that chronic infection and inflammation are important cancer risk factors (Karin et al. 2006; Mantovani et al. 2008). In addition, the long-term use of non-steroidal anti-inflammatory drugs (NSAIDs) can protect from certain cancers and in some animal models inhibition of inflammatory responses results in reduced tumor growth. The link between malignancy and inflammation can occur through either extrinsic or intrinsic pathways (Mantovani et al. 2008). The extrinsic pathway refers to cases in which an inflammatory pathology or infection can lead to neoplastic transformation (Mantovani et al. 2008). On the other hand, the intrinsic pathway is characterized by an oncogenic mutation that manifests itself in the tumor microenvironment with features of inflammation that contribute to tumor growth and progression (Mantovani et al. 2008). Macrophages play an important role in innate immune responses but the same molecules that can contribute to efficient pathogen clearance can also cause significant tissue damage during chronic infection/inflammation. These mediators include reactive oxygen and nitrogen intermediates (ROI and RNI) and macrophage migration inhibitory factor (MIF), which in the setting of chronic inflammation can not only cause tissue damage but can also be mutagenic and lead to defective p53 activity. These effects can promote the appearance of pre-malignant cells, further supported by classic pro-inflammatory cytokines that can act as tumor growth factors, for instance TNF-a, IL-6 and IL-1b, all of which are produced by activated macrophages (Pollard 2004; Biswas et al. 2008). In order to study the role of inflammatory cytokines in tumor initiation and progression, one approach has been to target components of signaling pathways required for the transcription of these soluble mediators. For example, the nuclear factor k-B (NF-kB) family of transcription factors not only plays an important role in the activation of macrophages in response to infection and inflammation, but has also been implicated in cancer (Hagemann et al. 2009). There are five different NF-kB proteins in mammals: RelA (p65), RelB, c-Rel, p50 (NF-kB1) and p52 (NF-kB2), and they can form different homo- or heterodimeric complexes (Perkins 2007). In the absence of the appropriate stimulus, NF-kB dimers are retained in the cytoplasm by IkB inhibitory proteins (IkBa, IkBb and IkBe) that mask the conserved nuclear localization sequence present in NF-kB proteins, thereby preventing their nuclear translocation (Perkins 2007). However, in response to inflammatory cytokines such as TNF-a and IL-1, or bacterial products like LPS that can bind to Toll-like receptors (TLR), IkB proteins are phosphorylated and targeted by ubiquitination for degradation in the proteasome. These phosphorylation events are mediated by IkB kinases, IKKa and IKKb, coupled to the regulatory subunit IKKg (NEMO). Once the IkB protein has been degraded, the NF-kB complex can then
376
M. Escorcio-Correia and T. Hagemann
translocate to the nucleus and bind its target genes to activate transcription. Importantly, positive feedback loops are often established as many NF-kB target genes, for example Tnf (TNF-a), encode proteins that will themselves lead to further NF-kB activation (Perkins 2007). The link between NF-kB and cancer was evident in studies of colitis-associated cancer in which this transcription factor was inhibited in myeloid cells, compromising their ability to induce, amongst other genes, TNF-a, IL-6 and IL-1b (Greten et al. 2004). In these experiments, ablation of IKKb resulted in a significant reduction in tumor incidence as well as tumor size, implicating myeloid cells in both tumor initiation and progression (Greten et al. 2004). In another model of chemically induced carcinogenesis, diethylnitrosamine (DEN)-induced liver cancer, deletion of IKKb and consequent ablation of NF-kB activation also inhibited tumor development (Maeda et al. 2005). Interestingly, this study showed that deletion of IKKb in hepatocytes actually resulted in increased tumor load, which was linked to increased hepatocyte death. DEN is a carcinogen that induces not only DNA damage but also cell death and the observed increase in apoptosis is explained by the lack of NF-kB signaling, which can protect cells against apoptotic cell death. In this case, instead of impeding the expansion of transformed hepatocytes, the increase in cell apoptosis actually induced compensatory cell proliferation, which in the presence of DEN resulted in increased malignancy (Maeda et al. 2005). However, ablation of NF-kB signaling in both hepatocytes and Kupffer cells (resident liver macrophages) had the opposite effect and a decrease in carcinogenesis was observed (Maeda et al. 2005). This decrease in tumor load correlated with decreased hepatocyte cell proliferation as these cells require cytokine growth factors such as TNF-a and IL-6, which are produced by Kupffer cells in an IKKb-dependent manner and are probably induced by factors released by dying hepatocytes (Maeda et al. 2005). Additional studies in a murine model of ovarian cancer also demonstrated that disruption of NF-kB signaling in TAMs resulted in several changes in their gene expression profile, including a reduction in arginase-1, IL-10 and TNF-a and an increase in IL12 and NO production (Hagemann et al. 2008). These changes in TAM phenotype correlated with overall reduced tumor growth, accompanied by an increase in the number of infiltrating NK cells which was dependent on the presence of IL-12 (Hagemann et al. 2008). Interestingly, the role of NF-kB activation in macrophages might be dependent on the stage of tumor development as there are also studies indicating that there is defective NF-kB activation in TAMs from established tumors due to the high expression of nuclear p50 homodimers (Biswas et al. 2006; Saccani et al. 2006).
Role in Angiogenesis and Metastasis An important aspect of tumor growth is the formation of new blood vessels, or neoangiogenesis, which is required to support the continuous expansion of the tumor mass beyond a certain size. TAMs can contribute to tumor angiogenesis
18 Macrophages in the Tumor Microenvironment
377
through the production of pro-angiogenic factors such as VEGF and platelet-derived endothelial cell growth factor. In addition, macrophage-derived cytokines can also promote angiogenesis, for example TNF-a induces matrix metalloproteinase-9 (MMP-9), a protease that cleaves latent VEGF from the extracellular matrix releasing it in its bioactive form (Pollard 2004). In fact, different studies have correlated high macrophage numbers with increased tumor vascularity in humans (Leek et al. 1996; Murdoch et al. 2008) and depletion of macrophages in some murine models of cancer has been shown to result in a significant decrease in the number of tumor blood vessels (Lin et al. 2006; Kimura et al. 2007). The role of macrophages in tumor angiogenesis provides yet another example of how the tumor microenvironment can influence the phenotype of infiltrating macrophages and polarize them to become tumor-promoting cells. Macrophages have been shown to preferentially accumulate in necrotic and hypoxic regions of solid tumors, in which the low oxygen microenvironment induces the upregulation of hypoxia-inducible transcription factors (HIF), HIF1 and HIF2 (Murdoch et al. 2008). In turn, the HIF transcription factors control the expression of several genes involved in the regulation of not only cell growth and metabolism but also angiogenesis, including Vegfa (VEGF) and Mmp7 (MMP-7), and can polarize macrophages to a pro-angiogenic phenotype (Murdoch et al. 2005; Burke et al. 2002; Talks et al. 2000). The cellular response to hypoxia and the expression of HIF transcription factors is also linked to the NF-kB signaling pathway. It has been shown that in the absence of IKKb, there is a defect in the upregulation of HIF-1a target genes, for example Vegfa, both in vitro and in vivo (Rius et al. 2008). The authors show that nuclear translocation of NF-kB complexes precedes the upregulation of HIF-1a, and the existence of a classical kB site on the Hif1a promoter further supports the hypothesis that Hif1a is an NF-kB target gene (Rius et al. 2008). A particular subpopulation of circulating monocytes has been implicated in the promotion of neoangiogenesis in the tumor microenvironment. These monocytes are identified by their expression of the angiopoietin receptor TIE2 and can be found in the peripheral blood of both humans and mice (Venneri et al. 2007; De Palma et al. 2005). One study demonstrated that the selective depletion of TIE2expressing monocytes (TEMs) in murine cancer models, despite not affecting recruitment of other monocyte subpopulations into the tumor, resulted in a significant decrease in tumor vascularization (De Palma et al. 2005). In addition, in the presence of TIE2+ human monocytes there was an increase in angiogenesis in a xenograft tumor model when compared to TIE2- cells (Venneri et al. 2007). TEMs do not express CCR2, the surface receptor for CCL2, and might be recruited to tumors by other factors. One such chemoattractant might be angiopoietin 2 itself, found to be highly expressed by tumor cells in hypoxic regions and endothelial cells in tumor blood vessels, both regions in which TEM accumulate in human and mouse tumors (Murdoch et al. 2008). Metastasis refers to the spreading of malignant cell colonies from the primary tumor where they originally developed to other tissues in the body. This process involves different steps in tumor progression including angiogenesis, breakdown of extracellular matrix proteins, and an increase in tumor cell motility, driven by the
378
M. Escorcio-Correia and T. Hagemann
expression of chemokine ligands and their receptors. Macrophages have also been implicated in tumor metastasis. Pollard and colleagues showed that when MMTV-PyMT transgenic mice, susceptible to mammary adenocarcinoma, were crossed to M-CSF deficient mice (Csf1op) there was a significant decrease in tumor metastasis, although there were no differences in the number and size of primary tumors (Lin et al. 2001). Conversely, the overexpression of M-CSF in the mammary epithelium of these mice led to an increase in lung metastasis, which correlated with an increase in the number of infiltrated macrophages in the primary tumor site (Lin et al. 2001). These observations support the role of TAMs in metastatic spread in this breast cancer model, as the expression of the M-CSF receptor in these tumors was restricted to the macrophage population (Lin et al. 2001). The same murine tumor model was used to demonstrate that in the absence of M-CSF there is a delay in the transition from premalignant lesions to malignant tumors, which was in turn accelerated when M-CSF overexpression was induced (Lin et al. 2006). It has also been demonstrated that both M-CSF and epidermal growth factor (EGF) can induce malignant cell and macrophage migration in vitro, despite the fact that the receptors for these molecules are not expressed on both cell types (Wyckoff et al. 2004). In fact, the EGF receptor was only found on tumor cells whereas the M-CSF receptor was present in macrophages but not tumor cells. However, the absence of either EGF or M-CSF resulted in the inhibition of migration from both cell types, revealing a collaborating interaction between them (Wyckoff et al. 2004). The communication between malignant cells and TAMs was also illustrated in a multiphoton microscopy study (Wyckoff et al. 2007). These experiments confirmed that, as previously observed in tissue sections, TAMs accumulate in the periphery of mammary tumors and the smaller numbers of TAMs present in the center of the tumor associate with blood vessels (perivascular). Time-lapse imaging showed that malignant cells have increased motility when they are in proximity to TAMs, towards which they were seen to migrate before binding to blood vessels and entering the circulation (Wyckoff et al. 2007). A recent study by DeNardo et al. (2009) further elucidates the role of TAMs in the promotion of tumor metastasis. The authors also used the MMTV-PyMT breast cancer transgenic mouse model, which they coupled to different transgenic mice deficient for both B and T cells (CD4+ and CD8+ subsets) and compared to mice that were deficient for each lymphocyte subset separately. It was observed that there were no differences in primary tumor incidence or phenotype, but there was a significant decrease in the lung tumor burden resulting from metastatic spread in the absence of CD4+ T cells (DeNardo et al. 2009). This effect was observed in MMTV-PyMT mice that had been crossed to RAG-/- mice (deficient for all B and T cells subsets), CD4-/-CD8-/- and CD4-/-, as well as MMTVPyMT mice treated with a depleting CD4 antibody, but not those selectively depleted of CD4+CD25+ T cells, a subpopulation enriched for Tregs. In the primary tumor, the numbers of infiltrating leukocytes remained the same but there were important differences in the TAM expression profile, which in the CD4-/- mice seemed to indicate a polarization towards an M1 phenotype. This included significantly increased mRNA expression of TNF-a, IL-6, IL-12p40, IL-1b and iNOS and
18 Macrophages in the Tumor Microenvironment
379
decreased expression of M2-associated mRNA products, for instance arginase-1 and TGF-b (DeNardo et al. 2009). The cytokine profile of CD4+ T cells in this tumor model was also indicative of Th2 polarization (induction of IL-4, IL-13 and IL-10) and in the presence of these CD4+ T cells the induction of M1 macrophages by LPS and IFN-g was inhibited. In fact, the authors showed that the neutralization of IL-4 resulted in a similar phenotype to that of CD4+ T cell deficient mice, including reduced metastatic tumor spread and increased expression of M1 markers in TAMs (DeNardo et al. 2009). Further studies have implicated macrophages in the promotion of tumor metastasis, for example in models of lung cancer that use the Lewis Lung Carcinoma (LLC) cell line. Culture of macrophages with conditioned media from this cell line was shown to induce high levels of IL-6 and TNF-a production by macrophages and versican, an extracellular matrix proteoglycan, was identified as being an important tumor-secreted factor for this response (Kim et al. 2009). Macrophages were shown to respond to versican through TLR2:TLR6 complexes by producing high amounts of TNF-a, a cytokine that promoted LLC tumor growth and metastatic spread. Silencing of versican in the LLC cell line resulted in reduced tumor multiplicity and reduced metastasis to lung, liver and adrenal glands when the tumor cells were implanted subcutaneously (Kim et al. 2009). Another study demonstrated that alveolar macrophages expressing high levels of VEGFR-1 responded to VEGF produced by tumor cells in the primary tumor site, which promoted the upregulation of MMP-9 by both macrophages and lung endothelial cells (Hiratsuka et al. 2002). This induction of MMP-9 only occurred in the presence of a primary tumor and correlated with a significant increase in the metastatic potential of systemically injected tumor cells.
Anti-tumor Potential/ Therapeutic Implications There is ample evidence to support that the abundance of TAMs in human tumors can correlate with poor prognosis, particularly when these macrophages are located in hypoxic/necrotic areas of the tumor (Lewis and Pollard 2006). On the other hand, large numbers of TAMs in the proximity of large, well vascularized areas of tumor has been found to correlate with a good prognosis (Forssell et al. 2007; Kawai et al. 2008; Ohno et al. 2004). In addition, despite the fact that the majority of experimental and clinical data points to macrophages as promoters of tumor growth, these cells also have the potential to lyse tumor cells (Koestler et al. 1987; Chen et al. 2002; Buhtoiarov et al. 2006; Lum et al. 2006; Bracher et al. 2007; Wu et al. 2009). William Coley, a late nineteenth century surgeon from New York, was probably one of the first to induce innate immune activation for the treatment of solid tumors, primarily sarcomas. He began by infecting cancer patients with isolates of bacteria (Coley 1893), and in later studies he used a preparation of “mixed toxins”, slightly less dangerous filtrates from cultures of Streptococcus pyogenes and gram-negative endotoxin-producing Serratia marcasens (Coley 1906). This mixture of bacterial
380
M. Escorcio-Correia and T. Hagemann
toxins probably resulted in the activation of the innate immune system and led to release of TNF, IL-1 and IL-12 by activated macrophages. Coley’s work was controversial and few were able to reproduce the beneficial effects that he had reported but, if the published case histories are to be believed, Coley was able to obtain rapid and sustained responses in patients. Nearly a century later, the use of autologous macrophages activated in vitro with either IFN-g alone or in combination with LPS was tested as a cancer treatment in clinical trials but this approach had only limited success (Hennemann et al. 1998; Andreesen et al. 1990). Much evidence for an anti-tumor potential of macrophages has come from the use of elegant transgenic murine tumor models in which modification of single activation components or inhibition of macrophage influx has been shown to alter the pro-tumor phenotype of TAMs (Hagemann et al. 2009; Hallam et al. 2009). Many cancers have a significant macrophage population, but despite their intrinsic anti-tumor potential, epidemiological and experimental data suggests that these cells are more likely to promote tumor growth rather than inhibit it. The concept of targeting TAMs and harnessing their anti-tumoral properties has clear potential for the treatment of cancer however, the experimental approaches used so far are not immediately applicable to the clinic and the systemic impact of a general inhibition of macrophage function is also not fully understood.
References Andreesen R, Scheibenbogen C, Brugger W, Krause S, Meerpohl HG, Leser HG, Engler H, Lohr GW (1990) Adoptive transfer of tumor cytotoxic macrophages generated in vitro from circulating blood monocytes: a new approach to cancer immunotherapy. Cancer Res 50:7450 Balkwill F, Mantovani A (2001) Inflammation and cancer: back to Virchow? Lancet 357:539 Bingle L, Brown NJ, Lewis CE (2002) The role of tumour-associated macrophages in tumour progression: implications for new anticancer therapies. J Pathol 196:254 Biswas SK, Gangi L, Paul S, Schioppa T, Saccani A, Sironi M, Bottazzi B, Doni A, Vincenzo B, Pasqualini F, Vago L, Nebuloni M, Mantovani A, Sica A (2006) A distinct and unique transcriptional program expressed by tumor-associated macrophages (defective NF-{kappa}B and enhanced IRF-3/STAT1 activation). Blood 107:2112 Biswas SK, Sica A, Lewis CE (2008) Plasticity of macrophage function during tumor progression: regulation by distinct molecular mechanisms. J Immunol 180:2011 Bracher M, Gould HJ, Sutton BJ, Dombrowicz D, Karagiannis SN (2007) Three-colour flow cytometric method to measure antibody-dependent tumour cell killing by cytotoxicity and phagocytosis. J Immunol Methods 323:160 Buhtoiarov IN, Lum HD, Berke G, Sondel PM, Rakhmilevich AL (2006) Synergistic activation of macrophages via CD40 and TLR9 results in T cell independent antitumor effects. J Immunol 176:309 Burke B, Tang N, Corke KP, Tazzyman D, Ameri K, Wells M, Lewis CE (2002) Expression of HIF-1alpha by human macrophages: implications for the use of macrophages in hypoxiaregulated cancer gene therapy. J Pathol 196:204 Chen GG, Chu YS, Chak EC, Leung BC, Poon WS (2002) Induction of apoptosis in glioma cells by molecules released from activated macrophages. J Neurooncol 57:179 Coley WB (1893) The treatment of malignant tumors by repeated inoculations of erysipelas: with a report of ten original cases. Am J Med Sci 105:487
18 Macrophages in the Tumor Microenvironment
381
Coley WB (1906) Late results of the treatment of inoperable sarcoma by the mixed toxins of erysipelas and bacillus prodigiosus. Am J Med Sci 131:375 Curiel TJ, Wei S, Dong H, Alvarez X, Cheng P, Mottram P, Krzysiek R, Knutson KL, Daniel B, Zimmermann MC, David O, Burow M, Gordon A, Dhurandhar N, Myers L, Berggren R, Hemminki A, Alvarez RD, Emilie D, Curiel DT, Chen L, Zou W (2003) Blockade of B7-H1 improves myeloid dendritic cell-mediated antitumor immunity. Nat Med 9:562 De Palma M, Venneri MA, Galli R, Sergi Sergi L, Politi LS, Sampaolesi M, Naldini L (2005) Tie2 identifies a hematopoietic lineage of proangiogenic monocytes required for tumor vessel formation and a mesenchymal population of pericyte progenitors. Cancer Cell 8:211 DeNardo DG, Barreto JB, Andreu P, Vasquez L, Tawfik D, Kolhatkar N, Coussens LM (2009) CD4+ T cells regulate pulmonary metastasis of mammary carcinomas by enhancing protumor properties of macrophages. Cancer Cell 16:91 Dunn GP, Bruce AT, Ikeda H, Old LJ, Schreiber RD (2002) Cancer immunoediting: from immunosurveillance to tumor escape. Nat Immunol 3:991 Forssell J, Oberg A, Henriksson ML, Stenling R, Jung A, Palmqvist R (2007) High macrophage infiltration along the tumor front correlates with improved survival in colon cancer. Clin Cancer Res 13:1472 Gordon S (2003) Alternative activation of macrophages. Nat Rev Immunol 3:23 Gordon S, Taylor PR (2005) Monocyte and macrophage heterogeneity. Nat Rev Immunol 5:953 Greten FR, Eckmann L, Greten TF, Park JM, Li ZW, Egan LJ, Kagnoff MF, Karin M (2004) IKKbeta links inflammation and tumorigenesis in a mouse model of colitis-associated cancer. Cell 118:285 Hallam S, Escorcio-Correia M, Soper R, Schultheiss A, Hagemann T (2009) Activated macrophages in the tumour microenvironment-dancing to the tune of TLR and NF-kappaB. J Pathol 219:143 Hagemann T, Lawrence T, McNeish I, Charles KA, Kulbe H, Thompson RG, Robinson SC, Balkwill FR (2008) “Re-educating” tumor-associated macrophages by targeting NF-{kappa} B. J Exp Med 205:1261 Hagemann T, Biswas SK, Lawrence T, Sica A, Lewis CE (2009) Regulation of macrophage function in tumors: the multifaceted role of NF-{kappa}B. Blood 113:3139 Hennemann B, Beckmann G, Eichelmann A, Rehm A, Andreesen R (1998) Phase I trial of adoptive immunotherapy of cancer patients using monocyte-derived macrophages activated with interferon gamma and lipopolysaccharide. Cancer Immunol Immunother 45:250 Hiratsuka S, Nakamura K, Iwai S, Murakami M, Itoh T, Kijima H, Shipley JM, Senior RM, Shibuya M (2002) MMP9 induction by vascular endothelial growth factor receptor-1 is involved in lung-specific metastasis. Cancer Cell 2:289 Karin M, Lawrence T, Nizet V (2006) Innate immunity gone awry: linking microbial infections to chronic inflammation and cancer. Cell 124:823 Kawai O, Ishii G, Kubota K, Murata Y, Naito Y, Mizuno T, Aokage K, Saijo N, Nishiwaki Y, Gemma A, Kudoh S, Ochiai A (2008) Predominant infiltration of macrophages and CD8(+) T cells in cancer nests is a significant predictor of survival in stage IV nonsmall cell lung cancer. Cancer 113:1387 Khazaie K, von Boehmer H (2006) The impact of CD4+CD25+ Treg on tumor specific CD8+ T cell cytotoxicity and cancer. Semin Cancer Biol 16:124 Kim R, Emi M, Tanabe K, Arihiro K (2006) Tumor-driven evolution of immunosuppressive networks during malignant progression. Cancer Res 66:5527 Kim S, Takahashi H, Lin W-W, Descargues P, Grivennikov S, Kim Y, Luo J-L, Karin M (2009) Carcinoma-produced factors activate myeloid cells through TLR2 to stimulate metastasis. Nature 457:102 Kimura YN, Watari K, Fotovati A, Hosoi F, Yasumoto K, Izumi H, Kohno K, Umezawa K, Iguchi H, Shirouzu K, Takamori S, Kuwano M, Ono M (2007) Inflammatory stimuli from macrophages and cancer cells synergistically promote tumor growth and angiogenesis. Cancer Sci 98:2009 Koestler TP, Johnson WJ, Rieman D, Dalton BJ, Greig RG, Poste G (1987) Differential expression of murine macrophage-mediated tumor cytotoxicity induced by interferons. Cancer Res 47:2804
382
M. Escorcio-Correia and T. Hagemann
Kusmartsev S, Gabrilovich DI (2005) STAT1 signaling regulates tumor-associated macrophage-mediated T cell deletion. J Immunol 174:4880 Leek RD, Lewis CE, Whitehouse R, Greenall M, Clarke J, Harris AL (1996) Association of macrophage infiltration with angiogenesis and prognosis in invasive breast carcinoma. Cancer Res 56:4625 Lepique AP, Daghastanli KRP, Cuccovia IM, Villa LL (2009) HPV16 tumor associated macrophages suppress antitumor T cell responses. Clin Cancer Res 15:4391 Lewis CE, Pollard JW (2006) Distinct role of macrophages in different tumor microenvironments. Cancer Res 66:605 Lin EY, Nguyen AV, Russell RG, Pollard JW (2001) Colony-stimulating factor 1 promotes progression of mammary tumors to malignancy. J Exp Med 193:727 Lin EY, Li J-F, Gnatovskiy L, Deng Y, Zhu L, Grzesik DA, Qian H, Xue X-N, Pollard JW (2006) Macrophages regulate the angiogenic switch in a mouse model of breast cancer. Cancer Res 66:11238 Loercher AE, Nash MA, Kavanagh JJ, Platsoucas CD, Freedman RS (1999) Identification of an IL-10-producing HLA-DR-negative monocyte subset in the malignant ascites of patients with ovarian carcinoma that inhibits cytokine protein expression and proliferation of autologous T cells. J Immunol 163:6251 Lum HD, Buhtoiarov IN, Schmidt BE, Berke G, Paulnock DM, Sondel PM, Rakhmilevich AL (2006) In vivo CD40 ligation can induce T cell-independent antitumor effects that involve macrophages. J Leukoc Biol 79:1181 Maeda S, Kamata H, Luo J-L, Leffert H, Karin M (2005) IKK[beta] couples hepatocyte death to cytokine-driven compensatory proliferation that promotes chemical hepatocarcinogenesis. Cell 121:977 Mantovani A, Sozzani S, Locati M, Allavena P, Sica A (2002) Macrophage polarization: tumorassociated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol 23:549 Mantovani A, Allavena P, Sica A, Balkwill F (2008) Cancer-related inflammation. Nature 454:436 Murdoch C, Muthana M, Lewis CE (2005) Hypoxia regulates macrophage functions in inflammation. J Immunol 175:6257 Murdoch C, Muthana M, Coffelt SB, Lewis CE (2008) The role of myeloid cells in the promotion of tumour angiogenesis. Nat Rev Cancer 8:618 Ohno S, Ohno Y, Suzuki N, Kamei T, Koike K, Inagawa H, Kohchi C, Soma G-I, Inoue M (2004) Correlation of histological localization of tumor-associated macrophages with clinicopathological features in endometrial cancer. Anticancer Res 24:3335 Perkins ND (2007) Integrating cell-signalling pathways with NF-[kappa]B and IKK function. Nat Rev Mol Cell Biol 8:49 Pollard JW (2004) Tumour-educated macrophages promote tumour progression and metastasis. Nat Rev Cancer 4:71 Rius J, Guma M, Schachtrup C, Akassoglou K, Zinkernagel AS, Nizet V, Johnson RS, Haddad GG, Karin M (2008) NF-[kgr]B links innate immunity to the hypoxic response through transcriptional regulation of HIF-1[agr]. Nature 453:807 Saccani A, Schioppa T, Porta C, Biswas SK, Nebuloni M, Vago L, Bottazzi B, Colombo MP, Mantovani A, Sica A (2006) p50 nuclear factor-{kappa}B overexpression in tumor-associated macrophages inhibits M1 inflammatory responses and antitumor resistance. Cancer Res 66:11432 Sica A, Bronte V (2007) Altered macrophage differentiation and immune dysfunction in tumor development. J Clin Invest 117:1155 Solinas G, Germano G, Mantovani A, Allavena P (2009) Tumor-associated macrophages (TAM) as major players of the cancer-related inflammation. J Leukoc Biol jlb.0609385 Stout RD, Watkins SK, Suttles J (2009) Functional plasticity of macrophages: in situ reprogramming of tumor-associated macrophages. J Leukoc Biol jlb.0209073
18 Macrophages in the Tumor Microenvironment
383
Talks KL, Turley H, Gatter KC, Maxwell PH, Pugh CW, Ratcliffe PJ, Harris AL (2000) The expression and distribution of the hypoxia-inducible factors HIF-1{alpha} and HIF-2{alpha} in normal human tissues, cancers, and tumor-associated macrophages. Am J Pathol 157:411 Taskinen M, Karjalainen-Lindsberg ML, Nyman H, Eerola LM, Leppa S (2007) A high tumorassociated macrophage content predicts favorable outcome in follicular lymphoma patients treated with rituximab and cyclophosphamide-doxorubicin-vincristine-prednisone. Clin Cancer Res 13:5784 Tiemessen MM, Jagger AL, Evans HG, van Herwijnen MJC, John S, Taams LS (2007) CD4+CD25+Foxp3+ regulatory T cells induce alternative activation of human monocytes/ macrophages. Proc Natl Acad Sci U S A 104:19446 Tsai C-S, Chen F-H, Wang C-C, Huang H-L, Jung S-M, Wu C-J, Lee C-C, McBride WH, Chiang C-S, Hong J-H (2007) Macrophages from irradiated tumors express higher levels of iNOS, arginase-I and COX-2, and promote tumor growth. Int J Radiat Oncol Biol Phy 68:499 Venneri MA, Palma MD, Ponzoni M, Pucci F, Scielzo C, Zonari E, Mazzieri R, Doglioni C, Naldini L (2007) Identification of proangiogenic TIE2-expressing monocytes (TEMs) in human peripheral blood and cancer. Blood 109:5276 Wu Q-L, Buhtoiarov IN, Sondel PM, Rakhmilevich AL, Ranheim EA (2009) Tumoricidal effects of activated macrophages in a mouse model of chronic lymphocytic leukemia. J Immunol 182:6771 Wyckoff J, Wang W, Lin EY, Wang Y, Pixley F, Stanley ER, Graf T, Pollard JW, Segall J, Condeelis J (2004) A paracrine loop between tumor cells and macrophages is required for tumor cell migration in mammary tumors. Cancer Res 64:7022 Wyckoff JB, Wang Y, Lin EY, Li J-f, Goswami S, Stanley ER, Segall JE, Pollard JW, Condeelis J (2007) Direct Visualization of macrophage-assisted tumor cell intravasation in mammary tumors. Cancer Res 67:2649 Zhou J, Ding T, Pan W, Zhu LY, Li L, Zheng L (2009) Increased intratumoral regulatory T cells are related to intratumoral macrophages and poor prognosis in hepatocellular carcinoma patients. Int J Cancer 125:1640 Zippelius A, Batard P, Rubio-Godoy V, Bioley G, Lienard D, Lejeune F, Rimoldi D, Guillaume P, Meidenbauer N, Mackensen A, Rufer N, Lubenow N, Speiser D, Cerottini J-C, Romero P, Pittet MJ (2004) Effector function of human tumor-specific CD8 T cells in melanoma lesions: a state of local functional tolerance. Cancer Res 64:2865 Zou W (2005) Immunosuppressive networks in the tumour environment and their therapeutic relevance. Nat Rev Cancer 5:263
Chapter 19
The Prognostic Significance of Tumor-Infiltrating Lymphocytes Ping Yu and Yang-Xin Fu
Abstract Tumors have been known to contain variable numbers of lymphocytes, referred to as tumor infiltrating lymphocytes (TILs). The degree of the lymphocytic infiltrate has been shown to correlate with positive outcome in some types of cancers. However, there are conflicting reports regarding the prognosis value of TIL even with the same types of cancers. Much research has gone into classifying TILs with respect to antigen receptor structure and the antigen to which the tumor-specific T cells react. However, these studies for the most part did not immunophenotype TILs, and recent data has revealed that the composition of TIL is not homogenous, but rather represents varying contributions from many lymphocytic subsets. Furthermore, the function of TILs is often compromised as a result of the accumulation of immunoregulatory cells and various tumor escape mechanisms. These recent findings stress the need to collect more data on the composition and function of TIL infiltrates before definitive conclusions about the prognostic significance of TILs can be drawn. This chapter summarizes the functional significance of different subsets of lymphocytes infiltrating the tumors.
Introduction The malignant tumors were noticed to contain variable numbers of lymphocytes since more than 100 years ago (Balkwill and Mantovani 2001). These lymphocytes have since been known as tumor-infiltrating lymphocytes (TIL). At the beginning, the TIL were thought to reflect the origin of cancer at the sites
P. Yu (*) The Committee on Immunology and Department of Pathology and Section of Dermatology/ Department of Medicine, University of Chicago, 5841 S. Maryland Ave, Chicago, IL 60637, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_19, © Springer Science+Business Media, LLC 2010
385
386
P. Yu and Y.-x. Fu
of chronic inflammation (Balkwill and Mantovani 2001), and later it was debated whether TIL provided a favorable environment for cancer growth or were evidence of the host’s attempt to eliminate cancer (Rollins 2006). This debate continues until today. The relationship between the extent of immune cell infiltration and prognosis might give us some hint what these cells are doing at the site of cancers. This relationship was first identified in 1949 in cases of breast cancer (Moore and Foote 1949). Since then, it is the significance of such inflammatory cells within or at the periphery of solid tumors that has long been the subject of conflicting reports (Clark et al. 1989; Clemente et al. 1996; Curiel et al. 2004; Zhang et al. 2003). Several studies have demonstrated that the increase of TIL is associated with better prognosis (Clark et al. 1989; Clemente et al. 1996; Zhang et al. 2003). However a sweeping assumption that the influx of lymphocytes to the tumor site is invariably beneficial to the patient may be a bit presumptuous. Recent studies suggest that the type, not the quantity, of tumor-infiltrating cells may be more of a critical determinant for the prognosis (Curiel et al. 2004). For example, in the case of infiltrating regulatory CD4+ cells to the tumor site, it has been shown to be more deleterious than favorable for the patient (Curiel et al. 2004). This subpopulation of TILs may behave in a more sinister manner to the detriment of the host’s defense against the malignant cells (Yu et al. 2005; Turk et al. 2004; Tawara et al. 2002; Li et al. 2003; Golgher et al. 2002; Casares et al. 2003; Machiels et al. 2001; Sutmuller et al. 2001; Steitz et al. 2001; Jones et al. 2002; Shimizu et al. 1999; Onizuka et al. 1999). Another example is the tumor-induced Th17 cells that are producing cytokine IL17. A recent report showed that IL-17 production by CD4+ cells led to myeloid cell recruitment into the tumor microenvironment and enhanced tumor growth in the ovarian cancers (Charles et al. 2009). Conversely, the antitumor lymphocytes migrating to the tumor site may become compromised once within the tumor milieu or become adversely adapted to the suppressive environment to promote growth instead of regression. In this chapter, we will strive to illustrate the type of T lymphocytes that behave either as friends or as foes in antitumor immunity. By understanding the functional peculiarities of the lymphocytic subtypes, it may serve to better explain the observable paradox that the evidence of TIL does not always correlate with better prognosis for the patient (Zhang et al. 2003; Marrogi et al. 1997; Naito et al. 1998; Schumacher et al. 2001; Saiki et al. 1996; Nakano et al. 2001; Banner et al. 1990; Kolbeck et al. 1992; Kowalczyk et al. 1997; Igarashi et al. 1992).
Antitumor Functions of T Lymphocytes The mature T cell population is composed of ab T cells expressing CD4 or CD8 and the CD4-/CD8- gd T cell receptor (TCR)-expressing cells. The unique function of CD4+ and CD8+ T cells is dictated by the expression of these coreceptors, CD4
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
387
or CD8, for which the ligand is b2 domain of the major histocompatibility complex (MHC) class II molecule (Cammarota et al. 1992) and the a3 domain of MHC class I molecules (Potter et al. 1989), respectively. Due to these specificities, the ab TCR of CD8+ T cells is restricted to the recognition of antigens presented by MHC class I molecules and the ab TCR of CD4+ T cells to antigens presented by MHC class II molecules. However, relatively little is known about the function of gdT in antitumor immunity so far, thus this review will mainly focus on the ab T cells. The important role of T cells as effectors in antitumor immunity was first shown in numerous murine models. For instance, UV light-induced regressor tumors are rejected regularly by normal mice but grow progressively in the absence of T cells (Ward et al. 1990; Kripke 1974; Spellman and Daynes 1981). It has also been demonstrated convincingly that T cell-mediated immunity is essential for the rejection of virally and chemically induced tumors (Leclerc et al. 1972; Tevethia et al. 1974; Cheever et al. 1986; Cheever et al. 1986; Klein et al. 1960; Boyse et al. 1962; Rouse et al. 1972). For example, in the model of murine chemical carcinogen methylcholanthrene-induced tumors, it was demonstrated that intravenous injection of immune cells, but not of immune serum, could transfer systemic tumor-specific immunity into sublethally irradiated mice (Boyse et al. 1962). These results are consistent with studies showing that the protective immunity against a plasma-cell tumor was abolished by prior depletion of T-lymphocytes via anti-T cell antibodies and complement (Rouse et al. 1972). The importance of T cells in tumor immunity has also been implicated in human studies, albeit with discernable limitations. While most murine tumor models have the advantage of utilizing antigen-specific T cells generated from tumor-free syngeneic mice, generating a human homologue is not feasible. Thus, in human studies, T-lymphocytes are isolated from peripheral blood (Mukherji and MacAlister 1983; Knuth et al. 1984) or from the tumor (Klein et al. 1980) of cancer patients. Such T cells can react in vitro with autologous cancer cells (Fossati et al. 1987). Utilizing adoptive transfer of in vitro expanded TIL in combination with chemotherapy, recent clinical trials have shown up to 50% positive response rate in selected patients with late stage aggressive cancers (Rosenberg 2001a, b; Rosenberg et al. 2004; Dudley et al. 2002; Dudley and Rosenberg 2003).
CD8+ T Cells While the significance of lymphocytes in tumor immunity is rarely disputed, the relative importance of various T cell subsets in tumor rejection is the subject of great controversies (Robins 1986). Broadly speaking, since cancer is a disease caused by an array of mutations in various types of cells, differences in the T cell subsets required for mediating its rejection is not altogether surprising. One such subset is the CD8+ cytotoxic T lymphocytes (CTLs). Most tumors are positive for MHC class I but negative for MHC class II, and CTLs are able to induce tumor killing upon direct recognition of peptide antigens, presented by the tumor’s MHC
388
P. Yu and Y.-x. Fu
class I molecules. Thus, the initial attention to antitumor immune responses was preferentially given to CD8+ T cells. That the CD8+ T cells are critical effectors against tumor cells is further supported by numerous studies in murine models. For instance, for certain tumors, such as UV light-induced tumors, the CD8+ cytolytic T cell subset appears to be regularly required for rejection (Ward et al. 1990). Murine CTLs that kill tumor targets in vitro can be freshly isolated from mice after repeated intraperitoneal injection of antigenic tumor cells or can be generated ex vivo in a 7-day mixed lymphocyte-tumor cell culture. Elimination of CD8+ T cells from mice either via depleting antibodies or using genetic knockout can, at least partially, abrogate the antitumor immunity induced by most cancer vaccines (Fearon et al. 1990; Golumbek et al. 1991; Dranoff et al. 1993; Lin et al. 1996). Consistent with the murine models, clinical data from the cancer patients uphold the importance of CD8+ T cells in bringing forth an antitumor response. Adoptive transfer studies of in vitro-stimulated CD8+ T cell lines and CD8+ clones specific for tumor antigens effectively mediate antitumor immunity when transferred back into tumor-bearing hosts (Rosenberg 1999, 2001a, b). Furthermore, recent reports suggest that immunization, using either adjuvant or dendritic cells (DCs) with pure tumor peptides, can result in productive antitumor immunity that is restricted by MHC class I (Noguchi et al. 1995; Feltkamp et al. 1993). Taken together, CD8+ T cells in tumor immunity can be unquestionably heralded as one of the principal subsets of T cells that constructively mediate an effective antitumor response.
CD4+ T Cells Undeniably, CD4+ T cells are an integral part of adaptive immunity, but the specific role they play in mounting an antitumor response remains dubious and somewhat conflicting. The critical function of CD4+ T cells in promoting immunity has been consistently demonstrated by vaccine and challenge experiments employing antibody-mediated depletion of CD4+ T cells or by using CD4-knockout mice (Fearon et al. 1990; Golumbek et al. 1991; Dranoff et al. 1993; Lin et al. 1996; Levitsky et al. 1994; Ostrand-Rosenberg 1994; Pulaski et al. 1993; Cavallo et al. 1992; Hock et al. 1991; Pan et al. 1995). More importantly, similar to CD8+ T cells, tumor-specific CD4+ T cell that can recognize tumor antigens do exist and evidence of migration to the tumor site has been shown in both murine models and human cancer tissues (Pardoll and Topalian 1998; Beck-Engeser et al. 2001; Monach et al. 1995). However, complications arise when the accumulation of CD4+ T cells inside the tumor microenvironment during tumor progression seemingly hinder the effector function of CD8+ T cells (Yu et al. 2005; Awwad and North 1988; Berendt and North 1980; Bursuker and North 1984; Wang et al. 2004). The hallmark phenotype of the CD4+ T cell that impedes the antitumor response seems to harbor characteristics akin to regulatory cells (Curiel et al. 2004; Dieckmann et al. 2001; Liyanage et al. 2002; Woo et al. 2001, 2002). This paradoxical dualism of CD4+ T cells obligates further differentiation of this subtype into helper and regulatory CD4+ T cells.
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
389
Although CD4+ T cells have been shown to be sufficient to eliminate tumor cells in the absence of CD8+ T cells in some tumor models (Monach et al. 1995; Beatty and Paterson 2001; Fujiwara et al. 1984; Greenberg et al. 1981; Mumberg et al. 1999), it is more often the case that both CD4+ and CD8+ are required for effective tumor rejection to occur (Beatty and Paterson 2000). In part, this is due to a substantial portion of tumor cells expressing only the class I MHC molecules, not the class II molecules, thereby limiting direct recognition by the CD4+ T cells. Moreover, the predominant effector mechanism in tumor immunity is by direct lysis of tumor cells by the MHC class I recognizing CD8+ CTL, the role of CD4+ T cells in antitumor responses is often to aid in the activation of CD8+ T cells leading to the destruction of the tumor by CD8+ CTL. The CD4+ T cell help of the CD8+ CTLs in tumor immunity can be divided into three phases: early activation, effector, and memory. CD4+ T Cells Help for Cytotoxic T Lymphocytes Induction It has been debated for decades if CD4+ T cells are required for the priming of CTL. Accumulating evidence has indicated that for the induction of tumor- specific CD8+ T cell responses, cross-presentation of antigens that have been captured by professional antigen presenting cells (APCs) such as DCs plays a dominant role (Huang et al. 1994, 1996; Toes et al. 1996; Yu et al. 2003; Spiotto et al. 2002). CD4+ T cell help has been hinted to be essential for such crosspriming in the induction of CTL immunity. This requirement for CD4 help is believed at least in part to activate APC (Ridge et al. 1998), which in turn expresses costimulatory molecules such as ICAM-1 (Shinde et al. 1996), CD80, and CD86 (Cella et al. 1996), or to secrete cytokines including IL-12 (Cella et al. 1996; Koch et al. 1996). These factors are essential for better CD8+ T cell activation. Most T cell help for CTL priming is dependent on the interaction between CD40L expressed by CD4+ T cells and CD40 on APC (Lu et al. 2000; Schoenberger et al. 1998; Bennett et al. 1997, 1998). The CD40–CD40L interaction has also been proved important in the generation of protective T cell-mediated tumor immunity (Mackey et al. 1997, 1998). The requirement for CD4+ T cells to “license” APC for the priming of CLT helps explaining some scenarios in which the induction of CTL can be achieved in the absence of CD4+ T cells. The typical CTL priming that is independent of CD4+ T cells is via the direct activation of DCs by virus that provides the optimal inflammatory signal (Stevenson et al. 1998) to activate DCs, which can subsequently prime antigen-specific CTL responses in the absence of CD4+ T cells (MacDonald and Johnston 2000). Evidence supporting CD4 independent CTL induction via activated DC is further demonstrated by the data showing that CD40-mediated activation with soluble ligand or activating antibodies licensed DC for cross-priming in the absence of CD4+ T cells (Mackey et al. 1997, 1998). However, these observations have not been confirmed in all experimental models. One study has shown that injection of MHC class I-deficient tumor cells
390
P. Yu and Y.-x. Fu
into MHC class II knockout mice resulted in the induction of tumor cell-specific CTL responses (Wolkers et al. 2001), demonstrating that cross-priming can occur even in the absence of CD4+ T cells, albeit with decreased efficiency. It cannot be excluded that the CD4 help required for the cross-priming of CLT was bypassed in this model because DCs were activated by cellular components released during tumor cell apoptosis, which can serve as adjuvant as shown in one study (Shi et al. 2000). Other evidence seems to suggest that high level of antigen can bypass CD4 help for CTL induction. For example, RMA-S cells loaded with a MHC class I-restricted peptide can induce CTL responses in vitro. The observed in vitro CTL priming was independent of CD4+ T cells or MHC class II-expressing cells, but was dependent upon the level of MHC class I expression on the RMA-S cells (De Bruijn et al. 1991). A high expression level of MHC class I peptide complexes on the peptide presenting cells was the decisive requirement for the induction of CD4+ T cellindependent CTL response in this model. This hypothesis would explain other observations in the vaccination experiment using MHC class I-restricted peptides emulsified in noninflammatory incomplete Freund’s adjuvant, robust CTL function was induced in mice depleted of CD4+ T cells (Fayolle et al. 1996). Even with high level of antigen provided for CLT priming, other conditions may be required for CTL induction in the absence of CD4 help. Our own study has showed that CD4 help is dispensable for cross-priming of CTL when intact draining lymph nodes are available (Yu et al. 2003). In the absence of draining lymphoid tissue, this additional “help” becomes essential for the proliferation of naïve CD8+ T cells (Yu et al. 2003). Other observation appropriately illustrates that the CD4+ help even in the presence of activated DCs cannot be altogether excluded or ignored, suggesting that CD4+ T cells may play other roles in CTL induction than “licensing” DCs. For example, in experiments utilizing vaccinations with ovalbumin (OVA)-transduced, CD40-activated DCs, only in the presence of CD4+ T cells was there protection against formation of OVA-expressing tumors (Schnell et al. 2000). Furthermore, in some studies, vaccination with transduced DCs or DCs pulsed with peptides, tumor lysates, or tumor cell-derived exosomes has been shown to be at least partially CD4+ T cell-dependent, although, it is possible that DCs were not properly activated prior to injection (Wolfers et al. 2001; Tuting et al. 1999; Fields et al. 1998; Porgador and Gilboa 1995). In summary, the available data suggest that CD4+ T cells are generally required for CTL priming when DCs are not activated through another mechanism. However, other conditions, such as the level of antigen provided or the integrity of draining lymphoid tissues, may also impact on the requirement for CD4 help. Nevertheless, CD4+ T cell help has been considered essential for the induction of CTL responses against tumors in the most cases. Under the noninflammatory condition of the majority of the cancers in addition to the unavailability of tumor antigens for cross-presentation (Spiotto et al. 2002), DCs require activation by CD4+ T cells before they can induce the full activation and differentiation of naïive CD8+ T cells into CTLs.
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
391
CD4+ T Cells for Maintenance of a Cytotoxic T Lymphocytes Response While the necessity of helper CD4+ T cells in the induction of CTL remains conflicting, it is generally accepted that CD4+ T cells are critical for the maintenance of CTL in both virally directed and tumor-specific immune response (Battegay et al. 1994; Matloubian et al. 1994; Cardin et al. 1996; Zitvogel et al. 1996). As clearly demonstrated in a human cytomegalovirus (CMV) infection, the survival of adoptively transferred anti-CMV CTLs were dependent on the presence of CMV-specific CD4+ T cells persisting in the host (Schomig et al. 1996). For the host to perpetually sustain an effector CTL function to counter a persisting virus infection may require help from CD4+ T cell through either cytokine secretions or stimulations that is independent of DC help. In certain reports, the need for helper CD4+ T cell can be replaced by an exogenously provided IL-2 (Kast et al. 1986, 1989; Bear et al. 1988), suggesting that the requirement for CD4+ T cells may simply be to supplement CD8+ T cells with IL-2, a cytokine traditionally thought to be essential for promoting growth and proliferation of T cells. Whether IL-2 is the primary mechanism or a byproduct utilized by CD4+ to provide help for the maintenance of CTL remains to be determined. However, the contribution of CD4+ help in sustaining tumor-specific CTL response is unequivocally recognized as an important means to an end. There is no shortage of corroborating data from therapeutic vaccination to adoptive transfer studies that necessitates CD4+ help in sustaining viable CTL function. In some murine models, while tumor growth was prevented by vaccinations with peptide-pulsed DC-mediated CTL induction, the same therapeutic immunization against established tumor, a scenario mimicking chronic persistence of antigen, required CD4+ T cells help (Zitvogel et al. 1996). Similarly, while adoptive transfer of tumor-specific CTLs prevented tumor formation in mice with CD4+ T cells depleted or MHC class II knockout mice, systemic metastases, a condition may require persistence of effector CTLs, could not be cured unless the hosts bear CD4+ T cells (Hu et al. 2000). Thus, as the tumor matures, CD4+ help becomes vital to the persistence of the effector response. Published reports on clinical trials of adoptive immunotherapy further support this concept (Dudley and Rosenberg 2003). Co-transfer of CD4+ T cells with CD8+ T cells expanded from autologous tumor infiltrating T cells prolonged the survival of the adoptively transferred T cells (Dudley and Rosenberg 2003). Taken together, it is evident that the longevity of a tumor-specific CTL response is stimulated by the presence of CD4+ T cells. CD4+ T Cells for the Induction and Maintenance of CD8+ T Cell Memory Responses After the clearance of the antigen, the greater part of the effector CTLs apoptosis while a marginal portion convert into lymphocytes with a memory phenotype; thus upon a second antigen encounter, CTL response propagates with alacrity and potency to provide the host with protection. It is generally accepted that memory cells persist in circulation subsequent to effector response, however, the role of
392
P. Yu and Y.-x. Fu
CD4+ T cells in the induction and maintenance of memory CTLs has been subjected to great controversy. While recent publications seem to have reached a consensus that CD4+ T cells are indispensable for an intact CTL memory response (Janssen et al. 2003; Shedlock and Shen 2003; Sun and Bevan 2003; Sun et al. 2004; Bevan 2004), several studies have suggested that CD4+ T cells are required in the primary responses to “program” the CD8+ T cells to differentiate into long-lived functional memory cells (Janssen et al. 2003; Shedlock and Shen 2003); others have argued that CD4+ T cells are required after antigen is eliminated to maintain the number or normal functions of CD8+ memory T cells (Sun et al. 2004; Bevan 2004). One earlier study reported a CD4+ T cell and B cell-independent persistence of memory H-Y-specific CTLs (Di Rosa and Matzinger 1996). Two other groups also found that virus-specific peptide-tetramer positive memory cells persisted in CD4 or MHC class II knockout mice using viral infection models (Stevenson et al. 1998; Zajac et al. 1998). However, despite the clear presence of virus-specific CD8+ memory T cells, both groups were unable to revert memory T cells into effector T cells in the absence of CD4+ T cells (Stevenson et al. 1998; Zajac et al. 1998). In other words, the memory phenotype CD8+ T cells were nonfunctional. Similar results were obtained in the previously described mouse H-Y model, where exogenous cytokine addition was required to revert memory cells back into CTLs in the absence of CD4+ T cells (Di Rosa and Matzinger 1996). These earlier findings suggest that CD4+ T cells were essential for the conversion of memory CTLs into effectors upon secondary antigen encounter. Other groups have expanded on the aforementioned studies by utilizing adoptive transfer experiments to exemplify the necessity of CD4+ in sustaining the memory cells in circulation. They complemented earlier studies with MHC class II knockout mice in adoptive transfer of memory T cells to illustrate that in the absence of CD4+ help, reactivation into effectors could not occur (Hu et al. 2000). While these earlier studies did not address if CD4+ T cells were required in the priming phase of CLT to “program” them to later develop to long-lived functional memory cells, recent studies by Bevan’s group utilizing techniques that allowed tracking of antigen specific memory T cells concluded that CD4+ T cells are not required during the priming of CTL, but only required at the later stage for the maintenance of memory CD8+ T cells (Sun et al. 2004). However, contradicting debate ensued with two recent publications by Shedlock and Shen (2003) and Janssen et al. (2003), respectively, claiming the opposite of previous studies. They were able to show that adoptively transferred CTLs survive and function properly in an environment without CD4+ T cells after being primed in its presence. In addition, they show that they were not required for the secondary activation of CTLs. Taken together, the data available so far suggested a general requirement for CD4 help to generate a healthy CD8 memory. However, at which stage of CD8 T cell responses CD4 help is essential, or the nature of such help still await to be determined. T helper 1 Versus T helper 2 Responses for Antitumor Immunity The CD4+ T cell responses can be divided into different types depending on their cytokine profile (Mosmann and Sad 1996). The defined T helper 1 (Th1) cells are
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
393
characterized by the production of IFN-g among others, whereas T helper 2 (Th2) cells produce IL-4, IL-5, and others. The balance between Th1 and Th2 cytokines has definite influence on the outcome of various immune responses as Th1 preferentially induces cellular immunity and Th2 tends to elicit humoral immunity (Romagnani 1997). The cytokine IFN-g positively impact on antigen processing and presentation, because MHC class I and II and the expression of several other molecules such as TAP and proteasome components are under the control of this cytokine (York et al. 1999). Therefore, the Th1 response is generally correlated with a better cellular and CTL response. Because a cellular immune response is preferable for tumor destruction, a Th1 response has been proposed to be beneficial for antitumor immunity. Several reports have supported this idea and demonstrated the parallelism between the generation of a Th1 response and a stronger antitumor immunity (Tsung et al. 1997; Aruga et al. 1997; Lowes et al. 1997; Fallarino and Gajewski 1999; Kacha et al. 2000). A Th1 response has even been shown to be essential for antitumor immunity and Th2 cytokines downregulate antitumor immunity in some reports (Kacha et al. 2000; Kobayashi et al. 1998; Ostrand-Rosenberg et al. 2000). The concept of immune deviation, a shift from Th1 to Th2 cytokine profile, has been hinted to be one of the major contributors to the failure of T cell-mediated immunity against tumors. Indeed, the immune deviation to Th2 cytokine production has been reported in progressive cancer patients (Pellegrini et al. 1996). In contrast, an immunization-evoked Th2 to Th1 change was shown to induce tumor rejection in a murine tumor model (Hu et al. 1998). Additionally, Th2 cytokines have been shown to promote tumor growth in several experimental models (Hu et al. 1998). However, as in most immunological situations, one can never irrevocably conclude as one way or the other. There are plenty of data supporting the opposite. Th2 cytokines have been shown to be helpful for cancer gene therapy (Tepper et al. 1989; Allione et al. 1994), and tumor-specific Th2 clones have been demonstrated to exhibit strong antitumor activity in vivo (Shen and Fujimoto 1996; Nishimura et al. 1999; Fallarino et al. 2000; Klugewitz et al. 2000). The mechanisms of how Th2 helper T cells destruct tumors is not yet clearly illustrated, but there is some evidence to suggest that the antitumor effect is mediated through the activation of innate immune cells such as eosinophils and macrophages by Th2 CD4+ T cells, which in turn secretes superoxide and nitric oxide (Hung et al. 1998). Although an effective antitumor immunity would be preferentially mediated through a Th1 response resulting in a direct killing by CTL, the results published up to this point seems to be directed towards a cooperative balance between the Th1 and Th2.
Regulatory CD4+ Cells To put it succinctly, the immune system is constantly working to find a balance between Th1 and Th2, activation and apoptosis, and pro-inflammatory and antiinflammatory to name a few. It has recently come to light that during tumor progression the tumor microenvironment becomes host to such an event, specifically between the effector and the regulatory response. Although not much is known regarding the
394
P. Yu and Y.-x. Fu
regulatory function pertaining to the tumor immunity, there is some evidence to suggest that the subtype of T cells responsible for “regulating” the effector immune response within the tumor site is similar to the now well-characterized regulatory CD4+ T cells involved in autoimmunity (Curiel et al. 2004; Yu et al. 2005; Wang et al. 2004; Dieckmann et al. 2001; Liyanage et al. 2002; Woo et al. 2001, 2002).
Existence of Different Types of CD4+ Regulatory T Cells In Vivo Although during the past decades there had been much speculation and some evidence for suppressor T cells, such cells had not been phenotypically identified until recent years. A major advance resulted from the discovery by Sakaguchi and colleagues, and later confirmed by others that the small subset of CD4+ T cell, separate from CD4+ helper T cells by the expression of high levels of CD25 in the naïve mice, could induce organ-specific autoimmunity by its depletion followed by adoptive transfer or prevent its development (Sakaguchi et al. 1985, 1995; Sugihara et al. 1988). These data formed the basis into the regulatory nature of these cells, now defined as naturally occurring CD4+CD25+ T cells. This subset of cells has been established to be a powerful regulator of T cell responses in organ-specific autoimmunity and chronic infections (Bach and Francois Bach 2003; Belkaid et al. 2002; Maloy and Powrie 2001; Sakaguchi 2004; Shevach 2002). However, it is becoming increasingly clear that in many situations, CD4+CD25- T cells are as effective as CD4+CD25+ T cells in controlling T cell-mediated disease (Apostolou et al. 2002; Curotto de Lafaille and Lafaille 2002; Furtado et al. 2001). There are other subsets of CD4+ T cells that have been defined to manifest regulatory phenotypes. For example, CD4+CD45Rblow suppressor cells that secrete large quantities of either IL-10 and IL-4 (termed Tr1 cells) and other CD4+ suppressor T cells that secrete large quantities of TGF-b (termed Th3 cells) (Levings and Roncaralo 2000; O’Garra et al. 2004). The immune regulatory potential and functional significance of these cytokine-secreting CD4+ T cells is supported by the findings that TGF-b-deficient mice develop autoimmune disease (Gorelik and Flavell 2000) and that administration of neutralizing antibodies to IL-4 or TGF-b abrogates the in vivo prevention of autoimmunity or tolerance-inducing activity of CD4+ T cells in the tumor model and some autoimmunity models (Yu et al. 2005; Seddon and Mason 1999; Zhai and Kupiec-Weglinski 1999). It remains to be determined, however, to what extent do these subtypes of CD4+ cells diverge from the naturally occurring CD4+CD25+ T cells in function and significance.
Markers to Identify CD4+ Regulatory T Cells In Vivo There are no known cell surface molecules that uniquely distinguish the CD4+ regulatory T cells from conventional activated CD4+ cells. For example, the CD25 molecule, which is the a-chain of the IL-2 receptor, is expressed on all peripheral
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
395
antigen-reactive CD4+ T cells from one to several days following antigen activation. Moreover, many of the other cell-surface molecules in addition to CD25, including the TNF-family member GITR and cytotoxic T-lymphocyte antigen-4 (CTLA4) (Sakaguchi et al. 2001; Wood and Sakaguchi 2003), that seem to distinguish CD4+CD25+ from CD4+CD25- effector cells are upregulated on CD4+CD25- T cells following antigen activation. In this regard, it is of great interest that a recently cloned transcription factor, termed Foxp3, a member of the forkhead family of DNA binding transcription factors, is not expressed in naïve CD4+CD25- cells but is highly expressed in the naturally occurring CD4+CD25+ regulatory cells. More importantly, mutational defects in the Foxp3 gene result in the fatal autoimmune and inflammatory disorder of the scurfy mouse and in the clinical and molecular features of the immunodysregulation, polyendocrinopathy, enteropathy, and X-linked syndrome (IPEX syndrome) in humans (Bennett et al. 2001; Brunkow et al. 2001; Wildin et al. 2001). In Foxp3-overexpressing mice, both CD4+CD25- and CD4-CD8+ T cells show suppressive activity, which suggests that expression of Foxp3 is linked to suppressor functions (Khattri et al. 2003; Fontenot et al. 2003; Hori et al. 2003). Taken together, these data strongly support the idea that Foxp3 may uniquely define the subset of CD4+ regulatory T cells in mice (Khattri et al. 2003; Fontenot et al. 2003; Hori et al. 2003) and humans (Walker et al. 2003). However, the recent findings that Foxp3 can be expressed in CD4+CD25- cells following activation and are also expressed in activated CD8+ T cells suggest that Foxp3 is linked to functional suppression, but not necessarily as a specific lineage marker (Walker et al. 2003; Chen et al. 2003; Cosmi et al. 2003; Manavalan et al. 2004). Whether a specific lineage marker even exists for these CD4+ regulatory cells awaits to be determined, but what can be concluded with little doubt is that a subset of CD4+ T cells indeed exist that regulate an inflammatory immune response. It can be speculated that such characterization of an inflammatory immune response may be ascribed to cancer.
Tumor-Induced CD4+ Regulatory T Cells Cancers generally develop over a long period of time. In addition, the major pathophysiologic characteristics of malignant cancer, invasion across natural tissue barriers and metastasis, are often associated with the disruption of normal tissue architecture leading to the initiation of inflammatory responses. In this regard, the cancers can be very much resembled to a chronic inflammation response. In view of this, it is possible to speculate that the anti-inflammatory mechanisms that are turned on at the beginning to protect normal tissues during a chronic inflammation coupled with the internal mechanisms controlled by malignant cells to produce cytokine like TGF-b start a process within the tumor microenvironment to set off a regulatory reaction that may adversely inhibit antitumor immunity. Given such correlative properties, the loss of regulatory function by the depletion of this CD4+ T cell subset may prolong the effector response resulting in tumor rejection.
396
P. Yu and Y.-x. Fu
Recent findings attest to possible negative regulatory roles by CD4+ T cells within the tumor environment. In some spontaneous tumor models, the presence of CD4+ T cells seems to promote cancer development instead of inhibiting (Daniel et al. 2003). One study by Schreiber’s group has shown that active immunization with antigen-specific CD4+ T cells in cancer-prone mice carrying a germline mutant ras oncogene resulted in immune responses that fail to eradicate mutant oncogene-expressing tumor cells, instead induced a remarkable enhancement of tumor growth (Siegel et al. 2000). Similarly, the studies by Robert North and his colleagues have shown that suppression by CD4+ T cells led to the progressive growth of an immunogenic tumor (Berendt and North 1980; Bursuker and North 1984) and intravenous depletion of these cells reversed the suppression to elicit CD8+ T cell-mediated antitumor immunity (Awwad and North 1988). While the presence of CD4+ T cells in some models have shown to be deleterious to the onset of tumor immunity, the necessity of CD4+ T cell help in mounting an effector immune response cannot be negated. In some studies, the depleting antibody given during the early stages of tumor growth was damaging to the generation of an immune response against the tumor, especially if T cell help was obligatory (Turk et al. 2004; Sutmuller et al. 2001), strongly suggesting the existence of functionally distinct CD4+ T cells subsets.
CD4+CD25+ Regulatory T Cells in Mice and Human Recent studies have suggested that CD4+CD25+ suppressor cells are relevant in to tumor immunology, albeit more deleterious than favorable for the host. Much of the study, up to recent times, have utilized a depletion of the entire CD4+ population, resulting in enhancement of immunity against tumors via the depletion of the suppressive CD4+CD25+ T cell population. In vivo experiments in several murine tumor models demonstrate this case as depletion of CD4+CD25+ T cells by anti-CD25 antibody treatment prior to tumor challenge significantly enhances the efficacy of vaccineinduced antitumor immunity (Tawara et al. 2002; Li et al. 2002; Golgher et al. 2002; Casares et al. 2003; Machiels et al. 2001; Sutmuller et al. 2001; Steitz et al. 2001; Jones et al. 2002; Shimizu et al. 1999; Onizuka et al. 1999). Another study revealed that splenic cells depleted of CD4+CD25+ T cells can mediate tumor regression presumably through promoting autoreactivity because autoimmune diseases were also induced (Shimizu et al. 1999). These studies suggest that regulatory T cells may inhibit initial priming of CD8+ T cells, some of which recognize tumor antigens. One data demonstrated more clearly that the CD4+CD25+ T cells prevented priming of CTL, as depletion of CD25+ T cells had to be performed within the first 2 days after tumor inoculation, suggesting that suppression was ineffective once the priming was initiated (Onizuka et al. 1999). It was also shown that elimination of these regulatory T cells, despite causing increased autoreactivity in some cases, could increase immune responses to tumors such as melanomas over-expressing self-antigens (Sutmuller et al. 2001; Steitz et al. 2001; Jones et al. 2002; Shimizu et al. 1999). Even when the
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
397
host bears a poorly immunogenic cancer, concomitant immunity can be rescued by systemic depletion of the CD4+CD25+ regulatory T cell subset (Turk et al. 2004). These observations in the tumor models are consistent with the features that have been defined for CD4+CD25+ regulatory T cells in other disease models. It has been demonstrated that the equivalent of CD4+CD25+ regulatory T cells identified in mice also exist in humans. These cells, CD4+CD25+CD45RO+ T cells, occupy mean 6% of CD4+ T cells and are present in the blood of healthy adults (Dieckmann et al. 2001). Prevalence of regulatory T cells is increased in peripheral blood and tumor microenvironment of patients with pancreas or breast adenocarcinoma (Liyanage et al. 2002). The CD4+CD25+ T cells possessing regulatory properties have also been reported to be among the tumor-infiltrating T cells in different types of human cancers, such as lung, ovarian, pancreas, breast and gastrointestinal cancers, and lymphoma (Woo et al. 2001, 2002; Sasada et al. 2003; Marshall et al. 2004). Antigen-specific activation and cell–cell contact were required for these clones of Treg cells to exert suppressive activity on CD4+ effector cells. The finding that the presence CD4+ Treg cells at tumor sites suggests that they could have a profound effect on the inhibition of T cell responses against some human cancers (Curiel et al. 2004; Yu et al. 2005; Wang et al. 2004).
Suppression Occurred Inside Tumor Tissues It is possible that different subsets of CD4+ T cells, either providing help or regulating, dominate during various stages of tumor progression. A recent study by our group (Yu et al. 2005) has shown that suppression of immunity against tumor mainly occurs in the effector phase at tumor site and depletion of the regulatory T cells at the late stage of tumor progression did not mitigate the possible T helper function. We utilized a highly antigenic tumor that expresses a strong antigen on the surface, yet fails to regress in the host induced an accumulation of CD4+ regulatory T cells within the tumor microenvironment as the tumor persists and establishes, subsequently inhibiting CD8+ T cell function. In this model, local intratumoral depletion of these regulatory T cells unmasked the immunogenicity of tumor and reversed the CTL toleration leading to the rapid rejection of well-established tumors. From this study, we propose that CD4+ cells predominately play an enhancing helper role during the initial stages, but once tumors become chronically persistent and established the increased accumulation of CD4+ regulatory T cells inhibits CD8+ cell function and masks the immunogenicity of tumor. In fact, depletion of regulatory T cells unveils the immunogenicity of tumor cells and provides long-term protection against re-challenge of even the parental tumor cells lacking the strong antigen. This result suggests that the depletion of regulatory T cells promoted immunity against previously poorly immunogenic tumor antigens and expanded the tumor-reactive CD8+ T cell repertoire. Our study revealed that the population of tumor-infiltrating cells is skewed to favor regulatory CD4+ T cells over the helper CD4+ T cells within the tumor tissue, especially as tumor progresses and gets established in the host.
398
P. Yu and Y.-x. Fu
Origins of Tumor-Induced CD4+ Regulatory T Cells It is not very clear thus far if these CD4+CD25+ T cells inhibiting antitumor immunity in mice and human are naturally occurring regulatory T cells or others that generated in the periphery. It is possible that tumor environment preferentially recruits naturally occurring CD4+CD25+ T cells. One study published recently suggested that chemokine CCL22 abundantly expressed in the ovarian cancer tissues and tumor ascites cells preferentially attract the CD4+CD25+ T cells they identified in the ovarian cancers (Curiel et al. 2004). These cells were positive for the transcriptional factor Foxp3 and exhibited regulatory function similarly as the naturally occurring CD4+CD25+ T cells. It is not very clear so far if these tumor-infiltrating regulatory T cells can be defined as the counterpart of the naturally occurring CD4+CD25+ T cells in the mouse. What the specific trafficking clues are for the naturally occurring CD4+CD25+ T cells in the murine models still waits exploring. Another possibility is that tumor environment converts CD4+ T cells to CD4+CD25+ regulatory T cells or expands naturally occurring CD4+CD25+ T cells. There is evidence to indicate changes in the function of tumor-specific CD4+ T cells, from effectors to suppressors, during cancer progression (Zhou et al. 2004). Conversion from effector cells coincided with a substantial reduction in the antigen expression level, resulting in tumor persistence that ultimately led to T cell tolerance. They have evidence to further suggest that these antigen-specific T cells became CD4+CD25+ regulatory T cells (Zhou et al. 2004). An interesting model was suggested by this study that the processes of immunosurveillance and tumor editing coexist with a process in which the functional tumor-specific T cell repertoire is also edited by the tumor environment to benefit the tumor progression. Besides the anti-inflammatory cytokines, the CD4+CD25+ regulatory T cells inside the tumor may suppress antitumor immunity via other mechanisms. For example, these Tregs may inhibit immune response through their ability to control T cell numbers because they have been shown to regulate T cell proliferation in vitro (Sakaguchi 2004; Shevach 2002). Whether the regulatory cells that accumulate in the tumor site are ones that naturally exist in the host, or they initially migrate as helper CD4+ T cells, but become converted to regulatory cells by encountering the suppressive tumor environment is not altogether clear. It would be beneficial to better characterize the nature of CD4+ T cells isolated from or present in the tumor tissues by surface markers and cytokine profiles to utilize the CD4+ helper T cells instead of CD4+ regulatory T cells for adoptive transfer immunotherapy and association with better prognosis of cancer patients.
Th17 Cells CD4+ T cells differentiate into different lineages of helper T (TH) cells that are defined by distinct developmental regulation and biological functions following activation. TH17 cells producing IL17 have recently been identified as a new lineage of effector TH cells (Dong 2008) whose induction in mice is dependent on the
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
399
synergistic activation by TGF-b and IL-6 (Bettelli et al. 2006; Mangan et al. 2006; Veldhoen et al. 2006). TGF-b is often present in various tissues and tumors, possibly contributing to the induction of either regulatory FOXP3+ or Th17 phenotype (Li et al. 2007; Nam et al. 2008). Indeed, Th17 cells have been found in the tumor microenvironment in animal models and in patients, although their specificity and function remain unclear (Kryczek et al. 2007; Tartour et al. 1999). Th17 cells have been shown to be important in immune responses to infectious agents and inflammation leading to autoimmune diseases (Dong 2008). The importance of an inflammatory milieu in the initiation of cancer has been well described (Lin and Karin 2007), consistent with the facts that tumors often arise in pro-inflammatory conditions induced by infections, chemical irritants, or autoimmunity. The IL-6/ IL-23/IL-17 axis of inflammation has been shown to be associated with carcinogenesis (Langowski et al. 2006). IL-6 and IL-23 via STAT3 signaling might deliver prosurvival stimuli to some malignancies in an autocrine or paracrine mechanism, supported by the finding that the animals deficient in IL-23 were protected from cancer progression (Foulds et al. 2008; Kortylewski et al. 2005). IL-17 might promote tumor-associated neovascularization (Numasaki et al. 2003; Overwijk et al. 2006) and possibly recruit myeloid suppressor cells into the inflammatory environment, potentiating tumor growth (Charles et al. 2009). However, it has also been shown that systemic overexpression of IL-23 results in increased antitumor activity of T cells, involving both IFN-g and IL-17 secretion (Overwijk et al. 2006; Kaiga et al. 2007). In an adoptive transfer model, Th17-polarized cells were more effective in eradicating tumors than unpolarized T helper cells or Th1 cells. However, the treatment with tumor-specific Th17 was critically dependent on IFN-g (Muranski et al. 2008). The authors interpreted the data as that the potential duality of the adoptively transferred effector Th17 cells later turned on T-bet to express a Th1 phenotype to contribute to the antitumor effects (Muranski and Restifo 2009). Since T-bet-expressing Th1 cells might represent a terminally differentiated phenotype that is more prone to senescence and/or apoptosis than their Th17 counterparts (Intlekofer et al. 2007), and Th17-defining transcription factor ROR-gt might have anti-apoptotic and antiproliferative activity, analogous to its well-described role during the thymocyte development (He et al. 1998; Kurebayashi et al. 2000; Sun et al. 2000; Xi et al. 2006), the early acquisition of t-bet might impair the ability of T cells to persist resulting in the disadvantages over Th17 cells in adoptive transfer therapy. Thus, the ability of the transferred Th17 cells to evolve in vivo to Th1 phenotype is crucial for the therapeutic outcome.
Other Aspects That Complicate the Relationship Between the Tumor-Infiltrating Lymphocytes and Prognosis The association of TIL with better prognosis seems to be influenced by other factors besides the immunophenotypes of TIL. Melanomas are among the tumors that are often associated with TIL. Many correlation studies have been done since Clark et al. (1969) first described the lymphocytic infiltration of PCM.
400
P. Yu and Y.-x. Fu
Patients with a moderate-to-marked lymphocytic infiltrate within their primary melanoma had a significantly better prognosis and a three-time higher 5-year survival rate than patients with a sparse or absent lymphocytic infiltrate (Day et al. 1981). Elder et al. (1985) differentiated the lymphocytic infiltrate into brisk, nonbrisk, or absent, according to its intensity, and demonstrated that TIL were of prognostic significance only in vertical growth phase (VGP) melanoma. In contrast, the extent of lymphocytic infiltration had no prognostic influence in radial growth phase (RGP) melanomas, regardless of whether the melanoma was in situ or invasive. These findings were later confirmed by another study (Clemente et al. 1996). The 5- and 10-year survival rates were 77 and 55% in melanomas with brisk VGP infiltrates; 53 and 45% with nonbrisk VGP infiltrates; and 37 and 27% without VGP infiltrates (Clemente et al. 1996). Also, the number of TILs in the primary tumor has been found to be inversely correlated with the probability for lymph node metastases (Clemente et al. 1996). Patients with brisk TIL infiltrates in their primary tumors showed a 3.9% probability of a positive sentinel lymph node (SLN), compared to a 26.2% probability in patients with TILs absent from their primary melanoma (Taylor et al. 2007). Furthermore, of those patients with regional lymph node metastases, patients with more marked lymphocytic responses in their metastatic melanoma showed a significantly higher 30-month disease-free survival rate (81.3% for patients with a brisk TIL infiltrate; 46.8% for patients with a nonbrisk infiltrate; and, 29.3% for patients with TILs absent from their lymph node metastases) (Day et al. 1981; Mihm et al. 1996). However, other studies could not convincingly demonstrate that brisk TIL infiltrates were associated with improved survival in melanoma patients (Barnhill et al. 1996; Thorn et al. 1994). These discrepant results may in part be explained by differences in patient populations investigated, with particular reference to the thickness of patients’ melanomas (Taylor et al. 2007). The study by Clemente et al. (1996) found the impact of TILs most pronounced in patients with high-risk lesions, thicker than 1.7mm but less than 6mm in Breslow depth (Taylor et al. 2007). This suggested that the briskness of the TIL infiltrate was prognostic for T2–T4 (TMN system) primary cutaneous melanoma (PCM), though the prognostic significance of TILs was lost in very thick lesions (advanced T4). In contrast, Barnhill et al. (1996) did not find any survival advantage to be associated with brisk TIL infiltrates; however, patients with both RGP and VGP were included in this study (Barnhill et al. 1996), even though other studies did not demonstrate a prognostic significance of TILs in RGP PCM (Clemente et al. 1996; Elder et al. 1985). Furthermore, only 25.6% of patients in Barnhill’s study had lesions thicker than 1.7mm (Barnhill et al. 1996), while 82 and 71% of patients had lesions thicker than 1.7mm in the studies by Clemente et al. (1996) and Tuthill et al. (2002). Taylor et al.(2007) did not find an impact of TILs on survival (44% of patients had lesions thicker than 2mm); however, they did show that TILs are an independent predictor of SLN positivity, which by itself is the most important independent predictor of recurrence and survival in malignant melanoma patients (Balch et al. 2001). It is possible that there are distinct mechanisms for the
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
401
induction of TIL at RGP and VGP, for which might be unique in each cancer type and wait for further investigation.
Targeted Tumor Tissues to Recruit and Train T Cells Tumor often forms barrier to limit T cell infiltration and reduce sufficient tumor antigens to draining lymph nodes. Since the positive roles played by CD8+ T cells and CD4+ helper T cells in antitumor immunity, it is reasonable to speculate that increase of infiltration of T lymphocytes and initiation of antitumor immune responses inside the tumor tissue at early phase could increase T cell repertoire and more antigens to stimulate T cells inside tumor would be a strategy for cancer immunotherapy. Several cytokines or chemokines have been used to attract and activate T cells at tumor sites. In this review, we focus on our recent finding that stimulation of lymphotoxin b receptor (LTbR) inside tumor tissues will promote strong infiltration of immune cells leading to tumor rejection. The LTbR plays an important role in the formation of lymphoid structures (Ettinger 2000; Fu and Chaplin 1999; Kim et al. 2000). LTbR is activated by two members of the TNF family, membrane lymphotoxin ab and LIGHT (Mauri et al. 1998). Signaling via LTbR regulates the expression of chemokines and adhesion molecules within secondary lymphoid organs. Chemokines and adhesion molecules control the migration and positioning of DCs and lymphocytes in the spleen (Wu et al. 1999; Ngo et al. 1999). Therefore, it is possible that enhanced LTbR signaling inside tumor tissues may promote the formation of lymphoid-like structure for direct T cell sequestration. TNFR signaling may also play similar but less effective role in regulating chemokine expression (Ngo et al. 1999). However, TNFR signaling may have a more toxic effect as seen in other systemic TNF treatments without extra costimulation (Spriggs et al. 1988). Soluble LTa can signal through the TNFR resulting in the upregulation of chemokines. To avoid toxic effect, recombinant lymphotoxin a (LTa) has been conjugated with antibody targeting specifically to the tumor tissues resulted in an effective antitumor immune response associated with the induction of peripheral lymphoid-like tissue (Schrama et al. 2001). However, LTa lacks of costimulation function, which would result in a less effective activation of recruited naïve T cells in the lymphoid-like structure inside tumor tissues. LIGHT is a ligand for LTbR and herpes virus entry mediator (Mauri et al. 1998; Rooney et al. 2000). LIGHT is predominantly expressed on lymphoid tissues, especially on the surface of activated DCs and T cells. LIGHT is a strong costimulatory molecule (Mauri et al. 1998; Tamada et al. 2000; Zhai et al. 1998). Our data indicates that the interactions between LIGHT and LTbR restore lymphoid structures in the spleen of LTa-/- mice. In addition, upregulation of LIGHT causes T cell activation and migration into nonlymphoid tissues and formation of lymphoid-like structures (Wang et al. 2001, 2002). Therefore, LIGHT-mediated microenvironment inside tumor could be effective in both recruiting and activating naïve T cells into tumor tissues. The expression of LIGHT in the tumor environment induces a
402
P. Yu and Y.-x. Fu
massive infiltration of naïve T lymphocytes that correlates with an upregulation of both chemokine production and expression of adhesion molecules. Activation and expansion of these infiltrating T cells leads to the rejection of established, highly progressive tumors at local and distal sites (Yu et al. 2004). Our study indicates that indeed induction of infiltration of T lymphocytes and initiation of antitumor immune responses inside the tumor tissue may be an effective strategy for cancer immunotherapy.
Concluding Remarks Understanding the balance of antitumor effector T cells versus regulatory T cells may be more important to determine the outcome of immune responses inside tumors. Our recent study has demonstrated that rapid recruitment of naïve lymphocytes and expansion of CD8+ T cells inside the tumor may be a way of creating a dominant pro-inflammatory environment, leading to the rejection of tumor at local and distal sites (Yu et al. 2004). We later further demonstrated that the depletion of regulatory T cells inside the tumor is another efficient way of converting the antiinflammatory environment inside tumor to pro-inflammatory one (Yu et al. 2005). From clinical therapeutic point of the view, the local treatment to eliminate regulatory T cells has certain critical advantages over systemic treatment: First, local treatment may avoid side effects induced by systemic depletion of all CD4+ T cells, which may abrogate T helper-mediated protective immunity against pathogens; Second, it would not hinder effective priming of CD8+ T cell in the lymphoid tissues by the helper CD4+ T cells because depletion remains local. Third, the local treatment would be even expected to be more effective if the key suppression of CD4+CD25+ regulatory T cells resides inside tumor. Also intratumoral treatment would reduce the dose of antibody to be applied to the patients as well as be more affordable even for the developing countries. CD4+CD25+ T cells have been shown to be present in a variety of human cancer tissues (Woo et al. 2001, 2002; Sasada et al. 2003; Marshall et al. 2004) and these cells are negatively associated with the prognosis of the ovarian cancer patients in addition (Curiel et al. 2004). Therefore, regulatory T cells within the tumor environment represent an attractive target, and their depletion may lead to improvements in the current immunotherapy protocol in the future clinical trials. It is likely that a combination treatment which would rapidly expand the effector cells at the tumor site, while locally depleting the regulatory cells, could potentially provide a potent strategy for enhancing antitumor immunity and permitting a clinically desirable outcome for cancer patients. It is increasingly clear that some of TIL may be friend while others may be foes. Increase of TIL may not always be associated with better prognosis. Accumulation of regulatory T cells inside tumor tissue in both animal model and human patients suggest that aberrant immune response occurs inside growing tumor causing immune invasion. How to effectively reverse the immunologically suppressive environment or depleting regulatory T cells inside tumors may be a new avenue to
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes
403
generate effective immune responses against tumors at local and distal sites. Careful laboratory investigation is critical for us to better understand the role of TIL. It is possible that future immunotherapy and prognosis depend on better defining of TIL. When cancer tissue contains fewer TIL, targeting tumor tissues for the increase of TIL, especially CTL may result in positive responses. On the other hand, depletion of T regulatory cells may be important for patients with cancer tissues containing T regulatory cells. The pathologists may play critical diagnostic role in determining the type of TIL, which provide not only prognosis but also the guide for immunotherapy.
References Allione A et al (1994) Cancer Res 54(23):6022 Apostolou I, Sarukhan A, Klein L, von Boehmer H (2002) Nat Immunol 3(8):756 Aruga A et al (1997) J Immunol 159(2):664 Awwad M, North RJ (1988) J Exp Med 168(6):2193 Bach JF, Francois Bach J (2003) Nat Rev Immunol 3(3):189 Balch CM et al (2001) J Clin Oncol 19(16):3622 Balkwill F, Mantovani, A (2001) Lancet 357(9255):539 Banner BF et al (1990) Mod Pathol 3(2):129 Barnhill RL, Fine JA, Roush GC, Berwick M (1996) Cancer 78(3):427 Battegay M et al (1994) J Virol 68(7):4700 Bear HD, Susskind BM, Close KA, Barrett SK (1988) Cancer Res 48(6):1422 Beatty GL, Paterson Y (2000) J Immunol 165(10):5502 Beatty G, Paterson Y (2001) J Immunol 166(4):2276 Beck-Engeser GB et al (2001) J Exp Med 194(3):285 Belkaid Y et al (2002) Nature 420(6915):502 Bennett SR et al (1997) J Exp Med 186(1):65 Bennett SR et al (1998) Nature 393(6684):478 Bennett CL et al (2001) Nat Genet 27(1):20 Berendt MJ, North RJ (1980) J Exp Med 151(1):69 Bettelli E et al (2006) Nature 441(7090):235 Bevan MJ (2004) Nat Rev Immunol 4(8):595 Boyse EA, Old LJ, Stockert E (1962) Ann NY Acad Sci 99:574 Brunkow ME et al (2001) Nat Genet 27(1):68 Bursuker I, North RJ (1984) J Exp Med 159(5):1312 Cammarota G et al (1992) Nature 356(6372):799 Cardin RD, Brooks JW, Sarawar SR, Doherty PC (1996) J Exp Med 184(3):863 Casares N et al (2003) J Immunol 171(11):5931 Cavallo F et al (1992) J Immunol 149(11):3627 Cella M et al (1996) J Exp Med 184(2):747 Charles KA et al (2009) J Clin Invest Cheever MA, Thompson DB, Klarnet JP, Greenberg PD (1986) J Exp Med 163(5):1100 Chen W et al (2003) J Exp Med 198(12):1875 Clark WH Jr, From L, Bernardino EA, Mihm MC (1969) Cancer Res 29(3):705 Clark W H, Jr et al (1989) J Natl Cancer Inst 81(24):1893 Clemente C G et al (1996) Cancer 77(7):1303 Cosmi L et al (2003) Blood 102(12):4107 Curiel TJ et al (2004) Nat Med 10(9):942
404
P. Yu and Y.-x. Fu
Curotto de Lafaille MA, Lafaille JJ (2002) Curr Opin Immunol 14(6):771 Daniel D et al (2003) J Exp Med 197(8):1017 Day CL, Jr. et al (1981) Ann Surg 193(4):436 De Bruijn ML et al (1991) Eur J Immunol 21(12):2963 Dieckmann D et al (2001) J Exp Med 193(11):1303 Di Rosa F, Matzinger P (1996) J Exp Med 183(5):2153 Dong C (2008) Nat Rev Immunol 8(5):337 Dranoff G et al (1993) Proc Natl Acad Sci USA 90(8):3539 Dudley ME et al (2002) Science 298(5594):850 Dudley ME, Rosenberg SA (2003) Nat Rev Cancer 3(9):666 Elder DE et al (1985) Cancer 56(2):413 Ettinger R (2000) Curr Top Microbiol Immunol 251:203 Fallarino F, Gajewski TF (1999) J Immunol 163(8):4109 Fallarino F et al (2000) J Immunol 165(10):5495 Fayolle C et al (1996) Immunology 89(1):41 Fearon ER et al (1990) Cell 60(3):397 Feltkamp MC et al (1993) Eur J Immunol 23(9):2242 Fields RC, Shimizu K, Mule JJ (1998) Proc Natl Acad Sci USA 95(16):9482 Fontenot JD, Gavin MA, Rudensky AY (2003) Nat Immunol 4(4):330 Fossati G, Anichini A, Parmiani G (1987) Int J Cancer 39(6):689 Foulds KE et al (2008) J Immunol 180(2):842 Fu YX, Chaplin DD (1999) Annu Rev Immunol 17:399 Fujiwara H et al (1984) J Immunol 133(3):1671 Furtado GC et al (2001) Immunol Rev 182:122 Golgher D et al (2002) Eur J Immunol 32(11):3267 Golumbek PT et al (1991) Science 254(5032):713 Gorelik L, Flavell RA (2000) Immunity 12(2):171 Greenberg PD, Cheever MA, Fefer A (1981) J Exp Med 154(3):952 He YW, Deftos ML, Ojala EW, Bevan MJ (1998) Immunity 9(6):797 Hock H, Dorsch M, Diamantstein T, Blankenstein T (1991) J Exp Med 174(6):1291 Hori S, Nomura T, Sakaguchi S (2003) Science 299(5609):1057 Hu HM, Urba WJ, Fox BA (1998) J Immunol 161(6):3033 Hu HM, Winter H, Urba WJ, Fox BA (2000) J Immunol 165(8):4246 Huang AY et al (1994) Science 264(5161):961 Huang AY, Bruce AT, Pardoll DM, Levitsky HI (1996) Immunity 4(4):349 Hung K et al (1998) J Exp Med 188(12):2357 Igarashi T et al (1992) Eur Urol 22(4):323 Intlekofer AM et al (2007) J Exp Med 204(9):2015 Janssen EM et al (2003) Nature 421(6925):852 Jones E et al (2002) Cancer Immun 2:1 Kacha AK, Fallarino F, Markiewicz MA, Gajewski TF (2000) J Immunol 165(11):6024 Kaiga T et al (2007) J Immunol 178(12):7571 Kast WM, Bronkhorst AM, de Waal LP, Melief CJ (1986) J Exp Med 164(3):723 Kast WM et al (1989) Cell 59(4):603 Khattri R, Cox T, Yasayko SA, Ramsdell F (2003) Nat Immunol 4(4):337 Kim D et al (2000) J Exp Med 192(10):1467 Klein G, Sjogren HO, Klein E, Hellstrom KE (1960) Cancer Res 20:1561 Klein E et al (1980) Contemp Top Immunobiol 10:79 Klugewitz K, Scheffold A, Radbruch A, Hamann A (2000) Int J Cancer 87(5):673 Knuth A, Danowski B, Oettgen HF, Old LJ (1984) Proc Natl Acad Sci U S A 81(11):3511 Kobayashi M, Kobayashi H, Pollard RB, Suzuki F (1998) J Immunol 160(12):5869 Koch F et al (1996) J Exp Med 184(2):741 Kolbeck PC et al (1992) Mod Pathol 5(4):420
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes Kortylewski M et al (2005) Nat Med 11(12):1314 Kowalczyk D, Skorupski W, Kwias Z, Nowak J (1997) Br J Urol 80(4):543 Kripke ML (1974) J Natl Cancer Inst 53(5):1333 Kryczek I et al (2007) J Immunol 178(11):6730 Kurebayashi S et al (2000) Proc Natl Acad Sci U S A 97(18):10132 Langowski JL et al (2006) Nature 442(7101):461 Leclerc JC, Gomard E, Levy JP (1972) Int J Cancer 10(3):589 Levings MK, Roncarolo MG (2000) J Allergy Clin Immunol 106(1 Pt 2):S109 Levitsky HI, Lazenby A, Hayashi RJ, Pardoll DM (1994) J Exp Med 179(4):1215 Li J, Hu P, Khawli LA, Epstein AL (2003) Cancer Res 63(23):8384 Li MO, Wan YY, Flavell RA (2007) Immunity 26(5):579 Lin WW, Karin M (2007) J Clin Invest 117(5):1175 Lin KY et al (1996) Cancer Res 56(1):21 Liyanage UK et al (2002) J Immunol 169(5):2756 Lowes MA et al (1997) J Invest Dermatol 108(6):914 Lu Z et al (2000) J Exp Med 191(3):541 MacDonald GH, Johnston RE (2000) J Virol 74(2):914 Machiels JP et al (2001) Cancer Res 61(9):3689 Mackey MF et al (1997) Cancer Res 57(13):2569 Mackey MF et al (1998) J Immunol 161(5):2094 Maloy KJ, Powrie F (2001) Nat Immunol 2(9):816 Manavalan JS et al (2004) Int Immunol 16(8):1055 Mangan PR et al (2006) Nature 441(7090):231 Marrogi AJ et al (1997) Int J Cancer 74(5):492 Marshall NA et al (2004) Blood 103(5):1755 Matloubian M, Concepcion RJ, Ahmed R (1994) J Virol 68(12):8056 Mauri DN et al (1998) Immunity 8(1):21 Mihm MC, Jr., Clemente CG, Cascinelli N (1996) Lab Invest 74(1):43 Monach PA, Meredith SC, Siegel CT, Schreiber H (1995) Immunity 2(1):45 Moore OS Jr, Foote FW, Jr (1949) Cancer 2(4):635 Mukherji B, MacAlister TJ (1983) J Exp Med 158(1):240 Mumberg D et al (1999) Proc Natl Acad Sci U S A 96(15):8633 Muranski P, Restifo NP (2009) Curr Opin Immunol 21(2):200 Muranski P et al (2008) Blood 112(2):362 Mosmann TR, Sad S (1996) Immunol Today 17(3):138 Naito Y et al (1998) Cancer Res 58(16):3491 Nakano O et al (2001) Cancer Res 61(13):5132 Nam JS et al (2008) Cancer Res 68(10):3915 Ngo VN et al (1999) J Exp Med 189(2):403 Nishimura T et al (1999) J Exp Med 190(5):617 Noguchi Y, Richards EC, Chen YT, Old LJ (1995) Proc Natl Acad Sci U S A 92(6):2219 Numasaki M et al (2003) Blood 101(7):2620 O’Garra A, Vieira PL, Vieira P, Goldfeld AE (2004) J Clin Invest 114(10):1372 Onizuka S et al (1999) Cancer Res 59(13):3128 Ostrand-Rosenberg S (1994) Curr Opin Immunol 6(5):722 Ostrand-Rosenberg S, Grusby MJ, Clements VK (2000) J Immunol 165(11):6015 Overwijk WW et al (2006) J Immunol 176(9):5213 Pan ZK, Ikonomidis G, Pardoll D, Paterson Y (1995) Cancer Res 55(21):4776 Pardoll DM, Topalian SL (1998) Curr Opin Immunol 10(5):588 Pellegrini P et al (1996) Cancer Immunol Immunother 42(1):1 Porgador A, Gilboa E (1995) J Exp Med 182(1):255 Potter TA, Rajan TV, Dick RF, 2nd, Bluestone JA (1989) Nature 337(6202):73 Pulaski BA et al (1993) Cancer Res 53(9):2112
405
406
P. Yu and Y.-x. Fu
Ridge JP, Di Rosa F, Matzinger P (1998) Nature 393(6684):474 Robins RA (1986) Biochim Biophys Acta 865(3):289 Rollins BJ (2006) Eur J Cancer 42(6):760 Romagnani S (1997) Immunol Today 18(6):263 Rooney IA et al (2000) J Biol Chem 275(19):14307 Rosenberg SA (1999) Immunity 10(3):281 Rosenberg SA (2001a) Nature 411(6835):380 Rosenberg SA (2001b) J Intern Med 250(6):462 Rosenberg SA, Yang JC, Restifo NP (2004) Nat Med 10(9):909 Rouse BT, Rollinghoff M, Warner NL (1972) Nat New Biol 238(82):116 Saiki Y et al (1996) Lab Invest 75(1):67 Sakaguchi S (2004) Annu Rev Immunol 22:531 Sakaguchi S, Fukuma K, Kuribayashi K, Masuda T (1985) J Exp Med 161(1):72 Sakaguchi S et al (1995) J Immunol 155(3):1151 Sakaguchi S et al (2001) Immunol Rev 182:18 Sasada T et al (2003) Cancer 98(5):1089 Schnell S, Young JW, Houghton AN, Sadelain M (2000) J Immunol 164(3):1243 Schoenberger SP et al (1998) Nature 393(6684):480 Schomig A et al (1996) N Engl J Med 334(17):1084 Schrama D et al (2001) Immunity 14(2):111 Schumacher K, Haensch W, Roefzaad C, Schlag PM (2001) Cancer Res 61(10):3932 Seddon B, Mason D (1999) J Exp Med 189(2):279 Shedlock DJ, Shen H (2003) Science 300(5617):337 Shen Y, Fujimoto S (1996) Cancer Res 56(21):5005 Shevach EM (2002) Nat Rev Immunol 2(6):389 Shi Y, Zheng W, Rock KL (2000) Proc Natl Acad Sci U S A 97(26):14590 Shimizu J, Yamazaki S, Sakaguchi S (1999) J Immunol 163(10):5211 Shinde S et al (1996) J Immunol 157(7):2764 Siegel CT et al (2000) J Exp Med 191(11):1945 Spellman CW, Daynes RA (1981) Hum Pathol 12(4):299 Spiotto MT et al (2002) Immunity 17(6):737 Spriggs DR et al (1988) J Natl Cancer Inst 80(13):1039 Steitz J et al (2001) Cancer Res 61(24):8643 Stevenson PG, Belz GT, Altman JD, Doherty PC (1998) Proc Natl Acad Sci U S A 95(26):15565 Sugihara S et al (1988) J Immunol 141(1):105 Sun JC, Bevan MJ (2003) Science 300(5617):339 Sun Z et al (2000) Science 288(5475):2369 Sun JC, Williams MA, Bevan MJ (2004) Nat Immunol 5(9):927 Sutmuller RP et al (2001) J Exp Med 194(6):823 Tamada K et al (2000) Nat Med 6(3):283 Tartour E et al (1999) Cancer Res 59(15):3698 Tawara I et al (2002) Jpn J Cancer Res 93(8):911 Taylor RC et al (2007) J Clin Oncol 25(7):869 Tepper RI, Pattengale PK, Leder P (1989) Cell 57(3):503 Tevethia SS, Blasecki JW, Vaneck G, Goldstein AL (1974) J Immunol 113(5):1417 Thorn M et al (1994) J Natl Cancer Inst 86(10):761 Toes RE et al (1996) Cancer Res 56(16):3782 Tsung K et al (1997) J Immunol 158(7):3359 Turk MJ et al (2004) J Exp Med 200(6):771 Tuthill RJ et al (2002) Am J Clin Pathol 118(4):504 Tuting T et al (1999) J Gene Med 1(6):400 Veldhoen M et al (2006) Immunity 24(2):179 Walker MR et al (2003) J Clin Invest 112(9):1437
19 The Prognostic Significance of Tumor-Infiltrating Lymphocytes Wang J et al (2001) J Immunol 167(9):5099 Wang J et al (2002) Eur J Immunol 32(7):1969 Wang HY et al (2004) Immunity 20(1):107 Ward PL et al (1990) Cancer Res 50(13):3851 Wildin RS et al (2001) Nat Genet 27(1):18 Wolfers J et al (2001) Nat Med 7(3):297 Wolkers MC, Stoetter G, Vyth-Dreese FA, Schumacher TN (2001) J Immunol 167(7):3577 Woo EY et al (2001) Cancer Res 61(12):4766 Woo EY et al (2002) J Immunol 168(9):4272 Wood KJ, Sakaguchi S (2003) Nat Rev Immunol 3(3):199 Wu Q et al (1999) J Exp Med 190(5):629 Xi H et al (2006) Immunity 24(6):813 York IA, Goldberg AL, Mo XY, Rock KL (1999) Immunol Rev 172:49 Yu P et al (2003) J Exp Med 197(8):985 Yu P et al (2004) Nat Immunol 5(2):141 Yu P et al (2005) J Exp Med 201(5):779 Zajac AJ et al (1998) J Exp Med 188(12):2205 Zhai Y, Kupiec-Weglinski JW (1999) Curr Opin Immunol 11(5):497 Zhai Y et al (1998) J Clin Invest 102(6):1142 Zhang L et al (2003) N Engl J Med 348(3):203 Zhou G et al (2004) J Exp Med 200(12):1581 Zitvogel L et al (1996) J Exp Med 183(1):87
407
Chapter 20
The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation Adam Yagui-Beltrán, Qizhi Tang, and David M. Jablons
Abstract Increasing research from preclinical and clinical studies is demonstrating that inflammation with its myriad of heterogeneous mediators and cellular effectors plays a role in the malignant initiation of epithelial cells. Pre-existing inflammation may exist prior to the development of cancer, or on the other hand it may be that oncogenic changes may lead to a certain inflammatory microenvironment that in turn will promote the development of a tumor. Examples of how inflammation may lead to cancer formation include enhanced proliferation of initiated cells, through resistance to apoptosis, induction of genomic instability, alterations in epigenetic events and subsequent inappropriate gene expression, abnormal tumor neovascularization or angiogenesis, and the promotion of metastasis among others. At the time of malignant transformation of epithelial cells, the many components of the tumor microenvironment (TME), such as the tumor cells themselves, the stromal cells in surrounding tissue and the infiltrating leukocytes cells generate an intratumoral inflammatory state by the aberrant expression and secretion of a florid array of pro-inflammatory molecules that include chemokines, cytokines, cyclooxygenase-2, prostaglandins (PGs), iNOS, nitric oxide (NO), and a vast network of intracellular signaling molecules including upstream kinases and transcription factors that facilitate tumor promotion and progression. In this chapter, we will explore current knowledge of the various components of the pro-inflammatory milieu within the TME; we will, for example, highlight the major molecular studies that support how pro-inflammatory cytokines, chemokines, and PGs are able to regulate some of the angiogenic switches controlled by vascular endothelial growth factors inducing inflammatory angiogenesis. We will highlight some of the principal tumor cell–stroma dynamic relationships and how often these interplays are incrementally manipulated in ways that favor malignant epithelial initiation, resulting in tumor
A. Yagui-Beltrán (*) Department of Surgery, Division of Adult Thoracic Surgery, The Helen Diller Family Comprehensive Cancer Center Thoracic Surgery, University of California San Francisco, 1600 Divisadero, Room A-743, San Francisco, CA 94143-1724, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_20, © Springer Science+Business Media, LLC 2010
409
410
A. Yagui-Beltrán et al.
immune evasion, tumor growth, and metastasis. We hope that an increasing understanding of the specific molecular mechanisms associating inflammation, TME, and cancer will subsequently result in a better understanding of how these cancers to subvert the immune system and overcome response our wide array of chemoradiotherapeutic agents.
Introduction The relationship between chronic inflammation and cancer is not new and can be traced back to the days of Rudolf Virchow (Coussens and Werb 2002; Balkwill and Mantovani 2001). However, it is only in recent years that this link has seen a dynamic renaissance, fueled by increasing animal and human studies and strengthened by epidemiology (Balkwill et al. 2005; Mager et al. 2005; Mager 2006). As part of this strengthening, various pathogenic causes for chronic inflammation have been proposed as players in multiple tumor systems for malignant initiation, progression, and metastatic spread (Kuper et al. 2000; Fidler 1995; Karin and Greten 2005; Hold and El-Omar 2008); despite these clear advances, the specific mechanisms regulating their functional association are not fully understood. Examples of specific chronic inflammatory agents of infectious nature leading to cancer include Helicobacter pylori colonization in the stomach, which increases the risk of gastric cancer and mucosal lymphoma. Other examples of infectious agents leading to chronic inflammation and eventually cancer are Hepatitis B and C, Clonorchis sinensis, and Schistosoma mansoni among others. In certain pathologies, inflammatory such as pelvic inflammatory disease, Barretts’ esophagus, and chronic pancreatitis, the etiology of inflammation is noninfectious or idiopathic (Quante and Wang 2008), and this inflammation is responsible for increasing the risk of developing cancer. It is also well established that chronic inflammation associated with autoimmune inflammatory bowel disease may ultimately lead to the development of carcinoma of the colon. Interestingly, while defining the mechanisms regulating inflammation in malignancy it became evident that various oxidizing enzymes involved in DNA damage, such as nicotinamide adenine dinucleotide phosphate-oxidase (NADPH) and inducible nitric oxide synthase (iNOS), are also responsible for the malignant initiation of epithelial cells within the TME. In chronic inflammation, specifically where a direct effect of the pathogen on neoplastic transformation of epithelial cells has been shown but overwhelming evidence points to a more generalized role of the resultant inflammatory process as the primary mediator of the carcinogenic process, both eradication of infection and anti-inflammatory therapies (e.g., COX-2 inhibitors) have been successful in preventing progression and even curing some of these cancers (Koehne and Dubois 2004; Flossmann and Rothwell 2007; Chan et al. 2007). Recent research has demonstrated that not all inflammation results in cancer; for example, the adaptive immune response elicited by some cancers exerts a potentially effective role in immune surveillance (Raulet 2004). A fundamental question in the field is the role that specific immune and inflammatory cell populations play in the neoplastic
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
411
initiation of the epithelium. In general terms, it can be suggested that either the tumor and its microenvironment actively affects the immune response so that it becomes pro-tumorigenic [where, for example, cytokines or reactive oxygen species (ROS) maximize and propagate tumor growth] or it remains as sustained chronic inflammation exerting an active and primary role in initiating and transforming epithelial stem cells into tumor cells; there is increasing evidence that tumor-derived signals appear to play a role in recruiting bone marrow derived cells (BMDC) where they can give rise to stroma and perhaps tumor cells that can contribute to tumor progression within the context of chronic inflammation.
Acute Versus Chronic Inflammation in the Context of the Tumor Microenvironment Acute inflammation often becomes chronic in nature especially when the causing inflammatory agent or agents persist; however, there are circumstances where chronic inflammation takes place right from the start as in the case of gastric cancer and infection with H. pylori. In contrast to the faster onset and the mostly vascular changes associated with acute inflammation, chronic inflammation is characterized by significant leukocyte infiltration in the affected tissues. In contrast to chronic inflammation, both the innate and the adaptive immune systems play a role in acute inflammation. They are responsible for two functionally reversible and opposing functions of the acute inflammatory response: (1) a pro-inflammatory response (mainly consisting of organized damaged cell death) and (2) a post-inflammatory or anti-inflammatory process (consisting of tissue repair and ultimately a resolution of inflammation). Subsequent to exposure to inflammatory agents, a series of complex inflammatory paraphernalia explodes through precise highly regulated crosstalk mechanisms between the innate and adaptive immune systems in an organ specific manner followed by expression of further death signals. Furthermore, such a crosstalk ensures the availability of the required amounts of inflammatory mediators such as cytokines/chemokines and their receptors, and antibodies responsible for recruiting, attracting, and activating other inflammatory cells and mediator cells to the site of inflammation via the activation and engagement of vascular cell components. The principal physiological purposes of the acute inflammatory response are dual (1) Obliterate any exogenous cells (organisms or harmful materials) or any endogenous harmful cells or cell components (e.g., cancerous cells, cells with aberrant nuclei, protein machinery, etc.) that are created during the inflammatory process. (2) Neutralize and eliminate the various apoptotic signals generated in the inflamed tissue after their signal has become redundant as well as eliminate the causative toxins. This is particularly important so that cellular and tissue repair can commence upon successful resolution of the initial acute inflammatory response. It is important to note that not all acute inflammatory conditions (e.g., minor exposure to chemical irritation) that can elicit instant response from innate immune cells result in extensive involvement of adaptive immunity. On the other hand, chronic
412
A. Yagui-Beltrán et al.
exposure to allergens, other environmental factors, or infectious agents often may result in the precise activation of innate and adaptive immunity pathways that produce suitable responses for defending the host tissue (Lodoen and Lanier 2006; Valent et al. 2001; Bonasio and von Andrian 2006; Drayton et al. 2006; Kabelitz and Medzhitov 2007). Perhaps of significance to human disease is the fact that the various dynamic changes in immune cell function associated to unresolved inflammation can result in the development of a myriad of inflammatory pathologies affecting the majority of human organs, and those can vary from local allergic reactions to major life threatening vasculitis, syndromes, and from pre-neoplastic syndromes to full blown cancers (Balkwill and Mantovani 2001; Dvorak 1986; Ferrantini et al. 2008; Dalgleish and O’Byrne 2006; de Visser et al. 2006; Swann et al. 2008; Wang et al. 2008). Although the precise molecular mechanisms governing the often complex multifaceted processes initiating, perpetuating, and linking the innate and adaptive immune systems during an acute inflammatory process are yet to be fully elucidated, an outline of the most important known molecular and cellular phenomena of acute inflammation will serve useful in subsequent understanding of the pleiotropic functional significance (tumorigenic versus tumoricidal) of the various innate and adaptive immune cell subpopulations in subsequent persistent or chronic inflammation specifically in the context of carcinogenesis. 1. The pro-inflammatory response phase: The initial phase of acute inflammation is also known as the programmed cell death or the “apoptosis” phase of the acute inflammatory phase. Apoptotic cellular functions are specifically engaged subsequent to recognition and encountering of antigens, mitogens, or infective agents and infected host cells, which result in presentation of processed antigen/pathogen to effector T or B cells for additional cell-mediated and/or humoral responses. The characteristics of the injurious stimuli – pathogens, neoplastic cells with aberrant DNA and/or proteins – as well as the target organs and associated tissues (immune privileged, epithelial, lymphoid, or mucosal tissues) each with the ability to initiate specialized inflammatory cells, such as, for example, natural killer cells (NKs), macrophages (MFs), polymorphonuclear neutrophils (PMNs), dendritic cells (DCs), mast cells (MCs), basophils or eosinophils as well as T or B cells become stimulated and specifically differentiated in order to produce the required amounts of proinflammatory mediators. Acute pro-inflammatory apoptotic mediators may be produced and secreted from the various cytosolic, endoplasmic reticulum, nuclear or mitochondrial compartments of the activated cells. An example of a pro-inflammatory response phase process is demonstrated in the respiratory tract, subsequent to exposure of the lungs to certain pathogens or certain pathogenically secreted compounds such as bacterial lipopolysaccharides (LPS). A cascade of signals in the membranes of DCs are induced, which in turn instigate adequate pro-inflammatory responses such as subsequent expression of pathogen (LPS)-specific receptor molecules, [toll-like-receptors (TLRs) TLR-1–11 to additionally promote the production of mediators that are also termed alarmins (danger molecules) or death factors (DFs)]. Pro-inflammatory signals or DFs that are activated in response to TLRs include TNF-a and TNF receptor protein (TNF-R), cell surface antigen molecules and their respective receptors and/or recognition molecules [e.g., TNF-related
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
413
apoptosis inducing ligand (TRAIL), TCR, TNF-R/TNF-SF5, IgE-FceR, TGF-R, CCL13/monocyte chemoattractant protein 4 (mcp-4), phosphatase and tensin homolog (PTEN), B cell leukemia/lymphoma associated protein 2-associated X protein (Bax), bcl2-associated death promoter (BAD), Fas-associated via death domain (FADD), or CD2, CD11, CD18, CD25, CD50, CD54, CD63, CD69, CD88]. Apoptotic responses of the pro-inflammatory response phase also include expression of a myriad of preformed cytokines, chemokines, and associated factors such as macrophage inflammatory protein 1 (MIP-1), eosinophil-derived neurotoxin (EDN), eosinophil chemotactic factor of anaphylaxis (ECFA), CCL11/eotaxin-1, MPIF-2/eotaxin-2, TGF-a, b, ILs (e.g., IL-1, IL-3, IL-4, IL-6, IL-7), (capsases 1–10; extracellular matrix proteases or membrane metalloproteases (MMPs)). Importantly, many newly synthesized vasoactive mediators, [such as histamine (his)], heparin (hep)], and various enzymes (like elastase, tryptase, chymase, and perforin) MHC class I/II molecules and their co-stimulatory molecules (e.g., DC28, DC86), and chemokine receptor family CCRs (e.g., CX3CR1, ILBrA, XCR1/CCXCR1), are also synthesized. Moreover, during the apoptotic pro-inflammatory phase a number of oxidants [e.g., ROS, reactive nitrogen species (RNS), superoxide (SO), hydroxyl ions (OH–), hydrogen peroxide (H2O2), or hydroclorous acid (HOCl)] are generated from the myriad activated inflammatory MCs, DCs, MFs, NKs, basophils, PMNs, eosinophils, T or B cells as well as vascular cells (Bonasio and von Andrian 2006; Kabelitz and Medzhitov 2007; Wang et al. 2008; Banchereau and Steinman 1998; Fukata and Abreu 2008; Conroy et al. 2008; Wilson and Villadangos 2004). It is important to note that the programmed cell death process of this part of the acute inflammatory response is also associated with the mobilization of various cell membrane enzymes and/or enzyme components, such as the case of phospholipases A, C, or D, of the metabolism of arachidonic acid, activation of the COX and lipooxygenase pathways and the synthesis of a multitude of PGs and leukotrienes, of activation of sphingosine kinase binding proteins and calcium mobilization resulting in increased cellular energy NADP+, of and increased mitochondrial oxidative metabolism and the production of oxidants as already indicated. Mast cell-derived programmed cell death often requires antibody-induced activation of blood platelets and complement cascades to assist cell lysis as well as the expression of receptor molecules essential for B cell-derived antigen-specific antibody synthesis, e.g. IgG, IgE, IgM, or IgA, and antibody-receptor aggregation (Lodoen and Lanier 2006; Valent et al. 2001; Bonasio and von Andrian 2006; Drayton et al. 2006; Kabelitz and Medzhitov 2007; Banchereau and Steinman 1998; Conroy et al. 2008; Wilson and Villadangos 2004; Yano et al. 1997; Juliet et al. 2003; Khatami et al. 1984; Fukata et al. 2005; Selsing 2006; Caro and Cederbaum 2006; Ibrahim et al. 2004; Peerschke et al. 2006; Monteseirin et al. 2004; Nyakern et al. 2006; Fischetti and Tedesco 2006). In conclusion, during the programmed cell death or the “apoptosis” part of the acute inflammatory phase, DFs are orchestrated in a timely spatial-temporal manner that will ensure that subsequent to the recognition of the causing pathogen, its correct processing will occur, and that through the deployment of the adequate molecular and cellular mechanisms its presentation to the relevant innate immune cells or their counterparts adaptive immune subpopulations T or B lymphocytes for additional cellular responses will ensue.
414
A. Yagui-Beltrán et al.
2. The resolution phase of acute inflammation. The termination or resolution part of the acute inflammatory response ensures that any redundant toxicity remaining in target organs after the relevant cascade of apoptotic danger signals has exerted their function (to resolve and terminate the initial inflammatory response, while allowing normalization of functional relationship between the host and target organ so that tissue repair can also commence) is efficiently nullified by postinflammatory processes facilitated by a myriad of recruited anti-inflammatory and anti-apoptotic mediators. The mediators are specific to the type of stimuli, degree of tissue injury, and the quantity of death signals produced in the host during apoptosis. Examples of mediators involved in wound healing include growth factors, anti-apoptotic mediators including lymphocyte NF-kB and mitogen-activated kinases (MAPKs). Other anti-inflammatory mediators that are generated in this phase include IFNs: IFN-a, IFN-g: ILs (IL-5, IL-8, IL-10, IL-13) B cell leukemia/lymphoma associate protein 2 (BCL-2), PKC, Jun kinase (JNK), IkB kinase (IKK), membrane phosphotidylinositide protein kinases (PI3K)/Akt pathways; ataxia telangiectasia mutated homolog (ATM); fibroblast growth factor (FGF); platelet-derived growth factor (PDGF), nuclear expression of Ataxia Telangiectasia, and a variety of inhibitors of toxins, antioxidants, and reducing enzymes [e.g., superoxide dismutase (SOD), NADP+ reductase, catalase, thioredoxin reductase (Trx)]; NO; adaptor molecules and protein complexes [e.g., toll-interleukin 1 receptor domain containing adaptor protein (TRAP), translocation-associated membrane protein (TRAM), MyD88, interleukin-1 receptor-associated kinases (IRAKs), TNF receptor-associated factors (TRAFs)] (Lodoen and Lanier 2006; Ferrantini et al. 2008; de Visser et al. 2006; Wang et al. 2008; Banchereau and Steinman 1998; Yano et al. 1997; Juliet et al. 2003; Khatami et al. 1984; Selsing 2006; Caro and Cederbaum 2006; Ibrahim et al. 2004; Peerschke et al. 2006; Monteseirin et al. 2004; Nyakern et al. 2006; Fischetti and Tedesco 2006; Perez et al. 2006; Mantovani et al. 2008; Snook et al. 2006; Ledeen and Wu 2006; Hanukoglu 2006; Berger 2006; Ii et al. 2006; Stenmark et al. 2006).
Malignant Epithelial Initiation Within a Pro-inflammatory Milieu The process of tumor development is associated with the growth and expansion of not only tumor cells but also stroma, new blood vessels, and a myriad of infiltrating inflammatory cells, and it is the interaction between these different cell populations that helps disseminate tumor development and spread. Traditionally for many decades, studies supported and focused around the dogma that acute inflammation through its deployment of various aspects of the immune system exerted a major role to mainly eliminate cancer (Nauts and McLaren 1990). Nevertheless, acute inflammatory processes are by definition self-limiting, and often develop into chronic inflammation, with links to cancer initiation as explained above (Kuper
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
415
et al. 2000; Fidler 1995; Karin and Greten 2005; Hold and El-Omar 2008). Immune responses elicited against cancerous cells, despite some similitude to those deployed against pathogenic organisms are generally weak and not adequate to eradicate malignant initiated epithelial cells (Whiteside 2008). A plausible explanation is that the initial signals originated by the malignant epithelial cells are not recognized as equivalently strong as those danger signals emanating from exogenous pathogens (Gallucci and Matzinger 2001). Other explanation would be that the host harboring the cancerous cells subsequently downregulate the perceived immune self– responses, or alternatively that certain components of the TME act as barriers blocking access of the immune system that would normally function to suppress or eliminate the cancer cells (Zhang 2008). More recently, the NF-kB and STAT3 pathways have gained significant importance as major protagonists playing important roles ensuring the persistence of chronic inflammation through the regulation of pro-inflammatory cytokines and as pivotal mediators of the pro-inflammatory milieu that facilitates the growth and malignant transformation of initiated epithelial cells (Karin and Greten 2005; Karin 2006; Rius et al. 2008).
Tumor Immune Evasion and Progression Within Sustained Chronic Inflammation Studies have shown that in order for cancer development to occur successfully, suppression of T cell as well as humoral immune responses must occur. Various factors, including the often heterogeneous nature of tumors, as well as their still poorly understood ability to evade and suppress the attack of the host’s immunity, have been postulated as likely contributors to the inadequacy of an antitumor response. Autologous effector T cells that normally function to suppress or eliminate cancer cells are downregulated by regulatory T cells (Tregs) that infiltrate and accumulate the TME. These regulatory T cells are often activated or recruited by DCs infiltrating the TME (Bacchetta et al. 2007). The other population of cells that have been suggested to play an important role in suppression of tumor infiltrating lymphocytes are myeloid-derived suppressor cells or MDSCs (in the case of mouse models, the CD11b+/Gr-1+ myeloid progenitor cell populations) that are present in increased number both in the cancerous lesions and peripherally in lymph nodes and the spleens of the tumor bearing animals (Talmadge et al. 2007; Murdoch et al. 2008). Despite the abundant ongoing MDSC research across Europe and the United States; it is important to emphasise that the nature and functional significance of MDSCs in the field of tumor immunology that remains rather controversial and undefined. For example, MDSCs were originally believed to exert an immune suppressive function; however, newer research in the field proposes a pro-inflammatory as well as a pro-angiogenic function for these immune cell populations (Murdoch et al. 2008). It seems therefore that tumor-associated inflammation exerts both a pro-inflammatory and an anti-inflammatory set of signals; and when a polarization of the balance favors a pro-inflammatory milieu malignant initiated epithelial cells are then able to
416
A. Yagui-Beltrán et al.
evade the various mechanisms of immune surveillance, and are then able to grow and mature into cancers while developing their own blood vessels, eventually breaking the basement membrane and spreading into distal sites (metastasis).
Tumor Progression, Metastatic Potential, and Inflammation Recent research by various groups has indicated that inflammation also exerts an imperative function in the ability of cancer cells to spread and metastasize; this is thought to be due to the capacity of initiated epithelial cancer cells, with the acquired ability to spread, to express chemokine receptors (Muller et al. 2001). The expression of specific chemokine receptors is an essential part of the machinery necessary to support the signals elicited by paracrine secretion of pro-inflammatory cytokines, such as IL1-b, IL-6, and TNF-a as well as the signals elicited through the production of other cytokines produced in an autocrine manner (Kulbe et al. 2007). An example of how the pro-inflammatory milieu can influence epithelial malignant initiation and metastatic spread is demonstrated by a knockdown study of TNF-a by Burger et al. in 2006 (Burger and Kipps 2006); in their experiments, they were able to corroborate that mesenchymal or marrow-derived stromal cells, constituting a large proportion of the noncancer cell population within the TME, and secreting the chemokine stromal cell-derived factor-1 (SDF-1/CXCL12), attract cancer cells (through its receptor CXCR4, expressed by both hematopoietic and nonhematopoietic tumor cells). CXCR4 promotes tumor progression by direct and indirect mechanisms. CXCR4 is essential for metastatic spread to organs where CXCL12 is expressed, allowing tumor cells to access cellular niches, such as the marrow that favor tumor-cell survival and growth. Stromal-derived CXCL12 itself can stimulate survival and growth of neoplastic cells in a paracrine fashion. Moreover, CXCL12 can promote tumor angiogenesis by attracting endothelial cells to the TME. Burger and colleagues’ knockdown of TNF-a resulted in decreased cellular CXCR4 and CXCL12 expression and decreased ability of the tumor to spread, highlighting that at least in preclinical tumor models CXCR4 antagonists may have antitumor activity in patients with various malignancies. This may be potentially exploited to target the TME therapeutically in the treatment of neoplastic disease (Burger and Kipps 2006). Apart the specific cytokines released by the tumor cells from the various other elements of the microenvironment, from the stroma, endothelial cells, and from noncancer cells, studies have importantly shown that for cancer development to occur certain infiltrating leukocytes and particular tumor-associated macrophages or TAMs must exert their function. TAMs have been the focus of research of many groups across the globe in the last two decades with a multitude of data on their functional significance in cancer development. TAMs have been traditionally subdivided into pro-inflammatory M1 and anti-inflammatory M2, and their
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
417
polarization into either in the tumor dictates the behavior of the neoplasm, its ability to metastasize and its degree of malignancy and this has been corroborated in an organ specific manner (Andreu et al. 2010; DeNardo et al. 2009; Lin et al. 2001; Mantovani et al. 2008).
Soluble Mediators of the Immune Response in the Pro-inflammatory Milieu Responsible for Cancer Development as well as Maintenance of Chronic Inflammation Various animal and human studies have showed a dynamic interplay between initiated epithelial cells and the myriad components of the TME necessary for the development of cancer. Undoubtedly, the fields of tumor immunology and the TME have been rapidly expanding scientifically while gaining popularity as an attractive integral model that serves to explain the process of carcinogenesis and more importantly as a model system to design novel therapies targeting the various components of the TME in a specific manner, affecting the various hallmarks of cancer progression. Inevitably, such an expansion in the field means that many of the findings are not devoid of controversy and require further ongoing corroboration before scientific robustness is achieved. The following are the best-known inflammatory mediators exerting important roles within the TME initiating epithelial cells promoting their carcinogenesis.
Oxidative Stress Species and Their Functional Significance in Cancer Development A major source of endogenous reactive oxygen (RO) and RNS: O2-, NO, H2O2, OH, ONOO-, HOCl are innate immune cells. As explained above, during the acute phase of inflammation, the release of these compounds plays a beneficial role in the elimination of pathogens (Hussain and Harris 2007). However, a persistence of inflammation, as is the case in chronic inflammation, results in sustained exposure of epithelial cells to oxidative stress – an exposure that is eventually harmful, as it can lead to abnormal genomic alterations (DNA strand breaks and base modifications), lipid and protein peroxidation and aberrant activation of signal transduction by posttranslational modification (RNS is a mediator of MAPK and AP-1 signal transduction. NO can induce posttranslation modification of p53 via ATM or ATR, or specific induction of anti-apoptosis genes (AIF) (Wu 2006; Azad et al. 2008; Hussain et al. 2000). These mechanisms link ROS and RNS production, and hence the innate immune system, to the ability of tumors to induce angiogenesis, to proliferate, and to promote distant metastasis. Additionally, oxidative stress has also been
418
A. Yagui-Beltrán et al.
associated with impaired DNA methyltransferase activity (DNMT-1) or MBP binding, and therefore provides a link to a myriad of epigenetic alterations (Shahrzad et al. 2007; Lim et al. 2008).
The Role of Matrix Remodeling Proteases in the Tumor Microenvironment Matrix metalloproteinases are crucial in establishing crosstalk and co-regulation among the various cellular elements of the TME. In a nutshell, the majority of MMPs are expressed by stromal and immune cells in the pro-inflammatory milieu in the TME (MMP7 is an exception and is predominantly expressed by epithelial cells). For the majority of MMPs, it logically follows that in cancer, a higher MMP expression tends to correlate with poorer prognosis and worse patient outcome. Research in various study models has demonstrated that extracellular proteases in the TME, as well as having a function of remodeling of certain extracellular matrix (ECM) components, can also modulate a range of growth factors, cytokines and chemokines. In this respect, MMP9 has been shown to be important in the regulation of malignant epithelial proliferation, specifically in angiogenesis through the processing of pro-growth factors such as vascular endothelial growth factor (VEGF). Interestingly, studies have shown that MMP9 expression correlates with a more invasive and aggressive phenotype in breast and skin cancers (van ’t Veer et al. 2002), which may perhaps be related to its ability to promote vasculogenesis. Anh et al. in a study in 2008 demonstrated that MMP9 may be necessary for the recruitment of bone marrow derived progenitor cells (MDSCs), such as endothelial progenitor cells (Ahn and Brown 2008); they showed that intra-tumoral expression of MMP9 by MDSCs is sufficient to restore angiogenesis in irradiated mice (Ahn and Brown 2008). MMP7 is able to process RANKL, positively influencing metastasis through tumorinduced osteolysis (Ahn and Brown 2008). McCaig et al. showed in a study in 2006 that infection with H. pylori directly induced MMP7 in gastric cancer epithelial cells; they also showed that MMP-7 could act as an epithelial-derived signal increasing the bioavailability of IGF-II released from myofibroblasts and because IGF-II could function on both stromal and epithelial cells, their findings suggested that increased MMP-7 expression contributes to redefining the niche occupied by dividing cells and leading to hyperproliferation occurring in H. pylori infection (McCaig et al. 2006). The field of MMPs is regaining momentum despite the fact that their targeting has been often complicated because of their parallel functions in both normal tissue homeostasis and cancer development and their varying functions depending on tumor biology and stage; this is particularly true as the gap in the understanding of the molecular mechanisms governing the function of myriad components of the TME and their dynamic interplay decrease.
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
419
Functional Significance of Specific Transcription Factors and Primary Inflammatory Cytokines In the wide array of molecular factors in the pro-inflammatory milieu promoting cancer development, a number of principal or players have emerged through the consistent work of various research groups across the world. These include transcription factors such as nuclear factor-kappaB (NF-kB), signal transducer activator of transcription-3 (Stat3), and primary inflammatory cytokines such as IL-1b, IL-6, IL-23, and TNF-a (Karin 2006; Yu et al. 2007; Voronov et al. 2003; Langowski et al. 2006). NF-kB is a pivotal orchestrator of innate immunity/inflammation, and aberrant NF-kB regulation has been observed in a number of malignancies (Karin 2006). In both tumor and infiltrating inflammatory cells, NF-kB is activated downstream of the toll-like receptor (TLR)-MyD88 pathway (sensing pathogenic microbes and tissue damage) and of the inflammatory cytokines TNF-a and IL-1b. In addition, NF-kB activation can be the result of cell-autonomous genetic alterations (amplification, mutations or deletions) in cancer cells. Interestingly, NF-kB can be activated in response to hypoxia though to a lesser extent than hypoxia inducible factor (HIF)-1a (Balkwill 2009; Carbia-Nagashima et al. 2007; Taylor 2008). Accumulating evidence suggests that intersections and compensatory pathways may exist between the NF-kB and HIF-1a systems linking innate immunity to the hypoxic response. NF-kB induces the expression of inflammatory cytokines, adhesion molecules, key enzymes in the prostaglandin synthase pathway (COX-2), NO synthase, and angiogenic factors. Furthermore, by inducing anti-apoptotic genes (e.g., Bcl2), it promotes survival in tumor cells and in epithelial cells targeted by carcinogens. Research has unequivocally demonstrated that NF-kB is involved in malignant epithelial initiation and progression in tissues where sustained chronic inflammation has been classically described, for example, the gastrointestinal tract and the hepatobiliary tract (Karin 2006; Greten et al. 2004; Pikarsky et al. 2004). Similarly, as when discussing the ability of IL-6 to polarize T cell differentiation toward a pro-inflammatory phenotype, precise inactivation of NF-kB in tumor-infiltrating leukocytes, by a strategy targeting Ik-Bkinase beta, inhibited colitis-associated cancer, thus providing genetic proof for the specific function of NF-kB and infiltrating leukocytes in intestinal cancer development (Greten et al. 2004). Various inhibitors functioning at different levels of the NF-kB pathway ensure its adequate control. For example, Tir8 is an orphan member of the IL-1R family that is highly expressed in the mucosa of the gastrointestinal tract. In a murine model of intestinal cancer, Tir8-deficient mice, in response to dextran sulfate sodium salt and azoxymethane administration, demonstrated exaggerated inflammation with an associated increase for colon cancer risk compared to normal animals. Furthermore, this was associated with local production of PG E2, proinflammatory cytokines (IL-1, IL-6), and chemokines (KC/CXC, JE/CCL2 and CCL3) (Garlanda et al. 2007; Xiao et al. 2007). Thus, the lack of a checkpoint (Tir8) of NF-kB activation leads to increased carcinogenesis in the gastrointestinal tract, highlighting once more the association between chronic inflammation and
420
A. Yagui-Beltrán et al.
malignant initiation of epithelial cells. As indicated already, there is a role for STAT3 as a point of convergence and also quite importantly for perpetuation of several oncogenic signaling pathways (Yu et al. 2007; Lee et al. 2009). This transcription factor is constitutively activated both in malignant cells and also in immune cells and plays a role in carcinogenesis as well as in tumor immune evasion by halting DCs maturation (Yu et al. 2007; Kortylewski et al. 2005; Becker et al. 2004). Research on colon cancer revealed that STAT3 is a major controller of cell proliferation and survival, regulating the expression of c-Myc, Mcl-1, Cyclin D, and Bcl-2 (Becker et al. 2004). Genetic studies in animal and human experiments have demonstrated that cytokines play a key role in linking chronic inflammation and cancer. They are activated by both inflammatory and tumor cells and their effect is similarly important for both sustaining chronic inflammation, promoting the progression of malignant epithelial cells, and also in inhibiting immune-mediated tumor surveillance. Although, conveniently, cytokines have been divided into either pro-inflammatory – IL-1, 6, 8, 11, 12, 18, 23, IFN-g, TNF-a, and MIF (e.g., IL-1b overexpression alone, in the absence of Helicobacter infection, is sufficient to cause gastric cancer (Tu et al. 2008) – or anti-inflammatory (IL-4, 10, IFN-a and b, TGF-b), compounding research is showing that an increasing number of these molecules have dichotomous roles. TGF-b is an example of a pleiotropic cytokine that can function to suppress cancer in certain circumstances, while having the capacity to promote cancer progression through the manipulation of definite microenvironment pathways during others. For example, complete deletion of TGF-b has been linked with increased tumor incidence; nevertheless most cancers are also able to progress and spread in the presence of high levels of TGF-b (Gonda et al. 2009). TGF-b has also been associated with the recruitment of MDSCs into the TME, suppression of cytotoxic T lymphocytes (CTLS) CD8+ T cell and infiltration of NK, i.e., facilitating an inflammatory milieu that is more favorable to tumor growth and development. IL-6 has been linked to chronic inflammation and tumor development in the colon and stomach mainly through an aberrant activation of STAT 1 and 3 (Bromberg and Wang 2009). Like IL-6, TGF-b polarizes T cell differentiation toward a pro-inflammatory phenotype. This year, two separate research groups have shown that IL-6 deficient mice were not able to develop colitis-associated carcinoma in a DSS/AOM mouse model and that the IL-6 effect was regulated by STAT-3 in intestinal epithelial cells (Bollrath et al. 2009; Grivennikov et al. 2009). Interestingly, in this murine model bone marrow derived myeloid cells were shown to be the source of IL-6. Moreover, IL-6 in the cancers was demonstrated to be effective not only on intestinal epithelial cells, but also its effect was important in other components of the microenvironmental milieu, leading for example to activation of DCs and T helper cells. The overall result of this dynamic interplay between IL-6, malignant epithelial cells, and other components of the microenvironment is thought to be the perseverance of cytokine production and secretion and further propagation of the chronic inflammatory state exacerbated by a pro-tumorigenic cytokine soup (Bromberg and Wang 2009). This is an example of various integrative models advocating that a proinflammatory cytokine, IL-1b, exerts an essential role in the recruitment of bone
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
421
marrow-derived myeloid cells, which in turn function as a source for IL-6 production. Moreover, IL-6 induction of STAT3 in epithelial cells may result in further malignant epithelial transformation. It is important to highlight the potential importance that inhibitors of these cytokines and their signal transducers (STAT-3/JAK) may have as novel anticancer targeted therapies. We have also indicated that IL-6, a multifunctional cytokine with growthpromoting and anti-apoptotic activity, is considered to be a pivotal effector molecule of the NF-kB pathway (Lin and Karin 2007; Naugler and Karin 2008). We also highlighted some of the recent research on the NF-kB–IL-6–STAT3 cascade, where IL-6 was found to be produced by myeloid cells and to be a key to the development of gastrointestinal cancer. IL-6 protects normal and premalignant intestinal epithelial cells from apoptosis and favors the proliferation of tumor-initiating cells (Bollrath et al. 2009; Grivennikov et al. 2009). Interestingly, STAT3 also regulates the balance between IL-12 and IL-23 in the TME, and consequently the polarization of T-helper subsets (Kortylewski et al. 2009). New mechanistic insights into links between IL-6 and cancer have been shown in a mouse model hepatocellular carcinoma, where in male animals IL-6 was shown to promote liver inflammation, injury, compensatory cell proliferation, and cancer development. In females, estrogen steroid hormones had a cancer protective effect through the inhibition of IL-6 production (Naugler et al. 2007; Rakoff-Nahoum and Medzhitov 2007). IL-1 is a pro-inflammatory cytokine with a well-described ability to promote cancer development and spread by influencing important pro-cancer mediators in the pro-inflammatory milieu (Mantovani et al. 2008; Giavazzi et al. 1990; Luo et al. 2007). The IL-1R1 gene-targeted mouse model system have provided good evidence for the pro-tumor potential of IL-1 (Dinarello 2006). Particularly, in models of chemical carcinogen-induced tumors, IL-1B secreted by malignant cells or infiltrating leukocytes contributes to properties of malignancy such as angiogenesis, immune suppression increased tumor adhesiveness and invasion, whereas IL-1ra negatively controls these processes (Krelin et al. 2007). In diethyl-nitrosamineinduced hepatocarcinoma, the unique membrane-associated form of IL-1a acts as protumorigenic mediator; diethyl-nitrosamine-induced hepatocyte death results in the release of IL-1a and activation of IL-1R signaling, leading to IL-6 induction and compensatory proliferation, critical for hepatocarcinogenesis (Sakurai et al. 2008). The importance of IL-1a is also worth mentioning in 3-methylcholantrene-induced fibrosarcoma for its efficiency in activating antitumor innate and specific immune responses by acting as a focused adjuvant through binding to IL-1RI on cells deputed to immune surveillance (Marhaba et al. 2008; Elkabets et al. 2009). This “danger signal” has been successfully collected artificially for mounting antitumor immunity from dying cells (Chen et al. 2007). Stomach-specific expression of human IL-1b in transgenic mice lead to spontaneous gastric inflammation and cancer that correlated with early recruitment of MDSCs (Tu et al. 2008). Last, but not least, recent studies have demonstrated a relationship between steroid hormones, IL-1, and cancer; in carcinoma of the prostate, androgen-dependent tumor sensitivity to hormonal stimulation is regulated by selective androgen receptor modulators. The
422
A. Yagui-Beltrán et al.
inflammatory cytokine IL-1 produced by macrophages in the TME converts selective androgen receptor modulator from inhibitors to stimulators, thus inducing resistance to hormonal therapies (Zhu et al. 2006) and highlighting the complexity and potential for novel cancer therapy that a good understanding of the functional significance of the pro-inflammatory milieu and the microenvironment entails. TNF is a major cytokine and player in the pro-inflammatory milieu exerting a cancer-inducing role on malignant initiated epithelial cells. TNF, which originally was identified as a cytokine inducing hemorrhagic necrosis in tumors, was soon after discovered to also have pro-tumoral functions. The finding that TNF-deficient mice are protected from skin cancers offered genetic proof associating TNFmediated inflammation and cancer (Balkwill 2009; Moore et al. 1999). TNF has been shown to be versatile in how it is able to enhance tumor development through a number of mechanisms: it can successfully facilitate the infiltration of tumor promoting leukocytes; it promotes angiogenesis, metastatic spread, and also assists epithelial-to-mesenchymal transition or EMT (e.g., through the activation of the Wnt/beta-catenin pathway subsequent to TNF signaling from activated gastric cancer TAMs), processes that have the overall result of tumor enhancement and growth (Kulbe et al. 2007; Balkwill 2009; Popivanova et al. 2008; Oguma et al. 2008). Not to ignore is the immune suppressive functions of some members of the TNF family. An example is that of the decoy receptor-3 that has been linked to the downregulation of major histocompatibility complex class II in TAM (Chang et al. 2008). Like in the case of other components of the pro-inflammatory milieu in the TME, a better understanding of the molecular mechanisms underlying this cytokine and its dynamic interplay with others is serving as a platform for the design of novel anticancer therapies. Indeed, there have already been some phase I and II clinical cancer studies with TNF antagonists to assess their efficacy and safety in a variety of neoplasms (Harrison et al. 2007; Brown et al. 2008).
Functional Significance of Myeloid Cell Recruitment Within Tumors It is estimated that in any single tumor, half of the component cells are tumor epithelial cells and the other half is composed of noncancerous epithelial cells and a stroma containing fibroblasts, vessels, and leukocytes, also often referred to as the TME (Allavena et al. 2008). In the majority of published studies, TAMs seem to be the main subset of leukocytes driving an amplification of the inflammatory response in the tumor milieu; although recent studies are showing that the predominating infiltrating immune cell subpopulations may be other than TAMs and their presence may be organ specific (Jablons and Coussens, unpublished). In fact, part from TAMs, also mast cells, neutrophils, and even effectors of the adaptive immunity (especially in the form of antibodies) may activate inflammatory reactions that promote cancer progression (de Visser et al. 2006; Galli et al. 2008). Chemokines have long been associated with the recruitment of TAM in tumors (e.g., CCL2 and
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
423
CCL5) (Mantovani et al. 2008; Balkwill 2004). For their phenotypic and functional properties, TAM resembles M2-polarized macrophages, though there are some specific characteristics (Mantovani et al. 2002; Sica et al. 2006). In the majority of published animal and human studies, accumulation of TAM correlates with the angiogenic switch, poorer patient outcome, and reduced survival (Balkwill et al. 2005; Mantovani et al. 2008; Bingle et al. 2002; Pollard 2004). TAMs have been shown to be pro-tumorigenic because as they preferentially accumulate in hypoxic regions of tumors they facilitate the release of pro-inflammatory growth factors, cytokines, and matrix-degrading enzymes, such as MMPs (described above) (Wyckoff et al. 2007; Mantovani et al. 2006; Condeelis and Pollard 2006; Yang et al. 2008) while enriching the pro-tumorigenic niche with a florid array of angiogenic factors: VEGF, PDGF, FGF, and CXCL8 (Murdoch et al. 2008; Ahn and Brown 2008; Du et al. 2008; Seandel et al. 2008; Dineen et al. 2008; Kusmartsev et al. 2008; Duluc et al. 2007; Morandi et al. 2007; Robinson-Smith et al. 2007; Stearman et al. 2008; Lewis and Pollard 2006) facilitating the growth of further tumor cells while perpetuating an environment of sustained chronic inflammation with everything that entails in terms of malignancy and metastasis. Proving some of the mechanisms behind secondary localization and seeding of cancer are some recent studies showing that VEGF1R+ hematopoietic cells home to tumor-specific pre-metastatic sites that favors secondary localization of cancer (monocytes express VEGF receptors and VEGF being a recognized chemoattractant of myeloid cells in cancers) (Fischer et al. 2007; Kaplan et al. 2005; Witz 2008). Chemokines (e.g., CXCL5 and CXCL12) are also involved in the attraction of MDSCs (Sawanobori et al. 2008; Sica and Bronte 2007). MDSCs, like TAM, are important effectors in tumor angiogenesis (Gonda et al. 2009; Bromberg and Wang 2009) and Gr+Mac1+ cells, presumably MDSC, have been shown to mediate resistance to anti-angiogenic therapy (Shojaei et al. 2007). Subsequent to the malignant initiation of epithelial cells within the rich proinflammatory pro-cancer milieu, further sustainment of chronic inflammation and cancer development and spread is principally mediated by the host’s failure to mount a protective antitumor immune response. TAM and MDSCs express an of immunosuppressive armamentum responsible for this. As well as inhibiting the CTL response, which in physiological circumstances is mediated by CD8 T cell activation through the expression of NOS2 and Arg1, MDSCs seem to promote the expansion of CD4+FOXP3+ T-regulatory cells and an M2 polarization of TAM (Sica and Bronte 2007; Huang et al. 2006; Sinha et al. 2007; Nagaraj and Gabrilovich 2008). The immunoregulatory activity of TAM is mostly influenced by tissue specific local cues. In the tumor milieu, a number of immunosuppressive factors (e.g., IL-10 and transforming growth factor-b) have been postulated by several studies to be important in the differentiation of infiltrating monocytes toward an M2 macrophage phenotype (Mantovani et al. 2002; Pollard 2004; Mantovani et al. 2006; Allavena et al. 2008). NF-kB, an important tumorigenic mediator of the pro-inflammatory milieu and a pivotal player connecting inflammation to many of the essential and known processes required to drive malignant transformation and already highlighted earlier, has also been recently associated to
424
A. Yagui-Beltrán et al.
the M2 polarization of TAM (Hagemann et al. 2008). Increasing evidence supports that metabolic changes in the tumor pro-inflammatory milieu, apart from providing the advantageous environment through the availability of a multitude of growth factors, can also have a more profound effect on infiltrating leukocytes (Tennant et al. 2009), a premise not difficult to comprehend. For example, several studies have demonstrated that tumor cells are able to secrete lactic acid that in turn stimulates the IL-23/IL-17 axis in TAM (Shime et al. 2008) at the expense of the immunoprotective IL-12-inducible Th1 pathway.
Relationship of Bone Marrow-Derived Cells and the Tumor Microenvironment This topic will be extensively discussed in another chapter. Cytokines and other tumor derived soluble factors not only recruit to tumors and well-defined inflammatory cells but also, as increasingly recognized, bone marrow-derived progenitor cell populations. BMDC that are recruited to the site of chronic inflammation have demonstrated remarkable plasticity and have been shown to differentiate into cells of diverse lineages. In vitro, this differentiation is governed by secreted factors. Similarly, in vivo their differentiation seems to be governed by the tissue (tumor) microenvironment (Gonda et al. 2009).
Conclusion Signals from the microenvironment have a profound influence on the maintenance and/or progression of epithelial cancers. Increasing research evidence from animal and human studies demonstrates that the pro-inflammatory milieu with its various components and soluble mediators plays a fundamental role in the initiation of malignant epithelial cells while sustaining and perpetuating chronic inflammation that in turn results in further malignant proliferation and distant malignant spread and seeding. As the fields of tumor immunology and tumor biology progress, the focus of investigation is starting to shift from merely identifying and describing the components of the protagonists of infiltrating immune cell populations in the proinflammatory milieu to a more careful definition of the functional significance and mechanisms behind their complex dynamic interplays. This is slowly helping the scientific and medical community to understand how chronic inflammation is exactly contributing to the genetic and epigenetic aberrant alterations that are observed in the pre-cancerous and cancerous cells dwelling within the pro-tumor milieu and cancer microenvironment. Of importance in this dynamic understanding of the TME, inflammation, and the neoplastic process is the recognition of the increasing evidence supporting that there may be small populations of pluripotent cells with the ability of self-renewal that may be recruited and forced by pathways
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
425
not yet fully understood into malignant initiation contributing to the development and spread of cancer. It is hoped that with better knowledge of the molecular mechanisms that link the stem-cell like populations and niches, with the proinflammatory microenvironment, we will be able to design novel targeted therapies aimed at specific components of these exciting pathways. Acknowledgment The authors are supported by grants from the National Institutes of Health (CA132566 and CA130980), Department of Defense Award for mesothelioma research (W81XWH-09-1-0342), an Era of Hope Scholar Award (W81XWH-06-1-0416), the Bonnie J. Addario Lung Cancer Research Foundation, the Kazan, McClain, Abrams, Fernandez, Lyons, Greenwood, Harley & Oberman Research Foundation Inc., and the Eileen D. Ludwig Endowed Chair Fund for Thoracic Oncology Research.
References Ahn GO, Brown JM (2008) Matrix metalloproteinase-9 is required for tumor vasculogenesis but not for angiogenesis: role of bone marrow-derived myelomonocytic cells. Cancer Cell 13(3):193–205. Allavena P et al (2008) Pathways connecting inflammation and cancer. Curr Opin Genet Dev 18(1):3–10. Allavena P et al (2008) The Yin-Yang of tumor-associated macrophages in neoplastic progression and immune surveillance. Immunol Rev 222:155–161. Andreu P et al (2010) FcRgamma activation regulates inflammation-associated squamous carcinogenesis. Cancer Cell 17(2):121–134. Azad N, Rojanasakul Y, Vallyathan V (2008) Inflammation and lung cancer: roles of reactive oxygen/nitrogen species. J Toxicol Environ Health B Crit Rev 11(1):1–15. Bacchetta R, Gambineri E, Roncarolo MG (2007) Role of regulatory T cells and FOXP3 in human diseases. J Allergy Clin Immunol 120(2):227–235; quiz 236–237. Balkwill, F (2004) Cancer and the chemokine network. Nat Rev Cancer 4(7):540–550. Balkwill, F (2009) Tumour necrosis factor and cancer. Nat Rev Cancer 9(5):361–371. Balkwill F, Mantovani A (2001) Inflammation and cancer: back to Virchow? Lancet 357(9255):539–545. Balkwill F, Charles KA, Mantovani A (2005) Smoldering and polarized inflammation in the initiation and promotion of malignant disease. Cancer Cell 7(3):211–217. Banchereau J, Steinman RM (1998) Dendritic cells and the control of immunity. Nature 392(6673):245–252. Becker C et al (2004) TGF-beta suppresses tumor progression in colon cancer by inhibition of IL-6 trans-signaling. Immunity 21(4):491–501. Berger SA (2006) Signaling pathways influencing SLF and c-kit-mediated survival and proliferation. Immunol Res 35(1–2):1–12. Bingle L, Brown NJ, Lewis CE (2002) The role of tumour-associated macrophages in tumour progression: implications for new anticancer therapies. J Pathol 196(3):254–265. Bollrath J et al (2009) gp130-mediated Stat3 activation in enterocytes regulates cell survival and cell-cycle progression during colitis-associated tumorigenesis. Cancer Cell 15(2):91–102. Bonasio R, von Andrian UH (2006) Generation, migration and function of circulating dendritic cells. Curr Opin Immunol 18(4):503–511. Bromberg J, Wang TC (2009) Inflammation and cancer: IL-6 and STAT3 complete the link. Cancer Cell 15(2):79–80. Brown ER et al (2008) A clinical study assessing the tolerability and biological effects of infliximab, a TNF-alpha inhibitor, in patients with advanced cancer. Ann Oncol 19(7):1340–1346.
426
A. Yagui-Beltrán et al.
Burger JA, Kipps TJ (2006) CXCR4: a key receptor in the crosstalk between tumor cells and their microenvironment. Blood 107(5):1761–1767. Carbia-Nagashima A et al (2007) RSUME, a small RWD-containing protein, enhances SUMO conjugation and stabilizes HIF-1alpha during hypoxia. Cell 131(2):309–323. Caro AA, Cederbaum AI (2006) Role of cytochrome P450 in phospholipase A2- and arachidonic acid-mediated cytotoxicity. Free Radic Biol Med 40(3):364–375. Chan AT, Ogino S, Fuchs CS (2007) Aspirin and the risk of colorectal cancer in relation to the expression of COX-2. N Engl J Med 356(21):2131–2142. Chang YC et al (2008) Epigenetic control of MHC class II expression in tumor-associated macrophages by decoy receptor 3. Blood 111(10):5054–5063. Chen CJ et al (2007) Identification of a key pathway required for the sterile inflammatory response triggered by dying cells. Nat Med 13(7):851–856. Condeelis J, Pollard JW (2006) Macrophages: obligate partners for tumor cell migration, invasion, and metastasis. Cell 124(2):263–266. Conroy H, Marshall NA, Mills KH (2008) TLR ligand suppression or enhancement of Treg cells? A double-edged sword in immunity to tumours. Oncogene 27(2):168–180. Coussens LM, Werb Z (2002) Inflammation and cancer. Nature 420(6917):860–867. Dalgleish AG, O’Byrne K (2006) Inflammation and cancer: the role of the immune response and angiogenesis. Cancer Treat Res 130:1–38. DeNardo DG et al (2009) CD4(+) T cells regulate pulmonary metastasis of mammary carcinomas by enhancing protumor properties of macrophages. Cancer Cell 16(2):91–102. de Visser KE, Eichten A, Coussens LM (2006) Paradoxical roles of the immune system during cancer development. Nat Rev Cancer 6(1):24–37. Dinarello CA (2006) The paradox of pro-inflammatory cytokines in cancer. Cancer Metastasis Rev 25(3):307–313. Dineen SP et al (2008) Vascular endothelial growth factor receptor 2 mediates macrophage infiltration into orthotopic pancreatic tumors in mice. Cancer Res 68(11):4340–4346. Drayton DL et al (2006) Lymphoid organ development: from ontogeny to neogenesis. Nat Immunol 7(4):344–353. Duluc D et al (2007) Tumor-associated leukemia inhibitory factor and IL-6 skew monocyte differentiation into tumor-associated macrophage-like cells. Blood 110(13):4319–4330. Du R et al (2008) HIF1alpha induces the recruitment of bone marrow-derived vascular modulatory cells to regulate tumor angiogenesis and invasion. Cancer Cell 13(3):206–220. Dvorak HF (1986) Tumors: wounds that do not heal. Similarities between tumor stroma generation and wound healing. N Engl J Med 315(26):1650–1659. Elkabets M et al (2009) Host-derived interleukin-1alpha is important in determining the immunogenicity of 3-methylcholantrene tumor cells. J Immunol 182(8):4874–4881. Ferrantini M, Capone I, Belardelli F (2008) Dendritic cells and cytokines in immune rejection of cancer. Cytokine Growth Factor Rev 19(1):93–107. Fidler IJ (1995) Modulation of the organ microenvironment for treatment of cancer metastasis. J Natl Cancer Inst 87(21):1588–1592. Fischer C et al (2007) Anti-PlGF inhibits growth of VEGF(R)-inhibitor-resistant tumors without affecting healthy vessels. Cell 131(3):463–475. Fischetti F, Tedesco F (2006) Cross-talk between the complement system and endothelial cells in physiologic conditions and in vascular diseases. Autoimmunity 39(5):417–428. Flossmann E, Rothwell PM (2007) Effect of aspirin on long-term risk of colorectal cancer: consistent evidence from randomised and observational studies. Lancet 369(9573):1603–1613. Fukata M, Abreu MT (2008) Role of toll-like receptors in gastrointestinal malignancies. Oncogene 27(2):234–243. Fukata S et al (2005) Levels of angiogenesis and expression of angiogenesis-related genes are prognostic for organ-specific metastasis of renal cell carcinoma. Cancer 103(5):931–942. Galli SJ, Grimbaldeston M, Tsai M (2008) Immunomodulatory mast cells: negative, as well as positive, regulators of immunity. Nat Rev Immunol 8(6):478–486.
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
427
Gallucci S, Matzinger P (2001) Danger signals: SOS to the immune system. Curr Opin Immunol 13(1):114–119. Garlanda C et al (2007) Increased susceptibility to colitis-associated cancer of mice lacking TIR8, an inhibitory member of the interleukin-1 receptor family. Cancer Res 67(13):6017–6021. Giavazzi R et al (1990) Interleukin 1-induced augmentation of experimental metastases from a human melanoma in nude mice. Cancer Res 50(15):4771–4775. Gonda TA, Tu S, Wang TC (2009) Chronic inflammation, the tumor microenvironment and carcinogenesis. Cell Cycle 8(13):2005–2013. Greten FR et al (2004) IKKbeta links inflammation and tumorigenesis in a mouse model of colitisassociated cancer. Cell 118(3):285–296. Grivennikov S et al (2009) IL-6 and Stat3 are required for survival of intestinal epithelial cells and development of colitis-associated cancer. Cancer Cell 15(2):103–113. Hagemann T et al (2008) “Re-educating” tumor-associated macrophages by targeting NF-kappaB. J Exp Med 205(6):1261–1268. Hanukoglu I (2006) Antioxidant protective mechanisms against reactive oxygen species (ROS) generated by mitochondrial P450 systems in steroidogenic cells. Drug Metab Rev 38(1–2):171–196. Harrison ML et al (2007) Tumor necrosis factor alpha as a new target for renal cell carcinoma: two sequential phase II trials of infliximab at standard and high dose. J Clin Oncol 25(29):4542–4549. Hold GL, El-Omar EM (2008) Genetic aspects of inflammation and cancer. Biochem J 410(2):225–235. Huang B et al (2006) Gr-1+CD115+ immature myeloid suppressor cells mediate the development of tumor-induced T regulatory cells and T-cell anergy in tumor-bearing host. Cancer Res 66(2):1123–1131. Hussain SP, Harris CC (2007) Inflammation and cancer: an ancient link with novel potentials. Int J Cancer 121(11):2373–2380. Hussain SP et al (2000) Increased p53 mutation load in noncancerous colon tissue from ulcerative colitis: a cancer-prone chronic inflammatory disease. Cancer Res 60(13):3333–3337. Ibrahim FB, Pang SJ, Melendez AJ, (2004) Anaphylatoxin signaling in human neutrophils. A key role for sphingosine kinase. J Biol Chem 279(43):44802–44811. Ii M et al (2006) Role of matrix metalloproteinase-7 (matrilysin) in human cancer invasion, apoptosis, growth, and angiogenesis. Exp Biol Med (Maywood) 231(1):20–27. Juliet PA et al (2003) Concomitant production of nitric oxide and superoxide in human macrophages. Biochem Biophys Res Commun 310(2):367–370. Kabelitz D, Medzhitov R (2007) Innate immunity – cross-talk with adaptive immunity through pattern recognition receptors and cytokines. Curr Opin Immunol 19(1):1–3. Kaplan RN et al (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438(7069):820–827. Karin M (2006) Nuclear factor-kappaB in cancer development and progression. Nature 441(7092):431–436. Karin M, Greten FR (2005) NF-kappaB: linking inflammation and immunity to cancer development and progression. Nat Rev Immunol 5(10):749–759. Khatami M et al (1984) Vernal conjunctivitis. Model studies in guinea pigs immunized topically with fluoresceinyl ovalbumin. Arch Ophthalmol 102(11):1683–1688. Koehne CH, Dubois RN (2004) COX-2 inhibition and colorectal cancer. Semin Oncol 31(2 Suppl 7):12–21. Kortylewski M et al (2005) Inhibiting Stat3 signaling in the hematopoietic system elicits multicomponent antitumor immunity. Nat Med 11(12):1314–1321. Kortylewski M et al (2009) Regulation of the IL-23 and IL-12 balance by Stat3 signaling in the tumor microenvironment. Cancer Cell 15(2):114–123. Krelin Y et al (2007) Interleukin-1beta-driven inflammation promotes the development and invasiveness of chemical carcinogen-induced tumors. Cancer Res 67(3):1062–1071.
428
A. Yagui-Beltrán et al.
Kulbe H et al (2007) The inflammatory cytokine tumor necrosis factor-alpha generates an autocrine tumor-promoting network in epithelial ovarian cancer cells. Cancer Res 67(2):585–592. Kuper H, Adami HO, Trichopoulos D (2000) Infections as a major preventable cause of human cancer. J Intern Med 248(3):171–183. Kusmartsev S et al (2008) Oxidative stress regulates expression of VEGFR1 in myeloid cells: link to tumor-induced immune suppression in renal cell carcinoma. J Immunol 181(1):346–353. Langowski JL et al (2006) IL-23 promotes tumour incidence and growth. Nature 442(7101):461–465. Ledeen RW, Wu G (2006) GM1 ganglioside: another nuclear lipid that modulates nuclear calcium. GM1 potentiates the nuclear sodium-calcium exchanger. Can J Physiol Pharmacol 84(3–4):393–402. Lee H et al (2009) Persistently activated Stat3 maintains constitutive NF-kappaB activity in tumors. Cancer Cell 15(4):283–293. Lewis CE, Pollard JW (2006) Distinct role of macrophages in different tumor microenvironments. Cancer Res 66(2):605–612. Lim SO et al (2008) Epigenetic changes induced by reactive oxygen species in hepatocellular carcinoma: methylation of the E-cadherin promoter. Gastroenterology 135(6):2128–2140, 2140, e1–e8. Lin EY et al (2001) Colony-stimulating factor 1 promotes progression of mammary tumors to malignancy. J Exp Med 193(6):727–740. Lin WW, Karin M (2007) A cytokine-mediated link between innate immunity, inflammation, and cancer. J Clin Invest 117(5):1175–1183. Lodoen MB, Lanier LL (2006) Natural killer cells as an initial defense against pathogens. Curr Opin Immunol 18(4):391–398. Luo JL et al (2007) Nuclear cytokine-activated IKKalpha controls prostate cancer metastasis by repressing Maspin. Nature 446(7136):690–694. Mager DL (2006) Bacteria and cancer: cause, coincidence or cure? A review. J Transl Med 4:14. Mager DL et al (2005) The salivary microbiota as a diagnostic indicator of oral cancer: a descriptive, non-randomized study of cancer-free and oral squamous cell carcinoma subjects. J Transl Med 3:27. Mantovani A et al (2002) Macrophage polarization: tumor-associated macrophages as a paradigm for polarized M2 mononuclear phagocytes. Trends Immunol 23(11):549–555. Mantovani A et al (2006) Role of tumor-associated macrophages in tumor progression and invasion. Cancer Metastasis Rev 25(3):315–322. Mantovani A et al (2008) Cancer-related inflammation. Nature 454(7203):436–444. Mantovani A et al (2008) Tumour immunity: effector response to tumour and role of the microenvironment. Lancet 371(9614):771–783. Marhaba R et al (2008) Opposing effects of fibrosarcoma cell-derived IL-1 alpha and IL-1 beta on immune response induction. Int J Cancer 123(1):134–145. McCaig C et al (2006) The role of matrix metalloproteinase-7 in redefining the gastric microenvironment in response to Helicobacter pylori. Gastroenterology 130(6):1754–1763. Monteseirin J et al (2004) Human neutrophils synthesize IL-8 in an IgE-mediated activation. J Leukoc Biol 76(3):692–700. Moore RJ et al (1999) Mice deficient in tumor necrosis factor-alpha are resistant to skin carcinogenesis. Nat Med 5(7):828–831. Morandi F et al (2007) Human neuroblastoma cells trigger an immunosuppressive program in monocytes by stimulating soluble HLA-G release. Cancer Res 67(13):6433–6441. Muller A et al (2001) Involvement of chemokine receptors in breast cancer metastasis. Nature 410(6824):50–56. Murdoch C et al (2008) The role of myeloid cells in the promotion of tumour angiogenesis. Nat Rev Cancer 8(8):618–631. Nagaraj S, Gabrilovich DI (2008) Tumor escape mechanism governed by myeloid-derived suppressor cells. Cancer Res 68(8):2561–2563. Naugler WE, Karin M (2008) The wolf in sheep’s clothing: the role of interleukin-6 in immunity, inflammation and cancer. Trends Mol Med 14(3):109–119.
20 The Pro-inflammatory Milieu and Its Role in Malignant Epithelial Initiation
429
Naugler WE et al (2007) Gender disparity in liver cancer due to sex differences in MyD88dependent IL-6 production. Science 317(5834):121–124. Nauts HC, McLaren JR (1990) Coley toxins – the first century. Adv Exp Med Biol 267:483–500. Nyakern M et al (2006) Synergistic induction of apoptosis in human leukemia T cells by the Akt inhibitor perifosine and etoposide through activation of intrinsic and Fas-mediated extrinsic cell death pathways. Mol Cancer Ther 5(6):1559–1570. Oguma K et al (2008) Activated macrophages promote Wnt signalling through tumour necrosis factor-alpha in gastric tumour cells. Embo J 27(12):1671–1681. Peerschke EI et al (2006) Blood platelets activate the classical pathway of human complement. J Thromb Haemost 4(9):2035–2042. Perez R et al (2006) Blockade of arachidonic acid incorporation into phospholipids induces apoptosis in U937 promonocytic cells. J Lipid Res 47(3):484–491. Pikarsky E et al (2004) NF-kappaB functions as a tumour promoter in inflammation-associated cancer. Nature 431(7007):461–466. Pollard JW (2004) Tumour-educated macrophages promote tumour progression and metastasis. Nat Rev Cancer 4(1):71–78. Popivanova BK et al (2008) Blocking TNF-alpha in mice reduces colorectal carcinogenesis associated with chronic colitis. J Clin Invest 118(2):560–570. Quante M, Wang TC (2008) Inflammation and stem cells in gastrointestinal carcinogenesis. Physiology (Bethesda) 23:350–359. Rakoff-Nahoum S, Medzhitov R (2007) Regulation of spontaneous intestinal tumorigenesis through the adaptor protein MyD88. Science 317(5834):124–127. Raulet DH (2004) Interplay of natural killer cells and their receptors with the adaptive immune response. Nat Immunol 5(10):996–1002. Rius J et al (2008) NF-kappaB links innate immunity to the hypoxic response through transcriptional regulation of HIF-1alpha. Nature 453(7196):807–811. Robinson-Smith TM et al (2007) Macrophages mediate inflammation-enhanced metastasis of ovarian tumors in mice. Cancer Res 67(12):5708–5716. Sakurai T et al (2008) Hepatocyte necrosis induced by oxidative stress and IL-1 alpha release mediate carcinogen-induced compensatory proliferation and liver tumorigenesis. Cancer Cell 14(2):156–165. Sawanobori Y et al (2008) Chemokine-mediated rapid turnover of myeloid-derived suppressor cells in tumor-bearing mice. Blood 111(12):5457–5466. Seandel M et al (2008) A catalytic role for proangiogenic marrow-derived cells in tumor neovascularization. Cancer Cell 13(3):181–183. Selsing E (2006) Ig class switching: targeting the recombinational mechanism. Curr Opin Immunol 18(3):249–254. Shahrzad S et al (2007) Induction of DNA hypomethylation by tumor hypoxia. Epigenetics 2(2):119–125. Shime H et al (2008) Tumor-secreted lactic acid promotes IL-23/IL-17 proinflammatory pathway. J Immunol 180(11):7175–7183. Shojaei F et al (2007) Tumor refractoriness to anti-VEGF treatment is mediated by CD11b+Gr1+ myeloid cells. Nat Biotechnol 25(8):911–920. Sica A, Bronte V (2007) Altered macrophage differentiation and immune dysfunction in tumor development. J Clin Invest 117(5):1155–1166. Sica A et al (2006) Tumour-associated macrophages are a distinct M2 polarised population promoting tumour progression: potential targets of anti-cancer therapy. Eur J Cancer 42(6):717–727. Sinha P et al (2007) Cross-talk between myeloid-derived suppressor cells and macrophages subverts tumor immunity toward a type 2 response. J Immunol 179(2):977–983. Snook CF, Jones JA, Hannun YA (2006) Sphingolipid-binding proteins. Biochim Biophys Acta 1761(8):927–946. Stearman RS et al (2008) A macrophage gene expression signature defines a field effect in the lung tumor microenvironment. Cancer Res 68(1):34–43.
430
A. Yagui-Beltrán et al.
Stenmark KR, Fagan KA, Frid MG (2006) Hypoxia-induced pulmonary vascular remodeling: cellular and molecular mechanisms. Circ Res 99(7):675–691. Swann JB et al (2008) Demonstration of inflammation-induced cancer and cancer immunoediting during primary tumorigenesis. Proc Natl Acad Sci U S A 105(2):652–656. Talmadge JE, Donkor M, Scholar E (2007) Inflammatory cell infiltration of tumors: Jekyll or Hyde. Cancer Metastasis Rev 26(3–4):373–400. Taylor CT (2008) Interdependent roles for hypoxia inducible factor and nuclear factor-kappaB in hypoxic inflammation. J Physiol 586(Pt 17):4055–4059. Tennant DA et al (2009) Metabolic transformation in cancer. Carcinogenesis 30(8):1269–1280. Tu S et al (2008) Overexpression of interleukin-1beta induces gastric inflammation and cancer and mobilizes myeloid-derived suppressor cells in mice. Cancer Cell 14(5):408–419. Valent P et al (2001) Variable expression of activation-linked surface antigens on human mast cells in health and disease. Immunol Rev 179:74–81. van ’t Veer LJ et al (2002) Gene expression profiling predicts clinical outcome of breast cancer. Nature 415(6871):530–536. Voronov E et al (2003) IL-1 is required for tumor invasiveness and angiogenesis. Proc Natl Acad Sci U S A 100(5):2645–2650. Wang RF, Miyahara Y, Wang HY (2008) Toll-like receptors and immune regulation: implications for cancer therapy. Oncogene 27(2):181–189. Whiteside TL (2008) The tumor microenvironment and its role in promoting tumor growth. Oncogene 27(45):5904–5912. Wilson NS, Villadangos JA (2004) Lymphoid organ dendritic cells: beyond the Langerhans cells paradigm. Immunol Cell Biol 82(1):91–98. Witz IP (2008) Tumor-microenvironment interactions: dangerous liaisons. Adv Cancer Res 100:203–229. Wu WS (2006) The signaling mechanism of ROS in tumor progression. Cancer Metastasis Rev 25(4):695–705. Wyckoff JB et al (2007) Direct visualization of macrophage-assisted tumor cell intravasation in mammary tumors. Cancer Res 67(6):2649–2656. Xiao H et al (2007) The Toll-interleukin-1 receptor member SIGIRR regulates colonic epithelial homeostasis, inflammation, and tumorigenesis. Immunity 26(4):461–475. Yang L et al (2008) Abrogation of TGF beta signaling in mammary carcinomas recruits Gr-1+CD11b+ myeloid cells that promote metastasis. Cancer Cell 13(1):23–35. Yano K et al (1997) Production of macrophage inflammatory protein-1alpha by human mast cells: increased anti-IgE-dependent secretion after IgE-dependent enhancement of mast cell IgEbinding ability. Lab Invest 77(2):185–193. Yu H, Kortylewski M, Pardoll D (2007) Crosstalk between cancer and immune cells: role of STAT3 in the tumour microenvironment. Nat Rev Immunol 7(1):41–51. Zhang B (2008) Targeting the stroma by T cells to limit tumor growth. Cancer Res 68(23):9570–9573. Zhu P et al (2006) Macrophage/cancer cell interactions mediate hormone resistance by a nuclear receptor derepression pathway. Cell 124(3):615–629.
Chapter 21
Natural Killer Cells for Adoptive Immunotherapy Jonathan E. Benjamin and Sally Arai
Abstract Natural killer (NK) cell-based immunotherapy in cancer is gaining prominence yet is still early in its development as we learn more about NK cell biology and the molecular mechanisms that drive NK cell expansion, direct trafficking, and determine target cell recognition. Clinical investigation of NK cell therapy has ranged from in vivo modulation of NK cell activity to ex vivo purification/expansion and adoptive transfer with modest results. Future trials will likely move toward refining the current techniques of NK expansion and overcoming the barriers of NK cell engraftment and host factors.
Introduction Natural Killer (NK) cells have attracted the attention of cancer immunologists for their ability to lyse a broad range of cancer cell types in vitro, to eradicate transplanted tumors, and to participate in the surveillance of spontaneously arising tumors in susceptible mouse strains. Their name reflects functional activity that was originally observed in vitro, namely, rapid, contact-dependent killing that did not require prior sensitization, was not genotypically restricted, and did not require the addition of antibodies (Herberman et al. 1975; Kiessling et al. 1975). Unlike T and B cells, NK cells do not rearrange germline encoded antigen receptor genes, and NK target recognition is not, strictly speaking, MHC restricted. Rather, NK cell function is governed by the integration of opposing signals from inhibitory and activating cell surface with diverse ligand-binding specificities. Moreover, cytotoxicity represents only one facet of NK cell function. Although previously considered to be a homogenous population, NK cells display phenotypic diversity that is matched by a broad repertoire of immune regulatory functions. It is likely that some of these functions will be relevant to the antitumor capabilities of NK cells. The J.E. Benjamin (*) Division of Blood and Marrow Transplantation, Stanford University Medical Center, Stanford, CA, USA e-mail:
[email protected]
R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_21, © Springer Science+Business Media, LLC 2010
431
432
J.E. Benjamin and S. Arai
increasing appreciation of the molecular mechanisms that drive NK cell expansion, trafficking, and target cell recognition improve the likelihood that NK cell-based therapies may be successfully employed to treat cancer. However, NK cell-based therapies must be considered to be at an early stage. In this chapter, we attempt to review the current understanding of NK cell biology as pertains to recent approaches toward their therapeutic utilization.
Immunophenotype The study of NK cell biology has lagged behind that of B and T cells because no surface marker has yet been identified that is expressed specifically and exclusively by all NK cells. Moreover, mouse and human NK cells do not necessarily express orthologous markers, making comparisons of NK cell subsets between these species difficult. In humans, NK cells are characterized by the expression of the adhesion glycoprotein CD56 (NCAM) and the absence of CD3. The function of CD56 on NK cells remains unknown, but its surface density further subdivides NK cells into two populations with distinct functional properties (Baume et al. 1992; Carson et al. 1997; Lanier et al. 1991). CD56dim cells account for 90% of circulating peripheral blood NK cells, are potently cytotoxic, and express high amounts of intracellular perforin (Jacobs et al. 2001). This population expresses FcgRIIIa (CD16), a receptor for IgG implicated in antibody-dependent cellular cytotoxicity (ADCC). In contrast, CD56bright cells account for approximately 10% of circulating CD3-CD56+ cells but comprise the vast majority of NK cells in the lymph nodes (Fehniger et al. 2003). There, they are found primarily in the para-follicular T-cell rich regions, colocalizing with dendritic cells. The CD56bright cells lack CD16, do not express high levels of perforin, and are minimally cytotoxic (Jacobs et al. 2001). Instead, they function as modulators of the immune system by virtue of their abundant cytokine secretion (Cooper et al. 2001). Depending on the specific activating signal, CD56bright cells have the capacity to secrete IFN-g, TNF-a, TNF-b, GM-CSF, IL-10, and IL-13. Recently, multiple groups have described a novel subset of CD56bright NK cells in tonsils and Peyer’s patches that secrete high levels of IL-22 (Cella et al. 2009; Sanos et al. 2009; Satoh-Takayama et al. 2008). It is likely that additional, functionally distinct populations of cells with NK-like phenotypes will be discovered.
Ontogeny NK cells are derived from Lin- CD34+ CD38+ common lymphoid progenitors (Galy et al. 1995; Miller et al. 1994), and bone marrow appears to be the primary, but not exclusive, site of NK cell development. Extramedullary NK cell development occurs in the lymph nodes and thymus (Ferlazzo et al. 2004; Freud et al. 2005). In the lymph node, a unique population of CD34dim NK cell precursors may develop into CD56bright NK cells under the influence of T cell-derived IL-2 and possibly
21 Natural Killer Cells for Adoptive Immunotherapy
433
myeloid-derived IL-15 (Freud et al. 2005). In the mouse, the thymus-derived NK cell population has functional properties similar to those of CD56bright cells, and is dependent upon IL-7R signaling (Vosshenrich et al. 2006). The developmental relationship between CD56dim and CD56bright NK cells remains controversial (Colucci et al. 2003). There is some evidence to suggest that the CD56bright subset represents a precursor population of the CD56dim (Chan et al. 2007; Romagnani et al. 2007). When cultured with IL-2, CD56brightCD16- lymph node NK cells upregulate perforin, CD16, and demonstrate potent cytoxicity (Ferlazzo et al. 2004). Conversely, in vitro cytokine stimulation of CD56dim cells leads to upregulation of CD56 expression, although this does not necessarily signify a switch to an immunoregulatory phenotype (Takahashi et al. 2007).
Localization and Trafficking NK cells are broadly distributed in lymphoid and non-lymphoid organs, consistent with their role as early responders to infection. In the peripheral blood, NK cells account for 10–25% of lymphocytes, yielding a concentration of 100–600 ml-1, and are predominantly CD56dim. In the spleen, NK cells constitute 25% of lymphocytes, are predominantly CD56dim, and are concentrated in the red pulp adjacent to antigen-presenting macrophages and dendritic cells (Witte et al. 1990). In the lymph node, NK cells represent 1–5% of cells (Fehniger et al. 2003). They represent the predominant lymphocyte subset in the liver, with a variable distribution of CD56dim versus CD56bright (Burt et al. 2009). In the lung, NK cells make up 10% of all lymphocytes, representing a large reservoir that is second only to the spleen in absolute numbers (Culley 2009). Interestingly, a large number of CD56bright NK cells are also found in the placenta and are essential to the formation of the decidua and may play a role in the development of maternal tolerance to the semiallogeneic fetus (Croy et al. 1997). The anatomic distribution of NK cell subsets is determined in part by soluble chemotactic factors and membrane-bound adhesion molecules. NK cell maturation occurs with the acquisition of S1P5, a G-protein coupled receptor for the lysophospholipid sphingosine-1-phosphate. Genetic ablation of this receptor in mice results in the accumulation of NK cells in bone marrow and lymph nodes, and their absence in the blood, spleen, and lung (Jenne et al. 2009; Walzer et al. 2007). It is unknown whether S1P5 agonists might function to mobilize NK cells into peripheral blood. Migration of CD56bright cells to lymph nodes is mediated by CCR7, the receptor for the CCL19 and CCL21 chemokines. CD62L, or L-selectin, interacts with ligands on high endothelial venules and mediates the entry of CD56bright cells into lymph nodes (Frey et al. 1998). CD56dim cells express CX3CR1 and ChemR23, allowing their homing to inflamed tissues (Nishimura et al. 2002; Parolini et al. 2007). Notably, tumor cell expression of CX3CL1, a ligand for CX3CR1, has been associated with NK cell infiltration and better prognosis (Hyakudomi et al. 2008; Ohta et al. 2005).
434
J.E. Benjamin and S. Arai
Activation of NK Cells by Cytokines and Accessory Cells Peripheral blood NK effector function can be augmented by treatment with pharmacologic doses of cytokines and by co-culture with APC, leading to a distinction between resting and activated NK cells (Fernandez et al. 1999; Trinchieri et al. 1984). IL-2 was the first cytokine demonstrated to enhance NK cytotoxicity and remains the most frequently used cytokine for NK cell activation in clinical trials (Caligiuri et al. 1993; Miller et al. 2005; Shi et al. 2008). However, IL-2 preferentially expands the CD56 bright subset (Baume et al. 1992; Soiffer et al. 1996). IL-15 is singularly important for NK cell development and survival (Cooper et al. 2002; Ranson et al. 2003). Deficiencies in IL-15 or IL-15R α result in profoundly reduced numbers of NK cells (Kennedy et al. 2000; Lodolce et al. 1998). Furthermore, adoptive transfer of NK cells into either IL-15 or IL-15R α recipients results in the rapid disappearance of NK cells, suggestive of an important role in NK cell survival (Cooper et al. 2002; Ranson et al. 2003). IL-15 is a member of the four alpha-helix family of cytokines. Its receptor includes a beta subunit shared with the IL-2R and a g-chain subunit shared by IL-2, IL-4, IL-7, IL-9, and IL-21. A feature apparently unique to IL-15 signaling is the requisite expression in cis of IL-15 and the IL-15R a subunit. As currently formulated, the transpresentation model posits that IL-15 binds to IL-15R a on the cell surface where it can stimulate the IL-15 signaling pathway of neighboring cells that express the IL-2/15R b and common g subunits (Burkett et al. 2004; Dubois et al. 2002; Koka et al. 2003; Lodolce et al. 2001). Both plasmacytoid and myeloid dendritic cells are important physiologic inducers of NK cell cytotoxicity and cytokine production, albeit by different mechanisms (Ferlazzo et al. 2002; Fernandez et al. 1999; Gerosa et al. 2002). Dendritic cell priming of NK cells is dependent on cell contact and IL-15Ra expression on the dendritic cell (Koka et al. 2004). Dendritic cell-derived type I IFN potentiates NK cell cytotoxicity. In contrast, NK cell IFN-g production is enhanced by direct cellular contact with IL-12 producing DC, usually in concert with additional cytokines such as IL-1, IL-2, IL-15, or IL-18 (Cooper et al. 2001).
Cytokine Secretion Interferon g is perhaps the most important of the cytokines secreted by NK cells, with higher levels produced by the CD56bright subset. Its effects are protean, including polarization of T cells to the Th1 phenotype, upregulation of MHC I molecules on APC, maturation and activation of dendritic cells, antiviral effects against herpes viruses, and antiproliferative effects on transformed cells (Boehm et al. 1997; Schroder et al. 2004). NK cells also secrete tumor necrosis factor-a and GM-CSF, which can promote dendritic cell maturation and phagocytosis (Gerosa et al. 2002; Pistoia et al. 1989).
21 Natural Killer Cells for Adoptive Immunotherapy
435
Cytotoxicity As has been observed with cytotoxic CD8 T cells, NK cells may kill targets via multiple mechanisms. Lytic granule exocytosis is the result of a multistep pathway resulting in cytoskeletal reorganization leading to localization of cytoplasmic perforin and granzyme-containing vesicles to the point of contact with the target cell (Bryceson et al. 2005). Perforin creates pores in the target cell membrane, allowing granzyme to enter and cleave substrates such as caspases. Downstream of caspase activation is DNAse that degrades target cell DNA, resulting in apoptosis. Fusion of the lytic granule membrane with the plasma membrane enables the lysosome-associated membrane protein LAMP-1 (CD107a) to appear on the cell surface, providing a marker of NK cell activity that can be detected by flow cytometry (Alter et al. 2004). NK cells can also kill targets in a perforin-independent mechanism that involves TNF family of ligands and receptors. Both FasL and TNF-related apoptosisinducing ligand (TRAIL) are expressed on the NK membrane. They bind to Fas (CD95) and TRAIL-R1 or TRAIL-R2, leading to trimerization and the recruitment of Fas-Associated Death Domain (FADD) proteins in the target, which triggers the caspase cascade (Arase et al. 1995; Oshimi et al. 1996). Studies with blocking antibodies and genetically modified mice have provided ample evidence that these pathways are important for tumor cell recognition and clearance mediated by NK cells (Kashii et al. 1999; Smyth et al. 2001).
Activating and Inhibitory Signals NK cells employ a panoply of transmembrane receptors to assess potential targets for evidence of infection, stress, or transformation. Of particular interest to the field of cellular therapy are the receptors, both activating and inhibitory, that recognize Major Histocompatibility Class (MHC) I molecules. These include the killer immunoglobulin-like receptors (KIR) in humans, Ly49 receptors in mouse, and CD94/NKG2A in both species. KIR expression is stochastic, generating an NK cell compartment with individual members expressing one or more inhibitory receptors. The Class I specificity of an NK cell is determined by the repertoire of receptors it expresses. Failure to engage the inhibitory receptors, known as missing-self recognition (Karre et al. 1986), is thought to contribute to the graft-versus-malignancy effect observed in Human Leukocyte Antigen (HLA) haploidentical hematopoietic cell transplants (Ruggeri et al. 2002). KIR expression is noted predominantly on the CD56dim subset and thus appears to be more relevant to the cytotoxic pathways. In humans, the KIR gene cluster is located on chromosome 19 and thus segregates independently of HLA genes on chromosome 6. KIR genotype, rather than HLA, is the major determinant of KIR expression pattern (Shilling et al. 2003). Fourteen KIR genes and two pseudogenes have been identified. KIR genes are
436
J.E. Benjamin and S. Arai
highly polymorphic, and growing evidence suggests that allelic diversity may correlate with functional differences in both ligand recognition and signaling strength (Bari et al. 2009; Hsu et al. 2002; Single et al. 2007). KIR gene nomenclature is determined by the number of extracellular immunoglobulin domains and the length of the cytoplasmic tails. For instance, a KIR gene with the 2D extracellular domain may have a long (2DL2) or short (2DS2) cytoplasmic tail. Long cytoplasmic tails contain Immunoreceptor Tyrosine-Based Inhibitory Motifs (ITIMs) that facilitate the docking of phosphatase molecules. In KIR-S molecules, the short cytoplasmic tails generally contain transmembrane charged residues that enable pairing to the DAP12 adapter molecule. DAP12 contains an Immunoreceptor Tyrosine-Based Activating Motif (ITAM). Crosslinking of these receptors results in tyrosine phosphorylation of the ITAM and recruitment of the Syk and Zap70 tyrosine kinases. The extracellular domains of activating KIR share sequence similarity with the corresponding inhibitory receptors and may share HLA ligand binding specificities, although perhaps with reduced affinity. Four KIR specificities have been determined. Based on motifs in the carboxy terminal of the Class I alpha helix, HLA-C molecules can be divided into two groups by an allelic dimorphism at positions 77 and 80. Group 1 has a serine at position 77 and asparagine at position 80 (HLA-Cw3 and related alleles), whereas group 2 has asparagine and lysine (HLA-Cw4). KIR2DL2/2DL3 binds C1 and KIR2DL1 binds C2. HLA-B allotypes with the Bw4 motif bind to KIR3DL1. KIR3DL2 recognizes HLA-A3 and A11. In the context of allogeneic transplantation, KIR ligand mismatching in the GVH direction refers to a situation in which the recipient lacks one or more KIR ligands expressed by the donor.
Activating Receptors Absence of self-MHC antigens is insufficient to trigger an NK attack. Rather, NK cells require stimulation via cell surface receptors that pair with ITAM-containing signaling molecules. A number of activating NK cell receptors have now been characterized, and most, if not all NK cells, express many of these receptors. Consequently, unlike T and B cells, NK cells do not have clonally distributed antigen specificity. With the exception of CD16, crosslinking of a single activating receptor does not trigger NK effector function. Rather, cytotoxicity or cytokine secretion requires crosslinking of pairs or multiple receptors (Bryceson et al. 2009). The repertoire of activating NK receptors is large and has been reviewed elsewhere (Bottino et al. 2005; Lanier 2009). Certain receptors merit discussion for the purpose of this review. CD16, the intermediate-affinity receptor for IgG, mediates ADCC and is essential to the therapeutic benefit of depleting monoclonal antibodies such as rituximab or trastuzumab. CD16 levels on NK cells can be augmented by pharmacologic doses of IL-2, and multiple clinical trials employing monoclonal antibodies with IL-2 have been published (Eisenbeis et al. 2004; Fleming et al. 2002;
21 Natural Killer Cells for Adoptive Immunotherapy
437
Repka et al. 2003). Another well-studied receptor is NKG2D, a C-type lectin expressed on all NK cells as well as certain T cell subsets. Ligands for NKG2D, including the human MIC-A/B, and ULBP proteins, are structurally similar to MHC-I molecules (Bauer et al. 1999; Cosman et al. 2001). As described below, target cells may be sensitized to NK cell attack by the upregulation of NKG2D ligands by drugs such as 5-azacytidine, bortezomib, and arsenic trioxide. Finally, the natural cytotoxicity receptors (NCR) NKp46, NKp30, and NKp44 have been implicated in the recognition and targeting of tumor cells, but information regarding their ligand specificities remains elusive.
Licensing The mechanism of self-tolerance by NK cells is a fascinating question, with significant implications for clinical trial design. It is known that subsets of NK cells express only inhibitory KIRs for absent Class I antigens. Rather than being autoreactive, as would be predicted by the missing-self model, these cells are hyporesponsive to activating stimuli (Wu and Raulet 1997). The licensing theory of NK cell maturation suggests that engagement of an inhibitory KIR by the cognate Class I molecule is necessary to acquire effector function (Anfossi et al. 2006; Kim et al. 2005). The requirement for NK licensing would appear to preclude the efficacy of KIR-ligand mismatched NK cells. However, it seems that licensing can be bypassed by inflammatory stimuli, such as might be anticipated after the administration of chemotherapy (Cooley et al. 2007; Kim et al. 2005).
NK Cells and Anti-tumor Response Many studies have demonstrated NK cell killing of a variety of human and mouse cancer cell lines in vitro, as well as NK-dependent eradication of transplanted tumor cell lines in vivo (Carlsten et al. 2009). To date, NKG2D is the best-studied activating NK cell receptor linked to tumor recognition and lysis. In the mouse, blockade, downregulation, or targeted deletion of NKG2D results in an increased incidence of experimentally induced sarcomas or spontaneous lymphoma and prostate cancer in susceptible mouse strains (Guerra et al. 2008; Oppenheim et al. 2005; Smyth et al. 2005). NKG2D ligand (NKG2D-L) expression is minimal on healthy cells but upregulated in cancerous cells, likely by multiple mechanisms including those involving the DNA-damage repair pathway (Gasser et al. 2005). However, chemotherapies with agents such as bortezomib, 5-azacytidine, and arsenic trioxide, i.e., whose primary effect is not DNA damage, also enhance NKG2D-L expression (Diermayr et al. 2008; Poggi et al. 2009; Soriani et al. 2009). Clinical trials employing these agents as a component of NK cell therapy are underway.
438
J.E. Benjamin and S. Arai
Tumor Infiltrating Lymphocytes Infiltration of NK cells into numerous different tumor types has correlated with better outcome (Coca et al. 1997; Ishigami et al. 2000; Takanami et al. 2001; Villegas et al. 2002). Preclinical studies with in vitro activated NK cells infused into tumor-bearing mice have suggested that the infused cells can accumulate preferentially into the tumor tissue (Daldrup-Link et al. 2005; Meier et al. 2008). The ability of NK cells to penetrate into tumor tissue likely depends on the location of the tumor (e.g., epithelial tissue vs. lymph node) as well as tumor-specific factors such as extracellular matrix components, tumor-derived chemokines, and adhesion molecules.
NK Cells and Haploidentical Transplantation Landmark studies by the Perugia group have generated greater interest in the therapeutic use of NK cells. Allogeneic HCT is a potentially curative therapy for hematolymphoid neoplasms including myeloid and lymphoid leukemias, lymphoproliferative disorders, and marrow failure syndromes including myelodsyplasia and aplastic anemia. The benefits of allogeneic HCT are derived from both the conditioning regimen of chemoradiotherapy and the donor alloimmune response, primarily mediated by donor T lymphocytes recognizing minor histocompatibility antigens. The curative potential of allogeneic HCT is offset by the toxicity of the therapy, the most important being graft-versus-host disease (GVHD), also mediated by donor T cells. GVHD risk increases with the degree of HLA mismatching between donor and recipient. HLA identical siblings are preferred donors. If a sibling donor is not available, then an HLA-matched unrelated donor is sought via registries of healthy volunteers. Many patients, particularly those from minority populations, lack a suitably matched donor. Haploidentical (i.e., sharing only one HLA haplotype) related donors, including parents and offspring, are available for most patients. Transplantation with T-cell containing grafts from haploidentical donors carries an unacceptable risk of GVHD. The Perugia group has pioneered the use of haploidentical transplantation for patients with high-risk leukemia lacking an HLA-matched donor. In order to avoid GVHD, the donor grafts are T cell depleted, and “megadose” CD34 infusions are required to allow for engraftment (Aversa et al. 1994; Aversa et al. 1998). Superior outcomes were observed when KIR ligand mismatching was predicted in the GVH direction, but only in the setting of myeloid and not lymphoid malignancies, possibly reflecting the lack of NKG2D-L expression on lymphoid leukemias (Romanski et al. 2005; Ruggeri et al. 2002). KIR ligand incompatibility was also protective against GVHD and graft rejection. Protection against GVHD has been explored in mouse models, where it was found that Ly49 ligand mismatching
21 Natural Killer Cells for Adoptive Immunotherapy
439
in the GVH direction resulted in the depletion of recipient antigen-presenting cells (Lundqvist et al. 2007; Ruggeri et al. 2002). As has been shown after myeloablative conditioning, NK cells are the first lymphocyte population to repopulate the peripheral blood. In the Perugia studies and in others, most NK cells isolated from the recipient after the transplant lacked KIR but expressed CD94/NKG2A, more consistent with the CD56bright immunoregulatory phenotype, although this may represent a developmental intermediate to the CD56dim cytotoxic population (Nguyen et al. 2005; Ruggeri et al. 2007; Schulze et al. 2008; Vago et al. 2008). A mature donor-type KIR repertoire emerges within 3 months to 3 years (Shilling et al. 2003). The potential benefits of NK cell alloreactivity have been explored in cohorts of patients receiving mismatched unrelated donor transplants, often with discordant conclusions. In an early review of the experience at the University of Minnesota, Davies et al. found no benefit of KIR ligand incompatibilities (Davies et al. 2002). The authors noted that a variety of transplant protocols were employed at that center over the 10 years of accrual, most differing from those used in Perugia by virtue of the presence of T cells in the graft and by the use of post-grafting immunosuppression, both of which may skew developing NK cells to the CD56bright lineage (Cooley et al. 2005; Lowe et al. 2003; Wang et al. 2007). In contrast, Beelen et al. reviewed transplant outcomes at Essen and found a marked reduction of AML relapse in patients receiving T-cell replete, KIR ligand incompatible grafts (Beelen et al. 2005). This benefit was offset by an increased risk of chronic GVHD; therefore, no survival advantage was observed with KIR ligand incompatibility. Not surprisingly, multi-institutional studies have not clarified matters. Giebel et al. analyzed the outcomes of 130 patients receiving URD transplants at three European centers (Giebel et al. 2003). A total of 61 patients received HLA-matched transplants, and of the remaining patients who received mismatched transplants, 20 had KIR ligand mismatching in the GVH direction. Approximately two-thirds of the patients had myeloid malignancies. Among patients without KIR ligand mismatches, survival at 4.5 years was not significantly different regardless of whether the patients were HLA matched or mismatched (48% vs. 58%). In contrast, patients receiving KIR-ligand mismatched transplants experienced 87% survival. The incidence of Grade III–IV acute GVHD was 0 and 15% in patients receiving KIRligand mismatched and matched transplants, respectively. The authors, in seeking to reconcile their results with those of Perugia and Minnesota, noted that although T cell-replete grafts were used, all patients received pre-transplant ATG, which, because of its persistence post-transplant, may have negatively influenced donor T cell recovery. Farag et al. reviewed international registry data of 1,571 unrelated bone marrow donor transplants for myeloid malignancies (Farag et al. 2006). Two-thirds of the patients were HLA matched. In the HLA-mismatched donor recipient pairs, 30% had KIR ligand incompatibility in the GVH direction, 24% in the
440
J.E. Benjamin and S. Arai
HVG direction, and the remainder had no predicted NK alloreactivity. T cell depletion was employed in approximately 20% of transplants in all groups. Overall survival and TRM were superior in patients receiving HLA-matched transplants. Among those receiving mismatched transplants, there were no significant differences in outcomes according to the presence or vector of KIR ligand mismatching.
Umbilical Cord Transplantation As with haploidentical transplantation, umbilical cord blood (UCB) transplants are frequently characterized by a high degree of HLA mismatching, allowing an evaluation of the role of KIR ligand mismatching. In a review of 218 transplants, Willemze et al. found that the recipients of KIR-ligand incompatible grafts had improved overall survival due to decreased relapse when compared with those without these incompatibilities (Willemze et al. 2009). As in the Perugia studies, benefits of KIR ligand incompatibility were most evident among patients with AML, although UCB recipients with ALL also had a trend toward improved leukemia-free survival. Brunstein et al., however, failed to observe any benefit of KIR-L mismatch after myeloablative or non-myeloablative conditioning regimens (Brunstein et al. 2009).
Non-myeloablative Transplantation More than 40 years ago, the demonstration of hybrid (a × b F1) resistance to parental (a or b) hematopoietic grafts contradicted the prevailing understanding of allograft rejection (Cudkowicz and Stimpfling 1964). In skin and solid organ transplants, F1 progeny tolerates parental grafts. In the subsequent decades, recipient NK cells were identified as the mediator of hybrid resistance. In the hybrid resistance model, the conditioning regimen employed is total body irradiation, which may allow the persistence of significant numbers of recipient NK cells. Likewise, non-myeloablative conditioning is marked by a period of mixed chimerism in which recipient NK cells may exert anti-donor effects. Sobecks et al. analyzed the impact of recipient anti-donor NK-cell alloreactivity in 31 patients conditioned with fludarabine and 2Gy TBI (Sobecks et al. 2008). Donor–recipient pairs were stratified by the number of recipient KIR genes for which the donor lacked the appropriate HLA Class I ligand. Those with the greatest number of unmatched KIR-L and KIR exhibited the highest incidence of graft rejection and the lowest incidence of complete donor chimerism.
21 Natural Killer Cells for Adoptive Immunotherapy
441
Killer Cell Immunotherapy Rosenberg and colleagues pioneered the field of adoptive transfer of killer cells with a series of clinical trials performed in the 1980s (Rosenberg et al. 1985). Autologous peripheral blood mononuclear cells (PBMC) from patients with advanced renal cell carcinoma or melanoma were treated in vitro with high-dose IL-2, yielding a population of lymphokine activated killer cells (LAK) that demonstrated broad cytotoxicity toward a variety of tumor cell lines. Following reinfusion of LAK cells, patients were treated with high-dose IL-2 with the hope of maintaining the cytotoxic activity of the cells. Objective responses were observed in approximately 20% of patients, although subsequent studies found similar response rates with high-dose IL-2 alone (Yang et al. 2003). From the current perspective of NK cell biology, the LAK method has inherent shortcomings. First, the LAK product, derived from unsorted mononuclear cells, consisted predominantly of T cells. Second was the use of autologous cells, which with current knowledge of KIR biology would be predicted to be inhibited by tumors that maintained Class I expression.
NK Cell Adoptive Therapy Following the observations in haploidentical transplantation, along with abundant preclinical data, multiple groups have attempted to infuse purified NK cells in patients with refractory malignancies or poor graft function after transplant. Pure NK cell products can be obtained from leukapheresis specimens by immunomagnetic selection using commercially available products under GMP conditions. The CliniMACS system can achieve purities of 40–99%, a 3–5 log reduction in T cell content, and yields of 20–80% (Iyengar et al. 2003; Koehl et al. 2005; McKenna et al. 2007; Meyer-Monard et al. 2009). It is expected that NK isolation from peripheral blood yields a predominantly CD56dim population. Haploidentical NK Cells A pilot study of NK donor lymphocyte infusion (DLI) was attempted in five patients as treatment for poor chimerism function or early relapse after haploidentical HCT. PBMC were collected from the original donor by 10-L leukapheresis, and then purified by immunomagnetic T cell depletion followed by positive selection with anti-CD56. NK cells were not activated in vitro, and no chemotherapy or antibody was administered to deplete recipient lymphocytes. NK cell infusion (doses 0.2–2.2 × 107 kg-1) was safe and no patients developed GVHD. The results were mixed, with some patients showing improved chimerism, but there was substantial loss of NK cells during the purification steps, and donor NK cell persistence was not measured (Passweg et al. 2004).
442
J.E. Benjamin and S. Arai
The Minnesota group has reported the largest study to date of haploidentical NK cell infusion. Cell processing with T cell depletion alone resulted in a reduction of T cells in the infused product to around 0.95%, whereas there remained a significant proportion of moncytes (25%) and B lymphocytes (19%). Cells were cultured overnight in the presence of IL-2, and doses up to 2 × 107kg-1 were administered to 43 patients with advanced malignancies, including AML, melanoma, and renal carcinoma. Minimal toxicity was observed. Patients received chemotherapy prior to infusion and high-dose subcutaneous IL-2 for 14 days after. The results from this study indicated that donor NK cells persist in the recipient only if a high-dose conditioning regimen is given to achieve significant lymphodepletion and high IL-15 levels. Achievement of a complete remission in AML patients correlated with post-infusion NK cell counts. Furthermore, a predicted KIR ligand mismatch in the graft-versus-host direction was associated with the ability to induce CR (Miller et al. 2005). The Arkansas group has tested haploidentical NK cell infusion as a means to achieve tumor cytoreduction prior to autologous hematopoietic transplantation in ten patients with multiple myeloma. All had progressed following previous highdose chemotherapy. Conditioning chemotherapy included fludarabine, melphalan, and dexamethasone. PBMC from haploidentical donors, each with KIR ligand mismatching in the GVH direction, were depleted of CD3 and cultured overnight with IL-2. These products were infused on days 0 and +2, with subcutaneous IL-2 on days +1 to +11. On day +14, patients were infused with unmanipulated autologous G-CSF mobilized PBMC. The dose of NK cells was highly variable, ranging from 0.2 to 9 × 107 kg-1. Allogeneic chimerism was detectable in recipients until day +14, and MLR analysis demonstrated potent anti-donor activity by residual host T cells. Following autologous transplant, all patients engrafted, and five achieved complete remission. Unfortunately, all but one patient progressed in the follow-up period. In Vitro NK Cell Expansion All the studies described above demonstrated that NK cell infusions are associated with few side effects, albeit at doses at which significant anti-tumor effects could not be convincingly demonstrated. The next stage in this evolving technology is optimizing NK expansion ex vivo to achieve higher yields of NK cells demonstrating an activated phenotype. Childs and colleagues have reported that co-culture of CD3 depleted, CD56 enriched PBMC with irradiated EBV transformed lymphoblastoid cell lines, and IL-2 yielded 3,000× expansion within 28 days. The expanded cells were mostly CD56+ CD16+, KIR+ with robust NKG2D and TRAIL surface expression. Compared with fresh NK cells, cytotoxicity toward renal cell carcinoma cells was increased by eightfold. Unfortunately, cryopreservation and thawing of expanded NK cells resulted in reduced cytotoxicity and viability that was only partially reversed by culture in the presence of IL-2 (Berg et al. 2009). NK cells can be stimulated to proliferate by the human leukemic cell line K562 (Robertson et al. 1996). Campana and colleagues modified K562 cells by the
21 Natural Killer Cells for Adoptive Immunotherapy
443
introduction of the ligand for the TNF receptor family member 4-1BB (CD137) and membrane bound IL-15 (Imai et al. 2005). Co-culture of unsorted peripheral blood mononuclear cells with irradiated K562-mb15-41BBL in the presence of low-dose IL-2 resulted in a 300–12,000-fold expansion of NK cells over three weeks. Importantly, no expansion of T cells was observed. CD56 expression was increased, as were KIR molecules and the natural cytotoxicity receptors NKp30 and NKp44. The NK repertoire, as assessed by KIR expression, reflected that of the input cells, suggesting that no repertoire skewing had occurred by this culture method system. Expanded NK cells were efficiently transduced with viral vectors encoding chimeric antigen receptors. Alternative NK Cell Sources Umbilical cord blood is increasingly used as a stem cell source for patients requiring hematopoietic cell transplantation. Advantages of UCB include the less stringent requirement for HLA matching, given the relative paucity of mature T lymphocytes capable of causing graft-versus-host disease. A significant drawback of UCB is the low cell dose available, with a resultant delay in hematopoietic engraftment and immune reconstitution. Therefore, investigators have sought to expand lymphocyte subsets in vitro. Cairo and colleagues have described a 7-day culture of thawed UCB mononuclear cells in the presence of anti-CD3, Il-2, Il-7, and IL-10 (Ayello et al. 2009). Under these conditions, the total cell number expanded tenfold, with a large proportion (approximately 70%) representing CD56+CD3+ cells (40-fold expansion) and the remainder CD56+CD3- cells (40-fold). Compared with UCB cells cultured under these conditions for 2 days, those cultured for 7 days had increased NKG2D, LAMP-1, and granzyme B expression, as well as moderately increased cytotoxicity toward target cell lines. The advantage of expanding CD56+CD3+ cells remains unclear, although these cells are believed to mitigate the development of GVHD. Functional NK cells may also be derived from human embryonic stem cell lines (hESC). Kaufman and colleagues have recently reported the functional competence of such cells in clearing multiple types of human tumors in xenografted mice (Woll et al. 2009). It is anticipated that similar advances will be made with inducible pluripotent stem cells that may be derived from the somatic tissue of any donor. NK Cell Lines The use of individual donors for the purpose of NK cell expansion and infusion has multiple disadvantages, including the need to perform screening and apheresis of healthy donors, the variability in NK cell expansion efficiency and kinetics, and the variable proportion of alloreactive NK subsets in both the starting material and final product. These problems are bypassed with the use of the cell line such as NK-92, derived from a patient with an NK cell lymphoma (Gong et al. 1994; Klingemann et al. 1996). NK-92 does not express KIR molecules and therefore is not inhibited
444
J.E. Benjamin and S. Arai
by Class I molecules on targets. It expresses the activating NCR molecules NKp30 and NKp44 as well as NKG2D. NK-92 is cytotoxic against a variety of cell lines, including those derived from patients with of leukemia, lymphoma, melanoma, and prostate and breast cancers. It can be propagated in vitro with IL-2 alone and can be efficiently transduced by retroviral vectors. A Phase I dose escalation trial of NK-92 cells has recently been reported (Arai et al. 2008). Of the 12 patients treated, 11 had renal cell carcinoma and 1 had metastatic melanoma. Patients were treated with three infusions of irradiated NK-92 cells, with one cohort reaching the highest predetermined cell dose of 3 × 109 m-2 cells per infusion. The infusions appeared to be safe, with some patients experiencing low-grade fevers. Two patients demonstrated objective evidence of a mixed or partial response, although all patients ultimately progressed. Engraftment and In Vivo Expansion of Adoptively Transferred NK Cells The ultimate therapeutic strategy of clinical protocols involving NK cell infusion may or may not require NK cell persistence; yet persistence is an important experimental readout, particularly for interpreting negative outcomes. As with many hematopoietic cell therapies, engraftment of infused NK cells requires “space,” likely involving as-yet undefined niche sites or cytokine availability. NK cell homeostasis and proliferative potential have been best investigated in the mouse (Jamieson et al. 2004). Transfer of congenic NK cells into a syngeneic, unmanipulated mouse does not result in meaningful expansion of the donor cells. Induction of lymphopenia in the recipient, however, promotes rapid IL-15-dependent donor NK cell proliferation (Koka et al. 2003). It is likely that the same conditions will be necessary in the human setting. Interestingly, and also with clinical import, donor cell expression of inhibitory receptors for recipient MHC molecules had no discernible impact on expansion. Adjunctive Strategies Genetic modification of cultured NK cells is technically feasible. NK cell target range may be broadened with expression of receptors for tumor-associated antigens not normally part of the NK receptor ligand repertoire (Kruschinski et al. 2008; Schirrmann and Pecher 2002; Uherek et al. 2002; Pegram et al. 2008). These chimeric receptors typically contain cytoplasmic tails that contain conserved NK cell signaling motifs. NK cell effector function may be enhanced by the enforced overexpression of signaling receptors, thereby potentially altering the balance of activating versus inhibitory receptors (Imai et al. 2005). The deeper understanding of the receptors that govern NK cell target recognition and cytolysis has also been followed by efforts to target them with antibodies. As has been shown in mouse models, NK cell recognition of tumors may be achieved by blockade of inhibitory MHC receptors. Interestingly, such inhibition does not
21 Natural Killer Cells for Adoptive Immunotherapy
445
break self-tolerance (Sola et al. 2009). Currently, an antibody that blocks KIR, 1-7F9, is being evaluated in multiple cancer trials (Romagne et al. 2009). Another approach has been to redirect NK cells to lyse otherwise-refractory targets by use of bispecific antibodies such as for HER2/neu and CD16 (Shahied et al. 2004). Host Factors It is necessary to consider recipient factors that may diminish NK cell efficacy. Immune escape mechanisms employed by cancer cells are particularly important. For example, soluble HLA class I molecules, present in the serum of healthy individuals, are increased in patients with malignant diseases (Campoli and Ferrone 2008). Such molecules have been reported to decrease NK cytotoxicity and trigger apoptosis (Ghio et al. 2009; Spaggiari et al. 2002). Likewise, soluble forms of the NKG2D ligands MICA and MICB and the ULBPs are also found in the serum of healthy individuals. Elevated levels, as has been shown in serum samples from patients with diverse malignancies, downregulate NKG2D expression (Doubrovina et al. 2003; Groh et al. 2002; Kaiser et al. 2007; Song et al. 2006; Wu et al. 2004). The recipient immune system may also present a barrier to the effective deployment of donor NK cells. Naturally occurring regulatory T (Treg) cells represent 5–10% of all CD4 T cells in humans and downregulate NK cell effector function by membrane-bound TGF-beta (Ghiringhelli et al. 2005; Smyth et al. 2006). Direct killing of NK cells by Tregs has been reported (Cao et al. 2007). Similarly, the newly described population of myeloid-derived suppressor cells has also been implicated in the downregulation of NK cell killing (Li et al. 2009). Elimination of these regulatory subsets may be an effective avenue to boost NK responses against cancer (Kottke et al. 2008).
Summary NK cell infusion appears to be safe, with encouraging therapeutic potential. It will be important to refine methods of ex vivo expansion and then to validate these techniques in both dose-escalation and efficacy trials. Growing knowledge of the functional heterogeneity of NK cells may lead to a conclusion that a particular subset of cells should be targeted for expansion and infusion. Likewise, the increasing ease with which lymphocytes may be genetically modified may allow for augmented NK cell effector function with a broadened range of potential targets. Ideally, future trials should track NK cell engraftment, persistence, and expansion, all of which appear to require an available niche as defined by lymphopenia and high levels of homeostatic cyokines, such as IL-15. The safest conditioning regimens to achieve these responses have yet to be defined. Adjunctive strategies to enhance activating ligand expression or to deplete immune barriers posed by regulatory cells will likely constitute essential components of the optimal preparative regimens.
446
J.E. Benjamin and S. Arai
References Alter G, Malenfant JM, Altfeld M (2004) CD107a as a functional marker for the identification of natural killer cell activity. J Immunol Methods 294:15–22 Anfossi N, Andre P, Guia S, Falk CS, Roetynck S, Stewart CA, Breso V, Frassati C, Reviron D, Middleton D et al (2006) Human NK cell education by inhibitory receptors for MHC class I. Immunity 25:331–342 Arai S, Meagher R, Swearingen M, Myint H, Rich E, Martinson J, Klingemann H (2008) Infusion of the allogeneic cell line NK-92 in patients with advanced renal cell cancer or melanoma: a phase I trial. Cytotherapy 10:625–632 Arase H, Arase N, Saito T (1995) Fas-mediated cytotoxicity by freshly isolated natural killer cells. J Exp Med 181:1235–1238 Aversa F, Tabilio A, Terenzi A, Velardi A, Falzetti F, Giannoni C, Iacucci R, Zei T, Martelli MP, Gambelunghe C et al (1994) Successful engraftment of T-cell-depleted haploidentical “three-loci” incompatible transplants in leukemia patients by addition of recombinant human granulocyte colony-stimulating factor-mobilized peripheral blood progenitor cells to bone marrow inoculum. Blood 84:3948–3955 Aversa F, Tabilio A, Velardi A, Cunningham I, Terenzi A, Falzetti F, Ruggeri L, Barbabietola G, Aristei C, Latini P et al (1998) Treatment of high-risk acute leukemia with T-cell-depleted stem cells from related donors with one fully mismatched HLA haplotype. N Engl J Med 339:1186–1193 Ayello J, van de Ven C, Cairo E, Hochberg J, Baxi L, Satwani P, Cairo MS (2009) Characterization of natural killer and natural killer-like T cells derived from ex vivo expanded and activated cord blood mononuclear cells: implications for adoptive cellular immunotherapy. Exp Hematol 37:1216–1229 Bari R, Bell T, Leung WH, Vong QP, Chan WK, Das Gupta N, Holladay M, Rooney B, Leung W (2009) Significant functional heterogeneity among KIR2DL1 alleles and a pivotal role of Arginine245. Blood 114:5182–5190 Bauer S, Groh V, Wu J, Steinle A, Phillips JH, Lanier LL, Spies T (1999) Activation of NK cells and T cells by NKG2D, a receptor for stress-inducible MICA. Science 285:727–729 Baume DM, Robertson MJ, Levine H, Manley TJ, Schow PW, Ritz J (1992) Differential responses to interleukin 2 define functionally distinct subsets of human natural killer cells. Eur J Immunol 22:1–6 Beelen DW, Ottinger HD, Ferencik S, Elmaagacli AH, Peceny R, Trenschel R, Grosse-Wilde H (2005) Genotypic inhibitory killer immunoglobulin-like receptor ligand incompatibility enhances the long-term antileukemic effect of unmodified allogeneic hematopoietic stem cell transplantation in patients with myeloid leukemias. Blood 105:2594–2600 Berg M, Lundqvist A, McCoy P, Jr., Samsel L, Fan Y, Tawab A, Childs R (2009) Clinical-grade ex vivo-expanded human natural killer cells up-regulate activating receptors and death receptor ligands and have enhanced cytolytic activity against tumor cells. Cytotherapy 11:341–355 Boehm U, Klamp T, Groot M, Howard JC (1997) Cellular responses to interferon-gamma. Annu Rev Immunol 15: 749–795 Bottino C, Castriconi R, Moretta L, Moretta A (2005) Cellular ligands of activating NK receptors. Trends Immunol 26:221–226 Brunstein CG, Wagner JE, Weisdorf DJ, Cooley S, Noreen H, Barker JN, DeFor T, Verneris MR, Blazar BR, Miller JS (2009) Negative effect of KIR alloreactivity in recipients of umbilical cord blood transplant depends on transplantation conditioning intensity. Blood 113: 5628–5634 Bryceson YT, Ljunggren HG, Long EO (2009) Minimal requirement for induction of natural cytotoxicity and intersection of activation signals by inhibitory receptors. Blood 114:2657–2666 Bryceson YT, March ME, Barber DF, Ljunggren HG, Long EO (2005) Cytolytic granule polarization and degranulation controlled by different receptors in resting NK cells. J Exp Med 202: 1001–1012
21 Natural Killer Cells for Adoptive Immunotherapy
447
Burkett PR, Koka R, Chien M, Chai S, Boone DL, Ma A (2004) Coordinate expression and trans presentation of interleukin (IL)-15Ralpha and IL-15 supports natural killer cell and memory CD8+ T cell homeostasis. J Exp Med 200:825–834 Burt BM, Plitas G, Zhao Z, Bamboat ZM, Nguyen HM, Dupont B, DeMatteo RP (2009) The lytic potential of human liver NK cells is restricted by their limited expression of inhibitory killer Ig-like receptors. J Immunol 183:1789–1796 Caligiuri MA, Murray C, Robertson MJ, Wang E, Cochran K, Cameron C, Schow P, Ross ME, Klumpp TR, Soiffer RJ et al (1993) Selective modulation of human natural killer cells in vivo after prolonged infusion of low dose recombinant interleukin 2. J Clin Invest 91:123–132 Campoli M, Ferrone S (2008) Tumor escape mechanisms: potential role of soluble HLA antigens and NK cells activating ligands. Tissue Antigens 72:321–334 Cao X, Cai SF, Fehniger TA, Song J, Collins LI, Piwnica-Worms DR, Ley TJ (2007) Granzyme B and perforin are important for regulatory T cell-mediated suppression of tumor clearance. Immunity 27:635–646 Carlsten M, Malmberg KJ, Ljunggren HG (2009) Natural killer cell-mediated lysis of freshly isolated human tumor cells. Int J Cancer 124:757–762 Carson WE, Fehniger TA, Caligiuri MA (1997) CD56bright natural killer cell subsets: characterization of distinct functional responses to interleukin-2 and the c-kit ligand. Eur J Immunol 27:354–360 Cella M, Fuchs A, Vermi W, Facchetti F, Otero K, Lennerz JK, Doherty JM, Mills JC, Colonna M (2009) A human natural killer cell subset provides an innate source of IL-22 for mucosal immunity. Nature 457:722–725 Chan A, Hong DL, Atzberger A, Kollnberger S, Filer AD, Buckley CD, McMichael A, Enver T, Bowness P (2007) CD56bright human NK cells differentiate into CD56dim cells: role of contact with peripheral fibroblasts. J Immunol 179:89–94 Coca S, Perez-Piqueras J, Martinez D, Colmenarejo A, Saez MA, Vallejo C, Martos JA, Moreno M (1997) The prognostic significance of intratumoral natural killer cells in patients with colorectal carcinoma. Cancer 79:2320–2328 Colucci F, Caligiuri MA, Di Santo JP (2003) What does it take to make a natural killer? Nat Rev Immunol 3:413–425 Cooley S, McCullar V, Wangen R, Bergemann TL, Spellman S, Weisdorf DJ, Miller JS (2005) KIR reconstitution is altered by T cells in the graft and correlates with clinical outcomes after unrelated donor transplantation. Blood 106:4370–4376 Cooley S, Xiao F, Pitt M, Gleason M, McCullar V, Bergemann TL, McQueen KL, Guethlein LA, Parham P, Miller JS (2007) A subpopulation of human peripheral blood NK cells that lacks inhibitory receptors for self-MHC is developmentally immature. Blood 110:578–586 Cooper MA, Bush JE, Fehniger TA, VanDeusen JB, Waite RE, Liu Y, Aguila HL, Caligiuri MA (2002) In vivo evidence for a dependence on interleukin 15 for survival of natural killer cells. Blood 100:3633–3638 Cooper MA, Fehniger TA, Turner SC, Chen KS, Ghaheri BA, Ghayur T, Carson WE, Caligiuri MA (2001) Human natural killer cells: a unique innate immunoregulatory role for the CD56(bright) subset. Blood 97:3146–3151 Cosman D, Mullberg J, Sutherland CL, Chin W, Armitage R, Fanslow W, Kubin M, Chalupny NJ (2001) ULBPs, novel MHC class I-related molecules, bind to CMV glycoprotein UL16 and stimulate NK cytotoxicity through the NKG2D receptor. Immunity 14:123–133 Croy BA, Ashkar AA, Foster RA, DiSanto JP, Magram J, Carson D, Gendler SJ, Grusby MJ, Wagner N, Muller W et al (1997) Histological studies of gene-ablated mice support important functional roles for natural killer cells in the uterus during pregnancy. J Reprod Immunol 35:111–133 Cudkowicz G, Stimpfling JH (1964) Deficient Growth of C57bl Marrow Cells Transplanted in F1 Hybrid Mice. Association with the Histocompatibility-2 Locus. Immunology 7:291–306 Culley FJ (2009) Natural killer cells in infection and inflammation of the lung. Immunology 128:151–163 Daldrup-Link HE, Meier R, Rudelius M, Piontek G, Piert M, Metz S, Settles M, Uherek C, Wels W, Schlegel J et al (2005) In vivo tracking of genetically engineered, anti-HER2/neu directed
448
J.E. Benjamin and S. Arai
natural killer cells to HER2/neu positive mammary tumors with magnetic resonance imaging. Eur Radiol 15:4–13 Davies SM, Ruggieri L, DeFor T, Wagner JE, Weisdorf DJ, Miller JS, Velardi A, Blazar BR (2002) Evaluation of KIR ligand incompatibility in mismatched unrelated donor hematopoietic transplants. Killer immunoglobulin-like receptor. Blood 100:3825–3827 Diermayr S, Himmelreich H, Durovic B, Mathys-Schneeberger A, Siegler U, Langenkamp U, Hofsteenge J, Gratwohl A, Tichelli A, Paluszewska M et al (2008) NKG2D ligand expression in AML increases in response to HDAC inhibitor valproic acid and contributes to allorecognition by NK-cell lines with single KIR-HLA class I specificities. Blood 111: 1428–1436 Doubrovina ES, Doubrovin MM, Vider E, Sisson RB, O’Reilly RJ, Dupont B, Vyas YM (2003) Evasion from NK cell immunity by MHC class I chain-related molecules expressing colon adenocarcinoma. J Immunol 171:6891–6899 Dubois S, Mariner J, Waldmann TA, Tagaya Y (2002) IL-15Ralpha recycles and presents IL-15 In trans to neighboring cells. Immunity 17:537–547 Eisenbeis CF, Grainger A, Fischer B, Baiocchi RA, Carrodeguas L, Roychowdhury S, Chen L, Banks AL, Davis T, Young D et al (2004) Combination immunotherapy of B-cell non-Hodgkin’s lymphoma with rituximab and interleukin-2: a preclinical and phase I study. Clin Cancer Res 10:6101–6110 Farag SS, Bacigalupo A, Eapen M, Hurley C, Dupont B, Caligiuri MA, Boudreau C, Nelson G, Oudshoorn M, van Rood J et al (2006) The effect of KIR ligand incompatibility on the outcome of unrelated donor transplantation: a report from the center for international blood and marrow transplant research, the European blood and marrow transplant registry, and the Dutch registry. Biol Blood Marrow Transplant 12:876–884 Fehniger TA, Cooper MA, Nuovo GJ, Cella M, Facchetti F, Colonna M, Caligiuri MA (2003) CD56bright natural killer cells are present in human lymph nodes and are activated by T cellderived IL-2: a potential new link between adaptive and innate immunity. Blood 101: 3052–3057 Ferlazzo G, Thomas D, Lin SL, Goodman K, Morandi B, Muller WA, Moretta A, Munz C (2004) The abundant NK cells in human secondary lymphoid tissues require activation to express killer cell Ig-like receptors and become cytolytic. J Immunol 172:1455–1462 Ferlazzo G, Tsang ML, Moretta L, Melioli G, Steinman RM, Munz C (2002) Human dendritic cells activate resting natural killer (NK) cells and are recognized via the NKp30 receptor by activated NK cells. J Exp Med 195:343–351 Fernandez NC, Lozier A, Flament C, Ricciardi-Castagnoli P, Bellet D, Suter M, Perricaudet M, Tursz T, Maraskovsky E, Zitvogel L (1999) Dendritic cells directly trigger NK cell functions: cross-talk relevant in innate anti-tumor immune responses in vivo. Nat Med 5:405–411 Fleming GF, Meropol NJ, Rosner GL, Hollis DR, Carson WE 3rd, Caligiuri M, Mortimer J, Tkaczuk K, Parihar R, Schilsky RL et al (2002) A phase I trial of escalating doses of trastuzumab combined with daily subcutaneous interleukin 2: report of cancer and leukemia group B 9661. Clin Cancer Res 8:3718–3727 Freud AG, Becknell B, Roychowdhury S, Mao HC, Ferketich AK, Nuovo GJ, Hughes TL, Marburger TB, Sung J, Baiocchi RA et al (2005) A human CD34(+) subset resides in lymph nodes and differentiates into CD56bright natural killer cells. Immunity 22:295–304 Frey M, Packianathan NB, Fehniger TA, Ross ME, Wang WC, Stewart CC, Caligiuri MA, Evans SS (1998) Differential expression and function of L-selectin on CD56bright and CD56dim natural killer cell subsets. J Immunol 161:400–408 Galy A, Travis M, Cen D, Chen B (1995) Human T, B, natural killer, and dendritic cells arise from a common bone marrow progenitor cell subset. Immunity 3:459–473 Gasser S, Orsulic S, Brown EJ, Raulet DH (2005) The DNA damage pathway regulates innate immune system ligands of the NKG2D receptor. Nature 436:1186–1190 Gerosa F, Baldani-Guerra B, Nisii C, Marchesini V, Carra G, Trinchieri G (2002) Reciprocal activating interaction between natural killer cells and dendritic cells. J Exp Med 195: 327–333
21 Natural Killer Cells for Adoptive Immunotherapy
449
Ghio M, Contini P, Negrini S, Boero S, Musso A, Poggi A (2009) Soluble HLA-I-mediated secretion of TGF-beta1 by human NK cells and consequent downregulation of anti-tumor cytolytic activity. Eur J Immunol 39:3459–3468 Ghiringhelli F, Menard C, Terme M, Flament C, Taieb J, Chaput N, Puig PE, Novault S, Escudier B, Vivier E et al (2005) CD4+CD25+ regulatory T cells inhibit natural killer cell functions in a transforming growth factor-beta-dependent manner. J Exp Med 202:1075–1085 Giebel S, Locatelli F, Lamparelli T, Velardi A, Davies S, Frumento G, Maccario R, Bonetti F, Wojnar J, Martinetti M et al (2003) Survival advantage with KIR ligand incompatibility in hematopoietic stem cell transplantation from unrelated donors. Blood 102:814–819 Gong JH, Maki G, Klingemann HG (1994) Characterization of a human cell line (NK-92) with phenotypical and functional characteristics of activated natural killer cells. Leukemia 8: 652–658 Groh V, Wu J, Yee C, Spies T (2002) Tumour-derived soluble MIC ligands impair expression of NKG2D and T-cell activation. Nature 419:734–738 Guerra N, Tan YX, Joncker NT, Choy A, Gallardo F, Xiong N, Knoblaugh S, Cado D, Greenberg NM, Raulet DH (2008) NKG2D-deficient mice are defective in tumor surveillance in models of spontaneous malignancy. Immunity 28:571–580 Herberman RB, Nunn ME, Holden HT, Lavrin DH (1975) Natural cytotoxic reactivity of mouse lymphoid cells against syngeneic and allogeneic tumors. II. Characterization of effector cells. Int J Cancer 16:230–239 Hsu KC, Chida S, Geraghty DE, Dupont B (2002) The killer cell immunoglobulin-like receptor (KIR) genomic region: gene-order, haplotypes and allelic polymorphism. Immunol Rev 190: 40–52 Hyakudomi M, Matsubara T, Hyakudomi R, Yamamoto T, Kinugasa S, Yamanoi A, Maruyama R, Tanaka T (2008) Increased expression of fractalkine is correlated with a better prognosis and an increased number of both CD8+ T cells and natural killer cells in gastric adenocarcinoma. Ann Surg Oncol 15:1775–1782 Imai C, Iwamoto S, Campana D (2005) Genetic modification of primary natural killer cells overcomes inhibitory signals and induces specific killing of leukemic cells. Blood 106:376–383 Ishigami S, Natsugoe S, Tokuda K, Nakajo A, Che X, Iwashige H, Aridome K, Hokita S, Aikou T (2000) Prognostic value of intratumoral natural killer cells in gastric carcinoma. Cancer 88:577–583 Iyengar R, Handgretinger R, Babarin-Dorner A, Leimig T, Otto M, Geiger TL, Holladay MS, Houston J, Leung W (2003) Purification of human natural killer cells using a clinical-scale immunomagnetic method. Cytotherapy 5:479–484 Jacobs R, Hintzen G, Kemper A, Beul K, Kempf S, Behrens G, Sykora KW, Schmidt RE (2001) CD56bright cells differ in their KIR repertoire and cytotoxic features from CD56dim NK cells. Eur J Immunol 31:3121–3127 Jamieson AM, Isnard P, Dorfman JR, Coles MC, Raulet DH (2004) Turnover and proliferation of NK cells in steady state and lymphopenic conditions. J Immunol 172:864–870 Jenne CN, Enders A, Rivera R, Watson SR, Bankovich AJ, Pereira JP, Xu Y, Roots CM, Beilke JN, Banerjee A et al (2009) T-bet-dependent S1P5 expression in NK cells promotes egress from lymph nodes and bone marrow. J Exp Med 206:2469–2481 Kaiser BK, Yim D, Chow IT, Gonzalez S, Dai Z, Mann HH, Strong RK, Groh V, Spies T (2007) Disulphide-isomerase-enabled shedding of tumour-associated NKG2D ligands. Nature 447: 482–486 Karre K, Ljunggren HG, Piontek G, Kiessling R (1986) Selective rejection of H-2-deficient lymphoma variants suggests alternative immune defence strategy. Nature 319:675–678 Kashii Y, Giorda R, Herberman RB, Whiteside TL, Vujanovic NL (1999) Constitutive expression and role of the TNF family ligands in apoptotic killing of tumor cells by human NK cells. J Immunol 163:5358–5366 Kennedy MK, Glaccum M, Brown SN, Butz EA, Viney JL, Embers M, Matsuki N, Charrier K, Sedger L, Willis CR et al (2000) Reversible defects in natural killer and memory CD8 T cell lineages in interleukin 15-deficient mice. J Exp Med 191:771–780
450
J.E. Benjamin and S. Arai
Kiessling R, Klein E, Pross H, Wigzell H (1975) “Natural” killer cells in the mouse. II. Cytotoxic cells with specificity for mouse Moloney leukemia cells. Characteristics of the killer cell. Eur J Immunol 5:117–121 Kim S, Poursine-Laurent J, Truscott SM, Lybarger L, Song YJ, Yang L, French AR, Sunwoo JB, Lemieux S, Hansen TH et al (2005) Licensing of natural killer cells by host major histocompatibility complex class I molecules. Nature 436:709–713 Klingemann HG, Wong E, Maki G (1996) A cytotoxic NK-cell line (NK-92) for ex vivo purging of leukemia from blood. Biol Blood Marrow Transplant 2:68–75 Koehl U, Esser R, Zimmermann S, Tonn T, Kotchetkov R, Bartling T, Sorensen J, Gruttner HP, Bader P, Seifried E et al (2005) Ex vivo expansion of highly purified NK cells for immunotherapy after haploidentical stem cell transplantation in children. Klin Padiatr 217:345–350 Koka R, Burkett P, Chien M, Chai S, Boone DL, Ma A (2004) Cutting edge: murine dendritic cells require IL-15R alpha to prime NK cells. J Immunol 173:3594–3598 Koka R, Burkett PR, Chien M, Chai S, Chan F, Lodolce JP, Boone DL, Ma A (2003) Interleukin (IL)-15R[alpha]-deficient natural killer cells survive in normal but not IL-15R[alpha]-deficient mice. J Exp Med 197:977–984 Kottke T, Galivo F, Wongthida P, Diaz RM, Thompson J, Jevremovic D, Barber GN, Hall G, Chester J, Selby P et al (2008) Treg depletion-enhanced IL-2 treatment facilitates therapy of established tumors using systemically delivered oncolytic virus. Mol Ther 16:1217–1226 Kruschinski A, Moosmann A, Poschke I, Norell H, Chmielewski M, Seliger B, Kiessling R, Blankenstein T, Abken H, Charo J (2008) Engineering antigen-specific primary human NK cells against HER-2 positive carcinomas. Proc Natl Acad Sci USA 105:17481–17486 Lanier LL (2009) DAP10- and DAP12-associated receptors in innate immunity. Immunol Rev 227:150–160 Lanier LL, Chang C, Azuma M, Ruitenberg JJ, Hemperly JJ, Phillips JH (1991) Molecular and functional analysis of human natural killer cell-associated neural cell adhesion molecule (N-CAM/CD56). J Immunol 146:4421–4426 Li H, Han Y, Guo Q, Zhang M, Cao X (2009) Cancer-expanded myeloid-derived suppressor cells induce anergy of NK cells through membrane-bound TGF-beta 1. J Immunol 182:240–249 Lodolce JP, Boone DL, Chai S, Swain RE, Dassopoulos T, Trettin S, Ma A (1998) IL-15 receptor maintains lymphoid homeostasis by supporting lymphocyte homing and proliferation. Immunity 9:669–676 Lodolce JP, Burkett PR, Boone DL, Chien M, Ma A (2001) T cell-independent interleukin 15Ralpha signals are required for bystander proliferation. J Exp Med 194:1187–1194 Lowe EJ, Turner V, Handgretinger R, Horwitz EM, Benaim E, Hale GA, Woodard P, Leung W (2003) T-cell alloreactivity dominates natural killer cell alloreactivity in minimally T-celldepleted HLA-non-identical paediatric bone marrow transplantation. Br J Haematol 123: 323–326 Lundqvist A, McCoy JP, Samsel L, Childs R (2007) Reduction of GVHD and enhanced antitumor effects after adoptive infusion of alloreactive Ly49-mismatched NK cells from MHC-matched donors. Blood 109:3603–3606 McKenna DH, Jr., Sumstad D, Bostrom N, Kadidlo DM, Fautsch S, McNearney S, Dewaard R, McGlave PB, Weisdorf DJ, Wagner JE et al (2007) Good manufacturing practices production of natural killer cells for immunotherapy: a six-year single-institution experience. Transfusion 47:520–528 Meier R, Piert M, Piontek G, Rudelius M, Oostendorp RA, Senekowitsch-Schmidtke R, Henning TD, Wels WS, Uherek C, Rummeny EJ et al (2008) Tracking of [18F]FDG-labeled natural killer cells to HER2/neu-positive tumors. Nucl Med Biol 35:579–588 Meyer-Monard S, Passweg J, Siegler U, Kalberer C, Koehl U, Rovo A, Halter J, Stern M, Heim D, Alois Gratwohl JR et al (2009) Clinical-grade purification of natural killer cells in haploidentical hematopoietic stem cell transplantation. Transfusion 49:362–371
21 Natural Killer Cells for Adoptive Immunotherapy
451
Miller JS, Alley KA, McGlave P (1994) Differentiation of natural killer (NK) cells from human primitive marrow progenitors in a stroma-based long-term culture system: identification of a CD34+7+ NK progenitor. Blood 83:2594–2601 Miller JS, Soignier Y, Panoskaltsis-Mortari A, McNearney SA, Yun GH, Fautsch SK, McKenna D, Le C, Defor TE, Burns LJ et al (2005) Successful adoptive transfer and in vivo expansion of human haploidentical NK cells in patients with cancer. Blood 105:3051–3057 Nguyen S, Dhedin N, Vernant JP, Kuentz M, Al Jijakli A, Rouas-Freiss N, Carosella ED, Boudifa A, Debre P, Vieillard V (2005) NK-cell reconstitution after haploidentical hematopoietic stemcell transplantations: immaturity of NK cells and inhibitory effect of NKG2A override GvL effect. Blood 105:4135–4142 Nishimura M, Umehara H, Nakayama T, Yoneda O, Hieshima K, Kakizaki M, Dohmae N, Yoshie O, Imai T (2002) Dual functions of fractalkine/CX3C ligand 1 in trafficking of perforin+/granzyme B+ cytotoxic effector lymphocytes that are defined by CX3CR1 expression. J Immunol 168:6173–6180 Ohta M, Tanaka F, Yamaguchi H, Sadanaga N, Inoue H, Mori M (2005) The high expression of Fractalkine results in a better prognosis for colorectal cancer patients. Int J Oncol 26: 41–47 Oppenheim DE, Roberts SJ, Clarke SL, Filler R, Lewis JM, Tigelaar RE, Girardi M, Hayday AC (2005) Sustained localized expression of ligand for the activating NKG2D receptor impairs natural cytotoxicity in vivo and reduces tumor immunosurveillance. Nat Immunol 6:928–937 Oshimi Y, Oda S, Honda Y, Nagata S, Miyazaki S (1996) Involvement of Fas ligand and Fasmediated pathway in the cytotoxicity of human natural killer cells. J Immunol 157:2909–2915 Parolini S, Santoro A, Marcenaro E, Luini W, Massardi L, Facchetti F, Communi D, Parmentier M, Majorana A, Sironi M et al (2007) The role of chemerin in the colocalization of NK and dendritic cell subsets into inflamed tissues. Blood 109:3625–3632 Passweg JR, Tichelli A, Meyer-Monard S, Heim D, Stern M, Kuhne T, Favre G, Gratwohl A (2004) Purified donor NK-lymphocyte infusion to consolidate engraftment after haploidentical stem cell transplantation. Leukemia 18:1835–1838 Pegram HJ, Jackson JT, Smyth MJ, Kershaw MH, Darcy PK (2008) Adoptive transfer of genemodified primary NK cells can specifically inhibit tumor progression in vivo. J Immunol 181:3449–3455 Pistoia V, Zupo S, Corcione A, Roncella S, Matera L, Ghio R, Ferrarini M (1989) Production of colony-stimulating activity by human natural killer cells: analysis of the conditions that influence the release and detection of colony-stimulating activity. Blood 74:156–164 Poggi A, Catellani S, Garuti A, Pierri I, Gobbi M, Zocchi MR (2009) Effective in vivo induction of NKG2D ligands in acute myeloid leukaemias by all-trans-retinoic acid or sodium valproate. Leukemia 23:641–648 Ranson T, Vosshenrich CA, Corcuff E, Richard O, Muller W, Di Santo JP (2003) IL-15 is an essential mediator of peripheral NK-cell homeostasis. Blood 101:4887–4893 Repka T, Chiorean EG, Gay J, Herwig KE, Kohl VK, Yee D, Miller JS (2003) Trastuzumab and interleukin-2 in HER2-positive metastatic breast cancer: a pilot study. Clin Cancer Res 9:2440–2446 Robertson MJ, Cameron C, Lazo S, Cochran KJ, Voss SD, Ritz J (1996) Costimulation of human natural killer cell proliferation: role of accessory cytokines and cell contact-dependent signals. Nat Immunol 15:213–226 Romagnani C, Juelke K, Falco M, Morandi B, D’Agostino A, Costa R, Ratto G, Forte G, Carrega P, Lui G et al (2007) CD56brightCD16- killer Ig-like receptor- NK cells display longer telomeres and acquire features of CD56dim NK cells upon activation. J Immunol 178:4947–4955 Romagne F, Andre P, Spee P, Zahn S, Anfossi N, Gauthier L, Capanni M, Ruggeri L, Benson DM, Jr., Blaser BW et al (2009) Pre-clinical characterization of 1–7F9, a novel human anti-KIR therapeutic antibody that augments NK-mediated killing of tumor cells. Blood 114: 2667–2677
452
J.E. Benjamin and S. Arai
Romanski A, Bug G, Becker S, Kampfmann M, Seifried E, Hoelzer D, Ottmann OG, Tonn T (2005) Mechanisms of resistance to natural killer cell-mediated cytotoxicity in acute lymphoblastic leukemia. Exp Hematol 33:344–352 Rosenberg SA, Lotze MT, Muul LM, Leitman S, Chang AE, Ettinghausen SE, Matory YL, Skibber JM, Shiloni E, Vetto JT et al (1985) Observations on the systemic administration of autologous lymphokine-activated killer cells and recombinant interleukin-2 to patients with metastatic cancer. N Engl J Med 313:1485–1492 Ruggeri L, Capanni M, Urbani E, Perruccio K, Shlomchik WD, Tosti A, Posati S, Rogaia D, Frassoni F, Aversa F et al (2002) Effectiveness of donor natural killer cell alloreactivity in mismatched hematopoietic transplants. Science 295:2097–2100 Ruggeri L, Mancusi A, Capanni M, Urbani E, Carotti A, Aloisi T, Stern M, Pende D, Perruccio K, Burchielli E et al (2007) Donor natural killer cell allorecognition of missing self in haploidentical hematopoietic transplantation for acute myeloid leukemia: challenging its predictive value. Blood 110:433–440 Sanos SL, Bui VL, Mortha A, Oberle K, Heners C, Johner C, Diefenbach A (2009) RORgammat and commensal microflora are required for the differentiation of mucosal interleukin 22-producing NKp46+ cells. Nat Immunol 10:83–91 Satoh-Takayama N, Vosshenrich CA, Lesjean-Pottier S, Sawa S, Lochner M, Rattis F, Mention JJ, Thiam K, Cerf-Bensussan N, Mandelboim O et al (2008) Microbial flora drives interleukin 22 production in intestinal NKp46+ cells that provide innate mucosal immune defense. Immunity 29:958–970 Schirrmann T, Pecher G (2002) Human natural killer cell line modified with a chimeric immunoglobulin T-cell receptor gene leads to tumor growth inhibition in vivo. Cancer Gene Ther 9:390–398 Schroder K, Hertzog PJ, Ravasi T, Hume DA (2004) Interferon-gamma: an overview of signals, mechanisms and functions. J Leukoc Biol 75:163–189 Schulze A, Schirutschke H, Oelschlagel U, Schmitz M, Fussel M, Wassmuth R, Ehninger G, Bornhauser M, Platzbecker U (2008) Altered phenotype of natural killer cell subsets after haploidentical stem cell transplantation. Exp Hematol 36:378–389 Shahied LS, Tang Y, Alpaugh RK, Somer R, Greenspon D, Weiner LM (2004) Bispecific minibodies targeting HER2/neu and CD16 exhibit improved tumor lysis when placed in a divalent tumor antigen binding format. J Biol Chem 279:53907–53914 Shi J, Tricot G, Szmania S, Rosen N, Garg TK, Malaviarachchi PA, Moreno A, Dupont B, Hsu KC, Baxter-Lowe LA et al (2008) Infusion of haplo-identical killer immunoglobulin-like receptor ligand mismatched NK cells for relapsed myeloma in the setting of autologous stem cell transplantation. Br J Haematol 143:641–653 Shilling HG, McQueen KL, Cheng NW, Shizuru JA, Negrin RS, Parham P (2003) Reconstitution of NK cell receptor repertoire following HLA-matched hematopoietic cell transplantation. Blood 101:3730–3740 Single RM, Martin MP, Gao X, Meyer D, Yeager M, Kidd JR, Kidd KK, Carrington M (2007) Global diversity and evidence for coevolution of KIR and HLA. Nat Genet 39:1114–1119 Smyth MJ, Cretney E, Takeda K, Wiltrout RH, Sedger LM, Kayagaki N, Yagita H, Okumura K (2001) Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) contributes to interferon gamma-dependent natural killer cell protection from tumor metastasis. J Exp Med 193:661–670 Smyth MJ, Swann J, Cretney E, Zerafa N, Yokoyama WM, Hayakawa Y (2005) NKG2D function protects the host from tumor initiation. J Exp Med 202:583–588 Smyth MJ, Teng MW, Swann J, Kyparissoudis K, Godfrey DI, Hayakawa Y (2006) CD4+CD25+ T regulatory cells suppress NK cell-mediated immunotherapy of cancer. J Immunol 176:1582–1587 Sobecks RM, Ball EJ, Askar M, Theil KS, Rybicki LA, Thomas D, Brown S, Kalaycio M, Andresen S, Pohlman B et al (2008) Influence of killer immunoglobulin-like receptor/HLA ligand matching on achievement of T-cell complete donor chimerism in related donor nonmyeloablative allogeneic hematopoietic stem cell transplantation. Bone Marrow Transplant 41: 709–714
21 Natural Killer Cells for Adoptive Immunotherapy
453
Soiffer RJ, Murray C, Shapiro C, Collins H, Chartier S, Lazo S, Ritz J (1996) Expansion and manipulation of natural killer cells in patients with metastatic cancer by low-dose continuous infusion and intermittent bolus administration of interleukin 2. Clin Cancer Res 2:493–499 Sola C, Andre P, Lemmers C, Fuseri N, Bonnafous C, Blery M, Wagtmann NR, Romagne F, Vivier E, Ugolini S (2009) Genetic and antibody-mediated reprogramming of natural killer cell missing-self recognition in vivo. Proc Natl Acad Sci USA 106:12879–12884 Song H, Kim J, Cosman D, Choi I (2006) Soluble ULBP suppresses natural killer cell activity via down-regulating NKG2D expression. Cell Immunol 239:22–30 Soriani A, Zingoni A, Cerboni C, Iannitto ML, Ricciardi MR, Di Gialleonardo V, Cippitelli M, Fionda C, Petrucci MT, Guarini A et al (2009) ATM-ATR-dependent up-regulation of DNAM-1 and NKG2D ligands on multiple myeloma cells by therapeutic agents results in enhanced NK-cell susceptibility and is associated with a senescent phenotype. Blood 113:3503–3511 Spaggiari GM, Contini P, Dondero A, Carosio R, Puppo F, Indiveri F, Zocchi MR, Poggi A (2002) Soluble HLA class I induces NK cell apoptosis upon the engagement of killer-activating HLA class I receptors through FasL-Fas interaction. Blood 100:4098–4107 Takahashi E, Kuranaga N, Satoh K, Habu Y, Shinomiya N, Asano T, Seki S, Hayakawa M (2007) Induction of CD16+ CD56bright NK cells with antitumour cytotoxicity not only from CD16- CD56bright NK Cells but also from CD16- CD56dim NK cells. Scand J Immunol 65: 126–138 Takanami I, Takeuchi K, Giga M (2001) The prognostic value of natural killer cell infiltration in resected pulmonary adenocarcinoma. J Thorac Cardiovasc Surg 121:1058–1063 Trinchieri G, Matsumoto-Kobayashi M, Clark SC, Seehra J, London L, Perussia B (1984) Response of resting human peripheral blood natural killer cells to interleukin 2. J Exp Med 160:1147–1169 Uherek C, Tonn T, Uherek B, Becker S, Schnierle B, Klingemann HG, Wels W (2002) Retargeting of natural killer-cell cytolytic activity to ErbB2-expressing cancer cells results in efficient and selective tumor cell destruction. Blood 100:1265–1273 Vago L, Forno B, Sormani MP, Crocchiolo R, Zino E, Di Terlizzi S, Lupo Stanghellini MT, Mazzi B, Perna SK, Bondanza A et al (2008) Temporal, quantitative, and functional characteristics of single-KIR-positive alloreactive natural killer cell recovery account for impaired graft-versusleukemia activity after haploidentical hematopoietic stem cell transplantation. Blood 112:3488–3499 Villegas FR, Coca S, Villarrubia VG, Jimenez R, Chillon MJ, Jareno J, Zuil M, Callol L (2002) Prognostic significance of tumor infiltrating natural killer cells subset CD57 in patients with squamous cell lung cancer. Lung Cancer 35:23–28 Vosshenrich CA, Garcia-Ojeda ME, Samson-Villeger SI, Pasqualetto V, Enault L, Richard-Le Goff O, Corcuff E, Guy-Grand D, Rocha B, Cumano A et al (2006) A thymic pathway of mouse natural killer cell development characterized by expression of GATA-3 and CD127. Nat Immunol 7:1217–1224 Walzer T, Chiossone L, Chaix J, Calver A, Carozzo C, Garrigue-Antar L, Jacques Y, Baratin M, Tomasello E, Vivier E (2007) Natural killer cell trafficking in vivo requires a dedicated sphingosine 1-phosphate receptor. Nat Immunol 8:1337–1344 Wang H, Grzywacz B, Sukovich D, McCullar V, Cao Q, Lee AB, Blazar BR, Cornfield DN, Miller JS, Verneris MR (2007) The unexpected effect of cyclosporin A on CD56+CD16- and CD56+CD16+ natural killer cell subpopulations. Blood 110:1530–1539 Willemze R, Rodrigues CA, Labopin M, Sanz G, Michel G, Socie G, Rio B, Sirvent A, Renaud M, Madero L et al (2009) KIR-ligand incompatibility in the graft-versus-host direction improves outcomes after umbilical cord blood transplantation for acute leukemia. Leukemia 23:492–500 Witte T, Wordelmann K, Schmidt RE (1990) Heterogeneity of human natural killer cells in the spleen. Immunology 69:166–170 Woll PS, Grzywacz B, Tian X, Marcus RK, Knorr DA, Verneris MR, Kaufman DS (2009) Human embryonic stem cells differentiate into a homogeneous population of natural killer cells with potent in vivo antitumor activity. Blood 113:6094–6101
454
J.E. Benjamin and S. Arai
Wu JD, Higgins LM, Steinle A, Cosman D, Haugk K, Plymate SR (2004) Prevalent expression of the immunostimulatory MHC class I chain-related molecule is counteracted by shedding in prostate cancer. J Clin Invest 114:560–568 Wu MF, Raulet DH (1997) Class I-deficient hemopoietic cells and nonhemopoietic cells dominantly induce unresponsiveness of natural killer cells to class I-deficient bone marrow cell grafts. J Immunol 158:1628–1633 Yang JC, Sherry RM, Steinberg SM, Topalian SL, Schwartzentruber DJ, Hwu P, Seipp CA, Rogers-Freezer L, Morton KE, White DE et al (2003) Randomized study of high-dose and low-dose interleukin-2 in patients with metastatic renal cancer. J Clin Oncol 21:3127–3132
Part V
Extracellular Matrix
Chapter 22
Fibronectin Andreas Menrad
Abstract Fibronectins multifunctionality is reflected by its complex modular molecular design consisting of distict domains which are highly conserved throughout the species. The domains are arranged in specific juxtapositions, sometimes controlled by highly regulated alternative splicing. The implication is that the complex and conserved architecture of fibronectin codes for specific information of high biological relevance. This chapter will review the role of fibronectin for tumor growth and discuss innovative therapeutic interventions.
The Tumor Stroma During early cancer development, neoplastic cells remodel their surrounding stroma, i.e. the basement membrane, immune cells, capillaries, fibroblasts and extracellular matrix (ECM) and form a microenvironment that is commonly referred to as “reactive stroma” which is similar to the stroma formed during wound healing (Dvorak et al. 1984; Ronnov-Jenssen et al. 1996). Until recently, most attention in cancer research was focused on the malignant tumor cells as the most important compartment of solid tumors. However, there is growing evidence that the complex stromal tumor compartment is equally important for the growth and progression of malignant disease (Wernert 1997; Pupa et al. 2002; Mueller and Fusenig 2004). The tumor-permissive matrix is generated both by proteolytic degradation of pre-existing ECM components by cancer cell-derived enzymes as well as the de-novo-synthesis of new tumor-associated ECM-molecules which may be secreted by tumor-, stromal- and endothelial cells. One of the most abundant ECM-molecules with critical regulatory roles in cellular processes such as adhesion, migration, differentiation and proliferation are the fibronectins (Hynes 1990). A. Menrad (*) Antibody Therapeutics Genzyme Europe Research, 310 Cambridge Science Park, Milton Road, Cambridge, CB4 OWG, UK e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_22, © Springer Science+Business Media, LLC 2010
457
458
A. Menrad
Fibronectin Molecular Structure of Fibronectin Fibronectin (FN) is a dimeric molecule consisting of similar, but not necessarily identical, chains of 220–250 kDa in size joined by a pair of disulphide bonds at the carboxy-terminus at the end of the polypeptide chain. Within each chain, there are approximately 30 intrachain disulphide bonds as well as two sulphydryl groups which can contribute to intermolecular bond-formation. FNs multifunctionality is reflected by a highly modular architecture consisting of a linear arrangement of repeating building blocks characterized as type I, II and III repeats. Type I repeats consist of ~45, type II of ~60 and type III of ~90 characteristic amino acids which have also been identified in molecules other than FNs. These modules, or domains, are resistant to proteolytic cleavage and contain binding sites for other ECMmolecules, FN-specific cell-surface receptors (integrins), the blood protein fibrin and glycosaminoglycans (heparin). Figure 22.1shows an overview on the molecular structure of FNs.
Gene Structure and FN-Knock Out FN is secreted after transcription of a single gene composed of 47 exons that spans over 90 kbp in the genome. Type III domains are encoded by two exons whereas Extra Domains A, B (EDA, EDB) and the ninth Type III repeat are encoded by a single exon. Alternative splicing of a single pre-mRNA can generate multiple isoforms including or excluding EDA and EDB (Schwarzbauer et al. 1983, 1987;
Cryptic binding sites Fibrin 1
NH2
2
3
4
5
6
7
B
8
9 10 11
A
12 13 14
III CS 15
COOH S
Matrix assembly Collagen Heparin Gelatine Fibrin
Cells Heparin Matrix assembly
type I module
B extradomain B
type II module
A extradomain A
Heparin Fibrin Chondroitin sulphate
type III module
Fig. 22.1 Modular structure of FN. The modular structure of FN, showing the relative positions of type I, II, and III repeat modules. The positions of the EDA, EDB and IIICS alternatively spliced regions are shown, as are binding regions for known biologic interaction partners. Two disulphide bridges at the COOH-termini link two monomers
22 Fibronectin
459
Kornblihtt et al. 1984; Gutman and Kornblihtt 1987). The Type III connecting segment (IIICS or variable region) shows a more complex splicing pattern. Depending on the species, it can be included, partially included or entirely excluded from the molecule. The complexity of variable region alternative splicing and the total number of FN-isoforms seems to correlate with evolution: form Zebrafish producing 5 FN-isoforms, chicken up to 6, rat and mice with 12 and up to 20 isoforms characterized in humans (Umezawa et al. 1986; Norton and Hynes 1987; DeSimone et al. 1992).
Synthesis and Matrix Assembly FN is secreted by a broad variety of cells. Soluble dimeric FN is predominantly produced by hepatocytes and enriched in the blood (pFN). pFN has a compact conformation mediated by intramolecular interactions and lacks significant expression of EDA or EDB domains (Johnson et al. 1999). Cellular FN (cFN) is secreted by numerous cells (fibroblasts, epithelial cells and macrophages) and it may contain variable amounts of both EDA and EDB as well as variable-region isoforms. In addition to alternative splicing, extensive post-transcriptional modifications occur during synthesis. These modifications include glycosylation, phosphorylation, sulfation and disulfide bond-formation. The assembly of a fibrillar FN matrix is a complex process and an area of active research. A major step towards the fibril formation is the activation of sFN induced by the interaction with the heterodimeric transmembrane a5b1-integrin receptor. This highly specific interaction of the integrin with its ligand FN requires the RGDcell binding sequence in the Type III-10 domain (Ruoslahti and Pierschbacher 1987; Hynes 1992). This interaction is further enhanced by the adjacent PHSRNsequence on domain 9 (Nagai et al. 1991; Bowditch et al. 1994). Function-blocking antibodies against either the cell binding domain of FN or the a5b1-integrin receptor inhibit FN fibril formation (McDonald et al. 1987; Fogerty et al. 1990). Initially, integrin-bound FN is diffusely distributed on the cell surface. Integrin receptorclustering via dimeric FN organizes FN into initial short fibrils. During this process, the cytoplasmic domains of integrins associate with the actin cytoskeleton of the cell. The contractibility of the cytoskeleton mediated by actin-myosin interactions is essential for fibiril formation and is part of the dynamic fibrillar adhesion signalling complexes consisting of a5b1-integrin receptors, focal adhesion kinase (FAK), vinculin, paxillin, tensin and FN (Sechler and Schwarzbauer 1997; Zamir et al. 2000; Geiger et al. 2001; Miranti and Brugge 2002). As FN goes from solution to fibrils, the molecule undergoes a conformational change from a compact form into an extended form. During this process, cryptic binding sites are exposed along the length of the molecule allowing new intra- and intermolecular interactions required for FN incorporation in the ECM (Fig. 22.2). At the same time, other epitopes may disappear.
460
A. Menrad
outside
a
cytoplasm
b
outside
cytoplasm
c
Outside
cytoplasm
Fig. 22.2 Major steps in FN fibril assembly. (a) Complact FN in solution binds to the a5b1integrin (red/green) via the cell binding domains on domains III-9 and III-10. (b) FN binding re-arranges cell-surface receptors (dark green, pink) and organizes the actin cytoskeleton (red lines) and activates intracellular signaling complexes (light blue). (c) Fibril-formation through FN–FN interactions. FN also interacts with other ECM-molecules as part of the ECM (black lines) and binds growth factors (yellow triangles)
Fibronectin Knock Out Mice and Phenotype George et al. (1993, 1997) showed the essential nature of FN in vivo when they demonstrated embryonic lethality at E8.5 in a knock-out mouse model. Embryonic death was associated with massive defects in mesoderm-derived structures, including the absence of somites and developmental defects in the heart and the vascular system. Similar defects were observed when the central cell binding sequence was mutated from RGD to the inactive RGE sequence. Interestingly, this inactivating mutation did not have a severe impact on matrix formation both in the mutated embryos and cells (Takahashi et al. 2007). Embryonic lethality was evident at E10 with severe vascular defects resembling the phenotype of the a5-integrin-deficient mice which also die early in development from a collection of defects, which include a improperly formed
22 Fibronectin
461
vasculature (Yang et al. 1994; Goh et al. 1997; Taverna and Hynes 2001; Francis et al. 2002). These results clearly indicate that the interaction between FN and the a5integrin are fundamental for the process of angiogenesis, the development of new blood vessels during embryonic development, tissue repair, fertility, chronic inflammation and tumor growth and metastasis (Carmeliet 2005).
Fibronectin and Cancer Fibronectin and Tumor Growth Previously, it had been thought that cancer cells have the potential to metastasize and grow in any number of locations in the body. However, the “seed and soil” hypothesis suggests that it is the mutual interaction between the cancer cell and its specific microenvironment that contributes to tumor formation (Mendoza and Khana 2009). Lung cancers are mostly diagnosed in patients with underlying chronic disorders such as pulmonary fibrosis and chronic obstructive pulmonary disease where FN expression is greatly enhanced (Littmann et al. 2004). While there is no data demonstrating the ability of FN to initiate oncogenic transformation, the aberrant secretion and abnormal deposition of FN in lung tissue may render the host susceptible to the development of carcinoma. Furthermore, FN is also expressed by a variety of cancer cells (Hynes et al. 1979; Labat-Robert et al. 1981; Plantefaber and Hynes 1989; Charalabopoulos et al. 2005; Qian et al. 2005; Sato et al. 2005). The tumor derived FN may serve as an autocrine/paracrine mitogenic factor. As an example, nicotine-induced proliferation of lung carcinoma cells is associated with increased expression of tumor cell-derived FN. It is important to note that nicotine-induced tumor cell proliferation in vitro can be blocked by specifically interfering with the FN–FN receptor-interactions (Zheng et al. 2007). FN deposition triggers the formation of FN-fibers which are embedded into the tumor microenvironment. This process has become increasingly important in recruiting both fibroblasts and macrophages to the tumor area (Jin et al. 2006) where they become M2/tumor associated macrophages (TAMs). TAMs express a variety of molecules which affect tumor cell proliferation, angiogenesis and dissolution of connective tissue. A hallmark of early malignant transformation in solid human tumors is the fragmentation and decrease of pericellular FN concomitant with an increase of FN-deposition in the stroma surrounding the proliferating tumor nests. In human laryngeal and ectocervical carcinomas, the amount of FN-mRNA in the stroma was 7–13 times higher than in the control tissues (Moro et al. 1992). This profound stromal reaction, also known as desmoplasia, is associated with fibroblast proliferation, secretion of potent pro-angiogenic cytokines, growth factors (Yamashiro et al. 1994; Lewis et al. 2000; Sica and Bronte 2007) and proteases (Ueno et al. 2000; Niwa et al. 2001; Giraudo et al. 2004; Jodele et al. 2005), necessary to support angiogenesis, ECM degradation, invasion and migration.
462
A. Menrad
Tumor Angiogenesis Angiogenesis is the process by which new blood vessels develop from pre-existing vessels. The growth of new blood vessels promotes embryonic development, wound healing and the female reproductive cycle. It also plays an important role in the pathological development of solid cancers. Growth factors released by TAMs, fibroblasts and tumor cells within a hypoxic environment stimulates the growth of new blood vessels. While growth factors and their receptors play key roles in angiogenic sprouting, adhesion to the tumor-permissive ECM also is a prime regulator of tumor angiogenesis. Adhesion promotes endothelial cell survival as well as endothelial cell proliferation and migration (Kim et al. 2000a, b, 2002; Cheresh and Stupack 2002). Angiogenesis is specifically associated with the upregulation of the FN-receptor integrin a5b1. This FN specific receptor molecule it is not detectable in quiescent endothelium but expressed in proliferating endothelium as a response to angiogenic growth factors in vitro or within the angiogenic vasculature of a growing tumor (Kim et al. 2000a, b) in vivo. Another hallmark of the provisional pro-angiogenic tumor ECM is the secretion of EDA and EDB-containing fibronectin isoforms as a result of alternative splicing.
EDB-FN in Tumor Growth and Angiogenesis The expression of EDB-FN, which is a feature of myofibroblasts (Pujuguet et al. 1996; Berndt et al. 1998), endothelial cells and cancer cells (Midulla et al. 2000) is tightly regulated. Under physiological conditions, ED-B-expression is restricted to embryonic tissue and a few normal adult tissues associated with the process of tissue regeneration during the female reproductive cycle (De Candia and Rogers 1999), and during bone and wound healing (Ffrench-Constant et al. 1989; Dhase et al. 2004). High levels of EDB-FN expression were detected in primary lesions as well as metastatic sites of almost all human solid cancers, including breast (Kaczmarek et al. 1994; D’Ovidio et al. 1998), colorectal (Pujuguet et al. 1996; Santimaria et al. 2003), non-small cell lung (Santimaria et al. 2003), pancreatic (Wagner et al. 2008), hepatocellular (Menrad et al. 2004a), head and neck (Kosmehl et al. 1999; Mhawech et al. 2005), human skin cancer (Karelina and Eisen 1998), meningioma and glioblastoma (Ohnishi et al. 1998; Castellani et al. 2000, 2002) as well as hematologic tumors (Sauer et al. 2009). Based on extensive immunohistochemical data it has become clear that EDB-FN expression is uniquely associated with tumor growth and angiogenesis. Nevertheless, little is known about the mechanisms which control the expression of this molecule during tumor growth. Zardi and colleagues (Coltrini et al. 2009) have observed that EDB-FN secretion is strongly upregulated when tumor cells are grown as subcutaneous grafts in nude mice whereas the cells rapidly lose expression after isolation and ex vivo growth. This clearly indicates the importance of the host’s microenvironment for EDB-production by tumor cells.
22 Fibronectin
463
Conditions which shift the “angiogenic switch” in the tumor microenvironment from a pro-angiogenic towards an anti-angiogenic state by downregulation of FGF2 and/or VEGF-production are characterized by a significant loss of EDB-FN expression and subsequently by a decrease in tumor vascularity thus confirming EDB-FN as a specific marker of tumor angiogenesis.
Potential Function of the EDB-Domain The EDB-domain is highly conserved throughout the species: it is identical in human, dog, rabbit, rat and mouse and shows 96% homology with chicken and 80% with Xenopus laevis. This together with the tightly regulated and highly specific expression pattern could be indicative for a specific role of this domain for tumor growth, angiogenesis and tissue regeneration. Early in vitro studies (Chen and Culp 1996, 1998; Hashimoto-Uoshima et al. 1997) suggested that recombinant EDBfragments generally promote cell adhesion and spreading of various cells including endothelial cells (Menrad et al. 2004b, c; Menrad and Menssen 2005). These observations are supported by several authors who have described that in contrast to FN with low EDB concentrations, FN-preparations from WI38-VA13 cells, which secrete high amounts of EDB-FN (Castellani et al. 1986; Borsi et al. 1987; Carnemolla et al. 1989) were able to conserve the phenotype and support the growth of tumor-derived endothelial cells in vitro (Alessandri et al. 1998, 1999Allport and Weissleder 2003). Furthermore, the integrin a3b1was described as a potential binding partner for EDB-FN (Chirivi et al. 2001). EDB-/-mice (Fukuda et al. 2002) were also generated. Surprisingly, these mice developed normally and were fertile. No significant phenotype was observed in vivo, even after analysis of a number of different models where the EDB-domain was previously thought to play a role, such as angiogenesis, thrombosis, organogenesis, tumorigenesis and wound healing (Fukuda et al. 2002; Astrof et al. 2004; Matuskova et al. 2006). However, a mild in vitro effect in matrix assembly and proliferation was observed in EDB-/-embryonic fibroblasts (Fukuda et al. 2002). These cells grew more slowly and produced fibrils that were shorter and thinner than those secreted by control cells under in vitro culture conditions.
EDA-FN in Tumor Growth and Angiogenesis The expression of EDA-FN is, comparable to EDB-FN, tightly regulated and specifically expressed during tumor growth around tumor vasculature in both human and mouse tumors (Oyama et al. 1989, 1990; Kaczmarek et al. 1994; Inufusa et al. 1995; Lohi et al. 1995; Pujuguet et al. 1996; Ohnishi et al. 1998; Kosmehl et al. 1999; Matsumoto et al. 1999; Scarpino et al. 1999; Castellani et al. 2002; Astrof et al. 2004; Trefzer et al. 2006; Villa et al. 2008; Schliemann
464
A. Menrad
et al. 2009). In vitro, numerous functions have been ascribed to the EDA-domain including its specific binding to integrins a4b1and a9b1via an EDGIHEL-sequence as well as Toll-4 receptor (Okamura et al. 2001; Liao et al. 2002; Shinde et al. 2008). However, the biological relevance of these interactions for in vivo tumor growth and angiogenesis remains unclear. Thus the development of genetically modified animal models was a prerequisite to address these issues in vivo. The individual deletion of EDA exon form the FN gene did not have significant developmental effects nor did it have a significant effect on either tumor angiogenesis or tumor progression in the transgenic RIP1-Tag2 model of pancreatic islet carcinogenesis (Astrof et al. 2004).
EDA/EDB-Double Null Mutants To determine the in vivo function of EDA and EDB-domains, a mouse strain with the simultaneous deletion of both domains was generated (Astrof et al.2007). Mice with the EDA-/-/EDB-/-double null mutations were embryonically leathal (~80% penetrance). No defects in production or cell surface association of FN was observed suggesting that the lack of EDA and EDB together accounted for the lethal variety of cardiovascular defects which are comparable, but milder than the ones observed in FN-null embryos. During tumor progression, a thick cuff of perivascular ECM, which includes FN and its splice variants surrounds tumor blood vessels. Interestingly, a thick layer of pericytes also surrounds those vessels whereas tumor vessels not expressing FNs have a single pericyte layer or no pericytes at all (Astrof et al. 2004; Jain 2005). Therefore one may hypothesize that pericyte recruitment to proangiogenic vessels may be non-functional in EDA/EDB-double null mice where vessels are not embedded in a properly formed matrix (Astrof and Hynes 2009).
Cryptic Site Exposure as a Result of EDB Alternative Splicing The insertion of EDB between the Type-III repeats 7 and 8 induces conformational modifications that unmask at least two cryptic sequences. One epitope on repeat III-7 is only accessible after EDB is included in the FN molecule. A new epitope was recently discovered on repeat III-8 (Balza et al. 2009). Again, the epitope is only accessible in the presence of the EDB. Surprisingly, two monoclonal antibodies which specifically bind to these neo-epitopes (BC1 on III-7 and C6 on III-8) show a staining pattern that is similar, if not identical to the reactivity of antibodies which directly bind to EDB (Castellani et al. 1986; Balza et al. 2009). Interestingly, these conformational changes seem to be limited to human FN since both monoclonal antibodies are strictly human-specific and do not crossreact with rodent FN. This is in strong contrast to the
22 Fibronectin
465
highly crossreactive monoclonal antibodies to EDB. The biological relevance of these cryptic sites is entirely unknown. However, one may hypothesize that the domains may be involved in the process of angiogenesis and tissue regeneration during the formation of the tumor-permissive environment of a growing tumor.
FN as a Modulator of Tumor Invasion and Metastasis The process of tumor cell invasion and metastasis requires the degradation of connective tissue associated with vascular basement membranes and interstitial connective tissue (Stetler-Stevenson et al. 1993; Yu et al. 1996). Several lines of evidence strongly implicate matrix metalloproteinases (MMPs), particularly MMP2 in this process. These include a positive correlation between MMP2 expression and invasive potential and the inhibition of metastasis formation in vivo by MMP inhibitors (Stetler-Stevenson et al. 1993; Yu et al. 1996; Curran and Murray 1999). Interestingly, MMP2 is often associated with the benign tissue surrounding tumor nests rather than with the tumor cells themselves. This suggests that malignant cells use MMPs secreted by normal cells to facilitate migration into new tissue sites (Heppner et al. 1996). MMP2 and 9 adhere to collagen via their three FN type II repeats within their catalytic domain (Kohn and Liotta 1995; Massova et al. 1998; Steffensen et al. 1998; Olson et al. 1998). A rapid increase of MMP2 in the culture supernatant is detectable after co-culture of the breast carcinoma cell line MDA-MB-231 and bone marrow derived fibroblasts (Saad et al. 2000). Interestingly, MMP2 was bound to collagen on the cell surface of the normal fibroblasts and displaced by cell surface associated FN on the surface of the breast carcinoma cells. Removal of FN from the surface of breast cancer cells before co-culture eliminated the increase of MMP2 in the culture supernatant whereas re-addition of FN restored the effect (Saad et al. 2002). sFN was less effective in fibroblastassociated MMP2 than cell bound FN. It is very likely that conformational changes in the FN molecule after surface binding contribute to this effect. Indeed, the conformation of FN is crucial for collagen binding, and cell surface associated fibronectin results in conformational changes within the molecule which are key as the first steps in fibril formation. In addition to the FN-triggered attraction to TAMs as important contributors to tissued reorganization, FN seems to play an important role for controlling MMP2 levels in the microenvironment surrounding the growing tumor. Most likely, this is an important mechanism for many cancers since most malignant cells express surface associated FN.
Migration-Stimulating Factor Migration-stimulating factor (MSF) is an oncofetal protein constitutively expressed by foetal and cancer patient-derived fibroblasts under in vitro culture conditions
466
A. Menrad
(Schor and Schor 2009). MSF is secreted by tumor cells and tumor-associated stroma cells in the majority of tumors examined so far, including carcinomas, melanomas and glioblastoma. Furthermore, there is evidence that high MSF-expression by tumor cells and tumor-associated stroma correlates with poor survival in patients with breast cancer and cancer of the oral cavity suggesting MSF as an important driver of tumor progression. Structurally, MSF is a genetically truncated isoform of human fibronectin generated from the single-copy FN gene of 70 kDa identical to the N terminus of a fulllength fibronectin with the addition of an MSF-unique (intron-coded) 10 amino acid C terminus. MSF is a pleiotropic factor which triggers multiple responses on a broad range of potential target cells (fibroblasts, vascular endothelial cell, pericytes and epithelial cells) which are involved in the process of matrix remodelling and angiogenesis both in vitro and in vivo were attributed to secreted MSF. Site directed mutagenesis experiments suggest that the conserved IGD-motif within repeats I-7 and I-9 are responsible for motogenic activity of fibroblasts, whereas all four IGD motifs on I-4, I-5, I-7 and I-9 specifically trigger endothelial cell migration. MSF-realted bioactivities can be blocked by neutralising anti-MSF antibodies in vitro. Apoptosis of VEGF-stimulated endothelial cells is observed after exposure to neutralizing anti MSF antibodies whereas resting cells are not affected. This indicates that in addition to its pro-angiogenic effect, MSF expressed by sprouting endothelial cells acts as an autocrine survival factor under in vitro cell culture conditions. Until today, the MSF-specific receptor molecule and the respective signalling mechanisms have not been characterized. Furthermore, the importance of MSF for in vivo tumor growth remains to be demonstrated.
FN and Tumor Dormancy The mechanism of long-term survival of occult (or dormant) cancer cells at distant sites within the patients body are poorly understood. Breast cancer cells preferentially metastasize to the bone marrow early in the course of disease. Although most metastatic cells die upon reaching the marrow microenvironment, few malignant cells are found at this site at the time of diagnosis (Braun et al. 2000a, b). These cells can remain at a dormant state without loss of viability for a long time (Boyce et al. 1999), remain protected form cell death and, in fact survive multiple rounds of chemotherapy (Braun et al. 2000a, b). Years later, some of these cells start to proliferate and will eventually kill the patient (Braun et al. 2000a, b). Although the precise mechanisms of tumor dormancy remain poorly understood, some growth factors and ligands to cellular integrins in the bone marrow niche may influence the fate of dormant metastatic cells. FGF2 is endemic to the bone marrow. It can inhibit growth (Wang et al. 1988, 1997; McLesky et al. 1994; Fenig et al. 1997; Johnson
22 Fibronectin
467
et al. 1998) and induce differentiation in breast cancer cells (Korah et al. 2000a, b). Furthermore, it upregulates the integrin repertoire, especially the a5b1-FN receptor on breast cancer cells. FN is a significant component of the bone marrow stroma (Orr et al. 2000); ligation of the a5b1-FN receptor on breast cancer cells acts as a survival signal which protects FGF 2-responsive dormant tumor cells from apoptotic death. FGF 2 also controls the expression of EDB-FN which is a specific marker of angiogenesis associated with tumor growth (Coltrini et al. 2009). Further research will be necessary to investigate whether EDB-FN is already expressed in the FGF2-rich environment which supports the state of tumor dormancy. The paradigm of dormancy in the marrow microenvironment may differ from other metastatic sites and the mechanisms which protect the malignant resting cell population in these sites remain under investigation.
Fibronectin: Beyond Fibrils FNs multifunctionality is reflected by its complex modular molecular design consisting of distict domains which are highly conserved throughout the species. The domains are arranged in specific juxtapositions, sometimes controlled by highly regulated alternative splicing. The implication is that the complex and conserved architecture of FN codes for specific information of high biological relevance. Domain structure controls steric conformation, exposure of cryptic epitopes and the specific interaction with other ECM molecules. HGF and VEGF bind to FN and form complexes with their respective receptors and integrins on cell surfaces leading to enhanced migration and proliferation (Vaday and Lider 2000; Wijelath et al. 2002, 2006; Rahman et al. 2005). As organized, solid-phase ligands, FNs integrate complex, multivalent signals to cells in a spatially patterned and highly regulated fashion. The vast progress in understanding the biological roles of FN under physiological conditions as well as in malignant disease opens up new opportunities for the discovery of new therapeutic.
Therapeutic Interventions FN plays a central role for tumor angiogenesis. Two important findings contributed to the discovery of novel therapeutic approaches: 1 . The disease-specific expression of EDA and EDB-FN 2. The FN-specific, a5b1-integrin mediated anti-apoptotic effect on tumor cells and proliferating endothelial cells in tumor-angiogenesis Figure 22.3 provides an overview on therapeutic approaches which are currently investigated in clinical trials.
468
A. Menrad
Pro-angiogenic growth factors
FN-degradation Anti α5β1 integrin
Targeted delivery of effector molecules to EDB/EDA-FN
Activation of endothelial cells
α5β1 integrin binding to fibronectin Secretion of EDA /EDB-FN
New blood vessel formation
Fig. 22.3 Therapeutic mode of action on targeted delivery and a5b1integrin function blocking antibodies. The growing tumor secretes a multitude of pro-angiogenic growth factors which activate resting endothelial cells to form new blood vessels. During this process, the angiogenesisspecific EDA and EDB-FN splice variants are formed and deposited in the perivascular stroma. Furthermore, the interaction of a5b1integrin-expressing endothelial cells with FN is essential. Both targeted delivery of toxic payloads or a5b1integrin-blocking antibodies are currently in clinical trials
Targeted Delivery to FN Isoforms FN isoforms are specifically expressed during tumor angiogenesis and tumor growth. The high degree homology from mouse to man and the comparable tissue expression pattern facilitates animal experiments in immunocompetent syngeneic or autochtonous animal models. Furthermore, the stable and high concentration of both EDA- and EDB-FN surrounding tumor vessels makes these domain ideal targets for a targeted delivery approach. High-quality, domain-specific antibodies were generated as targeting vehicles using synthetic human antibody phage display libraries (Carnemolla et al. 1996; Neri et al. 1997; Pini et al. 1998; Giovannoni et al. 2001; Villa et al. 2008). Preclinical biodistribution studies in tumor-bearing mouse models clearly demonstrated a tumor-specific accumulation with neglectable normal tissue exposure, a key-prerequisite for a targeted delivery approach. Interleukin-2 (IL2) and Tumor Necrosis Factor-a (TNF-a) were chosen as effector molecules. Both cytokines are clinically approved drugs for certain types of cancers. However, a short
22 Fibronectin
469
serum half-life and poor tumor accumulation in combination with severe side effects never allowed to use optimal therapeutic doses. This is especially true for TNF-a, where the cytokine can only be applied locoregionally in the Isolated Limb Perfusion setting for the treatment of metastatic melanoma or soft tissue sarcoma (Lejeune et al. 2006). Both cytokines have shown significant therapeutic efficacy in various preclinical animal models for disease in mono- and combination therapy (Menrad and Menssen 2005; Kaspar et al.2006). Currently, the EDB-FN-specific fusion proteins L19-IL2 and L19-TNF-a are in clinical development. Furthermore, L19 in the Small Immunoprotein-format (SIP) labelled with 131I showed promising results for targeted radioimmunotherapy (RIT) in Phase I/II clinical trials (Sauer et al. 2009). Radiolabeled antibodies for RIT of cancer represent a class of targeted therapeutics of high interest since performance in vivo is predictable by biodistribution and imaging data. A minimal dose of 50 Gy delivered 131I-RIT is necessary to sanitize a tumor. On the contrary, bone marrow can only tolerate ~2.5 Gy. The systematic re-formatting and analysis of the scFv L19 revealed that the SIP-format is optimal for RITapplications (Hu et al. 1996; Berndorff et al. 2005; Tijink et al. 2006).
Anti-a5b1 Integrin Function-Blocking Antibody (Volociximab) The murine anti-a5b1integrin function-blocking antibody IIA-1 inhibits both growth factor induced and tumorangiogenesis in vivo (Kim et al. 2000a, b, 2002; Bhaskar et al. 2007, 2008). Monoclonal antibody IIA-1 was humanized resulting in a 82% human, 18% mouse chimeric IgG4 monoclonal antibody M200 (Volociximab). Further attempts to fully humanize M200 resulted in a dramatic loss of bioactivity. Despite these limitations, a Phase I study was performed with M200 in onclological indications (Ricart et al. 2008). Despite of its chimeric nature, the antibody was well tolerated at doses up to 10 mg/kg weekly without dose-limiting toxicities or hypersensitivity reactions. Prelininary, but encouraging efficacy was seen in a patient with sunitinib-refractory renal cell carcinoma (size-reduction of a pulmonary lesion) and in a patient with visceral metastasis from malignant melanoma (durable stable disease for 7 month). To fully evaluate the therapeutic potential of M200, either as a single agent, or in combination with standard of care, further clinical studies need to be undertaken.
Endogenous Inhibitors of Angiogenesis Endostatin and Tumstatin Endostatin and Tumstatin are two endogenous inhibitors of angiogenesis derived from precursors of human collagen molecules a3 chain of type IV collagen and a1 chain of type XVIII collagen, respectively. Both inhibitors showed efficacy in vitro
470
A. Menrad
under cell culture conditions and in preclinical tumor models in vivo (Sudhakar and Boosani 2008; Fu et al. 2009). Recently, it was demonstrated that both fragments block the RGD-dependent integrin binding to either FN or Vitronectin (Sudhakar et al. 2003). Despite a lot of controversial pre-clinical and clinical findings (Fu et al. 2009), a modified version of endostatin (Endostar) was approved in 2009 in China for the treatment of non small cell lung carcinoma.
Anastellin (III1-C) FN fibrils are important for cell adhesion and spreading in the ECM. The first type III-repeat in the FN molecule has been shown to play a pivotal role in this process: a fragment of the first type III repeat of FN (Anastellin) binds to FN and triggers the formation of polymers of “superfibronectins” with significantly enhanced adhesive properties which inhibit cell migration (Morla et al. 1994). In vivo administration of Anastellin in tumor-bearing mice suppressed tumor growth, angiogenesis, and metastasis (Yi and Ruoslahti 2001) by complex-formation with the RGDcontaining FN. Competition with the natural receptor, the a5b1-integrin which is upregulated during tumor growth and angiogenesis inhibits cell proliferation and induces apoptosis.
References Alessandri G, Chirivi RG, Catellani P et al (1998) Isolation and characterization of human tumorderived capillary endothelial cells: role of oncofetal fibronectin. Lab Invest 78:127–128 Alessandri G, Chirivi RG, Fiorentini S et al (1999) Phenotypic and functional characterization of tumor-derived microvascular endothelial cells. Clin Exp Metastasis 17:655–662 Allport JR, Weissleder R (2003) Murine Lewis Lung Carcinoma-derived endothelium express markers of endothelial activation and requires tumor-specific extracellular matrix in vitro. Neoplasia 5:205–217 Astrof S, Hynes RO (2009) Fibronectins in vascular morphogenesis. Angiogenesis 12:165–175 Astrof S, Crowley D, George EL et al (2004) Direct test of potential EIIIA and EIIIB alternatively spliced segments of fibronectin in physiological and tumor angiogenesis. Mol Cell Biol 24:8662–8670 Astrof S, Crowley D, Hynes RO (2007) Multiple cardiovascular defects caused by the absence of alternatively spliced segments of fibronectin. Dev Biol 311:11–24 Balza E, Sassi F, Ventura E et al (2009) A novel human fibronectin cryptic sequence unmasked by the insertion of the angiogenesis-associated extra type III domain B. Int J Cancer 125: 751–758 Berndorff D, Borkowski S, Sieger S et al (2005) Radioimmunotherapy of solid tumors by targeting extra domain B fibronectin: identification of the best-suited radioimmunoconjugate. Clin Cancer Res 11:7053s–7063s Berndt A, Borsi L, Xinmei L et al (1998) Evidence of ED-B+ fibronectin synthesis in human tissues by non-radioactive RNA in situ hybridization. Investigation on carcinoma (oral squamous cell and breast cancer) chronic inflammation (rheumatoid synovitis) and fibromatosis (Morbus Dupyten). Histochem Cell Biol 109:249–255
22 Fibronectin
471
Bhaskar V, Zhang D, Fox M et al (2007) A function blocking anti-mouse integrin a5b1 antibody inhibits angiogenesis and impedes tumor growth in vivo. J Translat Med 5:61 Bhaskar V, Fox M, Breinberg D et al (2008) Volociximab, a chimeric integrin a5b1 antibody, inhibits the growth of VX2 tumors in rabbits. Invest New Drugs 26:7–12 Borsi L, Carnemolla B, Castellani P et al (1987) Monoclonal antibodies in the analysis of fibronectin isoforms generated by alternatively splicing of mRNA pre-cursors in normal and transformed human cells. J Cell Biol 104:595–600 Bowditch RD, Hariharan M, Tominna EF et al (1994) Identification of a novel integrin binding site in fibronectin: differential utilization by b3 integrins. J Biol Chem 269:10856–10863 Boyce BF, Yoneda T, Guise TA (1999) Factors regulating the growth of metastatic cancer in the bone. Endocr Relat Cancer 6:333–47 Braun S, Pantel K, Muller P et al (2000a) Cytokeratin-positive cells in the bone marrow and survival of patients with stage I, II, or III breast cancer. N Engl J Med 342:525–533 Braun S, Kentenich C, Janni W et al (2000b) Lack of an effect of adjuvant chemotherapy on the elimination of single dormant tumor cells in bone marrow of high risk breast cancer patients. J Clin Oncol 18:80–86 Carmeliet P (2005) Angiogenesis in life, disease and medicine. Nature 42:932–936 Carnemolla B, Balza E, Siri A et al (1989) A tumor-associated fibronectin isoform generated by alternatively splicing of messenger RNA precursors. J Cell Biol 108:1139–1148 Carnemolla B, Neri D, Castellani P et al (1996) Phage antibodies with pan-species recognition of the oncofetal angiogenesis marker fibronectin ED-B domain. Int J Cancer 68:379–405 Castellani P, Siri A, Rosellini C et al (1986) Transformed human cells release different fibronectin variants than do normal cells. J Cell Biol 103:1671–1677 Castellani P, Dorcaratto A, Pau A et al (2000) The angiogenesis marker ED-B+fibronectin isoform in intracranial meningioma. Act Neurochir (Wien) 142:277–282 Castellani P, Borsi L, Carnemolla B et al (2002) Differentiation between high- and low-grade astrocytoma using a recombinant antibody to the extra domain-B of fibronectin. Am J Pathol 161:1695–1700 Charalabopoulos K, Mittari E, Karakosta A et al (2005) Integrin adhesion molecules and some of their ligands in laryngeal cancer. Exp Oncol 27:86–90 Chen W, Culp LA (1996) Adhesion mediated by fibronectin’s alternatively spliced EDB (EIIB) and its neighboring type III repeats. Exp Cell Res 223:9–19 Chen W, Culp LA (1998) Adhesion to fibronectin’s EDB domain induces tyrosine phosphorylation of focal adhesion proteins in Balb/c 3T3 cells. Clin Exp Metastasis 16:30–42 Cheresh DA, Stupack DG (2002) Integrin-mediated death: and explanation of the integrin-knockout phenotype? Nat Med 8:193–194 Chirivi RG, Puente S, Storgard CM et al (2001) Angiogenesis induced by alternatively spliced oncofetal fibronectin binding to endothelial cell integrin a3b1. Proceedings of the AACR 92nd annual meeting, 24–28 March, Suppl 84, p. LB60 Coltrini D, Ronca R, Belleri M et al (2009) Impact on VEGF-dependent tumor micro-environment on EDB fibronectin expression by subcutaneous human tumor xenografts in nude mice. J Pathol 219:455–462 Curran S, Murray GI (1999) Matrix metalloproteinases in tumor invasion and metastasis. J Pathol 189:300–308 DeCandia LM, Rogers RJ (1999) Characterization of the expression of the alternative splicing of the ED-A, ED-B, and V-regions of fibronectin mRNA in bovine ovarian follicles and corporea lutea. Reprod Fertil Dev 11:367–377 DeSimone DW, Norton PA, Hynes RO (1992) Identification and characterization of alternative spliced fibronectin mRNAs expressed in early Xenopus embryos. Dev Biol 149:357–369 Dhase OK, Alt V, Zardi L et al (2004) Expression of ED-A+ and ED-B+ fibronectin splice variants in bone. Bone 35:1334–1345 D’Ovidio M, Mastraccio A, Marzullo A et al (1998) Intratumoral microvessel density and expression of ED-A/ED-B sequences of fibronectin in breast carcinoma. Eur J Cancer 34:1081–1085
472
A. Menrad
Dvorak HF, Form DM, Manseau EJ et al (1984) Pathogenesis of desmoplasia. J Natl Cancer Inst 73:1195–1205 Fenig E, Wieder R, Paglin S et al (1997) Basic fibroblast growth factor confers growth inhibition and MAP kinase activation in human breast cancer cells. Clin Cancer res 3:135–142 Ffrench-Consatant C, Van De Water L, Dvorak HF et al (1989) Reappearance of an embryonic pattern of fibronectin splicing during wound healing in the adult rat. J Cell Biol 109:903–914 Fogerty FJ, Akijama SK, Yamada KM et al (1990) Inhibition of binding of fibronectin to matrix assembly sites by anti-integrin (alpha5 beta1) antibodies. J Cell Biol 111:699–708 Francis SE, Goh KL, Hodivala-Dilke K et al (2002) Central roles of alpha5beta1 integrin and fibronectin in vascular development in mouse embryos and embryoid bodies. Arterioscler Thromb Vasc Biol 22:927–933 Fu Y, Tank H, Huang Y et al (2009) Unraveling the mysteries of endostatin. IUBMB Life 61:613–626 Fukuda T, Yoshida N, Kataoka Y et al (2002) Mice lacking the EDB segment of fibronectin develop normally but exhibit reduced cell growth and fibronectin matrix assembly in vitro. Cancer Res 62:5603–5610 Geiger B, Breshadsky A, Pankov R et al (2001) Transmembrane crosstalk between the extracellular matrix-cytoskeleton crosstalk. Nat Rev Mol Cell Biol 2:793–805 George EL, Georges-Labouesse EN, Patel-King RS et al (1993) Defects in mesoderm, neural tube and vascular development in mouse embryos lacking fibronectin. Development 119:1079–1091 George EL, Baldwin HS, Hynes RO (1997) Fibronectins are essential for heart and blood vessel morphogenesis but dispensable for initial specification of precursor cells. Blood 90:3073–3081 Giovannoni L, Viti F, Zardi L et al (2001) Isolation of anti-angiogenesis antibodies from a large combinatorial repertoire by colony filter screening. Nucleic Acids Res 29:E27 Giraudo E, Inoue M, Hanahan D (2004) An amino-bisphosphate targets MMP-9-expressing macrophages and angiogenesis to impair cervical carcinogenesis. J Clin Invest 114:623–33 Goh KL, Yang JT, Hynes RO (1997) Mesodermal defects and cranial neural crest apoptosis in alpha5 integrin-null embryos. Development 124:4309–4319 Gutman A, Kornblihtt AR (1987) Identification of a third region of cell-specific alternative splicing in human fibronectin mRNA. Proc Natl Acad Sci U S A 87:7179–7182 Hashimoto-Uoshima M, Dorcaretto A, Nicolo G et al (1997) The alternatively spliced domains EIIIB and EIIIA of human fibronectin affect cell adhesion and spreading. J Cell Sci 110:2271–2280 Heppner KJ, Matrisian LM, Jensen RA et al (1996) Expression of most matrix metalloproteinase family members in breast cancer represents a tumor-induced host response. Am J Pathol 149:273–282 Hu S, Shivley L, Raubitschek A et al (1996) Minibody: a novel engineered anti-carcinoembryonic antigen antibody fragment (single-chain Fv-CH3) which exhibits rapid, high-level targeting of xenografts. Cancer Res 56:3055–3061 Hynes RO (1990) Fibronectins. In: Rich A (ed) Springer series in molecular biology, Springer, New York, pp 1–538 Hynes RO (1992) Integrins: versatility, modulation and signalling in cell adhesion. Cell 69:11–25 Hynes RO, Destree T, Perkis ME et al (1979) Cell surface fibronectin and oncogenic transformation. J Supramol Struct 11:95–104 Inufusa H, Nakumura M, Adachi T et al (1995) Localization of oncofetal and normal fibronectin in colorectal cancer. Correlation with histologic grade, liver metastasis, and prognosis. Cancer 75:2802–2808 Jain RK (2005) Normalization of tumor vasculature: an emerging concept in angiogenic therapy. Science 307:58–62 Jin H, Aiyer A, Su J et al (2006) A homing mechanism for bone marrow-derived progenitor cell recruitment to the neovasculature. J Clin Invest 116:652–662
22 Fibronectin
473
Jodele S, Chantrain CF, Blavier L et al (2005) The contribution of bone marrow-derived cells to the tumor vasculature in neuroblastoma is matrix metalloprotease-9 dependent. Cancer Res 65:3200–3208 Johnson MR, Valentine C, Basilico C et al (1998) FGF signalling activates STAT1 and p21 and inhibits the estrogen response and proliferation of MCF-7 cells. Oncogene 16:2647–2656 Johnson KJ, Sage H, Briscoe G et al (1999) The compact conformation of fibronectin is determined by interamolecular ionic interactions. J Biol Chem 274:15473–15497 Kaczmarek J, Castellani P, Nicolo G et al (1994) Distribution of oncofetal fibronectin isoforms in normal, hyperplastic and neoplastic human breast tissues. Int J Cancer 59:11–16 Karelina TV, Eisen AZ (1998) Interstitial collagenase and the ED-B oncofetal domain of fibronectin as markers of angiogenesis in human skin tumors. Cancer Detect Prev 22:438–444 Kaspar M, Zardi L, Neri D (2006) Fibronectin as target for therapy. Int J Cancer 118:1331–1339 Kohn E, Liotta L (1995) Molecular insights into cancer invasion: strategies for prevention and internvention. Cancer Res 55:1856–1862 Korah R, Sysounthone V, Golowa Y et al (2000a) Basic fibroblast growth factor confers to a more differentiated phenotype in MDA-MB-231 human breast cancer cells. Cancer Res 60:733–740 Korah R, Sysounthone V, Scheff E et al (2000b) Intracellular FGF-2 promotes differentiation in T47-D breast cancer cells. Biochem Biophys Res Commun 277:255–260 Kornblihtt AR, Vibe-Pedersen K, Baralle FE (1984) Human fibronectin: molecular cloning evidence for two mRNA species differing by an internal segment coding for a structural domain. EMBO J 3:221–226 Kosmehl H, Berndt A, Strassburger S et al (1999) Distribution of laminin and fibronectin isoforms in oral squamous cell carcinoma. Br J Cancer 81:1071–1079 Kim S, Bell K, Mousa SA et al (2000a) Regulation of angiogenesis in vivo by ligation of integrin alpha5beta1 with the central cell-binding domain of fibronectin. Am J Pathol 156: 1345–1362 Kim S, Harris M, Varner JA (2000b) Regulation of integrin alpha vbeta3-mediated endothelial cell migration and angiogenesis by integrin alpha5beta1 and protein kinase A. J Biol Chem 275:33920–33928 Kim S, Bakre M, Yin H et al (2002) Inhibition of endothelial cell survival and angiogenesis by protein kinase A. J Clin Invest 110:933–941 Labat-Robert J, Birembaut P, Robert L et al (1981) Modification of fibronectin distribution pattern in solid human tumors. Diagn Histopathol 4:236–299 Lejeune FJ, Lienard D, Matter M et al (2006) Efficiency of recombinant human TNF in human cancer therapy. Cancer Immunity 6:6 Lewis JS, Landers RJ, Underwood JC et al (2000) Expression of vascular endothelial growth factor by macrophages is up-regulated in poorly vascularized areas of breast carcinomas. J Pathol 192:150–158 Liao YF, Gotwals PJ, Koteliansky VE et al (2002) The EIIIA segment of fibronectin is a ligand for integrins alpha9beta1 and alpha5beta1 providing a novel mechanism for regulating cell adhesion by alternative splicing. J Biol Chem 277:14467–14474 Littmann AJ, Thornquist MD, White E et al (2004) Prior lung disease and risk of lung cancer in a large prospective study. Cancer Causes Control 15:819–827 Lohi J, Tani T, Laitinen L et al (1995) Tenascin and fibronectin isoforms in human renal cell carcinomas, renal cell carcinoma cell lines and xenografts in nude mice. Int J Cancer 63:442–449 Massova I, Kotra LP, Fridman R et al (1998) Matrix metalloproteinases: structures, evolution, and diverisification. FASEB J 12:1075–1095 Matsumoto E, Yoshida T, Kawarada Y et al (1999) Expression of fibronectin isoforms in human breast tissue: production of extra domain A+/extra domain B+ by cancer cells and extra domain A+ by stromal cells. Jpn J Cancer Res 90:320–325 Matuskova J, Chauhan AK, Cambien B et al (2006) Decreased plasma fibronectin leads to delayed thrombus growth in injured arterioles. Arterioscler Thromb Vasc Biol 26:1391–1396
474
A. Menrad
McDonald JA, Quade BJ, Brockelmann TJ et al (1987) Fibronectin’s cell-adhesive domain and an amino-terminal matrix assembly domain participate in its assembly into fibroblast pericellular matrix. J Biol Chem 262:2957–2967 McLesky SW, Ding IYF, Lipman ME et al (1994) MDA-MB-134 breast cancer carcinoma cells overexpress fibroblast growth factor (FGF) receptors and are growth-inhibited by FGF ligands. Cancer Res 54:523–530 Mendoza M, Khana C (2009) Revisiting the seed and soil in cancer metastasis. Int J Biochem Cell Biol 41:1452–1462 Menrad A, Menssen HD (2005) ED-B fibronectin as a target for antibody-based cancer treatments. Expert Opin Ther Targets 9:491–500 Menrad A, Rosewicz S, Wagner K et al (2004a) Therapeutic efficacy of the recombinant fusion protein L19-IL2 in orthotopic animal models for pancreatic and hepatocellular carcinoma. Proceedings of the 21st international conference of advances in the application of monoclonal antibodies in clinical oncology, Suppl. p15 Menrad A, Cao YJ, Redlitz A et al (2004b) ED-B fibronectin as a target for antibody based therapeutics and the evaluation of phage angiomics for target identification and validation. Proceedings of the strategic research institute conference on vascular targeting agents, 8–9 November, Suppl. P1 Menrad A, Cao YJ, Redlitz A et al (2004c) Molecular and functional characterization of ED-B fibronectin selective function blocking antibodies and the evaluation of phage angiomics for target identification and validation. Human Antibodies 13:33 Mhawech P, Dulguerov P, Assalym M et al (2005) ED-B fibronectin expression in squamous cell carcinoma of the head and neck. Oral Oncology 41:82–88 Midulla M, Verma R, Pignatelli M et al (2000) Source of oncofetal ED-B-containing fibronectin: implication of production by both tumor and endothelial cells. Cancer Res 60:164–169 Miranti CK, Brugge JS (2002) Sensing the environment: a historical perspective on integrin signal transduction. Nat Cell Biol 4:E83–E90 Morla A, Zhang Z, Ruoslahti E (1994) Superfibronectin is a functionally distinct form of fibronectin. Nature 367:193–196 Moro L, Colombi M, Molinari-Tosatti MP et al (1992) Study of fibronectin mRNA in human laryngeal and ectocervical carcinomas by in situ hybridization and mage analysis. Int J Cancer 51:1–6 Mueller MM, Fusenig NE (2004) Friends and foes – bipolar effects of the tumour stroma in cancer. Nat Rev Cancer 4:839–849 Nagai T, Yamakawa N, Aota S et al (1991) Monoclonal antibody characterization of two distant sites required for function of the central cell-binding domain of fibronectin in cell adhesion, cell migration, and matrix assembly. J Cell Biol 114:1295–1305 Neri D, Carnemolla B, Nissim A et al (1997) Targeting by affinity-matured recombinant antibody fragments of an angiogenesis associated fibronectin isoform. Nat Biotechnol 15:1271–1275 Niwa Y, Akamatsu H, Niwa H et al (2001) Correlation of tissue and plasma RANTES levels with disease course in patients with breast or cervical cancer. Clin Cancer Res 7:285–289 Norton PA, Hynes RO (1987) Alternative splicing of chicken fibronectin in embryos and in normal and transformed cells. Mol Cell Biol 7:4297–4307 Ohnishi T, Hiraga S, Izumoto S et al (1998) Role of fibronectin-stimulated tumor cell migration in glioma invasion in vivo: clinical significance of fibronectin and fibronectin receptor expressed in human glioma tissues. Clin Exp Metastasis 16:729–741 Okamura Y, Watari M, Jerud ES et al (2001) The extradomain A of fibronectin activates Toll-like receptor 4. J Biol Chem 267:10229–10233 Olson MW, Toth M, Gervasi DC et al (1998) High affinity binding of latent matrix metalloproteinase-9 on the a2(IV) chain of collagen IV. J Biol Chem 273:10672–10681 Orr FW, Lee J, Duivenvoorden WCM et al (2000) Patophysiologic interactions in skeletal metastasis. Cancer 88 (Suppl 12):2912–2918 Oyama F, Hirohashi S, Shimosato Y et al (1989) Deregulation of alternative splicing of fibronectin pre-mRNA in malignant human liver tumors. J Biol Chem 264:10331–10334 Oyama F, Hirohashi S, Shimosato Y et al (1990) Oncodevelopmental regulation of the alternative splicing of fibronectin pre-messenger RNA in human lung tissue. Cancer Res 50:1075–1078
22 Fibronectin
475
Pini A, Viti F, Santucci A et al (1998) Human antibodies with subnanomolar affinity against a marker of angiogenesis eluted from a two-dimensional gel. J Biol Chem 273:21769–21776 Plantefaber LC, Hynes RO (1989) Changes in integrin receptors on oncogenically transformed cells. Cell 56:281–290 Pujuguet P, Hammann A, Moutet M et al (1996) Expression of fibronectin ED-A+ and ED-B+ isoforms by human and experimental colorectal cancer. Contribution of cancer cells and tumor-associated myofibroblasts. Am J Pathol 148:579–592 Pupa SM, Menard S, Forti S et al (2002) New insights into the role of extracellular matrix during tumor onset and progression. J Cell Physiol 192:259–267 Qian F, Zhang ZC, Wu XF et al (2005) Interaction between integrin alpha5 and fibronectin is required for metastasis of B16F10 melanoma cells. Biochem Biophys Res Comm 333:1269–1275 Rahman S, Patel Y, Murray J et al (2005) Novel hepatocyte growth factor (HGF) binding domains on fibronectin and vitronectin coordinate a distinct and amplified Met-integrin induced signalling pathway in endothelial cells. BMC Cell Biol 6:1–17 Ricart AD, Tolcher AW, Liu G et al (2008) Clin Cancer Res 14:7924–7929 Ronnov-Jenssen L, Petersen OW, Bissell MJ (1996) Cellular changes involved in conversion of normal to malignant breast:importance of the stromal reaction. Physiol Rev 76:69–125 Ruoslahti E, Pierschbacher MD (1987) New perspectives in cell adhesion: RGD and integrins. Science 238:491–497 Saad S, Bendall LJ, James A et al (2000) Coculture of breast cancer cells and bone marrow fibroblasts induce expression of the metalloproteinase MMP2. Breast Cancer Res Treat 63:105–115 Saad S, Gottlieb DJ, Kenneth F et al (2002) Cancer cell associated fibronectin induces release of matrix metalloproteinase-2 from normal fibroblasts. Cancer Res 62:283–289 Santimaria M, Moscatelli G, Viale GL et al (2003) Immunoscintigraphic detection of the ED-B domain of fibronectin, a marker of angiogenesis, in patients with cancer. Clin Cancer Res 9:571–579 Sato T, Yamochi T, Yamochi T, Aytac U et al (2005) CD26 regulates p38 mitogen-activated protein kinase-dependent phosphorylation of integrin beta1, adhesion to extracellular matrix, and tumorigenicity of T-anaplastic large cell lymphoma Karpas 299. Cancer Res 65:6950–6956 Sauer S, Erba PA, Petrini M et al (2009) Expression of oncofetal ED-B-containing fibronectin isoform in hematologic tumors enables ED-B-targeted 131I-L19-SIP radioimmunotherapy in Hodgkin lymphoma patients. Blood 113:2265–2274 Scarpino S, Stoppacciaro A, Pellegrini C et al (1999) Expression of EDA/EDB isoforms of fibronectin in papillary carcinoma of the thyroid. J Pathol 188:163–167 Schliemann C, Wiedmer A, Pedretti M et al (2009) Three clinical-stage tumor targeting antibodies reveal differential expression of oncofetal fibronectin and tenascin-C isoforms in human lymphoma. Leuk Res 33:1718–1722 Schor AM, Schor SL (2009) Angiogenesis and tumor progression: migration-stimulating factor as a novel target for clinical intervention. Eye 1–9 Schwarzbauer JE, Tamkun JW, Lemischka IR et al (1983) Three different fibronectin mRNAs arise by alternative splicing within the coding region. Cell 35:421–431 Schwarzbauer JE, Patel RS, Fonda D et al (1987) Multiple sites of alternative splicing of the rat fibronectin gene transcript. EMBO J 6:2573–2580 Sechler JL, Schwarzbauer JE (1997) Coordinated regulation of fibronectin fibril assembly and actin stress fiberformation. Cell Adhes Commun 4:413–424 Shinde AV, Bystroff C, Wang C et al (2008) Identification of the peptide sequence within the EIIA (EDA) segment of fibronectin that mediate integrin alpha9beta1-dependent cellular activities. J Biol Chem 283:2858–2870 Sica A, Bronte V (2007) Altered macrophage differentiation and immune dysfunction in tumor development. J Clin Invest 117:1155–1166 Steffensen B, Bigg HF, Overall CM (1998) The involvement of the fibronectin type II-like modules of human gelatinase A in the cell surface localization and activation. J Biol Chem 273:20622–20628
476
A. Menrad
Stetler-Stevenson W, Azanavoorian S, Liotta L (1993) Tumor cell interactions with the extracellular matrix during invasion and metastasis. Annu Rev Cell Biol 9:541–573 Sudhakar A, Boosani CS (2008) Inhibition of tumor angiogenesis by tumstatin: insights into signalling mechanisms and implications in cancer regression. Pharm Res 25:2731–2739 Sudhakar A, Sugimoto H, Yang C et al (2003) Human tumstatin and human endostatin exhibit distinct antiangiogenic activities mediated by avb3 and a5b1 integrins. Proc Natl Acad Sci U S A 100:4766–4771 Takahashi S, Leiss M, Moser M et al (2007) The RGD motif in fibronectin is essential for development but dispensable for fibril assembly. J Cell Biol 178:167–178 Taverna D, Hynes RO (2001) Reduced blood vessel formation and tumor growth in alpha5integrin-negative teratocarcinomas and embryoid bodies. Cancer Res 61:5255–5261 Tijink BM, Neri D, Leemans CR et al (2006) Radioimmunotherapy of head and neck cancer xenografts using 131I-labeled antibody L19-SIP for selective targeting of tumor vasculature. J Nucl Med 47:1127–1135 Trefzer U, Chen Y, Herberth G et al (2006) The monoclonal antibody SM5-1 recognizes a fibronectin varian which is widely expressed in melanoma. BMC Cancer 6:1–12 Ueno T, Toi M, Saji H et al (2000) Significance of macrophage chemoattractant protein-1 in macrophage recruitment, angiogenesis, and survival in human breast cancer. Clin Cancer Res 8:3282–3289 Umezawa K, Kornblihtt AR, Baralle FE (1986) Isolation and characterization of cDNA clones for human liver fibronectin. FEBS Lett 186:31–34 Vaday GG, Lider O (2000) Extracellular matrix moieties, cytokines, and enzymes: dynamics effects on immune cell behaviour and inflammation. J Leukoc Biol 67:149–159 Villa A, Trachsel E, Kaspar M et al (2008) A high-affinity human monoclonal antibody specific to the alternatively spliced EDA domain of fibronectin efficiently targets tumor neo-vasvulature in vivo. Int J Cancer 122:2405–2413 Wagner K, Schulz P, Scholz A et al (2008) The targeted immunocytokine L19-IL2 efficiently inhibits the growth of orthotopic pancreas cancer. Clin Cancer Res 14:4951–4960 Wang Q, Maloof P, Wang H et al (1988) Basic fibroblast growth factor (bFGF) downregulates Bcl-2 and promotes apoptosis in MCF-7 human breast cancer cells. Exp Cell Res 238:177–187 Wang H, Rubin M, Fenig E et al (1997) Basic FGF causes growth arrest in MCF-7 human breast cancer cells while inducing both mitogenic and inhibitory G1 events. Cancer Res 57: 1750–1757 Wernert N (1997) The multiple roles of tumor stroma. Virchows Arch 430:433–443 Wijelath ES, Murray J, Rahman S et al (2002) Novel vascular endothelial growth factor binding domains of fibronectin enhance vascular endothelial growth factor biologic activity. Circ Res 91:25–31 Wijelath ES, Rahman S, Namekata M et al (2006) Heparin-II domain of fibronectin is a vascular endothelial growth factor-binding domain: enhancement of VEGF biological activity by singular growth factor/matrix protein synergism. Circ Res 99:853–860 Yamashiro S, Takeya M, Nishi T et al (1994) Tumor-derived monocyte chemoattractant protein-1 induces intratumoral infiltration of monocyte-derived macrophage subpopulation in transplant rat tumors. Am J Pathol 145:856–67 Yang JT, Rayburn H, Hynes RO (1994) Embryonic mesodermal defects in alpha 5 integrindeficient mice. Development 119:1093–1105 Yi M, Ruoslahti E (2001) A fibronectin fragment inhibits tumor growth, angiogenesis, and metastasis. Proc Natl Acad Sci U S A 98:620–624 Yu AE, Hewitt RE, Kleiner DE et al (1996) Molecular regulationof cellular invasion-role of gelatinase A and TIMP2. Biochem Cell Biol 74:823–831 Zamir E, Katz M, Posen Y et al (2000) Dynamics and segregation of cell-matrix adhesions in cultured fibroblasts. Nat Cell Biol 2:191–196 Zheng Y, Ritzenthaler D, Roman J et al (2007) Nicotine stimulates human lung cancer cell growth by inducing fibronectin expression. Am J Respir Cell Mol Biol 37:681–690
Chapter 23
Collagen in Cancer Janelle L. Lauer and Gregg B. Fields
Abstract Collagen is a key structural component of the extracellular matrix (ECM), and also serves as a modular of diverse signaling pathways. Intact collagens may be upregulated in cancer to provide a rigid matrix that facilitates tumor growth. In turn, collagen catabolism by matrix metalloproteinases (MMPs) and other proteases reveals previously hidden binding sites that promote angiogenesis and tumor invasion. A variety of cell surface biomolecules (integrins, CD44, DDRs) and other ECM proteins and proteoglycans [fibronectin (FN), laminin (LNs), decorin] interact with collagen, and these interactions, along with the structural state of collagen, provide the foundation for tumorigenesis and metastasis.
The Collagen Family of Proteins The collagen family is a diverse group of proteins made up of at least 28 members (Myllyharju and Kivirikko 2001; Hashimoto et al. 2002; Boulégue et al. 2008; Gordon and Hahn 2009; Shoulders and Raines 2009). Collagens are composed of three a chains of primarily repeating Gly–Xxx–Yyy triplets, which induce each a chain to adopt a left-handed polyPro II helix. Three left-handed chains then intertwine to form a right-handed superhelix. The collagen triple helix (Fig. 23.1) is important for the integrity and workings of multiple connective tissues, including skin, bone, cartilage, tendon, and dentin. Collagens have been classified according to their a chains (Gordon and Hahn 2009). Homotrimeric collagens (i.e., types II and III) have three a chains of identical sequence. Heterotrimeric collagens have either two a chains of identical sequence (designated a1) and one a chain of differing sequence (designated a2), such as type I, or three a chains of differing sequence (designated a1, a2, and a3),
G.B. Fields (*) Department of Biochemistry, University of Texas Health Science Center, 7703 Floyd Curl Drive, San Antonio, TX 78229, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_23, © Springer Science+Business Media, LLC 2010
477
478
J.L. Lauer and G.B. Fields
Fig. 23.1 The structure of fibrillar collagens. Individual triple helices associate laterally to form collagen fibrils with a characteristic banded structure. Figure courtesy of JPK Instruments AG (www.jpk.com)
Table 23.1 Collagen types Collagen subfamily Fibrillar Fibril anchoring Network FACIT Multiplexin MACIT
Representative collagen(s) I, II, III, V, XI, XXIV, XXVII VII IV, VI, VIII, X IX, XII, XIV, XVI, XIX, XX, XXI, XXII, XXVI XV, XVIII XIII, XVII, XXIII, XXV
such as type VI (Cole 1994). Collagens are further classified into subfamilies based on their quaternary structure (Table 23.1). These subfamilies include fibrillar, fibril anchoring, fibril associated with interrupted triple helices (FACIT), network, multiplexin, and membrane associated with interrupted triple helices (MACIT) (Cole 1994). The most common collagens (types I, II, III, V, and XI) have fibrillar structures. Before collagen is properly folded into a triple helix, a series of posttranslational modifications on the central (Gly–Xxx–Yyy)n domain must occur, including hydroxylation of most Pro and some Lys residues in the Yyy position followed by glycosylation of certain hydroxylysines. Glycosylation also occurs on some Asn residues in the C-terminal propeptides. Disulfide bonds between the propeptides
23 Collagen in Cancer
479
are rearranged and isomerization of imino acids from cis to trans takes place (Baum and Brodsky 1999). C-terminal propeptides mediate interaction between three a chains and hold these chains in place, nucleating triple-helix formation. Propagation of the triple helix occurs in the C- to N-terminal direction. The triple-helical molecules are secreted from the cell and the N- and C-terminal propeptides that flank the central (Gly–Xxx–Yyy)n domain are removed (in the fibrillar collagens). Finally, the collagens assemble and cross-link, forming supramolecular structures such as fibrils, filaments, or networks. These quaternary structures often contain more than one type of collagen (Prockop and Kivirikko 1995; Olsen and Ninomiya 1999). Collagens most often implicated in cancer progression are types I and IV. Type I collagen, the most profuse and ubiquitous of the collagens, is found in most connective tissues and embryonic tissues [reviewed in Cole (1994)] (Fig. 23.1). Type I collagen is typically composed of two a1 and one a2 chain, although homotrimeric type I collagen containing three a1(I) chains has been reported in carcinomas and cancer cell lines (DeClerck et al. 1987; Rupard et al. 1988; Asokan et al. 1993). The nonfibrillar type IV collagen is the major collagen found in basement membranes (BMs), forming network-like structures (Fig. 23.2). Type IV collagen is made up of six genetically distinct a-chains (a1– a6), but most commonly contains two a1 and one a2 chains. The a-chains share 50–70% homology and produce similar domain structures (Kalluri 2003). Each chain contains an N-terminal 7S domain, a triple helical Gly–Xxx–Yyy repeating domain of ~1,400 residues, and a C-terminal noncollagenous 1 (NC1) domain of about 230 amino acids. Type IV collagen has 25–26 interruptions of the Gly–Xxx–Yyy repeat within the triplehelical domain (Timpl 1989; Linsenmayer 1991). Individual a-chains associate to form a triple helix known as a protomer. These protomers then dimerize end-to-end via NC1 domain interactions to form a dimer. Four of these dimers come together via interactions between the 7S domains to generate a loose cruciform structure. This octamer then associates into the sheet-like mesh found in the BM. The NC1 domain is thought to participate in a-chain selection, lateral chain association, and dimer formation (Hornebeck et al. 2002).
The Extracellular Matrix Extracellular matrix components include insoluble proteins such as collagens, LNs, and FN, as well as proteins modulating cell proliferation. These include secreted protein acidic and rich in cysteine (SPARC), thrombospondins, osteopontin, and tenascin, as well as matrix-associated proteins such as growth factors (Davis et al. 2000). The BM is a dense sheet-like structure of specialized ECM, approximately 50–100 nm thick (Kalluri 2003) (Fig. 23.3). The BM is generally associated with cells located basolateral to epithelium, endothelium, peripheral nerve axons, fat cells, and muscle cells, whereas the ECM is found more widely throughout the interstitium (Kalluri 2003). Type IV collagen predominates the BM, with other
480
J.L. Lauer and G.B. Fields
Fig. 23.2 The structure of BM (type IV) collagen (Kalluri 2003). Reproduced by permission of Nature Publishing Group
constituents including LN, heparan sulfate proteoglycans, perlecan, nidogen/ entactin, agrin, SPARC, and fibulins, as well as types XV and XVIII collagen (Kalluri 2003). The vascular BM surrounds all blood vessels and capillaries located between the layer of endothelial cells and specialized smooth muscle cells known as pericytes. Vascular BM is important for tumor cell metastasis for two reasons: (1) tumor cells must penetrate it to enter and leave the circulation; and (2) if the tumor is to grow larger than a few mm, it must produce angiogenic factors to assure adequate nutrient and gas exchange (Pluda 1997). Both the ECM and BM structures are extensively cross-linked (see below). However, tumor BM is significantly less cross-linked making it more susceptible to
23 Collagen in Cancer
481
Fig. 23.3 The structure of the BM showing the association of type IV collagen, laminin, perlecan, and nidogen, along with BM interaction with integrins from the cell surface. Reproduced by permission of the NCBI Bookshelf (http://www.ncbi.nlm.nih.gov/bookshelf)
proteolysis, remodeling, and turnover (Kalluri 2003). Type IV collagen and LN undergo independent intermolecular self-assembly to form a sheet-like structure, which then interact with each other (Kalluri 2003). The more minor BM constituents do not form independent structures, but merely associate with and facilitate the formation of the collagen/LN network. It appears that LN is deposited near the cell surface in a b1 integrin and dystroglycan-dependent process (Kalluri 2003). With assistance from nidogen, type IV collagen is then assimilated into the matrix.
Role of Collagen in Cancer: Overview Most collagens assist in anchoring cells to the ECM and some function in cellular regulation. For example, cell adhesion to collagen regulates cellular gene expression and inhibits cell death via apoptosis. Thus, the relative levels of collagen expression have significant implications for cell homeostasis. Given this role in cellular behaviors, collagen expression has been implicated in pathological conditions. Gene expression of the a chains involved in the formation of types I and III collagen [more specifically, a2(I) and a1(III), respectively] is elevated in selected,
482
J.L. Lauer and G.B. Fields
highly metastatic pulmonary melanomas, as compared to poorly metastatic tumors (Clark et al. 2000). It is presumed that these elevated collagen levels enhance melanoma survival. However, it has also been observed that collagen turnover is required for tumor invasion (Sabeh et al. 2004, 2009). As discussed below, collagen modulation of metastasis is complex, but further understanding of the process offers significant therapeutic potential. The cellular microenvironment provides not only signals but also mechanical properties. Matrix stiffness has been implicated in tumor progression, with collagen considered a significant contributor to changes in the cellular mechanical microenvironment (Ng and Brugge 2009). Increased matrix tension due to lysyl oxidase (LOX) cross-linking of collagen induces integrin signaling including phosphorylation of focal adhesion kinase (FAK) and Cas and enhanced PI3 kinase activity (Levental et al. 2009). In turn, inhibition of LOX activity impedes breast tumor progression (Levental et al. 2009). Mechanotransduction and oncogenic signaling pathways may be synergistic in promoting tumorigenicity (Ng and Brugge 2009). While intact collagen is required for signaling and matrix stiffness, MMP degradation of collagen facilitates tumorigenesis (Ng and Brugge 2009). LOX and MMPs most likely collaborate to create a dynamic microenvironment (Ng and Brugge 2009). MMPs and LOX may also regulate the activities of soluble factors such as transforming growth factor beta (TGF-b) that in turn modulate tumor cell behaviors (Levental et al. 2009). Angiogenesis, a process required for tumor cell growth, involves a transition of endothelial cells from their normally quiescent state to one of rapid proliferation as well as degradation of the BM to allow the sprouting of microvessels. During the process, endothelial cells are removed from the existing blood vessel and surrounded by a provisional matrix, eventually coming to rest upon the capillary BM (Kalluri 2003). The angiogenic process involves several temporally coordinated events including: (1) vasodilation and increased vascular permeability and subsequent endothelial cell adhesion and migration, endothelial cell stimulation by growth factors; (2) degradation of BM by proteolytic enzymes, including MMPs produced by stromal cells, endothelial cells, tumor cells, or immune cells associated with the tumor, in response to cytokines, growth factors, and cell–matrix interactions; (3) release, migration, and proliferation of endothelial cells in response to growth factors, MMPs, and integrins (detached endothelial cells and pericytes are now in contact with the provisional matrix, which supplies proliferative signals, in contrast to the antiproliferative signals generated by intact ECM); (4) formation and maturation of capillaries involving the differentiation of smooth muscle cells and pericytes; and (5) remodeling of vascular bed through apoptosis (Kalluri 2003; Bellon et al. 2004). Important regulators of angiogenesis include growth factors such as vascular endothelial growth factor (VEGF), basic fibroblast growth factor, platelet-derived growth factor, and chemokines, which are released from the ECM by hydrolysis of matrix components (Kalluri 2003; Kanematsu et al. 2004). Collagens are known to interact with numerous biomolecules, including cell adhesion receptors, glycoproteins, proteoglycans, enzymes, chaperones, etc. Mapping of ligand-binding sites in type I collagen has revealed several “hot spots” in the monomeric triple helix and discrete cell and matrix interaction domains in
23 Collagen in Cancer
483
the fibril (Di Lullo et al. 2002; Sweeney et al. 2008). Below we do not attempt to exhaustively cover all biomolecular interactions with collagen; instead, we have chosen to focus on those most relevant to cancer and/or representing relatively novel modes of action.
Adhesion Receptor Binding to Collagen A great variety of cell surface receptors have been recognized for their collagen binding activity (Leitinger and Hohenester 2007). The best-characterized cell surface adhesion molecules are integrins, which are heterodimeric proteins composed of one a and one b subunit. The roles of integrins in cancer are covered in more detail in Chap. 27. The collagen binding integrins include a1b1, a2b1, a3b1, a10b1, and a11b1 (Kühn and Eble 1994; van der Flier and Sonnenberg 2001). The collagen binding integrins associated with cancer are a1b1, a2b1, and a3b1 (Kramer and Marks 1989; Yoshinaga et al. 1993; Melchiori et al. 1995). However, upon proteolysis, fragments of collagen are bound by additional integrins (see below). The a2b1 integrin is critical for melanoma cell migration and matrix reorganization within three-dimensional type I collagen matrices (Friedl et al. 1997; Maaser et al. 1999). Engagement of the a2b1 integrin has also been linked to induction of MMP-1 and MMP-14 (Riikonen et al. 1995; Vogel 2001; Baronas-Lowell et al. 2004a) and implicated in MMP-9 regulation (Vo et al. 1998). Furthermore, MMP-1 binds to the a2b1 integrin via the I domain of the a2 subunit (Stricker et al. 2001). In an effort to better understand the roles of individual integrins, numerous triple-helical peptide (THP) binding sites for the collagen binding integrins have been identified (Table 23.2). The Gly–Phe–Hyp–Gly–Glu–Arg motif, in triple-helical
Table 23.2 Sequences of receptor binding sites within types I–IV collagen Receptor Binding site location Sequence GFOGER a1b1, a2b1, a11b1 a1(I)502–507 a1b1, a2b1 a1(III)64–69 GROGER a1b1, a2b1 a1(IV)382–393 GAOGFOGERGEK a1b1, a2b1 a1(I)127–135 GLOGERGRO a1b1 a1(IV)437–448 GPPGDQGPPGIP a2(IV)454–465 GAKGRAGFPGLP a2b1 a1(I)433–438 GADGEA a2b1 a1(III)115–120 GLOGEN a2b1 a1(III)522–528 GGPOGPR a3b1 a1(IV)531–543 GEFYFDLRLKGDK CD44 a1(IV)1263–1277 [IV-H1] GVKGDKGNPGWPGAP DDR2 GPRGQOGVMGFO a1(III)397–408 For type IV collagen, sequence numbers are based on the human a1(IV) and a2(IV) genes. (Hostikka and Tryggvason 1988) O = 4R-hydroxyproline
484
J.L. Lauer and G.B. Fields
conformation, has been shown to bind to the a2b1 and a11b1 integrins and recombinant a1 and a2 A-domains (Knight et al. 1998, 2000; Emsley et al. 2000; Xu et al. 2000; Sweeney et al. 2003; Zhang et al. 2003; Baronas-Lowell et al. 2004a, b; Khew and Tong 2008). The co-crystal structure of a THP incorporating Gly–Phe–Hyp–Gly–Glu–Arg and the a2 A-domain revealed that the Glu residue directly coordinated the a2 A-domain metal ion (Mg2+), the Arg residue formed a salt bridge with an Asp residue in the a2 A-domain, and Phe was involved in hydrophobic interactions with the receptor (Emsley et al. 2000). Changes in receptor conformation upon ligand binding appeared to be initially induced by changes in metal ion coordination (Bella and Berman 2000; Emsley et al. 2000). Binding also resulted in changes in the triple-helix main-chain conformation and bending of the triple helix (Emsley et al. 2004). The Gly–Phe–Hyp–Gly–Glu–Arg motif is found within type I collagen at a1(I)502–507 and type IV collagen at a1(IV)385–390. A THP model of a1(I)502–516 binds to platelets, a model of a1(I)496–507 binds to endothelial cells, and models of a1(IV)382–393 bind to HT-1080 and melanoma cells (Knight et al. 1998; Lauer-Fields et al. 2003b; Sweeney et al. 2003; Baronas-Lowell et al. 2004a, b). Other collagen binding motifs have been identified for the a2b1 integrin. Gly– Leu–Hyp–Gly–Glu–Arg from type I collagen has been identified as a ligand for the a1b1 integrin, a1 and a2 A-domains, and a2 I-domain (Xu et al. 2000; Sweeney et al. 2003), while the Gly–Arg–Hyp–Gly–Glu–Arg motif from type III collagen binds the a1 and a2 I-domains (Kim et al. 2005; Raynal et al. 2006) (Table 23.2). Gly–Gly–Pro–Hyp–Gly–Pro–Arg from a1(III)522–528 and Gly–Leu–Hyp–Gly– Glu–Asn from a1(III)115–120 binds the a2b1 integrin (Barnes et al. 1996; Morton et al. 1997; Raynal et al. 2006) (Table 23.1). Based on the model of type I collagen fibers, the Gly–Phe–Hyp–Gly–Glu–Arg and Gly–Leu–Hyp–Gly–Glu–Arg sequences are accessible for integrin binding (Herr and Farndale 2009). The a1b1 integrin, either isolated from human placenta or expressed recombinantly, simultaneously binds Asp441 from two a1(IV) chains and Arg458 from the a2(IV) chain (Eble et al. 1993; Golbik et al. 2000; Saccá et al. 2002). This type IV collagen region is believed to be a higher affinity a1b1 integrin binding site than the Gly–Phe–Hyp–Gly–Glu–Arg motif found in type IV collagen (Knight et al. 2000), although the KD value for integrin binding to the heterotrimeric THP was 20 mM while binding to homotrimeric Gly–Phe–Hyp–Gly–Glu–Arg occurred with KD= 1.45 mM (Saccá et al. 2002; Renner et al. 2004; Barth et al. 2009). Proline hydroxylation is required for a1b1 integrin binding (Perret et al. 2003). The melanoma cell a3b1 integrin binds to a1(IV)531–543 (Miles et al. 1994, 1995; Li et al. 1997; Lauer et al. 1998). Binding of the a3b1 integrin to collagenderived substrate is not dependent upon triple-helical conformation, nor specific for l- versus d-stereoisomers (Miles et al. 1994, 1995; Li et al. 1997) unlike the a1b1 and a2b1 integrins (Sweeney et al. 2003). Another collagen-binding receptor is the proteoglycan CD44. CD44s (also known as CD44H) is expressed by epidermal melanocytes in both an unmodified and chondroitin sulfate (CS) modified form (Herbold et al. 1996). Human CD44s is initially glycosylated via five potential N-linked and seven potential O-linked
23 Collagen in Cancer
485
carbohydrate sites, resulting in a core protein of 85–95 kDa (Screaton et al. 1992; Naor et al. 1997; Lesley and Hyman 1998). CD44 glycosylation may be differentially regulated between tumor and stromal cells (Matsuki et al. 2003). The CS modification further increases the CD44 molecular weight to 180–200 kDa (Naor et al. 1997). Several lines of evidence indicate that CD44 plays an important role in tumor progression. Upregulation of CD44s mRNA production and cell surface expression has been found in highly metastatic melanoma as compared to less metastatic melanoma or nontransformed melanocytes (Leigh et al. 1996, 1997; Ahrens et al. 2001; Ranuncolo et al. 2002). Enhanced expression of CD44 is also found in the vasculature of tumors (endothelial cells) compared with endothelial cells from normal tissue (Griffioen et al. 1997). The incorporation of CD44 isolated from metastatic cells into nonmetastatic tumor cells results in induction of metastatic behavior (Naor et al. 1997). In addition, tumor cells expressing CD44 display accelerated tumor growth and metastatic spread in immunodeficient mice compared with parental cells (Naor et al. 1997). The overexpression of CD44 is found on a variety of tumor cells (Naor et al. 2002), and elevated CD44 expression by four to six-fold is associated with tumor growth and metastasis (Goebeler et al. 1996). CD44 in the chondroitin sulfate proteoglycan (CSPG) form is among the receptors uniquely overexpressed in metastatic melanoma (Naor et al. 2002). Proteolytic removal of CD44 inhibits the growth of primary tumors and curtails metastasis in a mouse B16 melanoma model (Wald et al. 2001). Absence of CD44 gene products virtually eliminated osteosarcoma metastasis (Weber et al. 2002). Approaches that interfere with CD44 binding to ligand, such as administration of high molecular weight hyaluronic acid (HA), anti-CD44 mAb, or a CD44-receptor globulin, reduce tumor formation in the lung for animal models established from CD44-expressing tumor cell lines (Eliaz and Szoka 2001; Platt and Szoka 2008). Interestingly, CD44 has recently been revealed as a cancer stem cell marker for at least six different tumor types (Prince et al. 2007; Du et al. 2008; Ponnusamy and Batra 2008; Stamenkovic and Yu 2009). A theory is emerging that CD44 positive cells within a tumor display true stem cell properties such that an individual cell can give rise to an entire tumor (Prince et al. 2007). Additionally, CD44 interaction with hyaluronan stimulates the expression of and induces ankyrin binding to MDR1 (P-glycoprotein), resulting in the efflux of chemotherapeutic agents and chemoresistance in tumor cells (Bourguignon et al. 1998b; Misra et al. 2005). CD44 binds to types I, IV, VI, and XIV collagen, but it is not a primary receptor for cell adhesion to these collagens (Staatz et al. 1989; Faassen et al. 1992; Ehnis et al. 1996; Knutson et al. 1996). CD44 in the CS form facilitates melanoma motility on and invasion of type I collagen (Faassen et al. 1992). However, CD44 was not a contributor to melanoma cell migration and matrix reorganization within three-dimensional type I collagen matrices (Friedl et al. 1997; Maaser et al. 1999). CD44/CSPG is required for TGF-b stimulation of melanoma cell motility on type I collagen (Faassen et al. 1993). CD44 binding to type IV collagen is dependent upon CS (Knutson et al. 1996). In general, melanoma cell invasion of the BM is inhibited by removing cell surface
486
J.L. Lauer and G.B. Fields
CS (Knutson et al. 1996). The type IV collagen a1(IV)1263–1277 sequence (Table 23.2; designated [IV-H1]) promotes melanoma cell adhesion, spreading, and signaling (Chelberg et al. 1990; Mayo et al. 1991; Fields et al. 1993, 1998; Yu et al. 1997; Malkar et al. 2002; Lauer-Fields et al. 2003a). This sequence has been identified as a ligand for CD44 by a combination of methods, including (1) cell adhesion and spreading assays using triple-helical a1(IV)1263–1277 and an Asp1266Abu variant; (2) inhibition of cell adhesion and spreading assays; and (3) triple-helical a1(IV)1263–1277 affinity chromatography with whole cell lysates and glycosaminoglycans (Lauer-Fields et al. 2003a). Interaction of CD44 with this ligand was strongly dependent upon triple-helical conformation (Fields et al. 1993; Malkar et al. 2002) and CSPG content (Mickelson et al. 1991; Knutson et al. 1996; Lauer-Fields et al. 2003a). CD44 binding of other ligands has also been shown to be sensitive to the nature and levels of the CD44 glycosylation (Alves et al. 2009). Loss of triple-helical structure dramatically reduces melanoma cell adhesion, spreading, and signaling modulated by this ligand (Fields et al. 1993, 1998; Malkar et al. 2002). Melanoma cell responses to the triple-helical [IV-H1] sequence have been compared for the galactosylated versus nongalactosylated ligands (Lauer-Fields et al. 2003a). Galactosylation was found to strongly modulate adhesion and spreading, both of which were dramatically decreased due to the presence of a single sugar. This study was the first demonstration of the prophylactic effects of ligand glycosylation on tumor cell interaction with the BM, while related reports have shown that (1) tumor cell surface sialic acid reduced binding to type IV collagen (Dennis et al. 1982) and (2) decreased LN binding glycans results in increased prostate and breast carcinoma motility (Bao et al. 2009). Several studies have demonstrated CD44 interaction with MMPs. The CD44 splice variant CD44v3,8–10 in metastatic breast cancer cells is physically associated with active MMP-9 within invadopodia structures (Bourguignon et al. 1998a). Interaction of CD44v3–10 and MMP-9 is also found in prostate cancer cells (Desai et al. 2007). Association of CD44 with active MMP-9 in metastatic breast cancer and melanoma cells leads to the activation of TGF-b and the promotion of degradation of type IV collagen and invasion (Yu and Stamenkovic 1999, 2000). CD44v3–10 in the heparan sulfate proteoglycan form associates with active MMP-7 and heparin-binding epidermal growth factor precursor to form a complex on the surface of tumor cells (Yu et al. 2002). CD44 forms a complex with MMP-14/ MT1-MMP via the enzyme’s hemopexin-like (HPX) domain and directs it to lamellipodia (Mori et al. 2002; Nakamura et al. 2004). The CD44 cytoplasmic domain binds to members of the ezrin–radixin–moesin family of cytoskeletal proteins (Bourguignon et al. 1998b). Intracellular signaling by CD44 results from complex formation between the CD44 cytoplasmic tail and either Rho-family GTPases or members of the Src family of kinases (Thorne et al. 2004). One result of CD44 “outside-in” signaling is upregulation and activation of integrins (Fujisaki et al. 1999). Engagement of [IV-H1] by CD44 significantly stimulated melanoma cell production of MMP-8 and modulated lower levels of MMP-1, MMP2, MMP-13, and MMP-14/MT1-MMP (Baronas-Lowell et al. 2004a).
23 Collagen in Cancer
487
Shedding of CD44 has also been shown to be important in tumor cell migration, but the mechanism of shedding is controversial. CD44 shedding was first reported to be facilitated by a cell surface metalloproteinase whose activity was inhibited by tissue inhibitor of metalloproteinase 1 (TIMP-1) (Okamoto et al. 1999). CD44 shedding resulted in the release of tumor cells from HA and promotion of migration (Okamoto et al. 1999; Kajita et al. 2001). MMP-14/MT1-MMP was subsequently identified as the metalloproteinase responsible for CD44 shedding (Kajita et al. 2001; Nakamura et al. 2004). However, ADAM10 was also implicated as a CD44 sheddase (Murai et al. 2004). Further analysis revealed three sites of CD44 processing, Gly192–Tyr, Gly233–Ser, and Ser249–Gln (Nakamura et al. 2004). Shedding at Ser249–Gln (producing a 65–70 kDa fragment) was inhibited by TIMP-3 but not by TIMP-1 or TIMP-2, indicating the involvement of an ADAMs protease (Nakamura et al. 2004). Conversely, shedding at Gly192–Tyr and Gly233–Ser was inhibited by TIMP-2 and TIMP-3 but not by TIMP-1, indicating MT1-MMP involvement (Nakamura et al. 2004). Increased shedding at Gly192–Tyr was associated with malignant tumors (Nakamura et al. 2004). More recently, ADAM10 but not MT1MMP was deemed responsible for constitutive CD44 shedding in human melanoma cells (Anderegg et al. 2009; Stamenkovic and Yu 2009). Interestingly, shed CD44 can inhibit melanoma tumor growth by competing with cell surface CD44 for binding to HA (Ahrens et al. 2001). Following shedding, the intracellular CD44 fragment subsequently generated by presenilin/g-secretase action may function as a transcription factor (Seiki 2003; Thorne et al. 2004). Receptor tyrosine kinases (RTKs) are known to participate in numerous cellular behaviors and help regulate cell growth, differentiation, and migration. RTKs contain both intracellular and extracellular domains that allow binding of extracellular ligands to affect intracellular signal transduction. There is a family of RTKs that have extracellular domain motifs homologous to the Dictyostelium discoideium protein discoidin-I (Alves et al. 1995). The RTKs with discoidin-I homology have been named discoidin domain receptors (DDRs) (Vogel et al. 1997). Two distinct DDRs have been characterized (DDR1 and DDR2) (Vogel et al. 1997). Five isoforms of DDR1 are generated by alternative splicing (DDR1a-e). In situ analysis of several human primary mammary carcinomas has shown that the expression of DDR1 mRNA can be at least three-fold higher in tumor cells than in the adjacent normal epithelia (Barker et al. 1995). DDR1 is expressed in tumor cells themselves, while DDR2 has been detected in the stromal cells surrounding the tumor (Alves et al. 1995). Numerous breast carcinoma cell lines, as well as other carcinomas, have been shown to highly express DDR1 (Alves et al. 1995; Perez et al. 1996; Leitinger 2003). DDR1 and DDR2 are activated by fibrillar collagens, i.e., Tyr phosphorylation of DDR is substantially increased (Shrivastava et al. 1997; Vogel et al. 1997; Leitinger 2003). Triple-helical conformation is required for collagen to serve as a ligand for DDR1 and DDR2 (Shrivastava et al. 1997; Vogel et al. 1997; Agarwal et al. 2002). The specificity of DDR interaction with collagen is also dependent upon the ligand carbohydrate content. Deglycosylation of collagen by treatment with sodium m-periodate resulted in significant reduction of collagen-induced
488
J.L. Lauer and G.B. Fields
stimulation of DDR (Vogel et al. 1997; Wall et al. 2005). Thus, collagen requires both an intact triple-helical conformation and glycosylation for optimum induction of DDR. DDR1 is found as a disulfide-linked dimer, and dimerization is necessary for collagen binding and kinase activation (Abdulhussein et al. 2008). Fibrillar collagen induces DDR2-mediated tumor cell cycle arrest in the G0/G1phase (Wall et al. 2005). Binding to type I collagen ultimately results in DDR1 shedding, mediated by a member of the ADAM family (Slack et al. 2006). Stimulation of DDR1a and DDR1b overexpressing hepatocellular carcinoma cells by type I collagen induces the production of active MMP-2 (Park et al. 2007). Type I collagen-stimulated glioma cells that overexpress DDR1a also have increased levels of MMP-2 activity and enhanced invasion and migration capabilities (Ram et al. 2006). DDR2 stimulation by type II collagen upregulates MMP-1, MMP-2, and MMP-13 (Xu et al. 2005; Zhang et al. 2006). DDR2-mediated transactivation of the MMP-2 promoter requires Src activity (Ikeda et al. 2002). The type III collagen Toolkit was utilized to identify the Gly–Pro–Arg–Gly– Gln–Hyp–Gly–Val–Met–Gly–Phe–Hyp binding site for DDR2 (Farndale et al. 2008; Konitsiotis et al. 2008) (Table 23.2). Autophosphorylation of DDR2 was induced by this THP (Konitsiotis et al. 2008). X-ray crystallographic analysis of DDR2 interaction with acetyl–(Gly–Pro–Hyp)3–Gly–Pro–Arg–Gly–Gln–Hyp– Gly–Val–Nle–Gly–Phe–Hyp–(Gly–Pro–Hyp)2–Gly–NH2 indicated that the leading and middle chains of the triple helix accounted for 40 and 60% of the binding interface, respectively (Carafoli et al. 2009). The contribution of residues from differing chains explains why triple-helical structure is required for binding to DDR2. Interestingly, the THP bound to DDR2 and induced activation even though it does not contain a glycosylated Hyl residue, and thus appears at conflict with prior studies (see above). Preferential DDR2 binding sites in type I collagen have been noted at 35 ± 8, 65 ± 8, and 105 ± 8 nm from one end of the triple helix (Agarwal et al. 2002). Thus, other DDR binding motifs may exist.
Protein–Collagen Interactions Several components of the ECM play a role in collagen-mediated tumor cell behaviors. In this review, we address interactions between collagens and FN, LNs, or decorin. FN is a dimeric 440–550 kDa protein composed of highly homologous A and B chains [reviewed in Ruoslahti (1988)] (Fig. 23.4). It is found in two major forms, a soluble circulating form known as plasma FN (pFN), and an insoluble fibrillar form (fFN) associated with the ECM. FN is composed of three types of repeats, types I, II, and III. The type I repeat is ~45 residues long and makes up the N- and C-terminal domains of FN. Two 60-residue type II repeats interrupt a row of nine type I repeats near the N-terminus, whereas 15–17 type III repeats of ~90 residues in length, make up the middle portion of the molecule (Ruoslahti 1988). Fibrin, heparin, gelatin, collagen, and cellular binding sites are found within FN. Cell binding to FN is mediated by a variety of integrins, including a4b1, a5b1,
23 Collagen in Cancer
489
Fig. 23.4 The modular structure of FN showing the interaction regions. Reproduced by permission of the Creative Commons Attribution-Share Alike 3.0 Unported license and GNU Free Documentation License
a9b1, avb3, and avb6 (Cretu and Brooks 2007). Cryptic sites located between the first and tenth type III repeats are exposed by mechanical forces or protein–protein interactions to facilitate polymerization and incorporation into the ECM (Cretu and Brooks 2007). Gelatin and collagen binding sites are localized in a segment of type II repeats near the N-terminus of FN. While FN binds both native and denatured collagen, the affinity is much greater for denatured collagen (gelatin) compared with the intact native molecule. Type I collagen and FN have a complicated intertwining relationship, which helps modulate cell–matrix interactions. FN polymerization and incorporation into the ECM as insoluble material is required for the deposition of type I collagen and thrombospondin-1 as well as the localization of the a5b1 integrin (Sottile and Hocking 2002). Interestingly, the presence of soluble FN in vitro was necessary for the maintenance of the insoluble FN matrix. The RGD binding site located within the III10 module of FN was necessary but not sufficient for maintenance of the FN fibrils in the ECM (Sottile and Hocking 2002). FN binds to type I collagen at the same site where collagenolysis takes place, and binding to this site regulates fFN fibril formation (Dzamba et al. 1993). Interestingly, FN binding to collagen and FN fibril formation was decreased by the combination of Gln774Pro and Ala777Pro mutations in the a1(I) chain of type I collagen, but was not affected by Ile776Pro, Ile776Met, or Val782Ala + Val783Ala mutations (Dzamba et al. 1993). In contrast, all of the above mutations diminished or eliminated collagenolysis (Dzamba et al. 1993). FN appears to preferentially interact with the a1(I) collagen chain (Dzamba et al. 1993). The gelatin binding domain of FN located within 8–9FnI was shown to bind a peptide derived from the a1 chain of type I collagen as well as its analogous sequence in the a2 chain (Erat et al. 2009). Interestingly, kinetic unfolding experiments using a THP model of the a1(II)772–798 sequence demonstrated that the collagen triple helix was destabilized in a concentration dependent manner up to
490
J.L. Lauer and G.B. Fields
10°C by FN binding (Erat et al. 2009). Since the hydrolysis of triple-helical collagen requires a localized unwinding event and the FN binding site is located near the site of MMP hydrolysis, this suggests an important role for FN in regulating the proteolysis of type I collagen (Chung et al. 2004). The HPX domain of MMP-1 was shown to mediate this unwinding event, but the work of Erat et al. (2009) suggest that matrix proteins such as FN could participate as well. The gelatin-binding domain utilized in these studies was shown to simultaneously bind to multiple collagen peptides, suggesting that the protein could bind multiple chains within a collagen heterotrimer or might bind multiple chains between collagen heterotrimers. Cells of the immune system play an interesting role in modulating collagen in the ECM, in part through the expression of MMP-9. Monocytes secrete both FN and MMP-9, which is known to cleave types I, III, IV, V, and XI collagen as well as FN. The presence of native FN inhibits TNF-a induced proMMP-9 secretion by monocytes, whereas FN fragments reversed this inhibition (Marom et al. 2007). It was suggested that intact native FN present at the site of inflammation helps harness MMP-9 activity; however, the accumulation of fragmented FN alleviates inhibition allowing for faster migration of monocytes through the degraded ECM (Marom et al. 2007). Thus, the presence of intact FN plays a role in stabilizing collagen by inhibiting the expression of numerous collagenolytic and gelatinolytic enzymes. The LNs are large heterotrimeric extracellular glycoproteins composed of a, b, and g subunits. Currently, at least 16 different isoforms have been described (Cretu and Brooks 2007). This glycoprotein contains binding sites for a variety of receptors including the a1b1, a2b1, a3b1, avb3, a6b1, a6b4, and a7b1 integrins (Cretu and Brooks 2007). LN-5 (composed of a3, b3, and g2 subunits) is a BM component often implicated in cancer. LN-5 is found in the oral epithelium as well as the major isoform expressed by epidermal keratinocytes (Ziober et al. 2006). Type VII collagen helps attach the ECM to the underlying dermis via an interaction with LN-5 modulated by the FN-like region of the type VII collagen NC1 domain (amino acids 761–1,050) (Ortiz-Urda et al. 2005; Ziober et al. 2006). This interaction is critical for the invasion of squamous cell carcinoma cells, as cells devoid of type VII collagen were not tumorigenic (Ortiz-Urda et al. 2005). It was proposed that a physical interaction between the type VII collagen FN-like region and LN-5 was necessary and sufficient for invasion of squamous cell carcinoma cells (OrtizUrda et al. 2005). LN-5 supports the adhesion of melanoma cells, whereas hydrolysis of the g2 arm by MMP-2 or MT1-MMP, expressed by tumor cells and cells undergoing remodeling, supports cell migration (Koshikawa et al. 2000; Giles 2001; Tran et al. 2005). MMP-3, MMP-8, MMP-12, MMP-13, and MMP-20 also hydrolyze the same region in LN-5 in vitro, but their in vivo relevance has not been substantiated (Mott and Werb 2004). Interestingly, the site of hydrolysis appears to be critical, as MMP-8 cleavage, which occurs at a different site than the other enzymes, does not result in induction of cell migration (Pirila et al. 2003). Decorin is a member of the small Leu-rich proteoglycan family and functions in the regulation of collagen fibrillogenesis and the maintenance of tissue integrity (via its interactions with FN and thrombospondin), and acts as a reservoir for
23 Collagen in Cancer
491
TFG-b (Imai et al. 1997). Decorin binds types I, II, III, VI, and XIV collagen, thus delaying fibril assembly such that the average fibril diameter can be optimized to regulate fibrillogenesis (Imai et al. 1997). In the absence of decorin, collagen fibrils are irregular in size and shape and appear as loosely packed matrices (Imai et al. 1997; Reed and Iozzo 2003). Since it appears to be a vital mediator of collagen fibrillogenesis, the presence of decorin is essential in maintaining collagenous ECM. Several MMPs, including MMP-2, MMP-3, and MMP-7 have been shown to hydrolyze decorin, creating an inability to mediate fibrillogenesis as well as a release of TGF-b from the ECM (Imai et al. 1997). The absence of decorin alone was not sufficient to produce spontaneous tumorigenesis, whereas the combined ablation of decorin and p53 showed a faster rate of tumor development and nearly uniform death due to the development of thymic lymphomas compared with p53 knockouts (Iozzo et al. 1999). Decorin is known to bind the EGF receptor, which might mediate tumorigenesis; however, the analysis of collagen was not included in these studies. Thus, it is unclear whether or not decorin-mediated fibrillogenesis was important in the development of thymic lymphomas.
Matrix Metalloproteinases Alterations in activities of one family of proteases, the MMPs, have been implicated in primary and metastatic tumor growth, angiogenesis, and pathological degradation of ECM components, such as collagen, FN, LN, and decorin (Egeblad and Werb 2002; Overall and Lopez-Otin 2002; Fingleton 2007). The roles that MMPs play in the metastatic process are diverse and include involvement in primary and metastatic tumor growth, angiogenesis, and degradation of BM barriers (for instance collagen) during tumor cell invasion (Nelson et al. 2000). Multiple studies have established correlation between MMP production and metastasis. For example, melanoma cells have been found to express MMPs with collagenolytic activity, including MMP-1, MMP-2, MMP-9, MMP-13, and MMP-14/MT1-MMP (Hofmann et al. 2000). When localized to invadopodia, MT1-MMP activates MMP-2 on the cell surface (Hofmann et al. 2000). Several membrane-type-MMPs (MT-MMPs), including MT1-MMP, MT2-MMP, and MT3-MMP, are upregulated in metastatic melanoma (Ohnishi et al. 2001). MT1-MMP promotes melanoma invasion and growth (Iida et al. 2004), and appears to be a critical protease for tumor cell invasion of the ECM and dissemination (Hotary et al. 2000; Sabeh et al. 2004; Szabova et al. 2007; Devy et al. 2009). Furthermore, MT1-MMP has been demonstrated to catalyze CD44 shedding and subsequently promote cell migration (Kajita et al. 2001). MMP-2 and MMP-9 are elevated in tumor tissue of brain, prostate, ovarian, and colorectal cancer patients, while MMP-9 is elevated in breast cancer tumor tissue and MMP-2 in pancreatic cancer tumor tissue (Roy et al. 2009). MMP-13 is elevated in breast and colorectal tumor tissue, while MT1-MMP is elevated in ovarian tumor tissue and MMP-1 in colorectal tumor tissue (Roy et al. 2009). Tumor
492
J.L. Lauer and G.B. Fields
progression has been inversely correlated to MMP-3, MMP-8, and MMP-12 expression, making these MMPs antitargets of tumor therapy (Roy et al. 2009). The normal physiological roles of these MMPs may be antitumorigenic or critical for patient survival (Page-McCaw et al. 2007). The roles of MMPs in cancer are covered in more detail in Chap. 28. The “collagenolytic” MMPs [MMPs that catalyze the hydrolysis of one or more of the interstitial collagens (types I-III) within their triple-helical domain] include the secreted proteases MMP-1, MMP-2, MMP-8, MMP-9, and MMP-13 and the membrane-bound proteases MT1-MMP and MT2-MMP (McCawley and Matrisian 2001; Overall 2002; Morrison and Overall 2006; Bigg et al. 2007). In the cases of MMP-1, MMP-8, MMP-13, MT1-MMP, and MT2-MMP, efficient collagenolytic activity for the isolated enzyme requires both the catalytic (CAT) and HPX domains (Clark and Cawston 1989; Murphy et al. 1992; Knäuper et al. 1993, 1997; Ohuchi et al. 1997; Hurst et al. 2004). The linker region between these domains also participates in collagenolysis, either by direct binding of substrate (De Souza et al. 1996) or by allowing for the proper orientation of the CAT and HPX domains (Iyer et al. 2006). The gelatinase members of the MMP family (MMP-2 and MMP-9) possess three FN type II (FN II) inserts within their CAT domains, and these inserts possess similar type I collagen binding sites (Steffensen et al. 1995; Xu et al. 2005). While collagenolytic MMPs possess common domain organizations, there are subtle differences in their processing of triple-helical substrates. However, all collagenolytic MMPs hydrolyze types I–III collagen at the Gly775–Xxx776 bond, where Xxx is Ile or Leu (Fields 1991; Woessner and Nagase 2000). Within heterotrimeric type I collagen, the a2(I) chain is preferentially bound by MMPs (Chung et al. 2004). Homotrimeric type I collagen, which is found in tumors and cancer cells lines (see earlier discussion), is much more resistant to collagenolytic MMPs than heterotrimeric type I collagen (Narayanan et al. 1984). This may be due to the increased thermal stability, slower denaturation rate, and/or different microunfolding patterns of homotrimeric type I collagen compared with heterotrimeric type I collagen (Kuznetsova et al. 2003). Tumors containing homotrimeric type I collagen may possess enhanced matrix stiffness, which in turn upregulates cell signaling pathways (see earlier discussion). Members of the MMP family are also active against other collagens. For example, type IV collagen is processed by MMP-3 at a1(IV)Gly1369–Leu1370 and MMP-9 at a1(IV)Gly1429–Leu1430 (Mott et al. 1997). Several sites of MMP-2 hydrolysis of type IV collagen have been described, but these appear to occur only at higher temperatures (37°C) (Eble et al. 1996). Interestingly, hydrolysis of type IV collagen by MMP-2 does not impact the binding of the a1b1 and a2b1 integrins (Eble et al. 1996). Dimeric type IV collagen with intact disulfide bonds is not hydrolyzed by MMP-2 even at 37°C (Eble et al. 1996). MMP-9 cleaves types V and XI collagen at the Gly439–Val440 bond (Niyibizi et al. 1994). MMPs hydrolyze types IV, XV, and XVIII collagen to release anti-angiogenic fragments (see below). The hydrolysis of type IV collagen by MMP-9, which is produced by T cells, neutrophils, macrophages, monocytes, as well as a variety of tumor cells, releases sequestered VEGF, an event correlated with enhanced tumor progression (Bergers
23 Collagen in Cancer
493
et al. 2000; Belotti et al. 2003). VEGF produces positive feedback signals, which support tumor cell growth and recruit immune cells, fibroblasts, and other stromal cells to the site of angiogenesis. Cryptic sites within type IV collagen are also exposed by MMP action, providing further upregulation of angiogenesis (see below).
Collagen Fragments The physiological response to tissue injury is similar in some ways to that of angiogenesis. Both of these processes bear a resemblance to tumor cell metastasis, in part because neovascularization contributes to the successful growth and establishment of solid tumors as these processes require vascular support (Bellon et al. 2004). Thus, the responses to each of these stimuli will be addressed at once. Intact ECM molecules relay different signals to the surrounding cells compared with partially degraded components. During tissue injury, the composition of the ECM and its cellular interaction sites are altered (Davis et al. 2000). Once matrix proteins are compromised, cryptic sites are exposed. Collagen fragments have been referred to as cryptides (part of the cryptome), matrikines, and matricryptins (Davis et al. 2000; Bellon et al. 2004; Maquart et al. 2004; Autelitano et al. 2006) (Table 23.3). Downstream events resulting from exposure of cryptic sites include the recruitment of pFN, vitronectin, and fibrinogen into the ECM via increased vascular permeability as well as cellular release or synthesis of osteopontin, SPARC, thrombospondins, tenascins, or alternatively spliced FNs (Davis et al. 2000). Examples of cryptic sites and processes important for ECM remodeling include: (1) fibrin fragments that increase vascular permeability; (2) collagen, LN, FN, and elastin fragments that stimulate cell migration, cell proliferation, focal contact disassembly, or vasodilation; and (3) osteopontin, LN, or tenascin-induced cell adhesion (Davis et al. 2000). Exposure of new binding sites results in engagement of alternative integrins. Involvement of these integrins (such as avb3 or a5b1) induces new cell signaling pathways and alters the effects of ECM mechanical force (Geiger et al. 2001; Hood and Cheresh 2002). Binding to ECM ligands via the a5b1 and avb3 integrins leads to SHC-, PI3K (phosphatidylinositol 3-kinase)-, and FAK-mediated cell survival (Hood and Cheresh 2002). Hydrolysis of type I collagen by interstitial collagenases, followed by thermal denaturation, exposes Arg–Gly–Asp sites that are bound by the avb3 integrin (Davis 1992; Berman et al. 1993). Hydrolysis of type IV collagen by an apparent combination of serine proteases and MMPs (such as MMP-2) revealed a cryptic binding site that promoted avb3 integrin-mediated binding (Xu et al. 2001). The proteolytic action on type IV collagen removed the a1b1 integrin recognition site, but a2b1 integrin interactions were unaffected. The presence of the neoepitope was associated with angiogenic, but not quiescent blood vessels and was required for angiogenesis in vivo (Xu et al. 2001). Integrin-mediated cell attachment to native type I collagen is mainly via the a1b1, a2b1, a10b1, and a11b1 integrins. Denatured type I collagen, however,
Type XIX collagen
Type IV collagen Type VII collagen
NC1(XIX)
NC1 (a6 IV) FNC1
Promotes tumor cell invasion
Inhibits tumor cell growth, migration, and invasion. Inhibits MT1-MMP, VEGF expression
Inhibits adhesion Inhibits angiogenesis Inhibits EC proliferation Promote neutrophil migration
Type I collagen Type XV collagen Type VIII collagen Type I collagen
Restin Vastatin PGP peptide
Type I collagen
RGD site
Inhibits migration
Type I collagen
Type IV collagen
Hexastatin-3
Inhibits migration
Support adhesion and vasodilation Binds FN, MMP-2, MMP-9
Type IV collagen
Pentastatin-2
Table 23.3 Biological activities of collagen fragments Fragment Origin Bioactivity Endostatin Type XVIII Inhibits angiogenesis and collagen tumor cell growth Inhibits angiogenesis and Tumstatin Type IV collagen (a3) tumor progression (derived from different domains) Arrestin Type IV collagen Inhibits angiogenesis Canstatin Type IV collagen Inhibits angiogenesis Pentastatin-1 Type IV collagen Antiproliferative Target
avb3 integrin b3 chain of LN-5
CXCR1, CXCR2
avb3 integrin
avb3, b1 integrins
avb3, a1b1 integrins avb3, a3b1 integrins avb3, b1 integrins
avb3, avb5, a6b1 integrins
a5b1, avb3 integrins
Prolyl endopeptidase + MMP-8 and/or -9
MMP-1, elastase, thermal denaturation Protease, thermal denaturation Bacterial collagenase
Produced by Elastase, Cathepsin L, Cathepsin B, MMP-7 MMP-9 (primary), MMP-2, MMP-3, and MMP-13
Petitclerc et al. (2000) Ortiz-Urda et al. (2005)
Davis et al. (2000) Ramchandran et al. (1999) Xu et al. (2001) Adair-Kirk and Senior (2008) Ramont et al. (2007)
Davis et al. (2000)
Colorado et al. (2000) Kamphaus et al. (2000) Karagiannis and Popel (2007, 2008) Karagiannis and Popel (2007, 2008) Karagiannis and Popel (2007, 2008) Davis et al. (2000)
Maeshima et al. (2002)
Reference O’Reilly et al. (1997)
494 J.L. Lauer and G.B. Fields
23 Collagen in Cancer
495
supports RGD-dependent adhesion not found in the native molecule. The avb3 integrin binds denatured type I collagen, but a2b1 integrin does not. The converse is true for the native protein. This regulation of receptor binding has important implications for modulation of physiological response to collagen destruction. avb3 integrin-mediated binding has been shown to be important for tumor formation as well as mediating an interaction with MMP-2, which facilitates proteolysis of denatured collagen, FN, and LN at the cell surface (Brooks et al. 1996; Davis et al. 2000). MMP-2 and MMP-9-mediated hydrolysis of these matrix proteins produces additional biologically active cryptic peptides, both with and without RGD-binding sites. MMP-2 correlates with tumor angiogenesis, whereas MMP-9 correlates with retinal angiogenesis, suggesting different functions based on microenvironment (Mott and Werb 2004). Specifically, proteolytic fragments of denatured type I collagen was shown to induce avb3-mediated vasodilation of arterioles in a manner similar to synthetic RGD-containing peptides (Mogford et al. 1996). Interestingly, MMP-2 and the avb3 integrin were found functionally associated on the surface of angiogenic blood vessels. This interaction is presumably mediated by the MMP-2 C-terminal HPX domain because administration of exogenous HPX domain prevents MMP-2 activity (Brooks et al. 1998). Conversely, fragments of types IV, XV, and XVIII collagen can inhibit tumor cell growth and angiogenesis (O’Reilly et al. 1997; Colorado et al. 2000; Petitclerc et al. 2000; Myllyharju and Kivirikko 2001; Kim et al. 2002; Maeshima et al. 2002) (Table 23.3). MMP-mediated degradation produces collagen fragments with antiangiogenic activity such as endostatin, arrestin, canstatin, and tumstatin. Endostatin, the NC1 domain of type XVIII collagen, is generated by MMP-7-mediated hydrolysis or by a combination of elastase and MMP cleavage (Mott and Werb 2004). The type IV collagen fragment tumstatin can be generated in vitro by MMP-9, and MMP-9 null mice have decreased levels of circulating tumstatin combined with accelerated tumor growth (Hamano et al. 2003). A bioinformatics approach identified pentastatin-1 from the a5(IV) chain as an inhibitor of tumor cell proliferation and angiogenesis (Karagiannis and Popel 2008; Koskimaki et al. 2009). Pentastatin-1 binds to b1 integrins (Karagiannis and Popel 2008). Additionally, tetrasratin-1, -2, and -3 from the a4(IV) chain, penstastatin-1, -2, and -3 from the a5(IV) chain, and hexastatin-1 and -2 from the a6(IV) chain inhibit human umbilical vein endothelial cell proliferation and migration (Karagiannis and Popel 2008). It has not been determined if any of the tetrasratin, pentastatin, or hexastatin species are actually released from the a(IV) chains by proteolytic activity. The hydrolysis of type XVIII collagen results in the release of a 20 kDa C-terminal fragment named endostatin, which inhibits angiogenesis (Wen et al. 1999). Subsequently, MMPs such as MMP2, MMP-3, MMP-9, MMP-12, MMP-13, MMP-20, and MT1-MMP were shown to generate endostatin-containing fragments ranging in size from 20 to 30 kDa, which also inhibited the proliferation and migration of human endothelial cells (Ferreras et al. 2000). As the tumor grows and MMP production accelerates, what is left in the BM is mainly composed of MMP-resistant products such as those listed above. Thus, MMP hydrolysis generates pro-angiogenic signals in the early stages of tumor cell
496
J.L. Lauer and G.B. Fields
growth and metastasis, whereas it produces anti-angiogenic signals during the later stages of the process (Kalluri 2003). Another important downstream event is the enhanced binding of FN to denatured type I collagen (Johansson and Hook 1980). It is possible that FN-denatured collagen complexes present cryptic peptides such that a signal is transmitted for the surrounding cells to differentiate between intact and damaged ECM proteins (Davis et al. 2000). It is important to keep in mind that the presentation of cryptic sites happens in response to a variety of processes, not merely tissue destruction. Cryptic sites can be exposed upon enzymatic degradation, adsorption of heterotypic molecules, adsorption of homotypic molecules, cell-mediated mechanical force, and thermal denaturation (as produced by fever, localized inflammation, or tissue burns) (Davis et al. 2000). Heterotypic binding is especially important during acute inflammation or tumorigenesis. Both of these events generate a provisional ECM containing larger amounts of fibrin and FN (Davis et al. 2000). In some cases, homotypic multimerization stimulates the deposition of insoluble ECM, such as the exogenous addition of a 14-kDa FN fragment, which induced the accumulation of insoluble FN via covalent cross-links (Morla et al. 1994).
Fibronectin Fragments Fragments derived from FN hydrolysis have numerous critical physiological and pathological roles in modulating collagen function. Cryptic fragments of FN inhibit cell proliferation, induce proteinase gene expression, and induce cellular migration (Mott and Werb 2004). MMP-2 cleaves FN into numerous fragments, including ones of 120, 70, 45, and 11 kDa, and ovarian carcinoma cell adhesion increased with FN fragments compared to intact FN (Kenny et al. 2008). The integrins contributing to adhesion to FN fragments were a5b1 and avb3 (Kenny et al. 2008). A FN fragment containing the gelatin-binding domain inhibited movement of cells through a three-dimensional type I collagen matrix (Schor et al. 1996). This behavior was seen only when cells were migrating through native type I collagen, as the FN fragment did not stimulate migration in the absence of substratum or in the presence of type I gelatin. Intact FN as well as peptides containing all of the other functional domains did not support migration in this assay. Interestingly, this stimulation of migration was witnessed in cells plated for 24 h in the presence of the FN fragment prior to the addition to a preparation of type I collagen matrix containing no additional fragment (Schor et al. 1996). Thus, the effects could be seen even after the fragment was removed, suggesting a proteolytic or molecular mechanism rather than a direct physical denaturation of the collagen matrix. This illustrates another means for FN fragments to alter cellular behavior on collagen-containing ECM. Cell adhesion to FN fragments increased secretion of MMP-2, MMP-9, and MMP12, and pro-inflammatory cytokines IL-1, IL-6, and TNF-a, whereas intact FN did not (Werb et al. 1989; Homandberg et al. 1992; Al-Hazmi et al. 2007). It was proposed
23 Collagen in Cancer
497
that a5b1-mediated cell binding inhibited MMP gene expression. Two fragments derived from the N-terminal region of FN, the 29 and 50 kDa fragments, also caused enhanced gelatinase and collagenase activity as well as proteoglycan release from bovine articular cartilage explants. Reduced and alkylated fragments were inactive, suggesting an intact three-dimensional structure is important for cell binding. Both of these fragments contain the gelatin-binding domain, whereas intact FN and a larger (140 kDa) fragment containing the integrin-binding site had little or no activity (Homandberg et al. 1992). Using the ovarian cancer model, MMP-2-mediated hydrolysis of FN and the ECM protein vitronectin were shown to be critical for the initial step of metastasis, that of peritoneal adhesion (Kenny et al. 2008). Cells adhered more efficiently to hydrolyzed FN and VN compared with full-length proteins. In addition to supporting a variety of cell-modulating activities, FN fragments also contain protease activity. When FN is digested by cathepsin D, three fragments are initially produced: 70, 140 kDa single chain and 140 kDa double chain (Keildlouha and Planchenault 1986). The N-terminal 70-kDa fragment is able to hydrolyze native and fragmented FN, and thus termed fibronectinase (FNase). FNase is subsequently processed in the presence of Ca2+ into 27-kDa heparin binding and 45-kDa gelatin-binding domains (Vidmar et al. 1991). The 27-kDa fragment undergoes further processing to form an active serine proteinase that is able to cleave native and fragmented FN (Vidmar et al. 1991). FNase processes the 45-kDa fragment at two different sites to produce a 40-kDa protein known as FN-collagenase A or a 27-kDa protein called FN-collagenase B. These newly generated fragments have protease activity, with FN-collagenase A cleaving type IV collagen, type I gelatin, LN, a-casein, b-casein, and insulin, whereas FN-collagenase B cleaves types II and IV collagen (Vidmar et al. 1991; Schepel and Tschesche 2000). Interestingly, FN-collagenase A is a metalloproteinase with similarities in sequence to MMP-2 and MMP-9 active sites. The second and third His of the metzincin Zn-binding motif HExxHxxGxxH are not present in FN-collagenase A, but there are four additional His residues nearby which could potentially be involved in Zn binding (Schepel and Tschesche 2000). To date, none of these enzymes have been shown to be physiologically relevant, but it is intriguing to think that the hydrolysis of FN could modulate collagen on so many levels, such as binding, stability, and hydrolysis. Acknowledgments We gratefully acknowledge the National Institutes of Health (CA98799, EB000289, and MH078948), the Robert A. Welch Foundation, and the Texas Higher Education Science and Technology Acquisition and Retention (STAR) Award (all to GBF) for support of our research on collagen and metalloproteases and the National Institutes of Health NIDCR (DE14318) COSTAR Program (to JL).
References Abdulhussein R, Koo DHH, Vogel WF (2008) Identification of disulfide-linked dimers of the receptor tyrosine kinase DDR1. J Biol Chem 283:12026–12033. Adair-Kirk TL, Senior RM (2008) Fragments of extracellular matrix as mediators of inflammation. Int J Biochem Cell Biol 40:1101–1110.
498
J.L. Lauer and G.B. Fields
Agarwal G, Kovac L, Radziejewski C et al (2002) Binding of discoidin domain receptor 2 to collagen I: an atomic force microscopy investigation. Biochemistry 41:11091–11098. Ahrens T, Assmann V, Fieber C et al (2001) CD44 is the principal mediator of hyaluronicacid-induced melanoma cell proliferation. J Invest Dermatol 116:93–101. Al-Hazmi N, Thomas GJ, Speight PM et al (2007) The 120 kDa cell-binding fragment of fibronectin up-regulates migration of alpha v beta 6-expressing cells by increasing matrix metalloproteinase-2 and-9 secretion. Eur J Oral Sci 115:454–458. Alves CS, Yakovlev S, Medved L et al (2009) Biomolecular characterization of CD44-fibrin(ogen) binding. J Biol Chem 284:1177–1189. Alves F, Vogel W, Mossie K et al (1995) Distinct structural characteristics of discoidin I subfamily receptor tyrosine kinases and complementary expression in human cancer. Oncogene 10:609–618. Anderegg U, Eichenberg T, Parthaune T et al (2009) ADAM10 is the constitutive functional sheddase of CD44 in human melanoma cells. J Invest Dermatol 129:1471–1482. Asokan R, Reddy GK, Dhar SC (1993) Neoplastic association of enhanced type-V collagen production in rat fibrosarcoma. Mol Cell Biochem 120:25–32. Autelitano DJ, Rajic A, Smith AI et al (2006) The cryptome: a subset of the proteome, comprising cryptic peptides with distinct bioactivities. Drug Discov Today 11:306–314. Bao X, Kobayashi M, Hatakeyama S et al (2009) Tumor suppressor function of laminin-binding a-dystroglycan requires a distinct b3-N-acetylglucosaminyltransferase. Proc Natl Acad Sci USA 106:12109–12114. Barker KT, Martindale JE, Mitchell PJ et al (1995) Expression patterns of the novel receptor-like tyrosine kinase, DDR, in human breast tumors. Oncogene 11:569–575. Barnes MJ, Knight CG, Farndale RW (1996) The use of collagen-based model peptides to investigate platelet-reactive sequences in collagen. Biopolymers (Peptide Sci) 40:383–397. Baronas-Lowell D, Lauer-Fields JL, Borgia JA et al (2004a) Differential modulation of human melanoma cell metalloproteinase expression by a2b1 integrin and CD44 triple-helical ligands derived from type IV collagen. J Biol Chem 279:43503–43513. Baronas-Lowell D, Lauer-Fields JL, Fields GB (2004b) Induction of endothelial cell activation by a triple-helical a2b1 integrin ligand derived from type I collagen a1(I)496–507. J Biol Chem 279:952–962. Barth L, Sinner EK, Cadamuro SA et al (2009) Homotrimeric collagen peptides as model systems for cell adhesion studies. In: Escher E, Lubell WD, Del Valle S (eds) Peptides for youth: the proceedings of the 20th American Peptide Symposium. Springer Science, New York, pp 295–296. Baum J, Brodsky B (1999) Folding of peptide models of collagen and misfolding in disease. Curr Opin Struct Biol 9:122–128. Bella J, Berman HM (2000) Integrin-collagen complex: a metal-glutamate handshake. Structure 8:R121–R126. Bellon G, Martiny L, Robinet A (2004) Matrix metalloproteinases and matrikines in angiogenesis. Crit Rev Oncol/Hematol 49:203–220. Belotti D, Paganoni P, Manenti L et al (2003) Matrix metalloproteinases (MMP9 and MMP2) induce the release of vascular endothelial growth factor (VEGF) by ovarian carcinoma cells: implications for ascites formation. Cancer Res 63:5224–5229. Bergers G, Brekken RA, McMahon G et al (2000) Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat Cell Biol 2:737–744. Berman A, Morozevich G, Karmansky I et al (1993) Adhesion of mouse hepatocytes to type I collagen: role of supramolecular forms and effect of proteolytic degradation. Biochem Biophys Res Commun 194:351–357. Bigg HF, Rowan AD, Barker MD et al (2007) Activity of matrix metalloproteinase-9 against native collagen types I and III. FEBS J 274:1246–1255. Boulégue C, Musiol H-J, Götz MG et al (2008) Natural and artificial cystine knots for assembly of homo- and heterotrimeric collagen models. Antioxid Redox Signal 10:113–125. Bourguignon LYW, Gunja-Smith Z, Iida N et al (1998a) CD44v3,8–10 is involved in cytoskeletonmediated tumor cell migration and matrix metalloproteinase (MMP-9) association in metastatic breast cancer cells. J Cell Physiol 176:206–215.
23 Collagen in Cancer
499
Bourguignon LYW, Zhu D, Zhu H (1998b) CD44 isoform-cytoskeleton interaction in oncogenic signaling and tumor progression. Frontiers Biosci 3:637–649. Brooks PC, Silletti S, von Schalscha TL et al (1998) Disruption of angiogenesis by PEX, a noncatalytic metalloproteinase fragment with integrin binding activity. Cell 92:391–400. Brooks PC, Strömblad S, Sanders LC et al (1996) Localization of matrix metalloproteinase MMP-2 to the surface of invasive cells by interaction with integrin avb3. Cell 85:683–693. Carafoli F, Bihan D, Stathopoulos S et al (2009) Crystallographic insight into collagen recognition by discoidin domain receptor 2. Structure 17:1573–1581. Chelberg MK, McCarthy JB, Skubitz APN et al (1990) Characterization of a synthetic peptide from type IV collagen that promotes melanoma cell adhesion, spreading, and motility. J Cell Biol 111:261–270. Chung L, Dinakarpandian D, Yoshida N et al (2004) Collagenase unwinds triple helical collagen prior to peptide bond hydrolysis. EMBO J 23:3020–3030. Clark EA, Golub TR, Lander ES et al (2000) Genomic analysis of metastasis reveals an essential role for RhoC. Nature 406:532–535. Clark IN, Cawston TE (1989) Fragments of human fibroblast collagenase: purification and characterization. Biochem J 263:201–206. Cole WG (1994) Collagen genes: mutations affecting collagen structure and expression. Prog Nucleic Acid Res Mol Biol 47:29–80. Colorado PC, Torre A, Kamphaus G et al (2000) Anti-angiogenic cues from vascular basement membrane collagen. Cancer Res 60:2520–2526. Cretu A, Brooks PC (2007) Impact of the non-cellular tumor microenvironment on metastasis: potential therapeutic and imaging opportunities. J Cell Physiol 213:391–402. Davis GE (1992) Affinity of integrins for damaged extracellular matrix: avb3 binds to dematured collagen type I through RGD sites. Biochem Biophys Res Commun 182:1025–1031. Davis GE, Bayless KJ, Davis MJ et al (2000) Regulation of tissue injury responses by the exposure of matricryptic sites within extracellular matrix molecules. Am J Pathol 156:1489–1498. De Souza SJ, Pereira HM, Jacchieri S et al (1996) Collagen/collagenase interaction: does the enzyme mimic the conformation of its own substrate? FASEB J 10:927–930. DeClerck YA, Bomann ET, Spengler BA et al (1987) Differential collagen biosynthesis by human neuroblastoma cell variants. Cancer Res 47:6505–6510. Dennis J, Waller C, Timpl R et al (1982) Surface sialic acid reduces attachment of metastatic tumour cells to collagen type IV and fibronectin. Nature 300:274–276. Desai B, Rogers MJ, Chellaiah MA (2007) Mechanisms of osteopontin and CD44 as metastatic principles in prostate cancer cells. Mol Cancer 6:18. Devy L, HUang L, Naa L et al (2009) Selective inhibition of matrix metalloproteinase-14 blocks tumor growth, invasion, and angiogenesis. Cancer Res 69:1517–1526. Di Lullo GA, Sweeney SM, Körkkö J et al (2002) Mapping the ligand-binding sites and diseaseassociated mutations on the most abundant protein in the human, type I collagen. J Biol Chem 277:4223–4231. Du L, Wang H, He L et al (2008) CD44 is of functional importance for colorectral cancer stem cells. Clin Cancer Res 14:6751–6760. Dzamba BJ, Wu H, Jaenisch R et al (1993) Fibronectin binding site in type I collagen regulates fibronectin fibril formation. J Cell Biol 121:1165–1172. Eble J, Golbik R, Mann K et al (1993) The a1b1 integrin recognition site of the basement membrane collagen molecule a1(IV)2a2(IV). EMBO J 12:4795–4802. Eble JA, Ries A, Lichy A et al (1996) The recognition sites of the integrins a1b1 and a2b1 within collagen IV are protected against gelatinase A attack in the native protein. J Biol Chem 271:30964–30970. Egeblad M, Werb Z (2002) New functions for the matrix metalloproteinases in cancer progression. Nat Rev Cancer 2:161–174. Ehnis T, Dieterich W, Bauer M et al (1996) A chondroitin/dermatan sulfate form of CD44 is a receptor for collagen XIV (undulin). Exp Cell Res 229:388–397. Eliaz RE, Szoka J, F.C. (2001) Liposome-encapsulated doxorubicin targeted to CD44: A strategy to kill CD44-overexpressing tumor cells. Cancer Res 61:2592–2601.
500
J.L. Lauer and G.B. Fields
Emsley J, Knight CG, Farndale RW et al (2004) Structure of the integrin a2b1-binding collagen peptide. J Mol Biol 335:1019–1028. Emsley J, Knight CG, Farndale RW et al (2000) Structural basis of collagen recognition by integrin a2b1. Cell 101:47–56. Erat MC, Slatter DA, Lowe ED et al (2009) Identification and structural analysis of type I collagen sites in complex with fibronectin fragments. Proc Natl Acad Sci USA 106:4195–4200. Faassen AE, Mooradian DL, Tranquillo RT et al (1993) Cell surface CD44-related chondroitin sulfate proteoglycan is required for transforming growth factor-b-stimulated mouse melanoma cell motility and invasive behavior on type I collagen. J Cell Sci 105:501–511. Faassen AE, Schrager JA, Klein DJ et al (1992) A cell surface chondroitin sulfate proteoglycan, immunologically related to CD44, is involved in type I collagen-mediated melanoma cell motility and invasion. J Cell Biol 116:521–531. Farndale RW, Lisman T, Bihan D et al (2008) Cell-collagen interactions: the use of peptide Toolkits to investigate collagen-receptor interactions. Biochem Soc Trans 36:241–250. Ferreras M, Felbor, U., Lenhard, T., Olsen, B. R., Delaisse, J. (2000) Generation and degradation of human endostatin proteins by various proteinases. FEBS Lett 486(3): 247–251. Fields CG, Mickelson DJ, Drake SL et al (1993) Melanoma cell adhesion and spreading activities of a synthetic 124-residue triple-helical “mini-collagen”. J Biol Chem 268:14153–14160. Fields GB (1991) A model for interstitial collagen catabolism by mammalian collagenases. J Theor Biol 153:585–602. Fields GB, Lauer JL, Dori Y et al (1998) Proteinlike molecular architecture: biomaterial applications for inducing cellular receptor binding and signal transduction. Biopolymers 47:143–151. Fingleton B (2007) Matrix metalloproteinases as valid clinical targets. Curr Pharm Des 13: 333–346. Friedl P, Maaser K, Klein CE et al (1997) Migration of highly aggressive MV3 melanoma cells in 3-dimensional collagen lattices results in local matrix reorganization and shedding of a2 and b1 integrins and CD44. Cancer Res 57:2061–2070. Fujisaki T, Tanaka Y, Fujii K et al (1999) CD44 stimulation induces integrin-mediated adhesion of colon cancer cell lines to endothelial cells by up-regulation of integrins and c-Met activation of integrins. Cancer Res 59:4427–4434. Geiger B, Bershadsky A, Pankov R et al (2001) Transmembrane extracellular matrix-cytoskeleton crosstalk. Nat Rev Mol Cell Biol 2:793–805. Giles FJ (2001) The vascular endothelial growth factor (VEGF) signaling pathway: a therapeutic target in patients with hematologic malignancies. Oncologist 6:32–39. Goebeler M, Kaufmann D, Brocker EB et al (1996) Migration of highly aggressive melanoma cells on hyaluronic acid is associated with functional changes, increased turnover and shedding of CD44 receptors. J Cell Sci 109:1957–1964. Golbik R, Eble JA, Ries A et al (2000) The spatial orientation of the essential amino acid residues arginine and aspartate within the a1b1 integrin recognition site of collagen IV has been resolved using fluorescence resonance energy transfer. J Mol Biol 297:501–509. Gordon MK, Hahn RA (2009) Collagens. Cell Tissue Res 339:247–257. Griffioen AW, Coenen MJH, Damen CA et al (1997) CD44 is involved in tumor angiogenesis; an activation antigen on human endothelial cells. Blood 90:1150–1159. Hamano Y, Zeisberg M, Sugimoto H et al (2003) Physiological levels of tumstatin, a fragment of collagen IV alpha 3 chain, are generated by MMP-9 proteolysis and suppress angiogenesis via alpha V beta 3 integrin. Cancer Cell 3:589–601. Hashimoto T, Wakabayashi T, Watanabe A et al (2002) CLAC: a novel Alzheimer amyloid plaque component derived from a transmembrane precursor, CLAC-P/collagen type XXV. EMBO J 21:1524–1534. Herbold KW, Zhou J, Haggerty JG et al (1996) CD44 expression on epidermal melanocytes. J Invest Dermatol 106:1230–1235.
23 Collagen in Cancer
501
Herr AB, Farndale RW (2009) Structural insights into the interactions between platelet receptors and fibrillar collagen. J Biol Chem 284:19781–19785. Hofmann UB, Westphal JR, van Muijen GNP et al (2000) Matrix metalloproteinases in human melanoma. J Invest Dermatol 115:337–344. Homandberg GA, Meyers R, Xie DL (1992) Fibronectin fragments cause chondrolysis of bovine articular-cartilage slices in culture. J Biol Chem 267:3597–3604. Hood JD, Cheresh DA (2002) Role of integrins in cell invasion and migration. Nat Rev Cancer 2:91–100. Hornebeck W, Emonard H, Monboisse JC et al (2002) Matrix-directed regulation of pericellular proteolysis and tumor progression. Semin Cancer Biol 12:231–241. Hostikka SL, Tryggvason K (1988) the complete primary structure of the a2 chain of human type IV collagen and comparison with the a1(IV) chain. J Biol Chem 263:19488–19493. Hotary K, Allen E, Punturieri A et al (2000) Regulation of cell invasion and morphogenesis in a three-dimensional type I collagen matrix by membrane-type matrix metalloproteinases 1, 2, and 3. J Cell Biol 149:1309–1323. Hurst DR, Schwartz MA, Ghaffari MA et al (2004) Catalytic- and ecto-domains of membrane type 1-matrix metalloproteinase have similar inhibition profiles but distinct endopeptidase activities. Biochem J 377:775–779. Iida J, Wilhelmson KL, Price MA et al (2004) Membrane type-1 matrix metalloproteinase promotes human melanoma invasion and growth. J Invest Dermatol 122:167–176. Ikeda K, Wang LH, Torres R et al (2002) Discoidin domain receptor 2 interacts with Src and Shc following its activation by type I collagen. J Biol Chem 277:19206–19212. Imai K, Hiramatsu A, Fukushima D et al (1997) Degradation of decorin by matrix metalloproteinases: identification of the cleavage sites, kinetic analyses and transforming growth factor-beta 1 release. Biochem J 322:809–814. Iozzo RV, Chakrani F, Perrotti D et al (1999) Cooperative action of germ-line mutations in decorin and p53 accelerates lymphoma tumorigenesis. Proc Natl Acad Sci USA 96:3092–3097. Iyer S, Visse R, Nagase H et al (2006) Crystal structure of an active form of human MMP-1. J Mol Biol 362:78–88. Johansson S, Hook M (1980) Heparin enhances the rate of binding of fibronectin to collagen. Biochem J 187:521–524. Kajita M, Itoh Y, Chiba T et al (2001) Membrane-type 1 matrix metallproteinase cleaves CD44 and promotes cell migration. J Cell Biol 153:893–904. Kalluri R (2003) Basement membranes: structure, assembly and role in tumour angiogenesis. Nat Rev Cancer 3:422–433. Kamphaus GD, Colorado PC, Panka DJ et al (2000) Canstatin, a novel matrix-derived inhibitor of angiogenesis and tumor growth. J Biol Chem 275:1209–1215. Kanematsu A, Yamamoto S, Ozeki M et al (2004) Collagenous matrices as release carriers of exogenous growth factors. Biomaterials 25:4513–4520. Karagiannis ED, Popel AS (2007) Identification of novel short peptides derived from the alpha 4, alpha 5, and alpha 6 fibrils of type IV collagen with anti-angiogenic properties. Biochem Biophys Res Commun 354:434–439. Karagiannis ED, Popel AS (2008) A systematic methodology for proteome-wide identification of peptides inhibiting the proliferation and migration of endothelial cells. Proc Natl Acad Sci USA 105:13775–13780. Keildlouha V, Planchenault T (1986) Potential proteolytic activity of human-plasma fibronectin. Proc Natl Acad Sci USA 83:5377–5381. Kenny HA, Kaur S, Coussens LM et al (2008) The initial steps of ovarian cancer cell metastasis are mediated by MMP-2 cleavage of vitronectin and fibronectin. J Clin Invest 118: 1367–1379. Khew ST, Tong YW (2008) Template-assembled triple-helical peptide molecules: mimicry of collagen by molecular architecture and integrin-specific cell adhesion. Biochemistry 47:585–596. Kim JK, Xu Y, Xu X et al (2005) A novel binding site in collagen type III for integrins a1b1 and a2b1. J Biol Chem 280:32512–32520.
502
J.L. Lauer and G.B. Fields
Kim Y-M, Hwang S, Kim Y-M et al (2002) Endostatin blocks vascular endothelial growth factormediated signaling via direct interaction with KDR/Flk-1. J Biol Chem 277:27872–27879. Knäuper V, Cowell S, Smith B et al (1997) The role of the C-terminal domain of human collagenase-3 (MMP-13) in the activation of procollagenase-3, substrate specificity, and tissue inhibitor of metalloproteinase interaction. J Biol Chem 272:7608–7616. Knäuper V, Osthues A, DeClerk YA et al (1993) Fragmentation of human polymorphonuclearleukocyte collagenase. Biochem J 291:847–854. Knight CG, Morton LF, Onley DJ et al (1998) Identification in collagen type I of an integrin a2b1binding site containing an essential GER sequence. J Biol Chem 273:33287–33294. Knight CG, Morton LF, Peachey AR et al (2000) The collagen-binding A-domains of integrin a1b1 and a2b1 recognize the same specific amino acid sequence, GFOGER, in native (triplehelical) collagens. J Biol Chem 275:35–40. Knutson JR, Iida J, Fields GB et al (1996) CD44/chondroitin sulfate proteoglycan and a2b1 integrin mediate human melanoma cell migration on type IV collagen and invasion of basement membranes. Mol Biol Cell 7:383–396. Konitsiotis AD, Raynal N, bihan D et al (2008) Characterization of high affinity binding motifs for the discoidin domain receptor DDR2 in collagen. J Biol Chem 283:6861–6868. Koshikawa N, Giannelli G, Cirulli V et al (2000) Role of cell surface metalloprotease MT1-MMP in epithelial cell migration over laminin-5. J Cell Biol 148:615–624. Koskimaki JE, Karagiannis ED, Rosca EV et al (2009) Peptides derived from type IV collagen, CXC chemokines, and thrombospondin-1 domain-containing proteins inhibit neovascularization and suppress tumor growth in MDA-MB-231 breast cancer xenografts. Neoplasia 11:1285-U1245. Kramer RH, Marks N (1989) Identification of intracellular collagen receptors on human melanoma cells. J Biol Chem 264:4684–4688. Kühn K, Eble J (1994) The structural bases of integrin-ligand interactions. Trends Cell Biol. 4: 256–261. Kuznetsova NV, McBride DJ, Leikin S (2003) Changes in thermal stability and microunfolding pattern of collagen helix resulting from the loss of alpha2(I) chain in osteogenesis imperfecta murine. J Mol Biol 331:191–200. Lauer JL, Gendron CM, Fields GB (1998) Effect of ligand conformation on melanoma cell a3b1 integrin-mediated signal transduction events: implications for a collagen structural modulation mechanism of tumor cell invasion. Biochemistry 37:5279–5287. Lauer-Fields JL, Malkar NB, Richet G et al (2003a) Melanoma cell CD44 interaction with the a1(IV)1263–1277 region from basement membrane collagen is modulated by ligand glycoslyation. J Biol Chem 278:14321–14330. Lauer-Fields JL, Sritharan T, Stack MS et al (2003b) Selective hydrolysis of triple-helical substrates by matrix metalloproteinase-2 and -9. J Biol Chem 278:18140–18145. Leigh CJ, Palechek PL, Knutson JR et al (1996) CD44 expression in benign and malignant nevomelanocytic lesions. Hum Pathol 27:1288–1294. Leitinger B (2003) Molecular analysis of collagen binding by the human discoidin domain receptors, DDR1 and DDR2: identification of collagen binding sites sites in DDR2. J Biol Chem 278:16761–16769. Leitinger B, Hohenester E (2007) Mammalian collagen receptors. Matrix Biol 26:146–155. Lesley J, Hyman R (1998) CD44 structure and function. Frontiers Biosci 3:616–630. Lesley J, Hyman R, English N et al (1997) CD44 in inflammation and metastasis. Glycoconjugate J 14:611–622. Levental KR, Yu H, Kass L et al (2009) Matrix crosslinking forces tumor progression by enhancing integrin signaling. Cell 139:891–906. Li C, McCarthy JB, Furcht LT et al (1997) An all-D amino acid peptide model of a1(IV)531–543 from type IV collagen binds the a3b1 integrin and mediates tumor cell adhesion, spreading, and motility. Biochemistry 36:15404–15410. Linsenmayer TF (1991) Collagen. In: Hay ED (ed) Cell biology of extracellular matrix. Plenum Press, New York, pp 7–44.
23 Collagen in Cancer
503
Maaser K, Wolf K, Klein CE et al (1999) Functional hierarchy of simultaneously expressed adhesion receptors: integrin a2b1 but not CD44 mediates MV3 melanoma cell migration and matrix reorganization within three-dimensional hyaluronan-containing collagen matrices. Mol Biol Cell 10:3067–3079. Maeshima Y, Sudhakar A, Lively JC et al (2002) Tumstatin, an endothelial cell-specific inhibitor of protein synthesis. Science 295:140–143. Malkar NB, Lauer-Fields JL, Borgia JA et al (2002) Modulation of triple-helical stability and subsequent melanoma cellular responses by single-site substitution of fluoroproline derivatives. Biochemistry 41:6054–6064. Maquart F-X, Pasco S, Ramont L et al (2004) An introduction to matrikines: extracellular matrixderived peptides which regulate cell activity. Implication in tumor invasion. Crit Rev Oncol/ Hematol 49:199–202. Marom B, Rahat MA, Lahat N et al (2007) Native and fragmented fibronectin oppositely modulate monocyte secretion of MMP-9. J Leukocyte Biol 81:1466–1476. Matsuki H, Yonezawa K, Obata K et al (2003) Monoclonal antibodies with defined recognition sequences in the stem region of CD44: detection of differential glycosylation of CD44 between tumor and stromal cells in tissue. Cancer Res 63:8278–8283. Mayo KH, Parra-Diaz D, McCarthy JB et al (1991) Cell adhesion promoting peptide GVKGDKGNPGWPGAP from the collagen type IV triple helix. Biochemistry 30:8251–8267. McCawley LJ, Matrisian LM (2001) Matrix metalloproteinases: they’re not just for matrix anymore! Curr Opin Cell Biol 13:534–540. Melchiori A, Mortarini R, Carlone S et al (1995) The a3b1 integrin is involved in melanoma cell migration and invasion. Exp Cell Res 219:233–242. Mickelson DJ, Faassen AE, McCarthy JB (1991) A cell surface chondroitin sulfate proteoglycan mediates melanoma cell motility and adhesion to a helical domain of type IV collagen. J Cell Biol 115:287a. Miles AJ, Knutson JR, Skubitz APN et al (1995) A peptide model of basement membrane collagen a1(IV)531–543 binds the a3b1 integrin. J Biol Chem 270:29047–29050. Miles AJ, Skubitz APN, Furcht LT et al (1994) Promotion of cell adhesion by single-stranded and triple-helical peptide models of basement membrane collagen a1(IV)531–543: evidence for conformationally dependent and corformationally independent type IV collagen cell adhesion sites. J Biol Chem 269:30939–30945. Misra S, Ghatak S, Toole BP (2005) Regulation of MDR1 expression and drug resistance by a positive feedback loop involving hyaluronan, phosphoinositide 3-kinase, and ErbB2. J Biol Chem 280:20310–20315. Mogford JE, Platts SH, Davies GE et al (1996) Vascular smooth muscle alpha v beta 3 integrin mediates vasodilation in response to RGD peptides and collagen fragments. FASEB J 10:3514–3514. Mori H, Tomari T, Koshifumi I et al (2002) CD44 directs membrane-type I matrix metalloproteinase to lamellipodia by associating with its hemopexin-like domain. EMBO J 21:3949–3959. Morla A, Zhang ZH, Ruoslahti E (1994) Superfibronectin is a functionally distinct form of fibronectin. Nature 367:193–196. Morrison CJ, Overall CM (2006) TIMP-independence of MMP-2 activation by MT2-MMP is determined by contributions of both the MT2-MMP catalytic and hemopexin domains. J Biol Chem 281:26528–26539. Morton LF, Peachey AR, Knight CG et al (1997) The platelet reactivity of synthetic peptides based on the collagen III fragment a1(III)CB4. J Biol Chem 272:11044–11048. Mott JD, Khalifah RG, Nagase H et al (1997) Nonenyzmatic glycation of type IV collagen and matrix metalloproteinase susceptibility. Kidney Int 52:1302–1312. Mott JD, Werb Z (2004) Regulation of matrix biology by matrix metalloproteinases. Curr Opin Cell Biol 16:558–564. Murai T, Miyazaki Y, Nishinakamura H et al (2004) Engagement of CD44 promotes Rac activation and CD44 cleavage during tumor cell migration. J Biol Chem 279:4541–4550.
504
J.L. Lauer and G.B. Fields
Murphy G, Allan JA, Willenbrock F et al (1992) The role of the C-terminal domain in collagenase and stromelysin specificity. J Biol Chem 267:9612–9618. Myllyharju J, Kivirikko KI (2001) Collagen and collagen-related diseases. Ann Med 33:7–21. Nakamura H, Suenaga N, Taniwaki K et al (2004) Constitutive and induced CD44 shedding by ADAM-like proteases and membrane-type 1 matrix metalloproteinase. Cancer Res 64: 876–882. Naor D, Nedvetzki S, Golan I et al (2002) CD44 in cancer. Crit Rev Clin Lab Sci 39:527–579. Naor D, Slonov RV, Ish-Shalom D (1997) CD44: structure, function, and association with the malignant process. Adv Cancer Res 71:241–319. Narayanan AS, Meyers DF, Page RC et al (1984) Action of mammalian collagenases on type-I trimer collagen. Collagen Rel Res 4:289–296. Nelson AR, Fingleton B, Rothenberg ML et al (2000) Matrix metalloproteinases: biological activity and clinical implications. J Clin Oncol 18:1135–1149. Ng MR, Brugge JS (2009) A stiff blow from the stroma: collagen crosslinking drives tumor progression. Cancer Cell 16:455–457. Niyibizi C, Chan R, Wu J-J et al (1994) A 92 kDa gelatinase (MMP-9) cleavage site in native type V collagen. Biochem Biophys Res Commun 202:328–333. O’Reilly MS, Boehm T, Shing Y et al (1997) Endostatin: an endogenous inhibitor of angiogenesis and tumor growth. Cell 88:277–285. Ohnishi Y, Tajima S, Ishibashi A (2001) Coordinate expression of membrane type-matrix metalloproteinases-2 and 3 (MT2-MMP and MT3-MMP) and matrix metalloproteinase-2 (MMP-2) in primary and metastatic melanoma cells. Eur J Dermatol 11:420–423. Ohuchi E, Imai K, Fujii Y et al (1997) Membrane type I matrix metalloproteinase digests interstitial collagens and other extracellular matrix macromolecules. J Biol Chem 272: 2446–2451. Okamoto I, Kawano Y, Tsuiki H et al (1999) CD44 cleavage induced by a membrane-associated metalloproteinase plays a critical role in tumor cell migration. Oncogene 18:1435–1446. Olsen BR, Ninomiya Y (1999) Collagens: overview of the family. In: Kreis T, Vale R (eds) Guidebook to the extracellular matrix, anchor, and adhesion proteins, 2nd edn. Oxford University Press, Oxford, pp 380–383. Ortiz-Urda S, Garcia J, Green CL et al (2005) Type VII collagen is required for Ras-driven human epidermal tumorigenesis. Science 307:1773–1776. Overall CM (2002) Molecular determinants of metalloproteinase substrate specificity. Mol Biotech 22:51–86. Overall CM, Lopez-Otin C (2002) Strategies for MMP inhibition in cancer: innovations for the post-trial era. Nat Rev Cancer 2:657–672. Page-McCaw A, Ewald AJ, Werb Z (2007) Matrix metalloproteinases and the regulation of tissue remodelling. Nat Rev Mol Cell Biol 8:221–233. Park HS, Kim KR, Lee HJ et al (2007) Overexpression of discoidin domain receptor 1 increases the migration and invasion of hepatocellular carcinoma cells in association with matrix metalloproteinase. Oncol Rep 18:1435–1441. Perez JL, Jing SQ, Wong TW (1996) Identification of two isoforms of the Cak receptor kinase that are coexpressed in breast tumor cells lines. Oncogene 12:1469–1477. Perret S, Eble JA, Siljander PR-M et al (2003) Prolyl hydroxylation of collagen type I is required for efficient binding to integrin a1b1and platelet glycoprotein VI but not to a2b1. J Biol Chem 278:29873–29879. Petitclerc E, Boutaud A, Prestayko A et al (2000) New functions for non-collagenous domains of human collagen type IV. J Biol Chem 275:8051–8061. Pirila E, Sharabi A, Salo T et al (2003) Matrix metalloproteinases process the laminin-5 gamma 2-chain and regulate epithelial cell migration. Biochem Biophys Res Commun 303: 1012–1017. Platt VM, Szoka FC, Jr. (2008) Anticancer therapeutics: targeting macromolecules and nanocarriers to hyaluronan or CD44, a hyaluronan receptor. Mol Pharm 5:474–486.
23 Collagen in Cancer
505
Pluda JM (1997) Tumor-associated angiogenesis: mechanisms, clinical implications, and therapeutic strategies. Semin Oncol 24:203–218. Ponnusamy MP, Batra SK (2008) Ovarian cancer: emerging concept on cancer stem cells. J Ovarian Res 1:4. Prince ME, Sivanandan R, Kaczorowski A et al (2007) Identification of a subpopulation of cells with cancer stem cell properties in head and neck squamous cell carcinoma. Proc Natl Acad Sci USA 104:973–978. Prockop DJ, Kivirikko KI (1995) Collagens: molecular biology, diseases, and potentials for therapy. Annu Rev Biochem 64:403–434. Ram R, Lorente G, Nikolich K et al (2006) Discoidin domain receptor-1a (DDR1a) promotes glioma cell invasion and adhesion in association with matrix metalloproteinase-2. J Neurooncol 76:239–248. Ramchandran R, Dhanabal M, Volk R et al (1999) Antiangiogenic activity of restin, NC10 domain of human collagen XV: comparison to endostatin. Biochem Biophys Res Commun 255:735–739. Ramont L, Brassart-Pasco S, Thevenard J et al (2007) The NC1 domain of type XIX collagen inhibits in vivo melanoma growth. Mol Cancer Ther 6:506–514. Ranuncolo SM, Ladeda V, Gorostidy S et al (2002) Expression of CD44s and CD44 splice variants in human melanoma. Oncol Rep 9:51–56. Raynal N, Hamaia SW, Siljander PR-M et al (2006) Use of synthetic peptides to locate novel integrin a2b1-binding motifs in human collagen III. J Biol Chem 281:3821–3831. Reed CC, Iozzo RV (2003) The role of decorin in collagen fibrillogenesis and skin homeostasis. Glycoconjugate J 19:249–255. Renner C, Saccá B, Moroder L (2004) Synthetic heterotrimeric collagen peptides as mimics of cell adhesion sites of the basement membrane. Biopolymers (Peptide Sci) 76:34–47. Riikonen T, Westermarck J, Koivisto L et al (1995) Integrin a2b1 is a positive regulator of collagenase (MMP-1) and collagen a1(I) gene expression. J Biol Chem 270:13548–13552. Roy R, Yang J, Moses MA (2009) Matrix metalloproteinases as novel biomarkers and potential therapeutic targets in human cancer. J Clin Oncol 27:5287–5297. Ruoslahti E (1988) Fibronectin and its receptors. Annu Rev Biochem 57:375–413. Rupard JH, Dimari SJ, Damjanov I et al (1988) Synthesis of type I homotrimer collagen molecules by cultured human lung adenocarcinoma cells. Am J Pathol 133:316–326. Sabeh F, Ota I, Holmbeck K et al (2004) Tumor cell traffic through the extracellular matrix is controlled by the membrane-anchored collagenase MT1-MMP. J Cell Biol 167:769–781. Sabeh F, Shimizu-Hirota R, Weiss SJ (2009) Protease-dependent versus -independent cancer cell invasion programs: three-dimensional amoeboid movement revisited. J Cell Biol 185:11–19. Saccá B, Sinner E-K, Kaiser J et al (2002) Binding and docking of synthetic heterotrimeric collagen type IV peptides with a1b1 integrin. ChemBioChem 9:904–907. Schepel J, Tschesche, H. (2000) The proteolytic activity of the recombinant cryptic human fibronectin Type IV collagenase from E. Coli expression. J Protein Chem 19:685–692. Schor SL, Ellis I, Dolman C et al (1996) Substratum-dependent stimulation of fibroblast migration by the gelatin-binding domain of fibronectin. J Cell Sci 109:2581–2590. Screaton GR, Bell MV, Jackson DG et al (1992) Genomic structure of DNA encoding the lymphocyte homing receptor CD44 reveals at least 12 alternatively spliced exons. Proc Natl Acad Sci USA 89:12160–12164. Seiki M (2003) Membrane-type 1 matrix metalloproteinase: a key enzyme for tumor invasion. Cancer Lett 194:1–11. Shoulders MD, Raines RT (2009) Collagen structure and stability. Annu Rev Biochem 78: 929–958. Shrivastava A, Radziejewski C, Campbell E et al (1997) An orphan receptor tyrosine kinase family whose members serve as nonintegrin collagen receptors. Mol Cell 1:25–34. Slack BE, Siniaia MS, Blusztajn JK (2006) Collagen type I selectively activates ectodomain shedding of the discoidin domain receptor 1: involvement of Src tyrosine kinase. J Cell Biochem 98:672–684.
506
J.L. Lauer and G.B. Fields
Sottile J, Hocking DC (2002) Fibronectin polymerization regulates the composition and stability of extracellular matrix fibrils and cell-matrix adhesions. Mol Biol Cell 13:3546–3559. Staatz WD, Rajpara SM, Wayner EA et al (1989) The membrane glycoprotein Ia–IIa (VLA-2) complex mediates the Mg++-dependent adhesion of platelets to collagen. J Cell Biol 108: 1917–1924. Stamenkovic I, Yu Q (2009) Shedding light on proteolytic cleavage of CD44: the responsible sheddase and functional significance of shedding. J Invest Dermatol 129:1321–1324. Steffensen B, Wallon UM, Overall CM (1995) Extracellular matrix binding properties of recombinant fibronectin type II-like modules of human 72-kDa gelatinase/type IV collagenase. High affinity binding to native type I collagen but not native type IV collagen. J Biol Chem 270: 11555–11566. Stricker TP, Dumin JA, Dickeson SK et al (2001) Structural analysis of the a2 integrin I domain/procollagenase-1 (matrix metalloproteinase-1) interaction. J Biol Chem 276: 29375–29381. Sweeney SM, DiLullo G, Slater SJ et al (2003) Angiogenesis in collagen I requires a2b1 ligation of a GFP*GER sequence, and possibly p38 MAPK and focal adhesion disassembly. J Biol Chem 278:30516–30524. Sweeney SM, Orgel JP, Fertala A et al (2008) Candidate cell and matrix interaction domains on the collagen fibril, the predominant protein of vertebrates. J Biol Chem 283:21187–21197. Szabova L, Chrysovergis K, Yamada SS et al (2007) MT1-MMP is required for efficient tumor dissemination in experimental metastatic disease. Oncogene 27:3274–3281. Thorne RF, Legg JW, Isacke CM (2004) The role of the CD44 transmembrane and cytoplasmic domains in co-ordinating adhesive and signalling events. J Cell Sci 117:373–380. Timpl R (1989) Structure and biological activity of basement membrane proteins. Eur J Biochem 180:487–502. Tran KT, Lamb P, Deng JS (2005) Matrikines and matricryptins: implications for cutaneous cancers and skin repair. J Dermatol Sci 40:11–20. van der Flier A, Sonnenberg A (2001) Function and interactions of integrins. Cell Tissue Res 305:285–298. Vidmar SL, Lottspeich F, Emod I et al (1991) Collagen-binding domain of human plasma fibronectin contains a latent type-IV collagenase. Eur J Biochem 201:79–84. Vo HP, Lee MK, Crowe DL (1998) a2b1 integrin signaling via the mitogen activated protein kinase pathway modulates retinoic acid-dependent tumor cell invasion and transcriptional downregulation of matrix metalloproteinase 9 activity. Int J Oncol 13:1127–1134. Vogel W, Gish GD, Alves F et al (1997) The discoidin domain receptor tyrosine kinases are activated by collagen. Mol Cell 1:13–23. Vogel WF (2001) Collagen-receptor signaling in health and disease. Eur J Dermatol 11: 506–514. Wald M, Olejár T, Sebková V et al (2001) Mixture of trypsin, chymotrypsin and papain reduces formation of metastases and extends survival time of C57Bl6 mice with syngeneic melanoma B16. Cancer Chemother Pharmacol 47:S16–S22. Wall SJ, Werner E, Werb Z et al (2005) Discoidin domain receptor 2 mediates tumor cell cycle arrest induced by fibrillar collagen. J Biol Chem 280:40187–40194. Weber GF, Bronson RT, Hagan J et al (2002) Absence of the CD44 gene prevents sarcoma metastasis. Cancer Res 62:2281–2286. Wen W, Moses MA, Wiederschain D et al (1999) The generation of endostatin is mediated by elastase. Cancer Res 59:6052–6056. Werb Z, Tremble PM, Behrendtsen O et al (1989) Signal transduction through the fibronectin receptor induces collagenase and stromelysin gene-expression. J Cell Biol 109:877–889. Woessner JF, Nagase H (2000) Matrix metalloproteinases and TIMPs. Oxford University Press, Oxford. Xu J, Rodriguez D, Petitclerc E et al (2001) Proteolytic exposure of a cryptic site within collagen type IV is required for angiogenesis and tumor growth in vivo. J Cell Biol 154:1069–1080.
23 Collagen in Cancer
507
Xu X, Chen Z, Wang Y et al (2005) Functional basis for the overlap in ligand interactions and substrate specificities of matrix metalloproteinases-9 and -2. Biochem J 392:127–134. Xu Y, Gurusiddappa S, Rich RL et al (2000) Multiple binding sites in collagen type I for the integrins a1b1 and a2b1. J Biol Chem 275:38981–38989. Yoshinaga IG, Vink J, Dekker SK et al (1993) Role of a3b1 and a2b1 integrins in melanoma cell migration. Melanoma Res 3:435–441. Yu Q, Stamenkovic I (1999) Localization of matrix metalloproteinase 9 to the cell surface provides a mechanism for CD44-mediated tumor invasion. Genes Dev 13:35–48. Yu Q, Stamenkovic I (2000) Cell surface-localized matrix metalloproteinase-9 proteolytically activates TGF-b and promotes tumor invasion and angiogenesis. Genes Dev 14:163–176. Yu WH, Woessner JF, Jr., McNeish JD et al (2002) CD44 anchors the assembly of matrilysin/ MMP-7 with heparin-binding epidermal growth factor precursor and ErbB4 and regulates female reproductive organ remodeling. Genes Dev 16:307–323. Yu Y-C, Pakalns T, Dori Y et al (1997) Construction of biologically active protein molecular architecture using self-assembling peptide-amphiphiles. Methods Enzymol 289:571–587. Zhang W, Matrisian LM, Holmbeck K et al (2006) Fibroblast-derived MT1-MMP promotes tumor progression in vitro and in vivo. BMC Cancer 6:52. Zhang W-M, Käpylä J, Puranen JS et al (2003) a11b1 integrin recognizes the GFOGER sequence in interstitial collagens. J Biol Chem 278:7270–7277. Ziober AF, Falls EM, Ziober BL (2006) The extracellular matrix in oral squamous cell carcinoma: friend or foe? Head Neck 28:740–749.
Chapter 24
Integrins and Cancer Laurie G. Hudson and M. Sharon Stack
Abstract Integrins are heterodimeric transmembrane glycoproteins that bind extracellular matrix (ECM) proteins and thereby functionally couple the cytoskeleton to the extracellular environment. The relationship between integrins and cancer has been widely studied as malignant cells alter not only integrin expression and organization, but properties of the surrounding ECM as well. Although integrin cytoplasmic tails lack direct kinase activity, interaction with multiple adaptor proteins enables the assembly of signaling complexes that result in the activation of focal adhesion kinase (FAK) or integrin-linked kinase (ILK). These phosphorylation events trigger downstream signaling pathways that activate a variety of cellular responses key to tumor progression including survival, proliferation, motility, and invasion. In addition, gene expression changes induced by integrin signaling have been implicated in epithelial-to-mesenchymal transition (EMT), most notably via downregulation of E-cadherin expression and function. Emerging evidence that integrin-mediated ECM adhesion is inherently a mechanosensory process also supports a role for integrin-mediated mechanotransduction in tumor progression and metastasis. Together these studies suggest that integrin-directed therapeutics combined with strategies designed to target the mechanical properties of the ECM and restore normal tissue architecture may provide new options as single agent or combined anticancer therapies.
M.S. Stack (*) Department of Pathology & Anatomical Sciences, Medical Physiology and Pharmacology, University of Missouri School of Medicine, 1 Hospital Drive, M214C Medical Sciences Bldg., Columbia, MO 65212, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_24, © Springer Science+Business Media, LLC 2010
509
510
L.G. Hudson and M.S. Stack
Integrin Structure and Function Background The ECM supplies a rich source of structural and biomechanical cues that influence cell behavior and ultimately dictate tissue architecture. In both normal and malignant tissues, cell interaction with the matrix substratum is mediated predominantly via the integrin family of adhesion receptors. Functionally, integrins form a link between the ECM and the actin-based cytoskeleton at focal adhesions, points of cell–matrix contact. These heterodimeric type I transmembrane glycoproteins are comprised of an a and a b subunit that associate noncovalently in a variety of combinations to promote interaction with multiple matrix ligands (Fig. 24.1a). To date, 24 distinct ab heterodimers have been identified, half of which contain b1 integrin (Humphries et al. 2006; Berrier and Yamada 2007; Luo and Springer 2006). Although matrix protein specificity is inherent in some integrin heterodimers, many integrins bind more than one matrix ligand, and many matrix proteins are ligated by more than one integrin (Fig. 24.1b). Integrins engage in bidirectional information transfer between the cytoskeleton and the extracellular environment. As integrins lack constitutive binding activity, the capacity for bidirectional signaling aids in conformational conversion of integrins from a low- to a high-affinity binding state. This transition can be induced via both “inside out” and “outside in” signaling. Integrin activation leads to conformational rearrangement of the heterodimer, resulting in exposure of ligand binding sites that are comprised of residues from both the a and b subunits (Al-Jemal and Harrison 2008; Luo and Springer 2006). Furthermore, integrin engagement by multivalent ECM proteins promotes lateral integrin clustering in the plasma membrane (Berrier and Yamada 2007). A distinct hierarchy of integrin engagement has been described wherein integrin occupancy by monomeric ligand leads to integrin redistribution, while occupancy in the presence of integrin aggregation results in the assembly of cytoskeletal and signaling complexes at the cytoplasmic face of the integrin (Miyamoto et al. 1995; Yamada and Cukierman 2007; Humphries et al. 2006). This enables structural and functional coupling of the extracellular environment to the cytoskeleton as well as to signal transduction pathways that modulate cellular responsiveness.
Focal Adhesion Kinase The majority of integrins contain short (~50 amino acids) cytoplasmic tails that lack direct kinase activity. However, focal adhesions contain a large variety of integrin binding proteins, adaptor proteins, and enzymes (kinases/phosphatases) that function to transduce intracellular signals (Berrier and Yamada 2007; Dubash et al. 2009;
24 Integrins and Cancer
511
a
b
α
c
β beta I
β propeller
hybrid thigh
PSI
genu calf-1
EGF 1-4
calf-2
β tail transmembrane
closed headpiece, bent
open headpiece, extended
closed headpiece, extended
d
α2 α1
α3
α5
β3
β5
αv
β1
β6
α6
β4
collagen fibrinogen laminin vitronectin fibronectin
Fig. 24.1 Integrin domain structure and conformational alteration. The integrin a and b subunits are noncovalently associated and are comprised of an extracellular “head” attached to a flexible “leg” anchored by a transmembrane domain. (a) The bent conformation represents the physiological low affinity state. Integrin activation results in conformational alteration to (b) an extended low affinity “closed” structure and (c) an extended high affinity “open” structure. The remodeling of the ligand binding site that occurs as a result of altering the orientation between the beta I and hybrid domains results in a structure with high ligand affinity (Luo and Springer 2006; Al-Jemal and Harrison 2008). (d) Integrin subunit pairing and matrix ligands. The diagram depicts integrin subunits, their binding partners, and matrix ligands. Note that there are 18a and 8b subunits identified to date, not all of which are depicted here
512
L.G. Hudson and M.S. Stack
Legate and Fässler 2009). Interaction of integrin cytoplasmic tails with these adaptor proteins activates kinases, leading to phosphorylation of cytoskeletal and signaling proteins (Hynes 2002). One of the most ubiquitous nonreceptor tyrosine kinases in focal adhesions is termed FAK (Zhao and Guan 2009; Kornberg et al. 1991; Guan et al. 1991). FAK is activated by integrin-mediated cell adhesion and initially undergoes autophosphorylation at Tyr397, generating a binding site for Src family kinases via the SH2 domain. FAK-Src binding initiates Src activation, which in turn leads to Src-dependent phosphorylation of additional sites on FAK as well as other substrates in focal adhesions such as Shc, paxillin, and p130Cas. These phosphorylation events trigger downstream signaling pathways that activate a variety of cellular responses to alter cell spreading, motility, cell cycle progression, survival, and gene expression (Mitra and Schlaepfer 2006; Zhao and Guan 2009) (Fig. 24.2a). Genetic depletion of FAK results in an embryonic lethal phenotype due to defects in the developing vasculature (Illic et al. 1995). Conditional depletion of FAK in endothelial cells reinforced the key role of FAK in vascular development, as FAKdeficient endothelial cells show defects in tubulogenesis, survival, proliferation, and migration (Shen et al. 2005). Early studies linking FAK activity to tumorigenesis were based on the observation that both integrin signaling and v-Src-induced oncogenic transformation converged at FAK (Schaller et al. 1992; Guan and Shalloway 1992). Subsequent studies have shown that elevated FAK expression and/or activity is frequently associated with malignancy. FAK activation is correlated with the capacity for anchorage-independent growth, a hallmark function of malignant cells (Frisch et al. 1996), and activated FAK protects cells from detachment-induced cell death or anoikis a
b
α β
α β
TAE226 PF-562, 271
ILKAP QLT0267 KP32
FAK Src
EMT
ILK
GSK3β
MLC Rac/Cdc42
Akt
snail
motility
survival
EMT
PI3K /Akt
angiogenesis Rac/Rho motility invasion
survival proliferation
Fig. 24.2 Summary of cellular events downstream of integrin signaling through (a) FAK and (b) ILK. Integrin signaling regulates proliferation, survival, motility, invasion, angiogenesis, and EMT. Additionally, ILK functions as a negative regulator of GSK3b, thereby modulating Wnt signaling as well. Inhibitors of FAK and ILK kinase activity are shown in boxes (McDonald et al. 2008; Mitra and Schlaepfer 2006; Berrier and Yamada 2007)
24 Integrins and Cancer
513
(Zhao and Guan 2009; Frisch et al. 1996). Activated FAK also interacts with PI3K leading to Akt activation and subsequent inhibition of apoptosis. In addition, FAK signaling may also provide an antiapoptotic signal through a novel interaction with p53 that is independent of FAK kinase activity, resulting in suppression of p53mediated transcriptional activation of proapoptotic genes (Lim et al. 2008a; Mitra and Schlaepfer 2006). This regulation is dependent on nuclear localization of FAK (Lim et al. 2008a, b), although the factors that control FAK nuclear translocation remain under investigation. Aside from inhibiting apoptosis, FAK may also promote tumorigenesis through stimulation of cell proliferation, as FAK activity can promote cell cycle progression via Erk and JNK-dependent pathways (Zhao and Guan 2009). Both FAK expression and FAK activation are enhanced in many latestage cancers, suggesting a potential role as a therapeutic target (Lim et al. 2008a; Cance et al. 2000; Lark et al. 2005; McLean et al. 2005).
Integrin-Linked Kinase An alternative approach to detect proteins that interact with the b1 integrin cytoplasmic tail used yeast two hybrid screening to identify the enzyme designated ILK, a Ser/Thr protein kinase prevalent in focal adhesions and fibrillar adhesions (McDonald et al. 2008). The ILK activity is stimulated by integrins as well as growth factors, and its activity is antagonized by ILK-associated protein (ILKAP), a protein phosphatase 2C (PP2C) family member. The domain structure of ILK is suited to promote interaction with a variety of proteins involved in cell signaling and cytoskeletal dynamics, and as a result, ILK has been shown to regulate diverse processes including cell survival, proliferation, EMT, and invasion (Fielding and Dedhar 2009). Genetic ablation of ILK results in embryonic lethality due to defects in adhesive and migratory mechanics (Sakai et al. 2003), necessitating the development of tissue-specific deletion models. Mice with ILK-deficient keratinocytes display defective integrin-mediated adhesion and impaired basement membrane integrity, while cultured ILK-/- keratinocytes show defects in polarity, attachment, spreading, and motility (Lorenz et al. 2007; Nakrieko et al. 2008). In normal epithelial cells, ILK basal kinase activity is low, but is stimulated by cell–matrix interaction, leading to phosphorylation of key substrates such as b1 integrin cytoplasmic domain, protein kinase B (PKB/Akt), glycogen synthase kinase 3b (GSK3b), and myosin light chain (Fig. 24.2b). As a result, ILK overexpression in normal epithelial cells leads to a loss of cell–cell adhesion due to downregulation of E-cadherin, increased nuclear translocation of b-catenin, suppression of apoptosis and anoikis, and altered motility (McDonald et al. 2008; Fielding and Dedhar 2009; Sakai et al. 2003; Grashoff et al. 2003; Terpstra et al. 2003; Holmbeck et al. 1999; Lorenz et al. 2007; Nakrieko et al. 2008; Wu and Dedhar 2001). Recent studies have highlighted a potential cross-talk between ILK and Wnt signaling pathways. Constitutively active and/or overexpressed ILK inhibits GSK3b, leading to nuclear translocation of b-catenin and enhanced b-catenin/Tcf transcriptional
514
L.G. Hudson and M.S. Stack
activity (Delcommenne et al. 1998; Novak et al. 1998; Persad et al. 2001). Furthermore, inhibition of ILK reduces Wnt-induced activation of b-catenin/Tcfmediated transcription (Oloumi et al. 2006). These data highlight a role for ILK as a modulator of canonical Wnt signaling.
Integrins, Motility, and Invasion A PubMed search for “integrins and cancer” identifies almost 10,000 citations, highlighting the complexity of the topic. Indeed, numerous changes in expression of integrins are associated with advanced cancers including persistent focalized expression, local loss of expression, or disorganized expression. Further, malignant cells alter both ECM deposition and remodeling as the provisional tumor stroma is generated and replaced with fibrotic tissue, resulting in aberrant tissue organization and mechanical properties. Successful metastatic dissemination often necessitates changes in integrin subunit expression as cells adapt to a new tissue microenvironment. In addition, focal adhesions in tumor cells may be enriched with matrix degrading proteinases that alter ECM integrity (Lim et al. 2008a, b). It has been speculated that modulation of integrin profiles represents an adaptive response to the presence of new integrin ligands in the form of degraded stromal matrix components (Hanahan and Weinberg 2000). This deviation from normal cell–matrix contacts alters adhesion, signal transduction, and cytoskeletal rearrangements that promote a migratory phenotype that, in turn, facilitates local and regional spread of metastatic cells. Integrins are intimately involved in the processes of tumor cell motility and invasion, an “acquired capability” that represents one of the hallmarks of cancer (Hanahan and Weinberg 2000). Cell motility is regulated by both the density of matrix ligands and matrix rigidity (Yamada and Cukierman 2007). Motility encompasses a complex set of processes that include protrusion of the leading edge of the cell to establish new integrin–matrix contacts, localized proteolysis, contraction of actomyosin fibers, and rear release of adhesive contacts to enable retraction of the trailing edge (Lauffenburger and Horwitz 1996). The traction forces that propel migratory cells require signal transduction from matrix-bound integrins to the actin cytoskeleton. Cells use filopodia and lamellipodia, sheet- or rod-like actin-rich protrusions, for migration on planar two-dimensional (2D) substrata (Yilmaz and Christofori 2009). However, tissue invasion and metastasis commonly require migration through three-dimensional (3D) matrices. Conversely, when embedded in 3D ECM, many cancer cells will polarize and acquire a migratory phenotype. In contrast to the flattened and spread morphology of cells participating in motility on 2D surfaces, cells migrating in 3D are more elongated and contain a cortical actin cytoskeleton (Friedl and Wolf 2009). Actin-rich protrusions formed by tumor cells migrating in 3D matrices may also express proteolytic activity and can mediate matrix metalloproteinase (MMP)-driven cleavage of ECM proteins. In addition to engaging matrix ligands, integrins can actively participate in recruitment and
24 Integrins and Cancer
515
localization of proteolytic activity to these actin-rich protrusions (Wolf and Friedl 2009; Yamada and Cukierman 2007). Both integrins and proteases interact with the ECM substratum at the leading edge, laterally in compression zones around the mid-region of the cell, and at the trailing edge where they participate in rear retraction and formation of proteolytic microtracks (Friedl and Wolf 2009; Wolf et al. 2007). Intact 3D ECM presents cells with a multivalent adhesive substratum and integrinmediated contact between cells and multivalent matrices leads to integrin clustering, or lateral aggregation, in the plane of the cell membrane (Miyamoto et al. 1995; Berrier and Yamada 2007). Integrin binding events that result in receptor occupancy vs. receptor aggregation induce distinct hierarchies of cellular signaling events that functionally couple the extracellular environment to the cytoskeleton and to specific signal transduction pathways that modulate cellular responses. One functional output of integrin aggregation is altered proteinase expression, suggesting that metastasizing tumor cells receive distinct signals as they encounter different matrix substrata that can dictate subsequent matrix-degrading behavior. This is supported by early studies showing that culturing cells in the presence of matrix proteins, matrix-derived peptides, or anti-integrin antibodies induce proteinase expression (reviewed in Munshi and Stack 2006). Perhaps the best studied example of integrinregulated proteinase expression is membrane type 1 matrix metalloproteinase (MT1-MMP). Early studies with fibrosarcoma and breast carcinoma cells showed that culture on 3D collagen matrices, but not denatured collagen, induced activation of proMMP-2, a substrate of MT1-MMP (Azzam and Thompson 1992; Gilles et al. 1997). Subsequent analyses demonstrated integrin-induced expression of MT1MMP in breast and ovarian cancer cells (Ellerbroek et al. 1999, 2001; Barbolina et al. 2006). Cell surface MT1-MMP has been shown to localize to sites of integrin clustering, providing evidence for integrin-mediated redistribution of active enzyme to sites of cell–matrix contact (Ellerbroek et al. 2001; Gálvez et al. 2002; Wolf et al. 2003). Furthermore, MT1-MMP catalyzes a3 integrin ectodomain shedding in ovarian carcinoma cells, providing a mechanism for reciprocal regulation of integrin function (Moss et al. 2009). Integrin engagement can regulate expression of a variety of genes that alter metastatic success, as demonstrated in recent studies using an epithelial ovarian cancer (EOC) metastasis model (Barbolina et al. 2010). An early event in EOC metastasis is exfoliation of cells from the primary tumor into the peritoneal cavity as single cells or multicellular aggregates, followed by diffuse seeding of peritoneal surfaces. Shed tumor cells interact with mesothelial cells lining the inner surface of the peritoneal cavity and induce mesothelial cell retraction, whereupon integrinmediated adhesion contributes to metastatic anchoring in the interstitial collagenrich submesothelial matrix (Burleson et al. 2004a, b; Kenny et al. 2007; Moser et al. 1996; Fishman et al. 1998; Cannistra et al. 1993; Ahmed et al. 2005). Ovarian cancer cells adhere preferentially to interstitial collagens, inducing multivalent cell–matrix contact (Moser et al. 1996; Fishman et al. 1998; Ellerbroek et al. 1999). The subsequent alterations in integrin signaling may then contribute to metastatic success (Burleson et al. 2004b, 2006; Moser et al. 2006; Fishman et al. 1998; Ellerbroek et al. 1999, 2001; Barbolina et al. 2007). Studies using 3D tissue culture models, in
516
L.G. Hudson and M.S. Stack
a
b
CTGF TGFβ Dkk Claudin nucleus
MMP9,14 Actn4 WT1 Cat L CXCR4 Rho
Fig. 24.3 (a) Multicellular aggregates of ovarian carcinoma cells proliferate within 3-dimensional (3D) collagen gels and extend invasive projections into the multivalent matrix (yellow box) (Moss et al. 2009) (photo courtesy of Dr. NM Moss). (b) Comparative cDNA microarray analysis of changes in gene expression downstream of integrin signaling. Analysis of cells in 3D collagen gels relative to those on planar collagen surfaces reveals altered gene expression (both upregulation and downregulation) effecting multiple cellular processes including invasion, motility, proliferation, and survival (Barbolina et al. 2007, 2008, 2009)
which cells are placed in contact with a 3D collagen gel followed by cDNA microarray analysis, identified alterations in a number of genes that may play an important role in regulation of EOC metastatic success (Fig. 24.3). For example, MT1-MMP expression is upregulated via a pathway that involves integrin-Src signaling to induce expression of the early growth response gene Egr-1, followed by Egr-1-mediated transcriptional activation of the MT1-MMP promoter (Barbolina et al. 2007; Barbolina and Stack 2008). The resulting increase in MT1-MMPcatalyzed collagenolytic activity may facilitate metastatic anchoring and remove barriers to proliferation in the collagen-rich submesothelial matrix. Integrincollagen contact also induces MMP-9 in EOC cells. In this model, MMP-9 catalyzes E-cadherin ectodomain shedding, providing a mechanism for dissolution of multicellular aggregates to facilitate intraperitoneal anchoring (Symowicz et al. 2007;
24 Integrins and Cancer
517
Cowden Dahl et al. 2008). In addition to proteinases, integrin engagement alters expression of EOC genes that regulate adhesion, motility, and proliferation. For example, actinin alpha-4 (ACTN4) expression is upregulated by 3D collagen contact and protein expression is enriched at the migratory front of cell clusters (Barbolina et al. 2008). Silencing of ACTN4 using siRNA reduces both motility and invasion. Downregulation of gene expression is also observed as a consequence of 3D collagen culture. For example, expression of connective tissue growth factor (CTGF), an adhesive ligand for EOC cells, is downregulated by 3D collagen culture (Barbolina et al. 2009). As a consequence, loss of CTGF expression promotes a more invasive phenotype. These data support the hypothesis that integrin regulation of gene expression may play an active role in promoting metastatic dissemination.
Integrins and Epithelial–Mesenchymal Transition The EMT occurs during embryonic development, wound repair, fibrosis, and tumor metastasis. This process is functionally characterized by decreased cell–cell adhesion, dramatic reorganization of the actin cytoskeleton, elevated production of ECM components, increased proteolysis, enhanced cell migration and invasive capacity, and changes in gene expression associated with the acquisition of a mesenchymal phenotype (reviewed in Hay 2005; Moreno-Bueno et al. 2008; Kalluri and Weinberg 2009; Thiery et al. 2009; Zeisberg and Neilson 2009). There are three classifications of EMT associated with distinct biological events (Kalluri and Weinberg 2009). Type 1 EMT occurs during embryonic development whereas type 2 EMT is associated with tissue regeneration, wound repair, and fibrosis. In contrast to the defined events that characterize type 1 and 2 EMT, the tumor-associated type 3 EMT is viewed as a continuum of phenotypic plasticity and gain of mesenchymal characteristics. In recent years, it has become clear that partial or incomplete EMT can occur in tumors whereby cells retain some epithelial features and migrate or invade while maintaining cell–cell contacts (Brabletz et al. 2005; Christiansen and Rajasekaran 2006; Kalluri and Weinberg 2009; Thiery et al. 2009; Zeisberg and Neilson 2009). Tumor phenotype likely reflects the particular complement of EMT regulatory factors expressed in cells or within the tumor microenvironment (Guarino et al. 2007; Tse and Kalluri 2007; Kalluri and Weinberg 2009). The functional consequences of this phenotypic plasticity in cancer are not fully understood, but may modulate anchorage independent cell survival, chemoresistance, and the multiple steps required to establish metastatic lesions. EMT may be triggered by numerous mechanisms including integrin engagement and integrin-mediated signal transduction. Differentiated epithelial cells are in contact with the basement membrane that is predominantly composed of collagen type IV, laminin, and proteoglycans. When invasive tumor cells breach the basement membrane, they make contact with stromal ECM components such as collagen types I and III, fibronectin, glycoproteins, and matrix-associated growth factors (Sherwood 2006; Larsen et al. 2006;
518
L.G. Hudson and M.S. Stack
Rowe and Weiss 2008; Yilmaz and Christofori 2009). Exposure to new ECM proteins may modify tumor cell behavior through integrin engagement and stimulation of integrin-mediated signaling pathways. Integrins assemble signaling complexes that contain signaling molecules such as ILK and FAK (Harburger and Calderwood 2009; Legate et al. 2009; Yilmaz and Christofori 2009). ILK is a multidomain adaptor protein that binds directly to b1 and b3 integrins, and through scaffolding functions links to the actin cytoskeleton (Mitra and Schlaepfer 2006; Harburger and Calderwood 2009; Legate et al. 2009) (Fig. 24.2). ILK also forms signaling complexes with additional protein kinases and small GTPases. Signaling molecules activated (directly and indirectly) by integrins include, but are not limited to, ILK, FAK, Src family kinases, glycogen synthase-kinase-3 b (GSK3 b), protein kinase B (AKT/PKB), phosphoinositol-3 kinase (PI3K), the small GTPases RhoA and Rac1, Ras, and the mitogen-activated protein kinases (MAPK) (reviewed in Kalluri and Weinberg 2009; Harburger and Calderwood 2009; Legate et al. 2009; Yilmaz and Christofori 2009). Certain downstream effectors of these signaling cascades include EMT regulatory transcription factors such as Snail1 (Snail), Snail2 (Slug), ZEB1 (dEF1), ZEB2 (Sip1), E47, and Twist (Kalluri and Weinberg 2009; Yilmaz and Christofori 2009). In addition, changes in cell–matrix adhesion have been associated with disruption of adherens junctions and modulation of b-catenin-dependent signaling. Based on the vast signaling network stimulated by integrins, changes in cellular integrin expression or the ECM environment can dramatically alter cell behavior. There is evidence that collagen engagement leads to cross-talk between two adhesive processes (cell–matrix and cell–cell) linked to EMT. Loss of E-cadherin gene expression or E-cadherin protein is a well recognized indicator of EMT and a frequent occurrence in epithelial cancers (Kalluri and Weinberg 2009; Thiery et al. 2009; Zeisberg and Neilson 2009). E-cadherin containing adherens junctions can be destabilized upon integrin engagement through Src-mediated phosphorylation of E-cadherin, followed by ubiquitylation by the E3 ligase Hakai leading to E-cadherin internalization and degradation. Furthermore, integrin-activated FAK can phosphorylate b-catenin leading to its ubiquitylation and degradation which represents another mechanism to disrupt adherens junctions. Pancreatic and ovarian tumor cells provide examples whereby engagement of integrins a1b1 or a2b1 by collagen type I results in a loss of E-cadherin-mediated cell–cell contacts and accompanied by release of b-catenin (Koenig et al. 2006; Imamichi et al. 2007; Symowicz et al. 2007; Giehl and Menke 2008). b-catenin participates in the Wnt signaling pathway when it is stabilized and accumulates in the nucleus. Nuclear b-catenin interacts with members of the Tcf/Lef family of transcription factors and modulates expression of a number of genes associated with EMT (Kalluri and Weinberg 2009). In addition to downregulation of E-cadherin function, EMT is characterized by a gain of mesenchymal N-cadherin expression that is associated with changes in cell adhesive properties and migration (reviewed in Wheelock et al. 2008). The induction of N-cadherin expression appears to be regulated by collagen I and integrin-dependent mechanisms. It has been reported that coordinated engagement of the collagen receptor discoidin domain receptor 1 (DDR1) and a2b1-integrin leads to N-cadherin induction (Shintani et al.
24 Integrins and Cancer
519
2006, 2008) and the EMT-associated transcription factor Twist appears to be involved in N-cadherin gene expression via an b1-integrin-dependent mechanism (Alexander et al. 2006; Yang et al. 2007; Niu et al. 2007). These examples illustrate how collagen engagement can dramatically alter cell–cell adhesive properties through regulation of E-cadherin and N-cadherin expression and function. Another hallmark of EMT is increased production of a spectrum of extracellular proteinases with activities directed against ECM proteins, adhesion molecules, and other cell surface and surface-associated proteins (reviewed in Munshi and Stack 2006). MMPs represent the major family of proteinases that are regulated by integrins. Integrin-stimulated signaling cascades correspond to those identified as regulators of MMP gene expression (Munshi and Stack 2006; Yan and Boyd 2007). The consequences of MMP expression in EMT are broad and include cleavage and activation of growth factors and growth factor receptors, and modulation of cell adhesion (Orlichenko and Radisky 2008). Certain MMPs (MMP-3, MMP-7, and MMP-9) cleave the extracellular domain of E-cadherin and generate a fragment referred to as soluble E-cadherin or E-cadherin ectodomain. This cleavage disrupts E-cadherin-mediated cell–cell adhesion and the soluble E-cadherin ectodomain has been linked with increased migration, invasion, proliferation, MMP activity, and disrupted adhesion in various cell lines (Wheelock et al. 1987; Nawrocki-Raby et al. 2003; Chunthapong et al. 2004; Lee et al. 2007; Symowicz et al. 2007; Cowden Dahl et al. 2008). Disruption of cell–cell junctions using a soluble 80 kDa E-cadherin ectodomain fragment also promoted MMP-2, MMP-9, and MT1-MMP expression in human lung tumor cells without affecting the expression of MMP-1, MMP-3, or MMP-7 (Nawrocki-Raby et al. 2003). Recent studies identify MMP-9 as a proteinase responsible for E-cadherin ectodomain shedding in ovarian tumor cells (Symowicz et al. 2007; Cowden Dahl et al. 2008). This relationship appears to be relevant to human disease because addition of exogenous-activated MMP-9 to ovarian cancer cells at a concentration representing the average level found in human ovarian cancer ascites decreased junctional E-cadherin staining, human ovarian tumors with high MMP-9 expression exhibit low or absent E-cadherin staining (Symowicz et al. 2007) and the shed full length E-cadherin ectodomain (sEcad) accumulates to high concentrations in ascites from women with ovarian cancer (Symowicz et al. 2007). Interestingly, integrin-stimulated MMP production may serve to reinforce an EMT. Chronic exposure to either MMP-3 or MMP-9 (but not MMP-2) mediates an EMT in mammary epithelial cells (Radisky et al. 2005) and expression of an autoactivating form of MMP-3 leads to the spontaneous development of premalignant and malignant lesions in the mammary glands of transgenic mice (Sternlicht et al. 2000) suggesting that elevation of certain MMPs may be an important component of the EMT cascade. The EMT program is a network of multiple inputs with integrins playing a pivotal role. Integrins regulate many functions that impinge on hallmarks of EMT, most notably downregulation of E-cadherin expression and function (Wells et al. 2008; Kalluri and Weinberg 2009; Thiery et al. 2009; Zeisberg and Neilson 2009). Integrins activate kinases responsible for phosphorylation and targeting of adherens junction components for degradation, stimulate signaling cascades that induce
520
L.G. Hudson and M.S. Stack
transcriptional regulators of E- and N-cadherin, and increase expression of MMPs responsible for E-cadherin cleavage and release or activation of matrix-associated growth factors and chemokines. Integrin-dependent localization of proteinases additionally contributes to coordinated cell migration and invasion within the matrix (Yilmaz and Christofori 2009). Thus, integrins are dynamic participants in the process of tumor EMT.
Integrins, Mechanotransduction, and Cancer Ample evidence suggests that integrin-mediated ECM adhesion is inherently a mechanosensory process, as mechanical strain may modify the balance of forces within a focal adhesion and the cytoskeleton, and thereby initiate new molecular interactions (Ingber 2002, 2003a, b; Bershadsky et al. 2003; Katsumi et al. 2004). Mechanosensitivity is a universal property of all cells across the evolutionary spectrum (Ingber 2006; Orr et al. 2006). The mechanistic basis of mechanotransduction, conversion of mechanical into chemical signals within the cell, remains largely unknown; however numerous studies have demonstrated that cell adhesion receptors including integrins play a central role in mechanotransduction (Wang et al. 2009; Schwartz and DeSimone 2008; Yamada and Cukierman 2007). In addition to activation of signaling pathways, structural rearrangements induced by mechanical force propagated through cytoskeletal elements can functionally couple altered nuclear architecture to the extracellular environment. While it has long been recognized that the normal functional state of tissues such as cartilage, bone, lung, bladder, and vasculature involves cell and tissue-level response to changes in mechanical forces, studies delineating a role for altered tensional homeostasis in the development or progression of malignancy are much less frequent (LeBeyec et al. 2007; Paszek et al. 2005; Huang and Ingber 2005; Suresh 2007). However, these emerging studies support the hypothesis that mechanotransduction is a critical component in tumor progression and metastasis (Jaalouk and Lammerding 2009). Tumors are generally stiffer (less compliant) than surrounding host tissues (Paszek et al. 2005; Makale 2008), due in large part to modified ECM composition, often termed as “desmoplastic response.” In many cancers, tumor detection is initially based on palpable tissue stiffening, and differences in mechanical compliance form the basis of tumor imaging methods such as ultrasound (Kumar and Weaver 2009). Tumor growth into the surrounding normal tissue also results in increased compressive forces on the tumor, and leaky vasculature coupled with defective lymphatic drainage lead to increased interstitial pressure (Kumar and Weaver 2009). This altered tissue tensional homeostasis translates into cellular changes that may promote malignancy (Butcher et al. 2009). At the cellular level, tensile forces (stretch and compression) are distributed across the cytoskeleton and are balanced by adhesion to the ECM at sites of cell–matrix contact (Ingber 2003a, b; Makale 2008). Integrins are principally responsible for the transmission of mechanical forces from the ECM to the cytoskeleton and through the interior of the cell to the
24 Integrins and Cancer
521
nucleus (Makale 2008; Schwartz and DeSimone 2008). Through integrins coupled to the cytoskeleton, cells can pull against the ECM as a means of sensing matrix stiffness (Yamada and Cukierman 2007). Matrix stiffness, in turn, can alter the surface distribution of integrins and the recruitment of focal adhesion proteins, leading to altered focal adhesion dynamics (Schwartz and DeSimone 2008; Ingber 2003a, b). Mechanical signals are rapidly converted into chemical signals, as evidenced by Src activation at cellular sites distant from the source of force application (Wang et al. 2009). Interestingly, integrin signaling downstream of mechanical stress is distinct from signals generated in response to integrin clustering (Meyer et al. 2000; Chicurel et al. 1998). Changes in external forces are conveyed to the interior of the cell through multiple mechanisms that can ultimately alter gene expression. Engagement of signaling networks downstream of adhesion and/or growth factor receptors in response to force can activate a variety of transcription factors including AP1, p53, STAT1, and NF-kB that, in turn, modify gene expression levels (Butcher et al. 2009; Reichelt 2007; Avvisato et al. 2007). In addition, mechanical coupling from the ECM through integrins that are linked to the nucleus via the cytoskeleton provides a tensional integrity (or tensegrity) model for transfer of applied force directly to the nucleus (Ingber 2003a, b). Interesting new studies have shown that nuclear organization changes in response to culturing cells in 2D vs. 3D environments, resulting in altered chromatin packing (LeBeyec et al. 2007; Giene and Henzel 2008; Lelievre 2009; Castello-Cross et al. 2009). This, in turn, effects gene expression profiles and may provide a mechanism whereby the mechanical environment can influence cellular response to biochemical signals by controlling gene accessibility (Lelievre 2009). Epigenetic regulation of gene expression occurs through promoter methylation, posttranslational modification of histones, and by displacement of nucleosomes along the DNA. Force-induced changes in nuclear organization may alter the availability of transcription factor binding sites on DNA, directly linking nuclear organization and control of gene expression (Lelievre 2009). A potential mechanism for altered chromatin organization is indicated by altered distribution of the nuclear structural protein NuMA when cells are cultured in 3D ECM, resulting in changes in chromosome organization. Additional studies have linked altered cytoskeletal organization to changes in intracellular tension and corresponding changes in histone acetylation levels (Lelievre 2009; Kim et al. 2005). These reports highlight the need for additional studies aimed at elucidating the role of integrin mechanobiology and its interrelationship with chemical diffusion-based signaling in disease regulation (Baker and Aman 2009).
Integrin Signaling as a Therapeutic Target As summarized above, dysregulated integrin expression and/or function has been observed in multiple tumor types, suggesting that integrin-directed therapeutics can have antitumor efficacy. Furthermore, via modulation of cellular signaling pathways
522
L.G. Hudson and M.S. Stack
that control proliferation, survival, and apoptosis, integrin/ECM interactions may undermine the response of tumors to chemotherapeutic agents (Zutter 2007). Antiadhesion strategies may target the surface integrins themselves or adhesionassociated signaling pathways such as those initiated through FAK or ILK. Altered FAK expression and/or activation in human tumor tissues has been described in hundreds of reports. Many studies have used paired normal and tumor tissue or paired primary and metastatic tumor to evaluate the potential role of FAK in tumor progression (Zhao and Guan 2009; van Nimwegen and van de Water 2007). In the majority of cases, invasive or metastatic tumors displayed increased FAK mRNA and protein levels relative to normal or benign tissues (Owens et al. 1995; Weiner et al. 1993). Combined western blot and immunohistochemical analyses have been used to examine concomitant FAK expression and FAK activation (phosphorylation) in a wide variety of tumors including breast, cervical, head and neck, lung, oral, ovarian, and prostate (reviewed in Zhao and Guan 2009), showing correlations between FAK expression, activity, metastasis, and poor prognosis. Although the molecular mechanisms that regulate aberrant FAK expression in cancer are largely unexplored, these studies highlight the potential therapeutic efficacy of targeting FAK kinase activity. Several small molecule inhibitors of FAK have been developed and assessed in preclinical studies and phase I clinical trials. These compounds predominantly function as ATP-competitive inhibitors of FAK activity and exhibit selectivity against FAK and related kinases. For example, the compound TAE226 inhibits FAK as well as insulin-like growth factor I receptor (IGF-1R) and the FAK family kinase Pyk2. Tested in a number of carcinoma cell lines, TAE226 inhibited ECMinduced FAK phosphorylation and IGF-1-induced IGF-IR phosphorylation. Inhibition of adhesion, migration, and viability as well as induction of apoptosis was also reported (Siu et al. 2007; Liu et al. 2007; Bierele et al. 2008; Golubovskaya et al. 2008; Halder et al. 2007; Watanabe et al. 2008; Shi et al. 2007). In a murine model of ovarian carcinoma, TAE226 administration alone reduced tumor burden but showed enhanced efficacy when combined with docetaxel (Halder et al. 2007). The compound PF-562,271 is an ATP-competitive low nanomolar reversible inhibitor of FAK and the related kinase Pyk2. In preclinical studies, the compound blocked FAK phosphorylation in glioma xenografts and dose-dependent inhibition of tumor growth was observed using murine models of prostate, breast, pancreatic, and colon cancers (Roberts et al. 2008). Integrin-mediated ECM adhesion may undermine response of tumors to conventional chemotherapeutics, suggesting combination therapy may enhance efficacy (Zutter 2007). In support of this idea, inhibition of FAK activity in tumors was shown to induce anoikis, providing a potential mechanism for inhibition of tumor growth. PF-562,271 is currently being administered to patients with pancreatic, head and neck, and prostate tumors in phase I clinical trials (www.clinicaltrials.gov). Overexpression or constitutive activation of ILK results in oncogenic progression, suggesting that inhibiting ILK function may also have therapeutic efficacy. Interestingly, recent reports have demonstrated that ILK translocates to the nucleus, where it may function to regulate transcription (Acconica et al. 2007; Goulioumis
24 Integrins and Cancer
523
et al. 2008; Nakrieko et al. 2008). High level ILK expression has been reported in a number of tumor types including ovarian, colon, pancreatic, lung, prostate, and melanoma (McDonald et al. 2008). Expression levels are correlated with stage and grade of tumor as well as with survival (Fielding and Dedhar 2009), suggesting that inhibition of ILK kinase activity may show cancer cell specificity. Inhibition of ILK expression using RNAi or antisense oligonucleotides results in slower growing xenografts in murine models of glioma and melanoma (Edwards et al. 2005; Wong et al. 2007), providing additional rationale for therapeutic targeting of ILK. Small molecule ATP analogue inhibitors of ILK, such as QLT-0267 and QLT0254 have been developed that are potent and selective, reduce proliferation, induce apoptosis, inhibit motility and invasion, and block nuclear b-catenin translocation (reviewed in Fielding and Dedhar 2009; McDonald et al. 2008). Preclinical studies in a variety of murine tumor models have shown that the compounds are well tolerated and exhibit no apparent toxicity (McDonald et al. 2008). Interestingly, recent data show that inhibition of ILK activity leads to mitotic spindle defects and mitotic arrest, suggesting that ILK inhibitors may effectively induce cell death as a singleagent therapy (Fielding et al. 2008). However, preclinical studies demonstrate increased efficacy in combination with standard chemotherapeutic agents such as cisplatin or gemcitabine in murine xenograft models of pancreatic, lung, and anaplastic thyroid cancers (Yau et al. 2005; Liu et al. 2006; Younes et al. 2007). In addition to blocking integrin signaling, a recent opinion paper raised the intriguing possibility that targeting the mechanical properties of the ECM that surrounds the neoplastic tissue may provide a novel approach to reverse cancer progression (Ingber 2008). This hypothesis is based on data, summarized above, showing that enhanced ECM rigidity contributes actively to tumor progression. The use of biomimetic scaffolds that mimic embryonic tissues to restore normal tissue architecture may provide a novel mechanism to reverse altered integrin-mediated mechanical signaling found in tumors and restore normal tissue architecture. Acknowledgments This work was supported by National Institutes of Health/National Cancer Institutes Research Grants CA109545 (MSS, LGH), CA086984 (MSS), CA085870 (MSS), and GM079381 (LGH).
References Acconica F, Barnes CJ, Singh RR et al (2007) Phosphorylation dependent regulation of nuclear localization and functions of integrin linked kinase. Proc Natl Acad Sci USA 104:6782–6787 Ahmed N, Riley C, Rice G et al (2005) Role of integrin receptors for fibronectin, collagen and laminin in the regulation of ovarian carcinoma functions in response to a matrix microenvironment. Clin Exp Metastasis 22:391–402 Al-Jemal R, Harrison DJ (2008) Beta1 integrin in tissue remodeling and repair: from phenomena to concepts. Pharmacol Therapeut 120:81–101 Alexander, NR, Tran, NL, Rekapally et al (2006). N-cadherin gene expression in prostate carcinoma is modulated by integrin-dependent nuclear translocation of Twist1. Cancer Res 66:3365–3369
524
L.G. Hudson and M.S. Stack
Avvisato CL et al (2007) Mechanical force modulates global gene expression and b-catenin signaling in colon cancer cells. J Cell Sci 120:2672–2682 Azzam HS, Thompson EW (1992) Collagen-induced activation of the M(r) 72,000 type IV collagenase in normal and malignant human fibroblatoid cells. Cancer Res 52:4540–4544 Baker EL, Zaman MH (2010) The biomechanical integrin. J Biomech 43:38–44 Barbolina MV, Adley BP, Ariztia EV et al (2007) Microenvironmental regulation of membrane type 1 matrix metalloproteinase activity in ovarian carcinoma cells via collagen-induced EGR1 expression. J Biol Chem 282:4924–4931 Barbolina MV, Stack MS (2008) Membrane type 1-matrix metalloproteinase: substrate diversity in pericellular proteolysis. Semin Cell Dev Biol 19:24–33 Barbolina MV, Adley BP, Kelly DL et al (2008) Motility-related actinin alpha-4 is associated with advanced and metastatic ovarian carcinoma. Lab Invest 88:602–614 Barbolina MV, Adley BP, Kelly DL et al (2009) Downregulation of connective tissue growth factor by three-dimensional matrix enhances ovarian carcinoma cell invasion. Int J Cancer 125:816–825 Barbolina MV, Moss NM, Westfall SD et al (2010) Microenvironmental regulation of ovarian cancer metastasis. Cancer Treat Res 149:319–334 Berrier AL, Yamada KM (2007) Cell-matrix adhesion. J Cell Physiol 213:563–573 Bershadsky AD, Balaban NQ, Geiger B (2003) Adhesion-dependent cell mechanosensitivity. Ann Rev Cell Dev Biol 19:677–695 Bierele EA, Trujillo A, Nagaram A et al (2008) TAE226 inhibits human neuroblastoma cell survival. Cancer Invest 26:145–151 Brabletz T, Jung A, Spaderna S et al (2005) Opinion: migrating cancer stem cells – an integrated concept of malignant tumour progression. Nat Rev Cancer 5:744–749 Burleson KM, Hansen LK and Skubitz APN (2004a) Ovarian carcinoma spheroids disaggregate on type I collagen and invade live human mesothelial cell monolayers. Clin Exp Mets 21:685–697 Burleson KM, Casey RC, Skubitz KM et al (2004b) Ovarian carcinoma ascites spheroids adhere to extracellular matrix components and mesothelial cell monolayers. Gynecol Oncol 93:170–181 Burleson KM, Boente MP, Pambuccian SE et al (2006) Disaggregation and invasion of ovarian carcinoma ascites spheroids. J Transl Med 4:1–16 Butcher DT, Alliston T, Weaver VM (2009) A tense situation: forcing tumor progression. Nat Rev Cancer 9:108–122 Cance WG, Harris JE, Iacocca MV et al (2000) Immunohistochemical analyses of focal adhesion kinase expression in benign and malignant human breast and colon tissues: correlation with preinvasive and invasive phenotypes. Clin Cancer Res 6:2417–2423 Cannistra SA, Kansas GS, Niloff J et al Binding of ovarian cancer cells to peritoneal mesothelium in vitro is partly mediated by CD44H (1993) Cancer Res 53:3830–3838 Castello-Cross R, Khan DR, Simons J et al (2009) Stages stromal extracellular 3D matrices differentially regulate breast cancer cell responses through PI3K and beta1-integrins. BMC Cancer 26:94 Chicurel M, Singer RH, Meyer C et al (1998) Integrin binding and mechanical tension induce moement of mRNA and ribosomes to focal adhesions. Nature 392:730–733 Christiansen JJ, Rajasekaran AK (2006) Reassessing epithelial to mesenchymal transition as a prerequisite for carcinoma invasion and metastasis. Cancer Res 66:8319–8326 Chunthapong J, Seftor E, Khalkhali-Ellis Z et al (2004) Dual roles of E-cadherin in prostate cancer invasion. J Cell Biochem 91:649–661 Cowden Dahl KD, Symowicz J, Ning Y et al (2008) Matrix metalloproteinase 9 is a mediator of epidermal growth factor-dependent e-cadherin loss in ovarian carcinoma cells. Cancer Res 68:4606–4613 Delcommenne M, Tan C, Gray V et al (1998) Phosphoinositide-3-OH kinase-dependent regulation of glycogen synthase kinase 3 and protein kinase B/AKT by the integrin-linked kinase. Proc Natl Acad Sci USA 95:11211–11216
24 Integrins and Cancer
525
Dubash AD, Menold MM, Samson T et al (2009) Focal adhesions: new angles on an old structure. Int Rev Cell Mol Biol 277:1–65 Edwards LA et al (2005) Inhibition of OLK in PTEN-mutant human glioblastomas inhibits PKB/ Akt activation, induces apoptosis and delays tumor growth. Oncogene 24:3596–3605 Ellerbroek SM, Fishman DA, Kearns AS et al (1999) Ovarian carcinoma regulation of matrix metalloproteinase 2 and membrane type 1 matrix metalloproteinase through beta 1 integrin. Cancer Res 59:1635–1641 Ellerbroek SM, Wu YI, Overall CM et al (2001) Functional interplay between type I collagen and cell surface matrix metalloproteinase activity. J Biol Chem 276:24833–24842 Fielding AB et al (2008) Integrin linked kinase localizes to the centrosome and regulates mitotic spindle organization. J Cell Biol 180:681–689 Fielding AB, Dedhar S. (2009) The mitotic functions of integrin linked kinase. Cancer Mets Rev 28:99–111 Fishman DA, Kearns AM, Chilikuri K et al (1998) Metastatic dissemination of human ovarian epithelial carcinoma is promoted by a2b1 integrin-mediated interaction with type I collagen. Invasion Mets 18:15–26 Friedl P, Wolf K (2009) Proteolytic interstitial cell migration: a five-step process. Cancer Metastasis Rev 28:129–135 Frisch SM, Vuori K, Ruoslahti E et al (1996) Control of adhesion-dependent cell survival by focal adhesion kinase. J Cell Biol 134:793–799 Gálvez BG, Matías-Román S, Yáñez-Mó M et al (2002) ECM regulates MT1-MMP localization with beta1 or alphavbeta3 integrins at distinct cell compartments modulating its internalization and activity on human endothelial cells. J Cell Biol 159:509–521 Giehl K, Menke A (2008) Microenvironmental regulation of E-cadherin-mediated adherens junctions. Front Biosci 13:3975–3985 Giene RS, Hendzel MJ (2008) Mechanotransduction from the ECM to the genome: are the pieces now in place? J Cell Biochem 104:1964–1987 Gilles C, Polette M, Seiki M et al (1997) Implication of collagen type I-induced membrane-type 1-matrix metalloproteinase expression and matrix metalloproteinase-2 activaiton in the metastatic progression of breast carcinoma. Lab Invest 76:651–660 Golubovskaya VM, Virnig C, Cance WG (2008) TAE226-induced apoptosis in breast cancer cells with overexpressed Src or EGFR. Mol Carcinog 47:222–234 Goulioumis AK, Bravou V, Varakis J et al (2008) Integrin linked kinase cytoplasmic and nuclear expression in laryngeal carcinomas. Virchows Arch 453:511–519 Grashoff C, Aszódi A, Sakai T et al (2003) Integrin-linked kinase regulates chondrocyte shape and proliferation. EMBO Rep 4:432–438 Guan JL, Trevithick JE, Hynes RO (1991) Fibronectin/integrin interaction induces tyrosine phosphorylation of a 120 kDa protein. Cell Regul 2:951–964 Guan JL, Shalloway D (1992) Regulation of focal adhesion-associated protein tyrosine kinase by both cellular adhesion and oncogenic transformation. Nature 358:690–692 Guarino M, Rubino B, Ballabio G (2007) The role of epithelial-mesenchymal transition in cancer pathology. Pathology 39:305–318 Halder J, Lin YG, Merritt WM et al (2007) Therapeutic efficacy of a novel focal adhesion kinase inhibitor TAE226 in ovarian carcinoma. Cancer Res 67:10976–10983 Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57–70 Harburger DS, Calderwood DA (2009) Integrin signalling at a glance. J Cell Sci 122:159–163 Hay ED (2005) EMT concept and examples from the vertebrate embryo. In: Savagner P (ed) Rise and fall of epithelial phenotype: concepts of epithelial-mesenchymal transition. Springer, Berlin, pp 111–134 Holmbeck K, Bianco P, Caterina J et al (1999) MT1-MMP-deficient mice develop dwarfism, osteopenia, arthritis, and connective tissue disease due to inadequate collagen turnover. Cell 99:81–92 Huang S, Ingber DE (2005) Cell tension, matrix mechanics, and cancer development. Cancer Cell 8:175–176
526
L.G. Hudson and M.S. Stack
Humphries JD, Byron A, Humphries MJ (2006) Integrin ligands at a glance. J Cell Sci 119:3901–3903 Hynes RO (2002) Integrins: bidirectional, allosteric signaling machines. Cell 110:673–687 Illic D, Furuta Y, Kanazawa S et al (1995) Reduced cell motility and enhanced focal adhesion contact formation in cells from FAK-defieicnt mice. Nature 377:539–544 Imamichi Y, König A, Gress T et al (2007) Collagen type I-induced Smad-interacting protein 1 expression downregulates E-cadherin in pancreatic cancer. Oncogene 26:2381–2385 Ingber DE (2002) Cancer as a disease of epithelial-mesenchymal interactions and extracellular matrix regulation. Differentiation 70:547–560 Ingber DE (2003a) Tensegrity I. Cell structure and hierarchical systems biology. J Cell Sci 116:1157–1173 Ingber DE (2003b) Tensegrity II. How structural networks influence cellular information processing networks. J Cell Sci 116:1397–1408 Ingber DE (2006) Cellular mechanotransduction: putting all the pieces together again. FASEB J 20:811–827 Ingber DE (2008) Can cancer be reversed by engineering the tumor microenvironment? Semin Cancer Biol 18:356–364 Jaalouk DE, Lammerding J (2009) Mechanotransduction gone awry. Nat Rev Mol Cell Biol 10:63–73 Kalluri R, Weinberg RA (2009) The basics of epithelial-mesenchymal transition. J Clin Invest 119:1420–1428 Katsumi A, Orr AW, Tzima E et al (2004) Integrins in mechano-transduction. J Biol Chem 279:12001–12004 Kenny HA, Krausz T, Yamada SD et al (2007) Use of a novel 3D culture model to elucidate the role of mesothelial cells, fibroblasts and extracellular matrices on adhesion and invasion of ovarian cancer cells to the omentum. Int J Cancer 121:1463–1472 Kim YB, Yu J, lee SY et al (2005) Cell adhesion status-dependent histone acetylation is regulated through intracellular contractility-related signaling activities. J Biol Chem 280:28357–28364 Koenig A, Mueller C, Hasel C et al (2006) Collagen type I induces disruption of E-cadherinmediated cell-cell contacts and promotes proliferation of pancreatic carcinoma cells. Cancer Res 66:4662–4671 Kornberg LJ, Earp HS, Turner CE et al (1991) Signal transduction by integrins: increased protein tyrosine phosphorylation caused by clustering of beta 1 integrins. Proc Nat Acad Sci USA 88:8392–8396 Kumar S, Weaver VM (2009) Mechanics, malignancy and metastasis: the force journey of a tumor cell. Cancer Metastasis Rev 28:113–127 Lark AL, Livasy CA, Dressler L et al (2005) High focal adhesion kinase expression in invasive breast carcinomas is associated with an aggressive phenotype. Mod Pathol 18:1289–1294 Larsen M, Artym VV, Green JA et al (2006) The matrix reorganized: extracellular matrix remodeling and integrin signaling. Curr Opin Cell Biol 18:463–471 Lauffenburger DA, Horwitz AF (1996) Cell migration: a physically integrated molecular process. Cell 84:359–369 LeBeyec J, Xu R, Lee SY et al (2007) Cell shape regulates global histone acetylation in human mammary epithelial cells. Exp Cell Res 313:3066–3075 Lee KH, Choi EY, Hyun MS et al (2007) Association of extracellular cleavage of E-cadherin mediated by MMP-7 with HGF-induced in vitro invasion in human stomach cancer cells. Eur Surg Res 39:208–215 Legate KR, Fässler R (2009) Mechanisms that regulate adaptor binding to beta-integrin cytoplasmic tails. J Cell Sci 122:187–198 Lelievre SA (2009) Contributions of extracellular matrix signaling and tissue architecture to nuclear mechanisms and spatial organization of gene expression control. Biochim Biophys Acta 1790:925–935 Lim ST, Chen XL, Lim Y et al (2008a) Nuclear FAK promotes cell proliferation and survival through FERM-enhanced p53 degradation. Mol Cell 29:9–22
24 Integrins and Cancer
527
Lim ST, Mikolon D, Stupack DG et al (2008b) FERM control of FAK function: implications for cancer therapy. Cell Cycle 7:2306–2314 Liu WF, Nelson CM, Pirone DM et al (2006) E-cadherin engagement stimulates proliferation via Rac1. J Cell Biol 173:431–441 Liu TJ, LaFortune T, Honda T et al (2007) Inhibition of both focal adhesion kinase and insulinlike growth factor I receptor kinase suppresses glioma proliferation in vitro and in vivo. Mol Cancer Ther 6:1357–1367 Lorenz K, Grashoff C, Torka R et al (2007) Integrin-linked kinase is required for epidermal and hair follicle morphogenesis. J Cell Biol 177:501–513 Luo BH, Springer TA (2006) Integrin structures and conformational signaling. Curr Opin Cell Biol 18:579–586 Makale M (2008) Cellular mechanobiology and cancer metastasis. Birth Defects Res 81:329–343 McDonald PC, Fielding AB, Dedhar S. (2008) Integrin linked kinase—essential roles in physiology and cancer biology. J Cell Sci 121:3121–3132 McLean GW, Carragher N, Avizienyte E et al (2005) The role of focal adhesion kinase in cancer – a new therapeutic opportunity. Nat Rev Cancer 5:505–515 Meyer CJ, Alenghat FJ, Rim P et al (2000) Mechanical control of cyclic AMP signaling and gene transcription through activated integrins. Nat Cell Biol 2:666–668 Mitra SK, Schlaepfer DD (2006) Integrin-regulated FAK-Src signaling in normal and cancer cells. Curr Opin Cell Biol 18:516–523 Miyamoto S, Teramoto H, Coso OA et al (1995) Integrin function: molecular hierarchies of cytoskeletal and signaling molecules. J Cell Biol 131:791–805 Moreno-Bueno G, Portillo F, Cano A (2008) Transcriptional regulation of cell polarity in EMT and cancer. Oncogene 27:6958–6969 Moser TL, Pizzo SV, Bafetti LM et al (1996) Evidence for preferential adhesion of ovarian epithelial carcinoma cells to type I collagen mediated by the alpha2beta1 integrin. Int J Cancer 67:695–701 Moss NM, Barbolina MV, Liu Y et al (2009) Ovarian cancer cell detachment and multicellular aggregate formation are regulated by membrane type 1 matrix metalloproteinase: a potential role in I.p. metastatic dissemination. Cancer Res 69:7121–7129 Munshi HG, Stack MS (2006) Reciprocal interactions between adhesion receptor signaling and MMP regulation. Cancer Metastasis Rev 25:45–56 Nakrieko KA, Vespa A, Mason D et al (2008) Modulation of integrin linked kinase nucleocytoplasmic shuttling by ILKAP and CRM1. Cell Cycle 15:2157–2166 Nawrocki-Raby B, Gilles C, Polette M et al (2003) Upregulation of MMPs by soluble E-cadherin in human lung tumor cells. Int J Cancer 105:790–795 Legate KR, Wickström SA, Fässler R (2009) Genetic and cell biological analysis of integrin outside-in signaling. Genes Dev 23:397–418 Niu, RF, Zhang L, Xi GM et al (2007) Up-regulation of Twist induces angiogenesis and correlates with metastasis in hepatocellular carcinoma. J Exp Clin Cancer Res 26:385–394 Novak A, Hus SC, Leung-Hagesteijn C et al (1998) Cell adhesion and the integrin linked kinase regulate the LEF1 and beta catenin signaling pathways. Proc Natl Acad Sci USA 95:4374–4379 Oloumi A, Syam S, Dedhar S. (2006) Modulation of Wnt3a-mediated nuclear beta-catenin accumulation and activation by integrin-linked kinase in mammalian cells. Oncogene 25:7747–7757 Orlichenko LS, Radisky DC (2008) Matrix metalloproteinases stimulate epithelial-mesenchymal transition during tumor development. Clin Exp Metastasis 25:593–600 Orr AW, Helmke BP, Blackman BR et al (2006) Mechanisms of mechanotransduction. Dev Cell 10:11–20 Owens LV, Xu L, Craven FJ et al (1995) Overexprssion of the focal adhesion kinase (p125FAK) in invasive human tumors. Cancer Res 55:2752–2755 Paszek MJ, Zahir N, Johnson KR et al (2005) Tensional homeostasis and the malignant phenotype. Cancer Cell 8:241–254
528
L.G. Hudson and M.S. Stack
Persad S, Troussard AA, McPhee TR et al (2001) Tumor suppressor PTEN inhibits nuclear accumulation of b-catenin and TCF1 mediated transcriptional activation. J Cell Biol 153:1161–1174 Radisky DC, Levy DD et al (2005) Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature 436:123–127 Reichelt J (2007) Mechanotransduction of keratinocytes in culture and in the epidermis. Eur J Cell Biol 86:807–816 Roberts WG, Ung E, Whalen P et al (2008) Antitumor activity and pharmacology of a selective focal adhesion kinase inhibitor PF-562,271. Cancer Res 68:1935–1944 Rowe RG, Weiss SJ (2008) Breaching the basement membrane: who, when and how? Trends Cell Biol 18:560–574 Sakai T, Li S, Docheva D, Grashoff C et al (2003) Integrin-linked kinase (ILK) is required for polarizing the epiblast, cell adhesion, and controlling actin accumulation. Genes Dev 17:926–940 Schaller MD, Borgman CA, Cobb BS et al (1992) pp125FAK a structurally distinctive proteintyrosine kinase associated with focal adhesions. Proc Natl Acad Sci USA 89:5192–5196 Schwartz MA, DeSimone DW (2008) Cell adhesion receptors in mechano-transduction. Curr Opin Cell Bol 20:551–556 Shen TL, Park AYJ, Alcaraz A et al (2005) Conditional knockout of focal adhesion kinase in endothelial cells reveals its role in angiogenesis and vascular development in late embryogenesis. J Cell Biol 169:941–952 Shi Q, Hjelmeland AB, Keir ST (2007) A novel low-molecular weight inhibitor of focal adhesion kinase, TAE226, inhibits glioma growth. Mol Carcinog 46:488–496 Sherwood DR (2006) Cell invasion through basement membranes: an anchor of understanding. Trends Cell Biol 16:250–256 Shintani Y, Fukumoto Y, Chaika N et al (2008). Collagen I-mediated up-regulation of N-cadherin requires cooperative signals from integrins and discoidin domain receptor 1. J Cell Biol 180:1277–1289 Shintani Y, Hollingsworth MA, Wheelock MJ et al (2006) Collagen I promotes metastasis in pancreatic cancer by activating c-Jun NH(2)-terminal kinase 1 and up-regulating N-cadherin expression. Cancer Res 66:11745–11753 Siu LL, Burris HA, Mileshkin L et al (2007) Phase 1 study of a focal adhesion kinase (FAK) inhibitor PF-00562271 in patients with advanced solid tumors. J Clin Oncol 25:3527 Sternlicht MsD, Bissell MJ, Werb Z (2000). The matrix metalloproteinase stromelysin-1 acts as a natural mammary tumor promoter. Oncogene 19:1102–1113 Suresh S (2007) Biomechanics and biophysics of cancer cells. Acta Biomater 3:413–438 Symowicz J, Adley BP, Gleason KJ et al (2007) Engagement of collagen-binding integrins promotes matrix metalloproteinase-9-dependent E-cadherin ectodomain shedding in ovarian carcinoma cells. Cancer Res 67:2030–2039 Terpstra L, Prud’homme J, Arabian A et al (2003) Reduced chondrocyte proliferation and chondrodysplasia in mice lacking the integrin-linked kinase in chondrocytes. J Cell Biol 162:139–148 Thiery JP, Acloque H, Huang RY et al (2009) Epithelial-mesenchymal transitions in development and disease. Cell 139:871–890 Tse JC, Kalluri R (2007) Mechanisms of metastasis: epithelial-to-mesenchymal transition and contribution of tumor microenvironment. J Cell Biochem 101:816–829 van Nimwegen MJ, van de Water B (2007) Focal adhesion kinase: a potential target in cancer therapy. Biochem Pharmacol 73:597–609 Wang N, Tytell JD, Ingber DE (2009) Mechanotransduction at a distance: mechanically coupling the extracellular matrix with the nucleus. Nat Rev Mol Cell Biol 10:75–83 Watanabe N, Takaoka M, Sakurama K et al (2008) Dual tyrosine kinase inhibitor for focal adhesion kinase and insulin-like growth factor I receptor exhibits anticancer effect in esophageal adenocarcinoma in vitro and in vivo. Clin Cancer Res 14:4631–4639
24 Integrins and Cancer
529
Weiner TM, Liu ET, Craven RJ et al (1993) Expression of focal adhesion kinase gene in invasive cancer. Lancet 342:1024–1025 Wells A, Yates C, Shepard CR (2008) E-cadherin as an indicator of mesenchymal to epithelial reverting transitions during the metastatic seeding of disseminated carcinomas. Clin Exp Metastasis 25:621–628 Wheelock MJ, Buck CA, Bechtol KB et al (1987) Soluble 80-kd fragment of cell-CAM 120/80 disrupts cell:cell adhesion. J Cell Biochem 34:187–202 Wheelock MJ, Shintani Y, Maeda M et al (2008) Cadherin switching. J Cell Sci 121:727–735 Wolf K, Friedl P (2009) Mapping proteolytic cancer cell-extracellular matrix interfaces. Clin Exp Metastasis 26:289–298 Wolf K, Mazo I, Leung H et al (2003) Compensation mechanism in tumor cell migration: mesenchymal-amoeboid transition after blocking of pericellular proteolysis. J Cell Biol 160:267–277 Wolf K et al (2007) Multi-step pericellular proteolysis controls the transition from individual to collective cancer cell invasion. Nat Cell Biol 9:893–904 Wong RPC et al (2007) The role of integrin-linked kinase in melanoma cell migration, invasion and tumor growth. Mol Cancer Ther 6:1692–1700 Wu C, Dedhar S (2001) Integrin linked kinase and its interactors: a new paradigm for the coupling of extracellular matrix to actin cytoskeleton and signaling complexes. J Cell Biol 12:505–510 Yamada KM, Cukierman E (2007) Modeling tissue morphogenesis and cancer in 3D. Cell 130:601–610 Yan C, Boyd DD (2007) Regulation of matrix metalloproteinase gene expression. J Cell Physiol 211:19–26 Yang Z, Zhang X, Gang H et al (2007) Up-regulation of gastric cancer cell invasion by Twist is accompanied by N-cadherin and fibronectin expression. Biochem Biophys Res Commun 358:925–930 Yau CY, Wheeler JJ, Sutton KL et al (2005) Inhibition of integrin linked kinase by a selective small molecule inhibitor QLT0254 inhibits the PI3K/PKB/mRTOR, Stat3 and FKHR pathways and tumor growth, and enhances gemcitabine-induec apoptosis in human orthotopic primary pancreatic cell exnografts. Cancer Res 654:1497–1504 Yilmaz M, Christofori G (2009) EMT, the cytoskeleton and cancer cell invasion. Cancer Metastasis Rev 28:15–33 Younes MN, Yigitbasi OG, Yazici YD et al (2007) Effects of the integrin-linked kinase inhibitor ZLT0267 on squamous cell carcinoma of the head and neck. Arch Otolaryngol Head Neck Surg 133:15–23 Zeisberg M, Neilson EG (2009). Biomarkers for epithelial-mesenchymal transitions. J Clin Invest 119:1429–1437 Zhao J, Guan JL (2009) Signal transduction by focal adhesion kinase in cancer. Cancer Metastasis Rev 28:35–49 Zutter MM (2007) Integrin-mediated adhesion: tipping the balance between chemosensitivity and chemoresistance. Adv Exp Med Biol 608:87–100
Chapter 25
Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis Stanley Zucker and Jian Cao
Abstract A wealth of knowledge has accumulated over the past four decades on the importance of matrix metalloproteinases (MMPs) in cancer induction, invasion, and metastasis. This chapter aims to provide the reader with recent information to help understand the disconnect between experimental observations implicating the crucial role of MMPs in cancer progression and the historic failure of several broadspectrum MMP inhibitors in clinical drug trials in advanced cancer. The chemistry and biology of the large MMP family and tissue inhibitors of MMPs (TIMPs) will be summarized. Complexity of MMP function in cancer will be described with an emphasis on pericellular cleavage of extracellular matrix (ECM) and non-ECM substrates. Production of MMPs by stromal cells within a tumor, as well as cancer cells, is well established. Anticancer effects of selected in MMPs are described. The study of cell migration within a three-dimensional collagen matrix has been responsible for broadening our understanding of cancer progression. The involvement of MMPs in the transition from noninvasive to invasive, metastatic cancer, and an emphasis on epithelial-to-mesenchymal transition (EMT) will be presented. New approaches to improve the specificity of MMP inhibitors for use in future clinical trials are discussed.
Introduction Cancer progression is recognized to be a complex, multistage process in which the transformation from normal to malignant cells involves genetic changes that lead to numerous phenotypic alterations. The characteristics and capabilities of cells following malignant transformation have been described to include the following (1) self-sufficiency in growth signaling – cancer cells produce their own growth factors S. Zucker (*) Veterans Affairs Medical Center, Northport, NY 11768, USA and Stony Brook University, Stony Brook, NY 11794, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_25, © Springer Science+Business Media, LLC 2010
531
532
S. Zucker and J. Cao
and autostimulate; (2) lack of sensitivity to natural antigrowth factors; (3) evasion of apoptosis (avoidance of destruction by immune mechanisms); (4) uncontrolled replication potential; (5) sustained angiogenesis; and (6) tissue invasion and metastasis (Hanahan and Weinberg 2000). Although a myriad of incompletely understood factors are involved in cancer invasion and metastasis, this review will focus on the important role of MMPs in these processes. Rather than reiterate the wealth of information published in numerous excellent reviews on the subject (Coussens et al. 2002; Martin and Matrisian 2007; Overall and Kleifeld 2006a; Page-McCaw et al. 2007; Pavlaki and Zucker 2003; Sabeh et al. 2009; Yamada and Cukierman 2007; Yamagucki et al. 2005; Zucker et al. 2001a), this chapter aims to provide the reader with recent information to help understand the disconnect between experimental observations implicating the crucial role of MMPs in cancer dissemination and the previous failure of clinical drug trials of MMP inhibitors in advanced cancer. We hope to convince the reader that the lack of clinical efficacy of MMP inhibitory drugs developed in the past does not disprove the causal relationship between MMPs and cancer invasion, but in hindsight reflects the initiation of the wrong type of drug trials, based on incomplete understanding of the complexity of cancer pathobiology.
Classification of Proteases Proteases are the efficient executioners of a common chemical reaction: the hydrolysis of peptide bonds. Based on the mechanism of catalysis, mammalian proteases are classified into five distinct classes: aspartic, cysteine, serine, threonine, and metalloproteases. Proteases of the different classes are further grouped into families on the basis of amino acid sequence comparison, and families are assembled into clans based on similarities of their 3D structure (Lopez-Otin and Bond 2008). The members of the metzincin family are the matrixins (aka: MMPs), astacins, adamalysins (ADAMs and ADAMTSs), serralysins, snapalysins, and leishmanolysins. Many proteases link their catalytic domains to a variety of specialized functional modules or domains that provide substrate specificity, guide their cellular localization, modify their kinetic properties, and change their sensitivity to endogenous inhibitors (Lopez-Otin and Bond 2008). Current emphasis is placed on the fact that proteases act in the context of complex cascades, pathways, circuits, and networks, comprising many protein partners that dynamically interact to form the so-called protease web (Lopez-Otin and Bond 2008).
MMP Biology Interstitial collagenase, the first MMP family member identified, was discovered (Gross and Lapiere 1962) in experiments designed to explain how the collagen-rich tail of the frog is resorbed during metamorphysis. Following secretion from cells,
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
533
MMPs have been demonstrated to participate in many physiologic processes including tissue turnover and repair during blastocyst implantation, ovulation, postlactational involution, and bone resorption. Since collagens represent the major structural proteins of all tissues and the chief obstacle to tumor cell migration, it was initially postulated that collagenolytic enzymes play pivotal roles in facilitating dissemination of cancer. A pathological role for MMPs in arthritis, nonhealing wounds, aortic aneurysms, congestive heart failure, and other disorders has also been recognized. More recently, MMPs have also been incriminated in more complicated processes including the activation or inactivation of other proteins through limited hydrolysis of selected bonds, as well as the shedding of membrane-anchored proteins. Uncovering protease substrates has been facilitated by the use of modern proteomic techniques, which reveal the dynamic interplay of proteases within the cellular microenvironment (Dean and Overall 2007). MMP substrates include other proproteases, protease inhibitors, antimicrobial peptides, clotting factors, chemotactic and adhesion molecules, hormones, growth factors, angiogenic factors, regulators of immunity, and cytokines as well as their receptors and binding proteins (Dean and Overall 2007). In these shedding functions, MMPs overlap in substrate specificity, spatial location, and in their proclivity to cleave upstream of hydrophobic residues (Nagase and Woessner 1999; Tallant et al. 2010).
MMP Chemistry MMPs are a family of Zn2+-dependent proteins and peptide hydrolases; 24 paralogs are present in humans. These enzymes have both a descriptive name typically based on a preferred substrate and an MMP numbering system based on the order of discovery (Fig. 25.1). The basic structure of MMPs consists of the following domains (1) a signal peptide which directs MMPs to the secretory pathway; (2) an ~80 residue prodomain that confers latency to the enzymes and is configured as a 3-helix globular domain; (3) a zinc containing catalytic domain which is compact, spherical, ~165 residues in length, which is divided by a shallow substrate-binding crevice into an upper and a lower sub domain. The catalytic domain has an extended zinc-binding motif, HEXXHXXGXXH, which contains three zinc-binding histidines and a glutamate that activates a zinc-bound H2O molecule providing the nucleophile that cleaves peptide bonds (Birkedal-Hansen 1995; Stoker and Bode 1995). A conserved methionine lying within the “Met-turn” provides a hydrophobic base for the zincbinding site, (4) an ~200 residue hemopexin (PEX) domain with a characteristic 3D disc-like β-propeller structure, which mediates interactions with substrates and confers specificity of the enzymes. Each propeller blade is constituted of four antiparallel β-strands and one α-helix; and (5) a 15–65 flexible residue linker (hinge) region which links the catalytic and the PEX domain. Further earmarks of MMP family members are three a-helices and a five-stranded-b-sheet, as well as at least two calcium binding sites and a second zinc site with structural functions. MMP-7, -23, and -26 lack hinge and PEX domains. MMP-2 and MMP-9 contain fibronectin type II repeats, which mediate binding to collagen and gelatins.
534
S. Zucker and J. Cao
Fig. 25.1 History and domain structure of matrix metalloproteinases (MMPs). MMP family members can be divided into eight subgroups based on their domain structures. Almost half of the 24 MMPs were initially identified following protein purification (first year listed in parenthesis). Subsequently, reverse transcriptase polymerase chain reaction became the routine approach for DNA cloning. Using degenerate primers based on highly conserve regions of amino acid sequences in the propeptide domain (PRCGVPD) and catalytic domain (HEXGHXXGXXH) of MMP-1, the cDNAs of known MMPs were identified (second year listing within parenthesis). Using this approach, new family members of MMPs, including MT-MMPs, were subsequently identified
Most MMPs are secreted as inactive zymogens with a cysteine embedded in a conserved PRCGXPD within the prodomain, folding into and inhibiting the catalytic zinc. Removal of the prodomain enables access of the catalytic solvent molecule and substrate molecules to the active site cleft, which harbors a hydrophobic S1 pocket as the main determinant of specificity. MMP-23 is the only MMP to lack a prodomain. MMPs were initially classified into groups according to preferred substrates (1) true collagenases, which cut triple helical collagen at a single site across the three chains; (2) gelatinases, which target denatured collagen and gelatin; (3) stromelysins, which have broad substrate specificity and degrade many proteoglycans; and (4) other MMPs whose biologic function is not well understood. Additional knowledge of structural domains has led to further division of MMPs into subgroups (Fig. 25.1). A breakthrough in understanding the role of MMPs in cancer came with the discovery that cell surface-bound MMPs displayed enhanced function in the pericellular environment (Zucker et al. 1985). Four of the six membrane type-MMPs (MT1-, MT2-, MT3-, and MT5-MMP) contain an additional 20 amino acid transmembrane domain and a small cytoplasmic domain (Seiki 2002). The two remaining MT-MMPs (MT4- and MT6-MMP) are tethered to the plasma membrane via a glycosylphosphatidyl inositol linkage, which tethers them to the cell surface. Other MMPs lacking transmembrane binding domains (MMP-1, -2, -7, -9, -13) are capable of binding to other molecules on the cell surface following secretion.
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
535
Natural Inhibitors of MMPs Despite the presence of several well characterized natural inhibitors, little is known about the physiologic mechanisms that inactivate MMPs in the pericellular environment (Fu et al. 2008). Tissue inhibitors of metalloproteinases (TIMP-1, -2, -3, and -4) comprise a four-member family of homologous MMP inhibitors that, with few exceptions, inhibit the activation and the function of all MMPs (Baker et al. 2002). The major exception is that TIMP-1 is relatively ineffective in inhibiting MT-MMPs. In addition to binding to the catalytic site of activated MMPs, TIMP-1 binds to the PEX domain of proMMP-9 (Goldberg et al. 1989) and TIMP-2 binds to the PEX domain of proMMP-2 (Howard et al. 1991; Morgunova et al. 2002). PEX domain binding serves to regulate MMP function. TIMPs have also been shown to have growth promoting activities, which are independent of their MMP inhibitory function (Sternlicht and Werb 2001); the biologic significance of these effects remains to be clarified. The transcription of TIMPs is regulated by similar cytokines and growth factors that control MMP expression, i.e., TGFb, TNFa, IL-1, IL-6, although in distinctive ways. Total TIMP concentrations in tissue and extracellular fluids generally far exceed the concentration of MMPs, thereby limiting proteolytic activity to focal pericellular sites. As an example of this imbalance, free activated MMPs are not generally detectable in plasma (Zucker et al. 2004a). Although the concentration of individual TIMPs and MMPs are readily measurable in tissues and fluids by immunoassay techniques, these static measurements may not provide an accurate assessment of activities in the pericellular and intracellular environment. Although one might expect high levels of TIMP to interfere with cancer progression, increased tissue or blood levels of TIMP-1 have been correlated with poor prognosis in cancer (Zucker et al. 2004b). Other endogenous inhibitors of MMPs of uncertain physiologic function include a2 macroglobulin (a very potent, high molecular weight inhibitor with high concentrations in blood), the reversion-inducing cysteine-rich protein with kazal motifs (RECK) (Takahashi et al. 1998), b-amyloid precursor protein, tissue factor pathway inhibitor-2, procollagen C-terminal proteinase enhancer (Tallant et al. 2010) secreted leukocyte protease inhibitor, and cystatins (Overall and Kleifeld 2006a).
Regulation of MMP Function In vivo activity of MMPs is under tight control at several levels including gene expression, compartmentalization, proenzyme activation, inhibition by protease inhibitors, and endocytosis. Most members of the MMP family share common cisacting elements in their promoters. As a result, they are often co-expressed in response to inductive stimuli or co-repressed by inhibitors of gene expression, such as glucocorticoids (see review by Vincenti and Brinkerhoff 2007). Promoters of all MMPs contain multiple elements that cooperate to either induce or repress gene
536
S. Zucker and J. Cao
expression. These include ETS, Sp1, NFkB, and AP-1 sites. Epigenetic mechanisms, e.g., histone acetylation, methylation also contribute to the regulation of MMPs (Vincenti and Brinkerhoff 2007). MMPs have recently been sorted into three groups that are distinguished based on the mechanism regulating expression. These enzymes are generally expressed in very low amounts and their transcription is tightly regulated either positively or negatively by specific signals that are temporally and spatially confined, e.g., cytokines and growth factors such as interleukins (IL-1, IL-4, IL-6), transforming growth factors, tumor necrosis factor alpha (TNF-a), and lysophosphatidic acid (Sternlicht and Werb 2001). Some of these regulatory molecules can be proteolytically activated or inactivated by MMPs (feedback effect). Some MMPs are stored in inflammatory cell granules, which restrict their action. Although numerous regulatory processes have been identified, alteration in in vivo control mechanisms in cancer has received scant attention. Activation of MMPs following secretion from cells depends on the disruption of the prodomain interaction with the catalytic site, which may occur by conformational changes, oxidants, or proteolytic removal of the prodomain. There are a few exceptions to the rule that allosteric interactions lead to autolytic cleavage of the prodomain (Cao et al. 1998; Fu et al. 2008). MMPs that contain furin-like recognition domains in their propeptides (MMP-11, MMP-28, MT-MMPs) can be activated in the trans Golgi network by members of the subtilisin family (primarily furin) of serine proteases. With the exception of MMP-2, the mechanism for in vivo activation of secreted MMPs is not well understood. Activation of secreted proMMP-2 and proMMP-13 is mediated by a cell-surface complex that consists of a homodimer of MT1-MMP as well as a single molecule of TIMP-2. TIMP-2 binds to the catalytic domain of one of the MT1-MMP molecules in the dimer and to the PEX domain of proMMP-2, thereby facilitating cleavage and activation of proMMP-2 by the second MT1-MMP molecule of the dimer (Fig. 25.2) (Butler et al. 1998; Sato et al. 1994; Strongin et al. 1995). Extracellular proteolytic activation of secreted MMPs can be mediated by serine proteases such as plasmin, which implies an interdependence of these two enzyme groups in ECM remodeling (Zucker et al. 2002). Although some active MMPs can activate other proMMPs (MMP-3 activates proMMP-9 and proMMP-1), the physiologic relevance of this mechanism has been called into question (Fu et al. 2008). Recent studies have emphasized the importance of endocytosis and exocytosis as a way of controlling the activity of MT1-MMP. The observations that tumor cells assemble specialized degradative structures (invadopodia) when cultured on a 2D matrix and that MT1-MMP mediates focal degradation at invadopodia have provided a powerful model for understanding these pericellular events (Poincloux et al. 2009). MT1-MMP is efficiently internalized by clathrin-mediated and caveolar endocytosis with trafficking through early- and late-endosomal and lysosomal compartments, where degradation can occur. Intracellular processing and activation of MT1-MMP depends on its partitioning into lipid domains (Mazzone et al. 2004). Recycling of MT1-MMP from endosomes has been proposed as a means of regenerating the active enzyme at the cell surface (Itoh and Seiki 2006; Poincloux et al. 2009).
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
537
Fig. 25.2 Hypothetical model of proMMP-2 activation by MT1-MMP at the cell surface. (a) Activation of proMMP-2 by MT1-MMP under stoichiometric TIMP-2:MT1-MMP conditions (protein ratio ~1:1). MT1-MMP is depicted as a homodimer (formed through the PEX domain) adjacent to the plasma membrane. TIMP-2 serves to both immobilizes proMMP-2 (binding to PEX domain) and bind to the catalytic domain of MT1-MMP, thus forming a trimer at the cell surface. In the absence of excess TIMP-2, an adjacent MT1-MMP molecule (free of TIMP-2) cleaves off the prodomain of MMP-2, leading to MMP-2 activation (exposed Zn in catalytic site). (b) Inhibition of proMMP-2 activation by excess TIMP-2. In the presence of excess TIMP-2, the catalytic sites of both MT1-MMP molecules are occupied by TIMP-2, thus preventing the activation of proMMP-2
Internalization of proteolytically inactive MT1-MMP–TIMP-2 complexes represents a mechanism to dissociate TIMP-2 from MT1-MMP within the endocytic pathway in order to regenerate the active protease (Maquoi et al. 2000; Zucker et al. 2004a). A regulatory role of Src in MT1P-MMP function, involving endophilin A2, has been emphasized (Poincloux et al. 2009). Cortactin appears to coordinate MT1MMP exocytosis and actin reorganization at invadopodia; IQGAP1 (a key polarity regulator that links microtubule and actin cytoskeletal networks) and the exocyst vesicle docking complex are required for this invadopodia complex (Poincloux et al. 2009). CD44 serves as a cell surface-docking molecule for MMP-9 and MT1-MMP (Suenaga et al. 2004) and has been implicated as important for cell function (Samanna et al. 2007; Yu and Stamenkovic 1999). Dufour et al. (2008) recently demonstrated that proMMP-9 can enhance epithelial cell migration independent of its proteolytic activity. Complex formation between CD44 and blade I of the PEX domain of proMMP-9 is critical to this cell migration, and involves EGFR and phosphorylation of FAK, AKT, and ERK1/2 pathways (Dufour et al. 2010).
538
S. Zucker and J. Cao
Participation of MMPs in Various Aspects of Cancer Overwhelming experimental evidence accumulated over the last 40 years has implicated proteases in cancer dissemination. Although serine, cysteine, aspartic, and metalloproteinases have been implicated in the invasive process, MMPs appear to exert the dominant effect and have been implicated in virtually all aspects of cancer progression and dissemination (Pavlaki and Zucker 2003). Initial emphasis was on the few MMPs that can cleave intact fibrillar collagen (types I, II, and III) that make up the ECM. Based on early in vitro studies of cancer cell invasion, Liotta et al. (1980) placed special emphasis on the degradation of type IV collagen, the dominant basement membrane (BM) protein, by the gelatinases (MMP-2 and MMP-9). Basement membranes form a dense lamina that underlies all epithelia and ensheaths endothelial cells, nerves, smooth muscle cells, and adipocytes. MMPs were subsequently shown to collectively cleave and degrade virtually all other ECM components including laminin, fibronectin, vitronectin, elastin, enactin, and proteoglycans. There is current interest in MMP cleavage of chemokines, which may have profound effects on numerous biologic processes, not just limited to inflammation (McQuibban et al. 2001). Biological function of MMPs has been described in multiple cellular processes, including proliferation, angiogenesis, migration, invasion, and host defense (Sternlicht and Werb 2001). ADAMs have also been implicated in various aspects of cancer progression (see review by Mochizuki and Okada 2007). Hundreds of studies in experimental animals by numerous laboratories have demonstrated (1) that disease progression (invasion and metastasis) of cancer correlate with enhanced production and secretion of MMPs by tumor cells and/or stromal cells; (2) reduction of tumor growth and metastasis in vivo using natural (TIMPs) and synthetic protease inhibitors, neutralizing antibodies or antisense oligonucleotides; (3) modulation of the invasive properties of cancer cells by transfection with the cDNA of MMPs and their inhibitors, and (4) alteration of tumor growth and metastasis in mice, genetically modified in terms of MMP production. The question vexing scientists for the past two decades is: Does the presence of high tumor concentrations of MMPs (capable of degrading extracellular matrix) indicate that they have a causative role in cancer progression/dissemination? Correlation does not necessarily prove causation. Opposing views on the importance of MMPs in cancer have been presented (Della Porta et al. 1999; Noel et al. 1997; Zucker et al. 1992). It is fair to say that our current understanding of mechanisms by which tumor cells digest highly complex mixtures of fibrillar proteins in vivo is far from complete. Despite structure and sequence homologies and overlapping substrates, different MMPs display diverse and sometimes opposite effects depending upon the cellular source, tissue localization, and stage of cancer. Over the years, many fascinating, but difficult to prove theories of MMP function in cancer have been proposed, e.g., (1) cleavage of type IV collagen by MMP-2/-9 exposes a cryptic site in the ECM, which displays affinity for avb3 integrin, leading to enhancement of
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
539
angiogenesis (Xu et al. 2001). (2) MMP-induced release from the cell surface (shedding) of heparin binding epithelial growth factor, insulin-like growth factor, and fibroblast growth factor enhance cell proliferation. On the other hand, release and activation of ECM sequestered TGFb by MMPs can exert either a positive or negative effect on cell proliferation (Akhurst and Derynck 2001). It needs to be emphasized that none of the functions of invasive and metastatic cells are unique to cancer cells since they can be demonstrated in nonmalignant cells during processes such as migration and inflammation. The major difference between physiologic and neoplastic invasion of cells is one of regulation; dissemination of cancer cells is not effectively regulated by the host, whereas in nonmalignant conditions, these processes are generally under tight physiologic control. In addition to the manifold functions of MMPs in cancer, Boire et al. (2005) recently reported that protease-activated receptor-1 (PAR-1) is a MMP-1 receptor that promotes invasion and tumorigenesis of breast cancer cells in vitro and in vivo. MMP-1 derived from stromal fibroblasts functions as a protease agonist of PAR-1, cleaving the receptor at the proper site to generate PAR-1-dependent Ca2+ signaling and tumor cell migration. Only a few MMPs are expressed exclusively by tumor cells; most are also produced by host cells. MMP-7 expression comes the closest to being restricted to tumor cells. MMP-7 is expressed in benign and malignant tumors that arise from glandular epithelium and its secretion is regulated in a polarized fashion. MMP-7 affects cell–cell interactions and controls cell migration by releasing soluble E-cadherin. By shedding the ectodomain of membrane-bound Fas-L, MMP-7 increases apoptosis in normal surrounding cells, cancer cells being themselves refractory (Powell et al. 1999). In contrast, other studies have shown that chronic exposure to MMP-7 in vivo selects for apoptosis-resistant cancer cells, thereby foiling the effect of cytotoxic chemotherapeutic drugs (Vargo-Gogola et al. 2002). Osteoclast-derived MMP-7 has also been shown to contribute to tumor-induced bone resorption (osteolysis), a common event in metastatic breast cancer (Thiolloy et al. 2009). Hillon et al. (2009) recently demonstrated that the MMP-2 gene is a downstream target that is upregulated by HMGA1 in large cell lung carcinomas. In chromatin immunoprecipitation experiments, HMGA1 binds directly to the MMP-2 promoter in these lung cancer cells and is required for the transformation phenotype in these cells. In a study by Qian et al. (2002), irradiation of tumor cells resulted in enhanced cell invasive potential due to increased MMP production that could be blocked by the use of a matrix metalloproteinase inhibitor. It has been assumed that these MMP enhancement effects of radiation therapy would be transient, as many of the tumor cells would subsequently undergo apoptosis as a result of the radiotherapy.
Gene Expression Signatures in Cancer Recent emphasis has been placed on revealing the gene expression signature of primary tumors and their metastases, followed by clinical correlations. Using an
540
S. Zucker and J. Cao
approach to identify genes (transcriptomic analysis) expressed in breast cancer cells that selectively mediate lung metastasis, Minn et al. (2005) reported that MMP-1 was included in the short list of implicated genes (also inhibitor of differentiation 1, VCAM1, and CXCL1); MMP-2 was expressed only in rare, virulent metastatic cells selected in vivo. Subsequent experimental studies demonstrated that silencing (shRNA) MMP-1, MMP-2, COX2, or EREG in different combinations delayed tumor progression, with nearly complete abrogation of tumor growth following silencing all four genes simultaneously. Of interest, addition of a broad-spectrum MMP inhibitor to drugs inhibiting COX2 and EREG was not additive, in terms of inhibition of cancer progression (Gupta et al. 2007). Several studies have also shown that that loss of specific genes among cancer cells lends to a greater propensity for metastasis. It has been reported that KiSS1, a gene found deleted or rearranged in breast cancer and melanoma, exerts its metastatic suppressive effect by reducing MMP-9 expression (Yan et al. 2001).
Anticancer Effects of MMPs The past decade has been witness to a major change in understanding of MMP function in cancer. Contrary to the dogma, some MMPs exert anticancer effects. Expression of these MMPs, either at the primary or metastatic site, provides a beneficial and protective effect in multiple stages of cancer progression (see recent reviews: Martin and Matrisian 2007; Vincenti and Brinkerhoff 2007). Many of the examples of protective effects of MMPs in cancer come from MMPs that are host (infiltrating inflammatory cells) rather than tumor-derived (Martin and Matrisian 2007). Most of this information has come from genetic studies in which MMPs have either been overexpressed or ablated in murine models of cancer progression. Transgenic mice with knockdown of MMP-8 display an increased incidence and shorter latency of cutaneous papiloma production following experimentallyinduced carcinogenesis (Balbin et al. 2003). Increased expression of MMP-8 has also been associated with cancer metastasis. In a mutational analysis of the MMP gene family in human melanoma, Palavalli et al. (2009) recently identified somatic mutations in 23% of malignant melanomas. Five mutations in one of the most commonly mutated genes, MMP-8, were associated with loss of heterozygosity and resulted in reduced MMP enzyme activity. The capacity of MMPs, especially MMP-12 and MMP-9, to cleave plasminogen and generate angiostatin, which is a powerful inhibitor of tumor angiogenesis in mouse cancer models, further supports the concept that multi-purpose proteases can display antitumor effects (O’Reilly et al. 1994). Overexpression of tumor MMP-12, as determined by Northern blot analysis, in patients with colorectal cancer correlated significantly with increased survival and decreased tumor neovascularization (Yang et al. 2001). However, protumorigenic effects of MMP-12 have also been reported. It has been proposed that the effects of MMP-12 on tumor progression in
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
541
squamous cell carcinomas are dependent upon which cell is expressing MMP-12 (Martin and Matrisian 2007). MMP-3 and MMP-9 are examples of metalloproteinases that can be either protumorigenic or protective in relation to cancer cell growth; speculation for these opposing effects has been proposed (Martin and Matrisian 2007). MMP-9 also can cleave type IV collagen to generate the peptide tumastatin, which is a powerful antiangiogenic agent (Kalluri and Nielson 2003; Martin and Matrisian 2007). Studies in chick embryos have shown that tumor-derived MMP-1, -2, and -9 may have protective functions in metastasis by preventing cancer cell intravasation into the bloodstream (Partridge et al. 2007). These authors pointed out that a deficiency of a single protease appeared unlikely to determine intravasation efficiency, indicating the involvement of multiple factors. MMP-19 displays unique structural features and tissue distribution. MMP-19 is expressed in normal human epidermis and is downregulated during malignant transformation and dedifferentiation. MMP-19 deficiency was associated with an acceleration of the angiogenic response after malignant keratinocyte transplantation. In contrast, MMP-19 knockout mice develop fewer fibrosarcomas and a longer latency period than wild-type mice when treated with a chemical carcinogen. These paradoxical results appear to reflect different roles of MMP-19 at different stages of cancer progression (Noel et al. 2008).
Stromal Cell Production of MMPs: Contribution to Cancer Progression The stromal cells that surround and sustain epithelia have long been viewed primarily as a source of oxygen, nutrients, and growth stimuli for tumors. On the other hand, normal stromal cells may prevent epithelia from becoming tumorigenic. Prior to 1990, it had been assumed that cancer cells were solely responsible for producing MMPs in human tumors. This concept underwent revision when Basset et al. (1990), employing in situ hybridization technology, reported that stromal fibroblasts within tumors, not the tumor cells themselves, were responsible for the production of MMP-11 in human breast cancer. Other studies demonstrated that the localization of MMP-1, -2, -3, -9, and MT1-MMP mRNA was primarily in fibroblasts, especially in proximity to invading cancer cells in breast, colorectal, lung, prostate, ovarian, and head and neck cancer (Nelson et al. 2000; Polette et al. 1996). However, other reports of human tumors described the localization of MMP mRNA in pancreatic, prostatic, and brain carcinoma cells rather than in fibroblasts (Still et al. 2000). Immunohistochemical examinations of human cancer tissues have reported the localization of MMP protein in both cancer and stromal cells. One explanation for the production of MMPs by reactive stromal cells in tumors came from the seminal discovery by Biswas of extracellular matrix metalloproteinase inducer (EMMPRIN). EMMPRIN, a plasma membrane glycoprotein prominently displayed in all epithelial cancer cells, stimulates local fibroblasts to synthesize MMP-1, -2, and -3.
542
S. Zucker and J. Cao
Following secretion of MMP-1 by stromal cells, the protease binds to EMMPRIN on the tumor cell surface, thus arming the cancer cells with invasive factors (Guo et al. 2000). Further support for the importance of EMMPRIN in cancer comes from the observation that cancer cells transfected with EMMPRIN cDNA resulted in marked enhancement of experimental tumor growth (Zucker et al. 2001b). In an experimental model, fibroblasts can support collective invasion by squamous carcinoma cells that are otherwise unable to invade (Gaggioli et al. 2007). Fibroblast-expressed MT1-MMP is essential for this process. Similarly, metastasis of MT1-MMP negative mammary tumors in a genetically induced cancer model is reduced in a MT1-MMP-deficient genetic background in comparison with normal recipient mice (Szabova et al. 2008). These data argue that MT1MMP-mediated ECM degradation by nonmalignant stromal cells promotes cancer metastasis. In vitro imaging of collectively invading co-cultures of carcinoma cells and stromal fibroblasts has reported that the leading cell is always a fibroblast and that carcinoma cells move within tracts in the ECM behind the fibroblast. The generation of these tracts by fibroblasts is sufficient to enable the collective invasion of the squamous carcinoma cells and requires both force-mediated (dependent on Rho, ROCK, myosin light chain, integrins) and protease-mediated matrix remodeling (Gaggioli et al. 2007). A different experimental model of collective cancer cell invasion implicated podoplanin-mediated remodeling of the actin cytoskeleton in an MMP-dependent process (Gaggioli et al. 2007). It should be noted that upregulation of MMPs is one of the physiological changes that occur when fibroblasts undergo senescence. This may be an important component of the generation of a pro-oncogenic tissue environment that contributes to the increased incidence of cancer that occurs with aging (Lui and Hornsby 2007). Emphasis has also been placed on the production of MMPs within tumors by macrophages, neutrophils, mast cells, adipocytes, vascular, and perivascular cells (Noel et al. 2008). Recent studies have demonstrated that adipocytes are highly active endocrine cells that secrete numerous factors including growth factors, cytokines, ECM proteins, and MMPs (Noel et al. 2008).
MMP Involvement in Tumor Angiogenesis Early studies demonstrated that the generation of thrombin by VEGF induction of tissue factor in endothelial cells resulted in pericellular activation of MMP-2 (Zucker et al. 1995; Zucker et al. 1998). Subsequently, MMPs have been demonstrated to be involved physiologically in postnatal vascular refinement, remodeling and neoangiogenesis, but not in the original construction of the embryonic vascular network (Page-McCaw et al. 2007). Current concepts suggest that tumor vasculature is derived from both sprouting of local vessels (angiogenesis) and bone marrow-derived circulating cells (vasculogenesis). MMPs likely contribute to vascular remodeling by proteolysis of types I and IV collagen, modification of platelet-derived growth factor signaling (Lehti et al. 2005), regulation of perivascular cells, and processing of VEGF; MT1-MMP and
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
543
MMP-9 are especially implicated in these processes. After secretion, VEGF is stored in and binds to the ECM, from where it must be released to initiate angiogenesis. MMP processing of VEGF has been identified in tumor angiogenesis. Studies in a landmark mouse pancreatic islet cell carcinoma model have demonstrated that MMP-9 mobilization of matrix-bound VEGF is required to initiate the angiogenic switch, leading to tumor growth and progression (Bergers et al. 2000). Other studies demonstrated that tumors were unable to grow in an irradiated tumor bed in MMP-9 knockout mice, but growth was restored by transplantation of bone marrow from wild-type mice. The explanation is that VEGFR-1 positive, MMP-9 expressing myelomonocytes from the transplanted bone marrow rather than endothelial progenitor cells, were responsible for the development of immature blood vessels in the marrow-restored knockout mice, a prerequisite for tumor growth (Ahn and Brown 2008). Similar to thrombin, MMP-1 also acts as a proangiogenic signaling molecule by proteolytically activating PAR-1 on endothelial cells, thereby facilitating angiogenesis and promoting tumor progression (Blackburn and Brinkerhoff 2008).
Involvement of MMPs in Transition to an Invasive/Metastatic Cancer Phenotype Dissecting the cellular mechanisms involved in transition from an in situ, noninvasive carcinoma to an invasive cancer has been a tremendous challenge. The early explanation that cancer cell secreted MMPs chew up surrounding ECM, leading to unimpeded cell migration, has undergone modifications as new technologies have been introduced. The fact that cell behavior within a matrix composed of natural ECM fibrillar proteins differs from cells proliferating on an artificial surface, has forced scientists to revise the dogma of MMP function in cancer.
Cancer Cell Invasion in a Three-Dimensional Matrix A growing body of literature involving cancer growth and metastasis has underscored the importance of the physical three-dimensionality versus twodimensionality of the ECM in regulating cell behavior. Proteolysis of 3D cross-linked collagen, which surrounds and imprisons cells, facilitates their release. However, it needs to be emphasized that 3D models are experimental tools; they are not intact animals and do not reproduce all of the in vivo microenvironment (Yamada and Cukierman 2007). In that regard, the limitations of cancer models employing laboratory animals are also well recognized in regard to predictions with human cancer. In elegant in vitro studies, Hotary et al. (2003) demonstrated that of a large number of MMPs tested, including MMP-1, -2 and -9, only MT1-MMP, MT2-MMP, and MT3-MMP can serve as direct-acting collagenases that are able to dissolve BM during cell migration and permit vigorous cell proliferation inside a 3D collagen gel
544
S. Zucker and J. Cao
(Sabeh et al. 2009). Yet MT1-MMP production conferred no growth advantage on cells imbedded on a 2D substrate. To date, the specific BM components cleaved by MT1-MMPs in situ remain to be characterized, as does their role in promoting BM transmigration (Rowe and Weiss 2008). MT1-MMP has also been shown by direct visualization of protease activity to accumulate at the invasive front of tumors (Packard et al. 2009) and are enriched at invadopodia, which are small dot shaped, specialized actin-based membrane protrusions that are prominent on invasive tumor and transformed cells grown on ECM (Poincloux et al. 2009). Figure 25.3 provides an example of phenotypic alterations resulting from overexpression of MT1-MMP
Fig. 25.3 MT1-MMP induces morphologic changes in cancer cells cultivated in type I collagen gels (three-dimensional), but not in cells cultured under 2D conditions. (a) No morphologic differences were observed between LNCaP (prostate cancer) cells, LNCaP cells expressing green fluorescent protein (GFP), and LNCaP cells expressing MT1-GFP chimeric cDNA under 2D culture conditions: Parental LNCaP cells, LNCaP cells stably expressing GFP or expressing MT1GFP chimeric cDNAs, were cultured on top of type I collagen coated dishes for 3 days, followed by phase contrast and fluorescent microscopic examination. As noted, GFP is diffusely distributed throughout the GFP-transfected LNCaP cells, whereas fluorescence is focalized on the cell surface of MT1-GFP-transfected cells. Bar, 20 mm. (b) Induction of cell morphologic changes (change from epitheliod to fibroblast-like) and cell scattering by expression of MT1-MMP in LNCaP cells cultured under 3D conditions: LNCaP cells (4 × 104/ml) stably expressing GFP or MT1-GFP chimeric cDNA were mixed and cultured within neutralized type I collagen gels (2.5 mg/ml). The cells were examined daily under fluorescent microscope for 9 days. At day 6, a set of gels were also fixed and frozen sections were prepared for hematoxylin/eosin staining. Bar, 20 mm
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
545
in prostate cancer (LNCaP) cells embedded within a fibrillar type I collagen gel. Opposing views have argued that pericellular collagenolysis is restricted to the posterior region of the leading pseudopodium, thereby enabling adhesive interactions with the ECM to drive efficient 3D cell migration (Wolf et al. 2007). Cortactin, an actin binding protein, has been proposed as an essential regulator of MMPs (MMP-2, -9, and MT1-MMP) involved in invadopodia-induced ECM degradation (Clark et al. 2007). Understanding the protrusive activity of invasive cancer cells demands the integration of factors required for chemotaxis involving genes coding for cofilin, capping protein, and Arp2/3 complex pathways leading to actin polymerization (Yamagucki et al. 2005). MMP-9 has also been shown to be localized to invadopodia; WASP is involved in the regulation of MMP-9 in invadopodia (Desai et al. 2008).
Protease-Independent Cell Invasion: Fact or Fantacy An alternative concept of cancer cell invasion arose following in vitro experiments demonstrating that total inhibition of all proteases, including MMPs, resulted in an efficient amoeboid mode of cell migration through collagen-lined pores, characterized by rounded cell morphology with no obvious cell polarity, driven by actin/myosin contractility (Sanz-Moreno et al. 2008; Wolf et al. 2003). Amoeboid migration is characterized by enhanced RhoA-, ROCK- and myosin II-dependent mechanical contractility, which allows cells to squeeze through gaps in 3D matrix. Rac-controlled interconversion between mesenchymal-type migration and RhoA- and ROCK dependent migration that is similar to amoeboid movement has been reported for tumor cells invading through 3D Matrigel, a simulated basement membrane matrix with high laminin content (Sanz-Moreno et al. 2008). As pointed out by other investigators (Packard et al. 2009; Sabeh et al. 2009), amoeboid motility experiments were carried out using weakly cross-linked reconstituted networks formed with pepsin-extracted, as opposed to acid-extracted (tightly cross-linked) type I collagen or employing Matrigel. Although based on fascinating in vitro experiments, the amoeboid motility theory remains to be proven relevant to cancer (Rowe and Weiss 2008).
Epithelial-to-Mesenchymal Transition in Cancer Most human cancers are epithelial in origin. Loss of integrity and change of phenotype of the epithelium are critical for initiation of tumor invasion and metastasis. Based on examination of human tumor samples, Brabletz et al. (2001) proposed that a driving force for progression of well-differentiated carcinomas is the specific tumor microenvironment, initiating two phenotypic transition processes: first epithelial-to-mesenchymal transition (EMT) dedifferentiation of carcinoma cells at the invasive front, favoring detachment, migration and dissemination, and subsequently
546
S. Zucker and J. Cao
redifferentiation with regain of epithelial capabilities (MET) which is necessary for proliferation at metastatic sites (Kang and Massague 2004; Thompson and Newgreen 2005). Turning an epithelial cell into a mesenchymal cell requires alterations in morphology, cellular architecture, adhesion, and migration. Loss/decrease of E-cadherin (hallmark of EMT), cytokeratins and acquisition of mesenchymal proteins like fibronectin, vimentin, and N-cadherin denotes a loss of the epithelial phenotype and switch toward a mesenchymal dedifferentiated phenotype. The multiplicity of distinct pathways and molecules, each of which can lead independently to EMT-like events, is astonishing. Several developmentally important genes that induce EMT have been shown to act as E-cadherin repressors. Included in this list are: Snail, Slug, the zinc finger protein SIPI (ZEB 2), E12/E47 [12], and Twist, a basic helixloop-helix transcription factor that initiates mesoderm development during gastrulation (Thiery and Sleeman 2006; Thompson and Newgreen 2005; Yang et al. 2004). These EMT-inducing factors bind DNA using similar E-box sequence motifs. Radisky et al. (2005) have reported that MMP-3 induces EMT and genomic instability in an experimental model through a signaling pathway involving Rac1b, followed by the generation of reactive oxygen species. A vexing issue in deciphering metastasis is that only a small minority of carcinoma cells in the primary tumors may undergo EMT and metastasize, thus alteration of gene expression in such cells can be masked by the bulk of nonmetastatic cells. Hence, the transcriptomic contribution of such a population would be diluted by the whole. Overexpression of MMP-1, -2, -3, -7, -9, 13, -28, and MT1-MMP has been associated with EMT. The E-cadherin-repressed hNanos1 gene has been demonstrated to induce tumor invasion by upregulation of MT1-MMP expression; hNanos is also overexpressed in aggressive human lung cancers (Bonnomet et al. 2008). Cao et al. (2008) recently demonstrated that expression of MT1-MMP in less aggressive epithelial cancer cells resulted in EMT-like phenotypic changes. Antagonism of canonical and noncanonical (Wnt5a) pathways (Cao et al. 2008) has been associated with the reversal of EMT, including decreased expression and activity of MMP-2 and MMP-9, decreased TCF transcription activity, and decreased expression of Twist and Slug (Zi et al. 2005). Snail-deficient fibroblasts exhibit global alterations in gene expression, including defects in MT1-MMP-dependent invasion and defects in angiogenesis (Rowe et al. 2009).
Premetastatic Niche An intriguing body of evidence suggest that certain primary tumor cells secret soluble factors that induce a specific population of nonmalignant hematopoietic cells to mobilize and engraft in distant organ tissues, thereby establishing a “premetastatic niche” that lay a foundation for incoming circulating cancer cells (Kaplan et al. 2006). This process includes proteolytic matrix turnover and secretion of soluble
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
547
growth factors and chemokines that create a permissive microenvironment for circulating cancer cells. Primary tumors release VEGF, TGF-b, and TNFa that in turn induce the expression of chemoattractants (S100A8 and S100A9) by lung endothelium and Mac1+ myeloid cells, thereby facilitating the homing of tumor cells to the premetastatic niche within lung parenchyma. S100A8 and S100A9 also increase the motility of circulating cancer cells by p38-mediated activation of invadopodia, which facilitates assembly of the metastatic focus (Hirasuka et al. 2006). MMP-9 induction by VEGFR-1 is involved in lung specific metastasis. Primary tumor cells also release soluble factors that induce a specific subset of bone marrow derived cells, which express VEGFR-1 VLA4+ (myelomonocytes), to colonize the target organ before the arrival of tumor cells. Recruitment of these hematopoietic cells is simultaneously associated with increased deposition of fibronectin in the lung, presumably resulting in the formation of a fertile habitat for adhesion of circulating tumor cells (Kaplan et al. 2006). Lysyl oxidase (LOX), an amine oxidase that cross links collagens and elastins in the ECM, is critical for the formation of premetastatic niche. LOX, secreted by hypoxic breast cancer cells, accumulates at premetastatic sites, cross links collagen IV in the BM, and is essential for Mac-1+ (myeloid-derived suppressor) cell recruitment. These cells adhere to cross-linked collagen IV and produce MMP-2, which cleaves collagen, enhancing invasion and further recruitment of bone marrow-derived cells (Erler et al. 2009). Generation of chemo-attractive collagen IV peptides is responsible, in part, for the enhanced cell invasion.
Inflammation and Cancer: Role of MMPs In many tumors, inflammation contributes to the progression to malignancy (Coussens and Werb 2002). Classic examples are cancer developing in a tubercular scar in the lung and in long-standing colitis in the colon. In mammary tumors, the microenvironment for metastasis depends heavily on macrophages, which provide signals for angiogenesis, tumor growth, tumor cell chemotaxis, and invasion. The communication between macrophages and cancer cells in experimental animals is paracrine, involving the production of EGF by macrophages and CSF-1 by tumor cells to stimulate cancer and macrophage chemotaxis and invasion, respectively. Chemotaxis in response to EGF is regulated by the cofilin pathway involving a balance between cofilin and LIMK activities (Wang et al. 2007). Chronic inflammation has also been proposed as a causative factor in prostate cancer.
MMPs as Therapeutic Targets in Cancer The decade of the 1990s was witness to a frenzy of activity in pharmaceutical and biotechnology companies to discover new forms of treatment for cancer. MMPs, especially those capable of cleaving type IV basement membrane collagen (MMP-2 and -9),
548
S. Zucker and J. Cao
were considered to be ideal targets for drug development (Overall and Kleifeld 2006b; Zucker et al. 2001a), Chemists achieved considerable success in designing numerous MMP inhibitors (MMPIs) that are orally active, achieve effective blood levels with twice-a-day dosing in healthy volunteers, and display specificity for metalloproteinases but not limited to MMPs, while generally sparing most other types of proteases (Pavlaki and Zucker 2003). Most synthetic MMP inhibitors were designed to consist of two parts: a zincbinding group to bind the catalytic metal ion and a peptidomitic backbone to interact noncovalently with specific subsites neighboring the active site of the MMP protein. The catalytic Zn2+ in the active site is surrounded by subsite pockets designated as S1, S2, S3, S1¢, S2¢, and S3¢. Of the different subsite pockets, targeting of the S1¢ pocket provided the basis of selectivity for many MMP inhibitors (Agrawal et al. 2008). These drugs essentially mimicked the major collagen substrate of MMPs, and thereby work as competitive potent, reversible inhibitors of enzyme activity. Subsequently, MMP inhibitor design became more structure-based, due to the abundance of NMR and X-ray structural data. A growing trend in the development of synthetic MMP inhibitors has been to target not only the enzyme catalytic site but also distal surface residues residing away from the catalytic machinery, as well as the other MMP accessory domains (see review by Sela-Passwell et al. 2010). It has recently been demonstrated that more selective MMP inhibitors can be designed based on the nature of their zinc binding group (Agrawal et al. 2008). A new generation of covalent-binding, mechanism-based MMP-2 inhibitors has also been designed (Ikejiri et al. 2006). Initial nonrandomized clinical trials, reporting impressive tumor regression, led to MMP inhibitors being tested in numerous Phase II–III trials involving thousands of patients with various cancers (lung, stomach, colon, pancreas, ovary, brain, prostate, breast); unfortunately none of these clinical trials provided positive results (Zucker et al. 2001a). The lack of efficacy of broad-spectrum MMP inhibitors has been attributed to the following (1) some MMPs display antitumor activities (MMP-8 and MMP-12). Hence, current recommendations favor development of MMP inhibitors directed at specific, well-defined MMPs (Overall and Kleifeld 2006a); (2) MMP inhibitors appear to be more active in early, rather than late stage cancer; (3) drug intolerance severely reduced drug compliance (Zucker et al. 2001b); (4) inadequate blood levels of drug in cancer patients (Sparano et al. 2004) as opposed to healthy volunteers; and (5) even specific MMP inhibition can result in unpredictable induction of systemic protease web-associated modulations, which can comprise metastasis-promoting molecules such as other proteases and cytokines (Kruger et al. 2010). The basis of future drug design needs to take into account that MMPs promote tumor progression not only through ECM degradation as originally recognized, but also through cleavage of many non-ECM proteins (Overall and Kleifeld 2006b; Overall et al. 2004). Today’s challenge is to distinguish the action of MMPs that contribute to tumor progression from those that are crucial for host defense, as blocking the latter will worsen clinical outcome. Recently, Devy et al. (2009) have reported that a highly selective and potent fully human MT1-MMP inhibitory antibody markedly slowed tumor progression/
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
549
metastasis and inhibited angiogenesis in mice with xenogenic human cancer implants. This antibody apparently reacts only with activated MT1-MMP, not the latent enzyme. Concurrent administration of MT1-MMP antibody and chemotherapy or MT1-MMP antibody and anti-VEGF antibody resulted in additive antitumor activity. Commercial development of this antibody has been initiated with the goal of clinical trial(s) in cancer.
Exocyte Binding and Alosteric Inhibitors of MMPs Based on limitations of broad-spectrum catalytic domain-directed MMP inhibitors in clinical trials, there is interest in developing drugs that target other domains of procancerous MMPs. It has been proposed that specificity might be achieved with substrate-specific exosite inhibitors. However, the binding affinity of exosite domains for substrates is typically low on featureless sites, rendering it potentially difficult to develop compounds that bind here (Overall and Kleifeld 2006b). The PEX domains of MMPs share less homology than the propeptide and catalytic domains, suggesting that unique binding properties of these PEX domains may be exploited to affect protease function. The MMP-9 PEX domain has been used as a drug target to inhibit various pathologies including neural injuries, angiogenesis, and cancer metastasis. Peptides identified from a random phage library targeting the MMP-9 catalytic domain, collagen binding domain, and PEX domain binding to integrins have been shown to inhibit cell migration in vitro and tumor xenografts in vivo (Bjorklund et al. 2004). Radjabi et al. (2008) used both a cyclic peptide and an antibody to the MMP-9 PEX domain to interfere with MMP-9/ β1-integrin in an osteosarcoma cancer cell line (Radjabi et al. 2008). Purified MMP-9 PEX domain has also been employed as a reagent to inhibit angiogenesis and retard glioblastoma growth in vivo (Ezhilarasan et al. 2009). Another subtle approach to inhibit MMP function is to target interactions with critical molecules that bind to the PEX domain of MMPs. CD44 is a prominent cell surface protein that serves as a docking molecule for the PEX domain of MMP-9 and MT1-MMP (Samanna et al. 2007; Yu and Stamenkovic 1999). Complex formation between CD44 and blade I of the PEX domain of proMMP-9 has been reported to be critical for protease-independent cell migration. Dufour et al. (2010) recently demonstrated that short peptides (8 amino acids) that mimic PEX propeller binding to CD44 and homodimer formation are able to specifically inhibit MMP-9enhanced cell migration.
RNA Interference (RNAi) Technology to Target MMPs in Cancer A targeted approach with potential to interfere with MMPs in cancer lies in RNAi technology, in which expression of a single gene of interest is knocked-down by
550
S. Zucker and J. Cao
short double-stranded RNAs directed at mRNA (see review by Vincenti and Brinkerhoff 2007). Experimental mouse models have used siRNAs directed against MMP-1, -2, and -9 to decrease tumor initiation and progression, and prevent metastasis. Although the technology is evolving rapidly, substantial difficulties involving stability and site-specific delivery remain. Application of siRNA technology to human disease remains a major challenge. Acknowledgments This work support was provided by a Merit Review Grant from the Department of Veterans Affairs, NIH grant (RO1 CA11355301A1), a Baldwin Breast Cancer Foundation grant and a Walk-for-Beauty grant from the Research Foundation, Stony Brook University.
References Agrawal A et al (2008) Zinc-binding groups modulate selective inhibition of MMPs. Chem Med Chem 3:812–820 Ahn G-O, Brown JM (2008) Matrix metalloproteinase-9 is required for tumor vasculogenesis but not for angiogenesis: role of bone marrow-derived myelomonocytes. Cancer Cell 13:193–205 Akhurst RJ, Derynck R (2001) TGF-beta signaling in cancer – a double-edged sword. Trends Cell Biol 11:S44–S51 Baker AH, Edwards DR, Murphy G (2002) Metalloproteinase inhibitors: biologic actions and therapeutic opportunities. J Cell Sci 115:3719–3727 Balbin M et al (2003) Loss of collagenase-2 confers increased skin tumor susceptibility to male mice. Nat Genet 35:252–257 Basset P et al (1990) A novel metalloproteinase gene specifically expressed in stromal cells of breast carcinomas. Nature 348:699–704 Bergers G et al (2000) Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat Cell Biol 2:737–744 Birkedal-Hansen H (1995) Proteolytic remodeling of extracellular matrix. Curr Opin Cell Biol 7:728–735 Bjorklund M, Heikkila P, Koivunen E (2004) Peptide inhibition of catalytic and noncatalytic activities of matrix metaalloproteinase-9 blocks tumor cell migration and invasion. J Biol Chem 279:29589–29597 Blackburn JS, Brinkerhoff CE (2008) Matrix metalloproteinase-1 and thrombin differentially activate gene expression in endothelial cells via PAR-1 and promote angiogenesis. Am J Pathol 173:1736–1746 Boire A et al (2005) PAR-1 is a matrix metalloproteinase-1 receptor that promotes invasion and tumorigenesis of breast cancer cells. Cell 120:303–313 Bonnomet A et al (2008) The E-cadherin-repressed hNanos1 gene induces tumor cell invasion by upregulating MT1-MMP expression. Oncogene 27:3692–3699 Brabletz T et al (2001) Variable beta-catenin expression in colorectal cancers indicates tumor progeression driven by the microenvironment. Proc Natl Acad Sci U S A 98:10356–10361 Butler GS et al (1998) The TIMP-2 membrane type 1 metalloproteinase “receptor” regulates the concentration and efficiency of progelatinase A. A kinetic study. J Biol Chem 273: 871–880 Cao J et al (2008) Membrane type I matrix metalloproteinase induces epithelial-to-mesenchymal transition in prostate cancer. J Biol Chem 283:6232–6240 Cao J et al (1998) The propeptide domain of membrane type I matrix metalloproteinase is required for binding of tissue inhibitor of metalloproteinases and for activation of pro-gelatinase A. J Biol Chem 273:34745–34752
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
551
Clark ES et al (2007) Cortactin is an essential regulator of matrix metalloproteinase secretion and extracellular matrix degradation in invadopodia. Cancer Res 67:4227–4235 Coussens L, Fingleton B, Matrisian L (2002) Matrix metalloproteinase inhibitors and cancer: Trials and tribulations. Science 295:2387–2392 Coussens LM, Werb Z (2002) Inflammation and cancer. Nature 420:860–867 Dean RA, Overall CM (2007) Proteomics discovery of metalloproteinase substrates in the cellular context by iTRACTM labeling reveals a diverse MMP-2 substrate degradome. Mol Cell Proteomics 6:611–621 Della Porta P et al (1999) Combined treatment with serine protease inhibitor aprotonin and matrix metalloproteinase inhibitor Batimastat (BB-94) does not prevent invasion of human esophageal and ovarian carcinoma cells in vitro. Anticancer Res 19:3809–3816 Desai B, Ma T, Chellaiah MA (2008) Invadopodia and matrix degradation: a new property of prostate cancer cells during migration and invasion. J Biol Chem 283:13856–13866 Devy L et al (2009) Selective inhibitor of matrix metalloproteinase-14 blocks tumor growth, invasion, and angiogenesis. Cancer Res 69: 1517–1524 Dufour A et al (2010) Critical role of MMP-9 dimers in cell migration (submitted for publication) Dufour A et al (2008) Role of the hemopexin domain of matrix metalloproteinases in cell migration. J Cell Physiol 217:643–651 Erler JT et al (2009) Hypoxia-induce lysyl oxidas is a critical mediator of bone marrow cell recruitment to form the premetastatic niche. Cancer Cell 15:35–44 Ezhilarasan E et al (2009) The hemopexin domain of MMP-9 inhibits angiogenesis and retards growth of intracranial glioblastoma xenografts in nude mice. Int J Cancer 124:306–315 Fu X, Parks WC, Heinecke JW (2008) Activation and silencing of matrix metalloproteinases. Cell Dev Biol 19:2–13 Gaggioli C et al (2007) Fibroblast-led collective invasion of carcinoma cells with differing roles for RhoGTPases in leading and following cells. Nat Cell Biol 9:1392–1400 Goldberg GI et al (1989) Human 72-kilodalton type IV collagenase forms a complex with a tissue inhibitor of metalloproteinases designated TIMP-2. Proc Natl Acad Sci U S A 86: 8207–8211 Gross J, Lapiere CM (1962) Collagenolytic activity in amphibian tissues; a tissue culture assay. Proc Natl Acad Sci U S A 48:1014–1022 Guo H et al (2000) EMMPRIN (CD147), an inducer of matrix metalloproteinase synthesis, also binds interstitial collagenase to the tumor cell surface. Cancer Res 60:888–891 Gupta GP et al (2007) Mediators of vascular remodelling co-opted for sequential steps in lung metastasis. Nature 446:765–770 Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57–70 Hillon J et al (2009) Upregulation of MMP-2 by HMGA1 promotes transformation in undifferentiated, large-cell lung cancer. Mol Cancer Res 7:1803–1812 Hirasuka S et al (2006) Tumor-mediated upregulation of chemoattractants and recruitment of myeloid cells predetermines lung metastasis. Nat Cell Biol 8:1369–1375 Hotary KB et al (2003) Membrane type I matrix metalloproteinase usurps tumor growth control imposed by the three-dimensional extracellular matrix. Cell 114:33–45 Howard EW, Bullen EC, Banda MJ (1991) Regulation of the autoactivation of human 72-kDa progelatinase by tissue inhibitor of metalloproteinases-2. J Biol Chem 266:13064–13069 Ikejiri M et al (2006) Design, synthesis, and evaluation of a mechanism-based inhibitor for gelatinase A. J Org Chem 70:5709–5712 Itoh Y, Seiki M (2006) MT1-MMP: a potent modifier of pericellular microenvironment. J Cell Physiol 206:1–8 Kalluri R, Nielson EG (2003) Epithelial–mesenchymal transition and its implications for fibrosis. J Clin Invest 112:1776–1784 Kang Y, Massague J (2004) Epithelial–mesenchymal transitions: twist in development and metastasis. Cell 118:277–279 Kaplan RN, Rafi S, Lyden D (2006) Preparing the soil: the premetastatic niche. Cancer Res 66:11089–11093
552
S. Zucker and J. Cao
Kruger A, Kates RE, Edwards DR (2010) Avoiding spam in the proteolytic internet: Future strategies for anti-metastatic MMP-inhibition. Biochim Biophys Acta 1803:95–102 Lehti K et al (2005) An MT-MMP-PDGF receptor-beta axis regulates mural cell investment of themicrovasculature. Genes Dev 19:979–991 Liotta LA et al (1980) Metastatic potential correlates with enzymatic degradation of basement membrane collogen. Nature 284:67–68 Lopez-Otin C, Bond JS (2008) Proteases: multifunctional enzymes in life and disease. J Biol Chem 283:30433–30437 Lui D, Hornsby PJ (2007) Senescent human fibroblasts increase the early growth of xenograft tumors via matrix metalloproteinase secretion. Cancer Res 67:3117–3126 Maquoi E et al (2000) Membrane type I matrix metalloproteinase-associated degradation of tissue inhibitor of metalloproteinase 2 in human tumor cell lines. J Biol Chem 275:11368–11378 Martin MD, Matrisian LM (2007) The other side of MMPs: protective roles in tumor progression. Cancer Metastasis Rev 26:717–724 Mazzone M et al (2004) Intracellular processing and activation of membrane type 1-matrix metalloprotease depends on its partitioning into lipid domains. J Cell Sci 117:6275–6287 McQuibban GA et al (2001) Matrix metalloproteinase activity inactivates the CXC chemokine stromal cell-derived factor-1. J Biol Chem 276:43503–43508 Minn AJ et al (2005) Genes that mediate breast cancer metastasis to lung. Nature 436:518–524 Mochizuki S, Okada Y (2007) ADAMs in cancer cell proliferation and progression. Cancer Sci 98:621–628 Morgunova E et al (2002) Structural insight into the complex formation of latent matrix metalloproteinase 2 with tissue inhibitor of metalloproteinase 2. Proc Natl Acad Sci 99:7414–7419 Nagase H, Woessner F (1999) Matrix metalloproteinases. J Biol Chem 274:21491–21494 Nelson AR et al (2000) Matrix metalloproteinases: biologic activity and clinical implications. J Clin Oncol 18:1135–1139 Noel A et al (1997) Emerging roles for proteinases in cancer. Invasion Metastasis 17:221–239 Noel A, Jost M, Maquoi E (2008) Matrix metalloproteinases at cancer tumor–host interface. Semin Cell Dev Biol 19:52–60 O’Reilly MS et al (1994) Angiostatin: a novel angiogenesis inhibitor that mediates the suppression of metastasis by a Lewis lung carcinoma. Cell 79:315–328 Overall CM, Kleifeld O (2006a) Toward third generation matrix metalloproteinase inhibitors for cancer research. Br J Cancer 94:941–946 Overall CM, Kleifeld O (2006b) Validating matrix metalloproteinases as drug targets and antitargets for cancer therapy. Nat Rev Cancer 6:227–239 Overall CM et al (2004) Protease degradomics: mass spectroscopy discovery of protease substrates and the CLIP-CHIP, a dedicated DNA microarray of all human proteases and inhibitors. Biol Chem 385:493–504 Packard BZ et al (2009) Direct visualization of protease activity on cells migrating in threedimensions. Matrix Biol 28:3–10 Page-McCaw A, Ewald AJ, Werb Z (2007) Matrix metalloproteinases and the regulation of tissue remodeling. Nat Rev Mol Cell Biol 8:221–233 Palavalli LH et al (2009) Analysis of the matrix metalloproteinase family reveals that MMP8 is often mutated in melanoma. Nat Genet 41:518–520 Partridge JJ et al (2007) Functional analysis of MMPs and TIMPs differentially expressed by variants of human HT-1080 fibrosarcoma exhibiting high and low levels of intravasation and metastasis. J Biol Chem Pavlaki M, Zucker S (2003) Matrix metalloproteinase inhibitors (MMPIs): the beginning of phase I or the termination of phase III clinical trials. Cancer Metastasis Rev 22:177–203 Poincloux R, Lizarraga F, Chavrier P (2009) Matrix invasion by tumour cells: a focus on MT1MMP trafficking to invadopodia. J Cell Sci 122:3015–3024 Polette M et al (1996) MT-MMP expression and localization in human lung and breast cancer. Virchows Arch 428:29–35
25 Matrix Metalloproteinases and Cancer Cell Invasion/Metastasis
553
Powell WC et al (1999) The metalloproteinase matrilysin proteolytically generates active soluble Fas ligand and potentiates epithelial cell apoptosis. Curr Biol 9:1441–1447 Qian LW et al (2002) Radiation induced increase in invasive potential of human pancreatic cancer cells and its blockade by matrix metalloproteinase inhibitor. Clin Cancer Res 8:1223–1227 Radisky DC et al (2005) Rac1b and reactive oxygen species mediate MMP-3 induced EMT and genomic instability. Nature 463:123–127 Radjabi A et al (2008) Thrombin induces tumor invasion through the induction and association of matrix metalloproteinase-9 and beta1-integrin on the cell surface. J Biol Chem 283:2822–2834 Rowe RG et al (2009) Mesenchymal cells reactivate Snail 1 expression to drive three-dimensional invasion programs. J Cell Biol 184:399–408 Rowe RG, Weiss SJ (2008) Breaching the basement membrane: who, when and how? Trends Cell Biol 18:560–574 Sabeh F, Shimizu-Hirota R, Weiss SJ (2009) Protease-dependent versus-independent cancer invasion programs: three dimensional amoeboid movement revisited. J Cell Biol 185:11–19 Samanna V et al (2007) Actin polymerization modulates CD44 surface expression, MMP-9 activation, and osteoclast function. J Cell Physiol 213:710–720 Sanz-Moreno V et al (2008) Rac activation and inactivation control plasticity of tumor cell movement. Cell 135:510–523 Sato H et al (1994) A matrix metalloproteinase expressed on the surface of invasive tumor cells. Nature 370:61–65 Seiki M (2002) The cell surface: the stage for matrix metalloproteinase regulation of migration. Curr Opin Cell Biol 14:624–632 Sela-Passwell N et al (2010) Structural and functional base for allosteric control of MMP activities. Biochim Biophys Acta 1803:29–38 Sparano JA et al (2004) Randomized phase III trial of marimastat versus placebo in patients with metastatic breast cancer who have responding or stable disease after first-line chemotherapy: An Eastern Cooperative Oncology Group trial (E2196). J Clin Oncol 22:4631–4638 Sternlicht MD, Werb Z (2001) How matrix metalloproteinases regulate cell behavior. Annu Rev Cell Dev Biol 17:463–516 Still K et al (2000) Localization and quantification of mRNA for matrix metalloproteinase-2 (MMP-2) and tissue inhibitor of matrix metalloproteinase-2 (TIMP-2) in human benign and malignant prostatic tissue. Prostate 42:18–25 Stoker W, Bode W (1995) Structural features of a superfamily of zinc-endopeptidases: the metzincins. Curr Opin Struct Biol 5:383–390 Strongin AY et al (1995) Mechanism of cell surface activation of 72-kDa type IV collagenase. Isolation of the activated form of the membrane metalloproteinase. J Biol Chem 270:5331–5338 Suenaga N et al (2004) CD44 binding through the hemopexin-like domain is critical for its shedding by membrane-type 1 matrix metalloproteinase. Oncogene 1–10 Szabova L et al (2008) MT1-MMP is required for efficient tumor dissemination in experimental metastatic disease. Oncogene 27:3274–3281 Takahashi C et al (1998) Regulation of matrix metalloproteinase-9 and inhibition of tumor invasion by the membrane-anchored glycoprotein RECK. Proc Natl Acad Sci U S A 95:13221–13226 Tallant C, Marrero A, Gomis-Ruth FX (2010) Matrix metalloproteinases: fold and function of their catalytic domains. Biochim Biophys Acta 1803:20–28 Thiery JP, Sleeman JP (2006) Complex networks orchestrate epithelial–mesenchymal transitions. Nat Rev Mol Cell Biol 7:131–142 Thiolloy S et al (2009) Osteoclast-derived matrix metalloproteinase-7, but not matrix metalloproteinase-9, contributes to tumor-induced osteolysis. Cancer Res 69:6747–6755 Thompson EW, Newgreen DF (2005) Carcinoma invasion and metastasis: a role for epithelial– mesenchymal transition? Cancer Res 65:5991–6001
554
S. Zucker and J. Cao
Vargo-Gogola T et al (2002) Matrilysin (matrix metalloproteinase-7) selects for apoptosis-resistant mammary cells in vivo. Cancer Res 62:5559–5563 Vincenti MP, Brinkerhoff CE (2007) Signal transduction and cell-type specific regulation of matrix metalloproteinase gene expression: can MMPs be good for you? J Cell Physiol 213:355–363 Wang W, Eddy R, Condeelis J (2007) The cofilin pathway in breast cancer invasion. Nat Rev 7:429–440 Wolf K et al (2003) Compensation mechanism in tumor cell migration: mesenchymal-amoeboid transition after blocking of pericellular proteolysis. J Cell Biol 160:267–277 Wolf K et al (2007) Multi-step pericellular proteolysis controls the transition from individual to collective cancer cell invasion. Nat Cell Biol 9:893–904 Xu J et al (2001) Proteolytic exposure of a cryptic site within collagen type IV is required for angiogenesis and tumor growth in vivo. J Cell Biol 154:1069–1079 Yamada KM, Cukierman E (2007) Modeling tissue morphogenesis and cancer in 3D. Cell 130:601–610 Yamagucki H, Wyckoff J, Condeelis J (2005) Cell migration in tumors. Curr Opin Cell Biol 17:559–564 Yan C, Wang H, Boyd DD (2001) KiSS-1 represses 92-kDa type IV collagenase expression by down-regulating NFkB binding to the promoter as a consequence of Ik B alpha induced block of p65/p50 nuclear translocation. J Biol Chem 276:1162–1172 Yang J et al (2004) Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 117:927–939 Yang W et al (2001) Human macrophage metalloelastase gene expression in colorectal carcinoma and its clinicopathologic significance. Cancer 91:1277–1283 Yu Q, Stamenkovic I (1999) Localization of matrix metalloproteinase 9 to the cell surface provides a mechanism for CD44-mediated tumor invasion. Genes Dev 13:35–48 Zi X et al (2005) Expression of Frzb/secreted frizzled-related protein 3, a secreted Wnt antagonist, in human androgen-independent prostate cancer PC-3 cells suppresses tumor growth and cellular invasion. Cancer Res 65:9762–9770 Zucker S, Cao J, Chen W-T (2001a) Critical appraisal of the use of matrix metalloproteinase inhibitors in cancer treatment. Oncogene 19:6642–6650 Zucker S et al (2001b) Tumorigenic potential of extracellular matrix metalloproteinase induce (EMMPRIN). Am J Pathol 158:1921–1928 Zucker S, Cao J, Molloy CJ (2002) Role of matrix metalloproteinases and plasminogen activators in cancer and metastasis. Therapeutic strategies. In: Baguley BC, Kerr DJ (eds) Anticancer drug development. Academic, San Diego, CA, pp 91–122 Zucker S et al (1995) Thrombin induces the activation of progelatinase A in vascular endothelial cells: Physiologic regulation of angiogenesis. J Biol Chem 270:23730–23738 Zucker S, Doshi K, Cao J (2004a) Measurement of matrix metalloproteinases (MMPs) and tissue inhibitors of metalloproteinases (TIMP) in blood and urine: potential clinical applications. Adv Clin Chem 38:37–85 Zucker S et al (2004b) TIMP-2 is released as an intact molecule following binding to MT1-MMP on the cell surface. Exp Cell Res 293:164–174 Zucker S et al (1985) Diversity of melanoma plasma membrane proteinases. Inhibition of collagenolysis and cytolytic activity by minocycline. J Natl Cancer Inst 75:517–525 Zucker S et al (1992) Type IV collagenase/gelatinase (MMP2) is not increased in plasma of patients with cancer. Cancer Epidemiol Biomarkers Prev 1:475–479 Zucker S et al (1998) Vascular endothelial growth factor and matrix metalloproteinase production in endothelial cells: conversion of prothrombin to thrombin results in progelatinase A activation and cell proliferation. Int J Cancer 75:780–786
Chapter 26
Tetraspanins and Cancer Metastasis Margot Zöller
Abstract Metastasis formation is the final result of a cascade of events that primary tumor cells pass through by changing their phenotype and the crosstalk with the tumor environment. Molecules involved in this process are besides others tetraspanins, which surprisingly can either inhibit or promote metastasis formation. These opposing activities are supposed to rely on the special feature of tetraspanins that mostly act via modulating the activity of a multitude of associating molecules. Tetraspanins assemble a web between themselves and other associating molecules in special glycolipid-enriched membrane microdomains, which function as signaling platform, but are also prone for internalization. Internalization of tetraspanins and associated molecules by itself can contribute to promotion or inhibition of tumor progression. Notably, the internalized tetraspanin web is abundantly recovered in exosomes, small vesicles that derive from internalized membrane microdomains. Thus, it appears reasonable to assume that exosomal tetraspanins are of major importance for the crosstalk between the metastasizing tumor cell, the tumor stroma, the vessel endothelium, and the premetastatic organ. I will briefly introduce the structure of tetraspanins and their presently known main functional activities as a starting point to appreciate the contribution of selective tetraspanins in metastasis promotion and inhibition.
M. Zöller (*) Department of Tumor Cell Biology, University Hospital of Surgery, Im Neuenheimer Feld 365, D-69120, Heidelberg, Germany and Department of Tumor Progression and Immune Defense, German Cancer Research Center, Heidelberg, Germany and Department of Applied Genetics, University of Karlsruhe, Karlsruhe, Germany e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_26, © Springer Science+Business Media, LLC 2010
555
556
M. Zöller
Structure, Organization, and Major Functions of Tetraspanins The Structure of Tetraspanins Tetraspanins have first been described in 1990 (Oren et al. 1990). They are evolutionary highly conserved small membrane proteins expressed in many species ranging from sponges to mammals, with each organism expressing a large number of tetraspanin family members (Garcia-España et al. 2008). Tetraspanins cross the membrane four times, such that the short N- and C-terminal tails are located within the cytoplasm. They have a small extracellular loop (ECL1) between the transmembrane (TM) regions TM1 and TM2, a small intracellular loop between TM2 and TM3, and a large extracellular loop (ECL2) between TM3 and TM4. The essential signature of tetraspanins, which differentiates them from other four-TM molecules, is highly conserved cysteines in this large extracellular loop. Furthermore, the ECL2 is subdivided in a more constant and a variable region. The more constant region likely accounts for tetraspanin dimerization. The variable region is important for interactions with nontetraspanin partner molecules. The tertiary structure of tetraspanins is stabilized by polar residues in the TM regions (reviewed in Stipp et al. 2003; Seigneuret 2006). Palmitoylation (S-acylation of proteins with palmitate) of intracellular, juxtamembrane cysteines likely is required for initiating tetraspanin–tetraspanin web formation, which can also be supported by palmitoylation of associating integrins. Protein palmitoylation in general regulates subcellular traffic and/or protein degradation (Percherancier et al. 2001) or can protect the molecules from lysosomal degradation and link them to cholesterol and gangliosides (Berditchevski et al. 2002; Charrin et al. 2002; Sharma et al. 2008; Todeschini et al. 2008). The DHHC2 (Asp-His-His-Cys) protein appears to be particularly important for tetraspanin palmitoylation (Sharma et al. 2008) (Fig. 26.1a).
The Tetraspanin Web Tetraspanins form complexes by interacting between themselves as well as a large variety of transmembrane and cytosolic proteins (Hemler 2005; Levy and Shoham 2005a, b; André et al. 2006) which are essential for the functional activity of tetraspanins. Integrins, particularly a3b1, a4b1, and a6b1, are the most prominent nontetraspanin partners, whereas the association with a6b4 is more restricted (Berditchevski 2001). Tetraspanins also associate with growth factor receptors (Sridhar and Miranti 2006; Murayama et al. 2008), G protein-coupled receptors (GPCR), seven-TM receptors that signal through small heterotrimeric GTP-binding proteins and activate multiple signaling pathways through second messengers such as cAMP, Ca2+, and inositol trisphosphate (IP3) (Little et al. 2004), several peptidases (Le Naour et al. 2006), claudins (Le Naour and Zöller 2008), a family of four-TM protein originally described as components of tight junctions (Tsukita and
26 Tetraspanins and Cancer Metastasis
557
a
CCG tetraspanin motif ECL2 ECL1
TM1
TM4
palmitoylation
internalization motif NH2
COOH
b
α2α6α 6-
β1 β4
TEM P PKC
α β γ
PI4KII
CD13 ITSN2 CD44v EWI-F EpCAM tetraspanin CD26 EWI-2 claudin-1 / -7
integrins
GPCR
Fig. 26.1 Tetraspanins and associated molecules. (a) Prototype structure of tetraspanins with four transmembrane (TM) regions, palmitoylation site (pink) in the cytoplasmic domains, a potential internalization motif (blue) in the C-terminal domain, and polar AA (yellow) in the cytoplasmic and transmembrane domains. The ECL2 contains 4–8 conserved cysteine residues (red) including the CCG motif (red–red–gray); disulfide bonds are indicated. (b) Molecules frequently associating with tetraspanins: some integrins, the dipeptidases CD13 and CD26, EWI-F and EWI-2, intersectin-2 (ITSN2), EpCAM and claudin-1 can directly or indirectly associate with some tetraspanins. For other molecules, like CD44 variant (CD44v) isoforms, GPCR, PKC, and PI4K only indirect associations have been described
Furuse 2000), molecules associated with tumor progression like CD44 and EpCAM (Le Naour et al. 2006), and Ig superfamily members including two members of a new class that contains a conserved Glu-Phe-Ile (EWI) motif (Dong et al. 2008), EWI-F/FPRP/CD9-P1 and EWI-2 (Charrin et al. 2001; Stipp et al. 2001a; Claas et al. 2005; André et al. 2006). Prominent cytosolic signal transduction molecules coimmunoprecipitating with tetraspanins are protein kinase C (PKC), a type II
558
M. Zöller
phosphatidylinositol 4 kinase (PI4KII), and phospholipase Cg (PLCg) (Berditchevski et al. 1997; Yauch and Hemler 2000; Claas et al. 2001; Zhang et al. 2001; Hemler 2005; André et al. 2006). Several isoforms of PKC mediate serine and threonine phosphorylation in many different protein substrates thereby propagating various signal transduction pathways, which lead to transcription factor activation (Walker 2008). PI4KII generates phosphoinositol (PtdIns) 4-phosphates, precursors of important regulatory phosphoinosites. Besides regulating signaling events, PI4KII is also important in vesicular trafficking and lipid transport (Guo et al. 2003). PLCg catalyzes hydrolysis of phospholipid PIP2 to generate the signaling molecules IP3 and diaglycerol (Wells and Grandis 2003). The interactions of tetraspanins with themselves and this large array of different nontetraspanin molecules are grouped into three categories according to the strength of the detergent required to disrupt the association. Direct protein–protein interactions, so-called type I interactions, resist stringent detergents. They are rare and comprise some tetraspanin-homodimers (tri- and tetramers) and few heterointeractions. Prominent examples are the interactions of CD151 with some integrins and claudin-1 (Yauch et al. 1998; Sterk et al. 2000; Kovalenko et al. 2007) as well as the interactions of CD9, CD81, and Tspan8 with EWI proteins (Stipp et al. 2001a, b; Claas et al. 2005). These direct interactions may proceed through ECL2 and/or TM regions 2–4 (Kovalenko et al. 2004). The majority of tetraspanin–integrin and tetraspanin–tetraspanin interactions are type II interactions maintained under milder lysis conditions. Palmitoylation of tetraspanins and possibly of the associating proteins is essential for this type of interaction. It may be initiated in the Golgi and provide a targeting sequence for coassociations (Berditchevski et al. 2002; Charrin et al. 2002; Zhou et al. 2004). Weak type III interactions, e.g., with several kinases, also are stabilized by palmitoylation (Hemler 2005) (Fig. 26.1b). Besides these primary interactions, tetraspanins also associate with cholesterol and gangliosides (Charrin et al. 2003; Miura et al. 2004; Odintsova et al. 2006), anionic glycosphingolipids characterized by sialic acid residues (Yates and Rampersaud 1998). Thereby higher order tetraspanin complexes are formed in microdomains, termed TEM (tetraspanin-enriched membrane microdomains) (Hemler 2003). TEM share several features with lipid rafts, but are independent membrane microdomains that, different to lipid rafts, are disrupted by Triton X-100 at 4°C. Tetraspanins also do not associate with glycosylphosphatidylinositolanchored proteins and caveolin, which are classical raft components (Hemler 2005; Le Naour et al. 2006; Todeschini et al. 2008). Tetraspanins are not only recovered in the cell membrane, but also abundantly in intracellular vesicles and exosomes (Hemler 2003; Berditchevski and Odintsova 2007; Pols and Klumperman 2009; Zöller 2009) (Fig. 26.2). Exosomes are small 30–100 nm vesicles that are released upon exocytic fusion of the limiting membranes of multivesicular bodies (MVB) with the plasma membrane (Simpson et al. 2009). Some tetraspanins possess a tyrosine-based sorting motif, a sequence of Tyr–Xaa–Xaa–f, where f stands for an AA with a bulky hydrophobic side chain, in the C-terminal cytoplasmic domain (Marks et al. 1997). By this sorting motif, tetraspanins become prone for delivery to intracellular compartments (Marks et al.
26 Tetraspanins and Cancer Metastasis
559
tumor cell-derived exosome integrins EpCAM
ICAMs EWI-F/EWI-2 CD13/ CD26
claudin
HSP Signaling mol. Anti-apoptotic mol. Enzymes Histones CD9 CD82 CD151 Tspan8
LAMP1/2 Rabs Annexins Arp2/3 MHC I
ERM actin tubulin mRNA
Fig. 26.2 The composition of exosomes: Exosomes are cup-shaped vesicles. Their composition varies with the originating cell, but exosomes contain some common components, including tetraspanins and tetraspanin-associated molecules like integrins, EpCAM, claudins, CD13/CD26 and EWI-F/EWI-2, the cytoskeletal proteins ezrin, radixin and moesin (ERM proteins), and actin. Rab proteins, annexins and Arp2/3 are involved in membrane transport and fusion. Exosomal signal transduction molecules that are involved in tetraspanin-initiated pathways are small heterotrimeric GTP-binding proteins, src proteins, ERK1/2, Rho family protein members, SH2 phosphatase, and catenin. Exosomes also contain mRNA and miRNA
1997; Berditchevski and Odintsova 2007). However, some tetraspanins enriched in exosomes do not possess a sorting motif (CD9) or an inappropriately located sorting motif (Tspan8) (Berditchevski and Odintsova 2007). Thus, it appears likely that individual tetraspanins follow different routes of internalization. For CD63 it was shown that the C-terminal domain interacts with µ4, µ2, µ3A, and µ3B subunits of adaptor protein (AP) complexes, linking the traffic of CD63 to the clathrin-dependent pathway (Rous et al. 2002). Alternatively, CD63 may be sorted via its association with syntenin-1 (Janvier and Bonifacino 2005). Whether trafficking of tetraspanins can also rely on AP-independent pathways or whether additional domains are required has not yet been comprehensively answered (Berditchevski and Odintsova 2007). To give an example, CD151 is involved in the internalization of a3b1. CD151 has a sorting motif (YRSL) and may promote internalization through the clathrin-dependent pathway (Sincock et al. 1999; Winterwood et al. 2006). Alternatively, CD151 may link the complex to PKCa, which regulates integrin traffic (Ng et al. 1999; Zhang et al. 2001) or CD151 may connect a3b1 to other
560
M. Zöller
TM proteins in TEM with an internalization sequence (Berditchevski et al. 2002; Winterwood et al. 2006). Another example is CD82, where high level surface expression correlates with low a6b1 and a6b4 expression (He et al. 2005), which affects integrin-mediated cell migration (Winterwood et al. 2006). Taken together, the enrichment of tetraspanins in exosomes is well known, the mode of recruitment is only partly defined and their impact on exosomal activity has not been unraveled. Possible pathways will be discussed with respect to exosomal activities in angiogenesis and tumor progression.
Major Functional Activities of Tetraspanins: Migration and Membrane Fusion One of the first reviews on tetraspanins from Shoshana Levy’s group “The tetraspanin superfamily: molecular facilitators” (Maecker et al. 1997) still hits the nail’s head, tetraspanin act via modulating, stabilizing, or preventing activities of their associated molecules. Though the same tetraspanin may associate with different molecules in distinct cells or in a different tetraspanin web, which further widens the range of activities, the main functional activities are well recapitulated by “cellular penetration, invasion, and fusion” (Hemler 2003). The first activities ascribed to tetraspanins relate to cell motility due to their preferential association with integrins. As there is no evidence that tetraspanins directly alter integrin affinity for their ligands and mostly do not affect static cell adhesion, it was proposed that tetraspanins promote cell migration by integrin compartmentalization or by integrin internalization and recycling or by modulating integrin signaling (Berditchevski 2001; Stipp et al. 2003; Hemler 2005; Levy and Shoham 2005a, b), thus promoting cell spreading and migration including growth along lines of mechanical tension forming a pattern of intersecting cellular cables when growing on gelatinous matrices like matrigel (Berditchevski 2001; Hemler 2001, 2005; Lekishvili et al. 2008). Matrigel, the extracellular matrix secreted by a mouse sarcoma line, resembles the basement membrane (Malinda 2009). Detailed biochemical studies uncovered tetraspanin-induced activation of integrin-mediated signaling pathways including PI3K, ERK1/2, and Cdc34 (reviewed in Sawada et al. 2003; Yunta and Lazo 2003; Wright et al. 2004a, b; Hemler 2005; Levy and Shoham 2005a, b). By regulating trafficking and biosynthesis of associated molecules, tetraspanins also can become important in cell adhesion events (He et al. 2005; Winterwood et al. 2006). Besides integrins, the association with EWI proteins influences cell polarity and migration. This is due to EWI proteins binding to ERM (ezrin–radixin– moesin) proteins that negatively regulate ERM protein phosphorylation (Yalaoui et al. 2008). The latter is required for a conformational change allowing the ERM proteins to link TM molecules with the actin cytoskeleton (Tsukita and Yonemura 1999). Thus, silencing of endogenous EWI-2 expression by short interfering RNA (siRNA) augmented cell migration, cellular polarity, and increased ERM phosphorylation (Sala-Valdés et al. 2006).
26 Tetraspanins and Cancer Metastasis
561
Several tetraspanins have been shown to regulate invasiveness, which activity could well depend on their association with peptidases (Potolicchio et al. 2005; Le Naour et al. 2006), ADAMs (A disintegrin and metalloproteinase), particularly ADAM10 (Arduise et al. 2008), matrix metalloproteinases (MMP) (Yanez-Mo et al. 2008) and uPAR (Bass et al. 2005), or could proceed through modulating MMP transcription and secretion (Hasegawa et al. 2007). Possibly the best known feature of tetraspanins is the deficit in egg–sperm fusion in CD9-knockout mice, where selectively the sperm–oocyte fusion is severely impaired. It has been hypothesized that the arrangement of adhesion molecules or of the cytoskeleton or of signaling complexes may be altered (Kaji et al. 2000). Besides CD9, a CD81 deficiency also is accompanied by reduced egg–sperm fusion (Rubinstein et al. 2006). CD9, CD81, and CD82 are involved in virus-induced syncytia formation. Additional examples of the involvement of tetraspanins in membrane fusions are the involvement of CD9 in the susceptibility to feline immunodeficiency virus, the accumulation of HIV virions in tetraspaninenriched endosomal compartments and the resistance of CD9- and CD81-deficient mice to infection with Plasmodium yoelii sporozites (de Parseval et al. 1997; Pileri et al. 1998; Silvie et al. 2003; Martin et al. 2005; Ardón-Alonso et al. 2006; Garcia et al. 2008a, b; Singethan et al. 2008; Yalaoui et al. 2008). However, distinct pathogens may use different mechanisms of entry, even when the same tetraspanin(s) is engaged, as has been demonstrated for Plasmodium and hepatitis C virus (Helle and Dubuisson 2008; Rocha-Perugini et al. 2009). For some viral pathogens, it has been proven that their invasion requires TEM and is independent of clathrin-coated pits and caveolae (Spoden et al. 2008). Morphogenic features of tetraspanins (Zhang et al. 2002; Hemler 2003; Loewen et al. 2003; Yamamoto et al. 2007) frequently have been associated with angiogenesis induction and cable formation of tumor cells, where an involvement of CD82, CD151, and CO-029 via their association with a3b1 or a6b1 has been described (Zhang et al. 2002; He et al. 2005; Gesierich et al. 2006; Hashimoto et al. 2007; Takeda et al. 2007a, b). It is likely that the involvement of tetraspanins in migration, membrane fusion events, and morphogenesis relates, at least in part, to their sorting into MVB and their enrichment in exosomes. In turn, exosomal tetraspanins may well account for the selectivity of defined target cells and the mode of the exosome–target cell interaction. In view of the importance of exosomes as intercellular communicators, it is demanding to answer these questions (Pap et al. 2009; Zöller 2009). Overall, via their associated molecules embedded in cholesterol-enriched membrane microdomains and their motifs for enrichment in intracellular vesicles, tetraspanins contribute to cell motility, adhesion, invasion, and fusion (Hemler 2003). Taking into account the reversibility of palmitoylation (Bijlmakers and Marsh 2003; Linder and Deschenes 2007) and the instability of membrane microdomains (Devaux and Morris 2004), it becomes obvious that the activities of tetraspanins may considerably vary depending on the activation state of the cell as well as the surrounding tissue. The fact that tetraspanins act via their laterally associated partner molecules and only exceptionally via ligand binding, adds to
562
M. Zöller
the functional divergence. It is, thus, not surprising that these environment-dependent facilitators can suppress or promote the metastatic process.
Tetraspanins, Metastasis, Angiogenesis, and Thrombosis Metastasis and Angiogenesis The metastatic cascade is initiated by the epithelial–mesenchymal transition (EMT) of cancer stem cells/cancer initiating cells (Brabletz et al. 2005; Hlubek et al. 2007; Yang and Weinberg 2008). The following steps of the metastatic cascade, evasion from the primary tumor, intravasation, extravasation, settlement and growth in distant organs, are well known (Stracke and Liotta 1992; Ahmad and Hart 1997; Birchmeier et al. 2003; Gassmann et al. 2004; Marhaba and Zöller 2004; Deryugina and Quigley 2006; Eble and Haier 2006). Molecules involved in tumor progression are cell–cell and cell–matrix adhesion molecules, matrix degrading enzymes and their inhibitors as well as chemotactic factors released from the degraded matrix and upregulated chemokine receptor expression by the metastasizing tumor cells, apoptosis resistance genes, and angiogenesis inducers (Kerbel and Folkman 2002; Friedl and Wolf 2003; Mott and Werb 2004; Albini et al. 2008). With progress in unraveling the biochemical processes underlying each step of the metastatic cascade and increasing awareness on the requirement of coordinate activities of all components, evidence was provided that at least two tetraspanins, CD151 and Tspan8, support tumor progression. Growth of a primary tumor requires formation of new blood vessels. Abundant angiogenesis, the process of new capillary formation from a pre-existing vasculature, is crucial to supply the growing tumor with nutrients and also facilitates metastatic spread. Tumor angiogenesis proceeds through several sequential steps. The common concept implies that the process is initiated by angiogenic factors produced by tumor cells that bind to endothelial cell (EC) receptors such that EC become activated (Ribatti et al. 2007). Tumor vessels frequently are leaky, which allows the extravasation of plasma proteins that constitute a scaffold for newly migrating EC and facilitate initiation of thrombus formation and spontaneously occurring focal hemorrhages (Sood 2009). The particular features of tumor vessels supporting thrombus formation and activation of the coagulation cascade provide a feedback for angiogenesis induction. Notably, exosomes play a crucial role in the hemostatic balance (Aharon and Brenner 2009) and tumor angiogenesis as well as tumorassociated thrombosis have also been associated with the two metastasis supporting tetraspanins, CD151 and Tspan8. Distinct to the requirement for coordinated activity in progression through the metastatic cascade, a blockade at any step of the cascade efficiently hampers metastasis formation. This fact has facilitated the discovery of metastasis suppressor genes, which are defined as genes that selectively interfere with metastasis formation, while the growth of the primary tumor is unimpaired (Steeg 2003;
26 Tetraspanins and Cancer Metastasis
563
Stafford et al. 2008). Metastasis suppressor genes belong to diverse classes of molecules, including cytoplasmic molecules like Nm23, nuclear proteins like p53, and membrane-bound molecules like cadherins and also the tetraspanin CD82/KAI1 (Stafford et al. 2008). Here it will be discussed that why some tetraspanins interfere with tumor progression, while other support progression through the metastatic cascade, angiogenesis, and thrombosis.
Tetraspanins and Metastasis Suppression The Metastasis Suppressor Gene CD82/KAI1 and Tumor Cell Migration CD82, also known as KAI1, R2, C33, IA4, 4F9, is widely expressed in solid tissues, hematopoietic cells, and very high in platelets (Horejsí and Vlcek 1991; Dong et al. 1995; Maecker et al. 1997). CD82 has first been described as a metastasis suppressor gene (reviewed in Malik et al. 2009). CD82 belongs to the tetraspanins with six cysteines in the LEL (Seigneuret et al. 2001). It associates with the tetraspanins CD9, CD63, and CD81 (Rubinstein et al. 1996). It binds several integrins and cadherins, immune response related molecules like CD4, CD8, CD19, CD21, MHC I and MHC II, L6-Ag, a tumor-specific antigen (Marken et al. 1992), signal transduction molecules, e.g., PKC and PI4K, growth factor receptors (Imai and Yoshie 1993; Hemler et al. 1996; Mannion et al. 1996; Szöllósi et al. 1996; Lagaudrière-Gesbert et al. 1997; Hammond et al. 1998; Horváth et al. 1998; Nakamura et al. 2000; Odintsova et al. 2000; Bienstock and Barrett 2001; Mohan et al. 2007). CD82 forms a direct complex with EWI-2 (Zhang et al. 2003a, b) and binds other proteins like KITENIN and g-glutamyl transpeptidase (Nichols et al. 1998; Lee et al. 2004). CD82/KAI1 has an internalization motif (Berditchevski and Odintsova 2007) and is recovered in endosomal/lysosomal compartments and exosomes (Lagaudrière-Gesbert et al. 1997; Hammond et al. 1998; Horváth et al. 1998; Berditchevski and Odintsova 2007). CD82 is frequently downregulated in advanced stages of cancer (Dong et al. 1995) and loss of CD82, first described in a rat prostate cancer (Ichikawa et al. 1991; Dong et al. 1995) has been strongly associated with a poor prognosis in patients with esophageal, prostate, gastric, colon, cervix, ovarian, breast, skin, bladder, endometrial, lung, pancreatic, liver and thyroid cancer, neuroblastoma, melanoma, myeloma, and others (Adachi et al. 1996; Ueda et al. 1996; Yang et al. 1997; Takaoka et al. 1998; White et al. 1998; Lombardi et al. 1999; Liu et al. 2000; Miyazaki et al. 2000; Schindl et al. 2000; Houle et al. 2002; Wu et al. 2003; Chen et al. 2004; Farhadieh et al. 2004; Guo et al. 2005; Son et al. 2005; Tsutsumi et al. 2005; Drucker et al. 2006; Leavey et al. 2006; Briese et al. 2008; Protzel et al. 2008; and reviewed in Jackson et al. 2005; Tonoli and Barrett 2005; Liu and Zhang 2006; Malik et al. 2009). CD82 was found to be downregulated in >80% of metastases and in >85% of studies downregulation was associated with a poor prognosis (Zöller 2009).
564
M. Zöller
In agreement with these findings, CD82 expression is able to suppress metastasis formation in animal models (Phillips et al. 1998; Takeda et al. 2007a, b). An opposing result on CD82-supported hematogenous spread of human lung cancer in SCID mice may be related to the xenogeneic system (Shinohara et al. 2001). In vitro, CD82 overexpression inhibits migration and invasion (Jackson et al. 2005; Tonoli and Barrett 2005; Liu and Zhang 2006). Studies using human tumor lines have revealed several pathways through which CD82 achieves this. In general, CD82 likely suppresses tumor progression by inhibiting cell motility via regulating signal transduction through associated proteins (reviewed in Liu and Zhang 2006). First, the association of CD82 with the a6 integrin chain and the epidermal growth factor receptor (EGFR) is accompanied by impaired laminin adhesion and migration, which is due to cointernalization with CD82, and is abolished by mutating the CD82 sorting motif (He et al. 2005; Odintsova et al. 2000, 2003). Similarly, the Ly6 antigen associates with CD82 and CD63 in TEM and promotes internalization of these tetraspanins. Accordingly, downregulation of Ly6 is accompanied by increased levels of CD82 and CD63 surface expression and reduced motility (Lekishvili et al. 2008). Second, the association of CD82 with EWI-2 strengthens the motility inhibitory activity of EWI-2 on laminin and fibronectin (Zhang et al. 2003a, b) by preventing the activation of ERM proteins, which is required for the linkage with actin (Devaux and Morris 2004). Third, uPAR colocalizes with a5b1 in focal adhesions only in the presence of CD82. The stable association between uPAR and a5b1 prevents binding of uPA to its receptor and pericellular proteolysis, a necessary step in invasion, is strikingly reduced (Bass et al. 2005). In multiple myeloma CD82, CD81 overexpression affects motility and invasive potential accompanied by reduced MMP9 secretion. In addition, a decrease in Ki67 and an elevated intracellular glutathione level are accompanied by reduced survival (Tohami et al. 2007). Fourth, CD82 interferes with c-Met signaling such that hepatocyte growth factor (HGF, also known as scatter factor)-induced cell migration, but not proliferation is reduced. In a nonsmall cell lung cancer line overexpressing CD82, phosphorylation of c-Met by HFG stimulation was unimpaired, but CD82 interfered with binding of growth factor receptor-bound protein 2 (Grb2). Grb2 is a key molecule in intracellular signal transduction, linking activated cell surface receptors to downstream targets by binding to specific phosphotyrosine-containing and proline-rich sequence motifs. CD82 interfering with Grb2 binding was accompanied by inhibition of downstream signaling via phosphoinositide 3-kinase (PI3K) and the ras–raf–MAPK (mitogen-activated protein kinase) pathway with activation of rac and Cdc42 GTPases. Lamellipodia formation and cell migration were severely impaired (Takahashi et al. 2007). Fifth, CD82 can recruit negative inhibitors that affect c-Met and src activation. Src are cytoplasmic tyrosine kinases, which are controlled by multiple membrane receptors and signal to a variety of downstream effectors (reviewed in Fuss et al. 2008) In a prostate cancer line, where CD82 interferes with integrin-dependent, HGF-induced activation of c-Met, src activation as well as c-Met phosphorylation
26 Tetraspanins and Cancer Metastasis
565
was impaired, although CD82 did not interact with c-Met. Instead, in the absence of CD82 c-Met activation was unimpaired and the formation of the focal adhesion kinase (FAK)–p130CAS–Crk complex downstream of src initiated the molecular switch for motility (Sridhar and Miranti 2006). p130CAS and FAK are downstream substrates of src. P130CAS is a multiadaptor and scaffold molecule that spatially and temporally controls signaling events through changes in phosphorylation and association with effector molecules like Crk, FAK, PTP-1B, 14-3-3, PI3K, and src family kinases (reviewed in Defilippi et al. 2006). Downregulation of the p130CAS– Crk complex by CD82/KAI1 has also consequences on integrin-mediated cell migration (Zhang et al. 2003a, b). Sixth, some activities of CD82 can only be explained by taking into account the organization of tetraspanin complexes in TEM and the contribution of gangliosides (Ono et al. 1999; Todeschini and Hakomori 2008; Hakomori 2010). For example, the impact of CD82 on EGFR activation varies depending on the presence of ganglioside GD1a, which is required for the relocalization of the CD82–EGFR complex in TEM. GD1a is important for the spatial organization of CD82-enriched microdomains, which can interfere with the capacity of CD82 to recruit molecules that negatively regulate EGFR activation, e.g., a tyrosine phosphatase (Odintsova et al. 2006). The CD82–a3–Met crosstalk also becomes regulated via formation of a complex of GM2/GM3 with CD82, which interferes with c-Met activation. Both Grb2 and Ras upstream of the MAPK pathway and Gab1 (Grb2-associated binder 1) upstream of PI3K do not become activated. This blockade impairs not only cell motility, but also proliferation (Todeschini et al. 2008). It is discussed that the CD82/GM2/GM3 complex-inhibited tumor cell proliferation proceeds similar to the PKCa-mediated inhibition of EGFR-induced proliferation, where GM3 together with CD82 account for translocation and phosphorylation of PKCa that induces EGFR phosphorylation and internalization (Wang et al. 2007a, b). Seventh, two pathways of CD82-mediated metastasis suppression are motility independent. KITENIN is a metastasis-supporting four-TM protein that does not belong to the tetraspanin family. It binds to the C-terminal tail of CD82. Over expression of KITENIN in a murine colon carcinoma line promotes adhesion to ECM elements, tumor cell migration, and metastasis formation (Rowe and Jackson 2006). This may, in part, be due to the expression of a CD82 splice variant, which does not bind KITENIN and, thus, does not interfere with its metastasis-promoting activity (Jackson et al. 2007). There exists some kind of a feedback loop, because KITENIN expression by itself supports expression of this CD82 splice variant (Cherukuri et al. 2004; Clark et al. 2004). Finally, CD82 interacts in trans with DARC (Duffy antigen receptor for chemokines) on vascular ECs. This induces tumor cell senescence via reduced expression of the senescence-related transcription factor TBX2 (T-box 2) gene and upregulation of the cyclin-dependent kinase inhibitor p21WAF1, which is repressed by TBX2 (Prince et al. 2004). Accordingly, the metastasis-suppressor activity of CD82 is significantly reduced in DARC-/- mice (Bandyopadhyay et al. 2006). Besides the latter two exceptions, CD82 is a convincing example that blocking one step of the metastatic cascade, migratory activity of primary tumor cells, is
566
M. Zöller
sufficient to prevent metastasis. CD82 inhibits migration and invasion by associating directly or via bridging integrins with a multitude of different molecules as well as by the recruitment of the partner molecules in TEM, where gangliosides support or counteract the metastasis-prohibiting activity of CD82. In this respect, it is particularly challenging that perturbation of TM interactions of CD82 significantly affect the molecule’s intrinsic activities (Bari et al. 2009). A therapeutic perturbation of TM interactions may open a new avenue to prevent cancer invasion, which could be far easier approached than a blockade of individual signaling pathways. Rescuing CD82 gene expression also should prevent tumor progression (see “Rescuing the Metastasis Suppressor Gene CD82” section).
CD9 Interferes with Distinct Steps of the Metastatic Cascade CD9, expressed by ECs, brain tissue, peripheral nerves, vascular smooth muscle cells, cardiac muscles, epithelial cells, some hematopoietic cells including platelets (Horejsí and Vlcek 1991; Maecker et al. 1997), is not a metastasis suppressor gene in the strict sense (Boucheix and Rubinstein 2001; Wright et al. 2004a, b), as CD9 expression can favor metastasis formation in some tumors like gastric cancer (Hori et al. 2004). However, a positive correlation between CD9 downregulation and tumor progression has been described in colon cancer (Mori et al. 1998; Ovalle et al. 2007), peritoneal dissemination of ovarian cancer (Houle et al. 2002; Drapkin et al. 2004; Furuya et al. 2005), endometrial cancer (Miyamoto et al. 2001), small cell lung carcinoma (Higashiyama et al. 1995, 1997; Adachi et al. 1998; Miyake et al. 1999; Funakoshi et al. 2003; Saito et al. 2006; Takeda et al. 2007a, b), breast cancer (Huang et al. 1998, 2005; Mimori et al. 2005), cervical carcinoma (Sauer et al. 2003a, b) and squamous cell carcinoma of head and neck (Kusukawa et al. 2001; Mhawech et al. 2004), bladder cancer (Mhawech et al. 2003), and brain tumors (Kawashima et al. 2002). For prostate cancer, different effects of CD9 expression were described (Bérubé et al. 1994; Zvieriev et al. 2005; Wang et al. 2007a). The opposing activities of CD9 may depend on the associating molecules in the tetraspanin web. CD9 homoclustering is promoted by a3b1, a6b4, and by palmitoylation of CD9 and b4. Instead EWI-F- and EWI-2-associated or unpalmitoylated CD9 forms heteroclusters, which particularly are seen on malignant epithelial tumors (Yang et al. 2006). Distinct to CD82, CD9 can interfere with tumor progression at several steps of the metastatic cascade. CD9 can have an impact on tumorigenicity. V-Jun transformation of chicken or mouse fibroblasts interferes by an unknown mechanism with GM3 synthase transcription, such that Jun-induced oncogenic transformation was accompanied by loss of CD9–GM3 association, leading to integrin activation, enhanced motility, and soft agar colony formation. Transfection with the GM3 synthetase gene reverted the oncogenic phenotype (Miura et al. 2004). Furthermore, ectopic expression in colon carcinoma cells has an impact on in vivo tumorigenicity in nude mice (Ovalle et al. 2007). Growth inhibition by CD9 can also rely on increased expression of tumor necrosis factor (TNF)a, where the inhibitory effect
26 Tetraspanins and Cancer Metastasis
567
of CD9 is mediated by membrane-bound TNFa (Ovalle et al. 2007), whose production is delayed in CD9-/- mice (Yamane et al. 2005). A similar phenomenon has been described in hepatic carcinoma (Li et al. 2006). How CD9 influences TM TNFa activity has not been clarified, but affecting the molecules stability by inhibiting the activity of ADAM in cleaving TNFa could provide a possible explanation (Moss and Bartsch 2004). CD9 might also interfere with the EMT, the initiating step of the metastatic cascade. CD9 expression induces downregulation of several Wnt family genes, such as Wnt1, Wnt2b1, and Wnt5a and their targets including WISP-1, WISP-3, c-Myc, VEGF-A, and MMP26. Wnt proteins are a large family of secreted glycoproteins that activate signal transduction pathways to control a wide variety of cellular processes such as determination of cell fate, proliferation, migration, and polarity. Wnts are capable of signaling through several pathways, the best characterized being the canonical b-catenin/Tcf-mediated pathway (Coombs et al. 2008). Notably, CD9 expression had no impact on expression of genes involved in cell motility-associated signaling or cell cycle regulation. Instead a CD9-specific antibody inhibited downregulation of Wnt genes, which suggests that the CD9 signal is located upstream of the Wnt signaling pathway (Huang et al. 2004). Thereby, downregulation of Wnt signaling by CD9 could well result in suppression of transformation and epithelial– mesenchymal transition (Huang et al. 2004) as Wnt1 stimulates the canonical Wnt/b-catenin signaling pathway that leads to changes in cell fate and/or transformation (You et al. 2002). The impact of CD9 on Wnt signaling may also affect later stages in the metastatic cascade-like cell motility, where Wnt5a stimulates the Wnt/ Ca2+ signaling pathway that directly affects actin reorganization (Ishitani et al. 2003) and invasiveness, the downregulated target genes WISP-1 and MMP26 being associated with aggressive tumor growth (Yamamoto et al. 2004). CD9 can hamper metastasis formation by prohibiting integrin-mediated motility. In ovarian carcinoma cells, expression levels of CD9 and b1, a2, a3, a5, and a6 integrin chains correlate, and downregulation of CD9 is accompanied by weaker matrix adhesion and diffuse growth in vitro (Ikeyama et al. 1993; Furuya et al. 2005). In addition, CD9 can associate with gangliosides that modulates the activity of CD9. A noninvasive bladder cancer line expresses the GM3–CD9 complex at a high level. This correlates with a strong association with a3b1 and low cell motility. The reverse is true for an invasive bladder cancer line. When GM3 is expressed at a low level, it activates Src, whereas a high level GM3 causes Csk (C-terminal Src kinase, an endogenous inhibitor of the Src-family protein tyrosine kinases) translocation into TEM microdomains with subsequent inhibition of Src by its inhibitor phoshatase 2. By the recruitment of Csk, GM3 interferes with Src activity with the consequence of impaired Rac and PI3K/Akt activation (Mitsuzuka et al. 2005). CD9 also can hamper the migration of the isolated metastasizing cells. CD9 associates with the b1 integrin chain and the EGFR. CD9 crosslinking or EGF stimulation in cells overexpressing CD9 promotes EGFR internalization, which results in a reduction in EGFR autophosphorylation after stimulation and reduced SHC phosphorylation and recruitment of Grb2 (Murayama et al. 2008). An additional pathway, whereby CD9 could regulate ligand-induced activation of the
568
M. Zöller
EGFR relies on its association with TM transforming growth factor (TGF)a, which binds, like proteolytically cleaved soluble TGFa, to the EGFR and activates EGFRinduced signaling pathways. Metalloproteinase ectodomain cleavage of TGFa is important in many cellular signaling pathways (Fan and Derynck 1999; Herrlich et al. 2008). Soluble TGFa activates the EGFR and induces receptor downregulation. This also accounts for CD9-associated TGFa, which enhances EGFR activation and proliferation in juxtacrine assays, but decreases autocrine growth stimulation (Shi et al. 2000). This is a sequel of persisting EGFR stimulation by CD9-stabilized TGFa, which leads to pronounced EGFR internalization, reorganization of the cytoskeleton, an increase in cell adhesion, and a decrease in cell migration (Imhof et al. 2008). Finally, by a not yet defined mechanism, though not proceeding through the Wnt signaling pathway (Takenawa and Suetsugu 2007), CD9 expression is accompanied by transcriptional downregulation of WAVE2 (Huang et al. 2006). WAVE2, a member of the WASP (Wiskott–Aldrich syndrome proteins) family of proteins, serves as a scaffold that links upstream signals to the activation of the ARP2/3 (actin-related proteins 2 and 3) complex, which leads to a burst of actin polymerization, where WAVE2 is crucial for lamellipodium and filipodium formation. CD9 also can interfere with transendothelial migration of tumor cells, another essential step in the metastatic cascade. CD9, CD81, and CD151 colocalize at the tumor cell–endothelial cell contact area, where CD9 promotes strong adhesion via the b1 integrin chain, which hampers transendothelial migration of the tumor cell (Longo et al. 2001). On the other hand, although downregulated in metastases, high level CD9 expression at tumor cones can support transendothelial migration in cervical carcinoma and recovery of these cone-localized CD9 “hot spots” is a highly significant indicator of lymphangiogenesis (Sauer et al. 2003a, b). Strong CD9 expression was also observed on myeloma cells in close contact to bone marrow ECs (De Bruyne et al. 2006). The reason(s) for these opposing observations may rely on differences in the CD9 web of individual tumor cells. Within the blood stream, the tumor cell can profit again from CD9 downregulation. CD9 associates with the platelet aggregation-inducing factor podoplanin. Ectopic expression of CD9 in podoplanin-expressing tumor cells leads to reduced lung metastasis formation accompanied by impaired platelet aggregation (Nakazawa et al. 2008). Platelets bind via CLEC-2 (C-type lectin-like receptor-2) to podoplanin that induces platelet degranulation (Suzuki-Inoue et al. 2006). Because CLEC-2 does not recognize CD9-associated podoplanin (Nakazawa et al. 2008), platelet aggregation is impaired. Tumor cell platelet aggregates facilitate embolization of the microvasculature and metastasis formation. Platelet aggregates embedded tumor cells may also be protected from an immune attack (Sierko and Wojtukiewicz 2007). Depending on the prevailing conditions, CD9 can mitigate the aggressiveness of the metastasizing tumor cell and of metastatic growth. Ectopic CD9 expression can suppress motility and neurite-like outgrowth and promotes apoptotic cell death when grown in vitro under starvation. This correlates with impaired adhesion-dependent Akt phosphorylation and suppression of MMP2 secretion (Saito et al. 2006). In contrast, under physiological conditions, CD9 overexpression revealed increased
26 Tetraspanins and Cancer Metastasis
569
MMP2 expression at the cone of invading cells that facilitates transendothelial migration (Hong et al. 2005). Inhibition of p38MAPK and JNK as well as siRNA transfection abrogated MMP2 expression, which points towards CD9-induced MMP2 expression to be mediated through an AP-1 site (Hong et al. 2005). Because metastasizing tumor cells less likely are exposed to starvation, CD9-induced MMP2 expression appears more likely to reflect the in vivo situation. Though partly opposing activities have been observed, which may be due to differences in the tetraspanin web, the potential of CD9 to suppress metastasis relies on inhibiting tumorigenicity and proliferation, EMT, motility, aggressiveness, and survival including the recruitment of a protective environment. CD81 and CD63 and Metastasis Suppression CD81 is widely expressed in human tissue with the exception of erythrocytes and platelets. Similar expression profiles were found in mouse tissue (Levy et al. 1998). CD81 downregulation has been observed in hepatocellular carcinoma (Inoue et al. 2001; Schöniger-Hekele et al. 2005), multiple myeloma (Drucker et al. 2006), melanoma (Xu and Hynes 2007), and glioblastoma (Staflin et al. 2009). In hepatocellular carcinoma, the interaction of CD81 with PI4KII suppresses cell motility by promoting the formation of CD81-enriched vesicles that sequestered actinin-4. It is discussed that the CD81 association-mediated redistribution of PI4KII to intracellular vesicles negatively affects actin-bundling activity of actinin (Fraley et al. 2003; Janmey and Lindberg 2004; Mazzocca et al. 2008). In melanoma, GPR56 forms a complex with Gaq and CD81. GPR56 binds tissue transglutaminase 2 (TG2), a major crosslinking enzyme in the ECM. The binding of the GPR56– Gaq–CD81 complex to TG2 may support adhesion, thereby interfering with tumor cell migration (Xu and Hynes 2007). CD81, but also CD82, have been described to cause multiple myeloma death. The antimyeloma effect of CD81 and CD82 involves downregulation of Akt, activation of FoxO (forkhead box, subgroup O) transcription, and a decrease in the active mechanistic target of rapamycin (mTOR). This is surprising because inhibitors of mTOR frequently activate Akt and vice versa (Lishner et al. 2008). The authors speculate that CD81/CD82 may act as a switch in signaling pathways rather than affect a single factor. Though the complexity of tetraspanin networks could allow for such switching, further studies are required before considering therapeutic application. Another pathway whereby CD81 is involved in metastasis inhibition proceeds via its association with EWI-2 in glioblastoma. EWI-2 expression causes diminished cell motility and invasion. EWI-2 affects the organization of CD9 and CD81 as well as of MMP2 and MT1-MMP. EWI-2 strengthens tetraspanin–tetraspanin associations and the association of CD81 with MMP-2 and MT1-MMP becomes weakened (Kolesnikova et al. 2009). CD63 is expressed rather ubiquitously. It was first described as a molecule expressed on activated platelets as platelet glycoprotein 40, and in early stage human melanoma cells as melanoma antigen 491. CD63 is present in TEM as well as in late endosomes and lysosomes. It is also enriched in exosomes. Endocytosis
570
M. Zöller
is supposed to be clathrin dependent, though not exclusively (reviewed in Pols and Klumperman 2009). CD63 interacts with CD9, CD81, CD82, CD151 (Radford et al. 1996; Cannon and Cresswell 2001; Berditchevski and Odintsova 2007), several integrins (Berditchevski et al. 1995; Radford et al. 1995; Rubinstein et al. 1996; Skubitz et al. 1996), cells surface receptors including the chemokine receptor CXCR4 (Kitani et al. 1991; Levy and Shoham 2005a, b; Yoshida et al. 2008), kinases, APs, L6, syntenin, tissue inhibitor of metalloproteinase (TIMP)1, H,K-ATPase, and MT1-MMP (Skubitz et al. 1996; Hirst et al. 1999; Duffield et al. 2003; Takino et al. 2003; Jung et al. 2006; Latysheva et al. 2006; Lekishvili et al. 2008). An inverse correlation between CD63 and tumor progression has been described for ovarian cancer (Zhijun et al. 2007), lung cancer (Kwon et al. 2007), breast and colon cancer (Sordat et al. 2002; Sauer et al. 2003a, b), and melanoma (Hotta et al. 1988; Radford et al. 1995; Jang and Lee 2003). However, for melanoma the opposite finding has also been reported (Lewis et al. 2005). As recently reviewed (Pols and Klumperman 2009), potential pathways, whereby CD63 prohibits metastasis formation, may rely on integrin endocytosis, lysosomal degradation of MT1-MMP (Takino et al. 2003), and recruitment of TIMP-1 (Jung et al. 2006). Another pathway, whereby CD63 becomes involved in tumor progression relies on its association with syntenin-1. Syntenin-1 regulates cell migration and promotes metastasis (Koo et al. 2002; Boukerche et al. 2005). Syntenin-1 directly interacts with CD63 and overexpression of syntenin-1 decreases the rate of CD63 internalization (Latysheva et al. 2006). Besides a report on Tspan13 (TM4SF13) to suppress breast carcinoma invasion in vitro (Huang et al. 2007), to my knowledge nothing is known about the involvement of additional tetraspanins in metastasis suppression.
Tetraspanins and Tumor Progression CD151 and Tumor Cell Motility CD151/PETA-3 has a broad tissue distribution. It is expressed in epithelial cells and high in ECs, platelets, megakaryocytes, and monocytes (Horejsí and Vlcek 1991; Hasegawa et al. 1997; Maecker et al. 1997; Sincock et al. 1997). First evidence of CD151 as a metastasis-promoting molecule was derived from a blinded study. An antibody for an unknown target inhibited metastasis formation of a human epidermoid carcinoma line. The antibody was found to recognize CD151. The antibody inhibited cell migration without having any effect on cell adhesion or cell growth (Testa et al. 1999). Meanwhile, overexpression of CD151 has been described in many tumor types like nonsmall cell lung cancer (Tokuhara et al. 2001), where the overall survival rate was much lower for patients with CD151+ as compared to patients with CD151- tumors. Overexpression of CD151 in squamous cell carcinoma and prostate cancer correlates with a poor prognosis, CD151 expression being a better predictor than histological grading (Ang et al. 2004).
26 Tetraspanins and Cancer Metastasis
571
In brain, mammary, pancreatic and colorectal cancer (Hashida et al. 2003; Bredel et al. 2005; Gesierich et al. 2005; Yang et al. 2008), too, high CD151 expression mostly is associated with a poor prognosis (Zöller 2006; Lazo 2007; Le Naour and Zöller 2008). CD151 also has been shown to inhibit tumor progression in several animal models (Testa et al. 1999; Kohno et al. 2002; Zijlstra et al. 2008). Several lines of evidence point towards a link between MMPs and CD151. CD151 contributes to pericellular activation of MMPs by associating with proMMP7, which results in activation of MMP7; this can be prevented by antiCD151 antibodies. Capture and activation of MMP7 at the cell membrane allows for focal proteolysis of the surrounding ECM (Shiomi et al. 2005). In addition, CD151 influences MMP9 expression, which was inhibited in the presence of FAK, Src p38, and JNK inhibitors and was mediated by c-Jun binding to AP1 sites in the MMP9 promoter. Signaling was initiated via CD151-associated a3b1 or a6b1 and was stimulated by CD151 homodimerization (Fitter et al. 1995; Hong et al. 2006; Yang et al. 2008). Reduced expression of MMP2, MMP7, and MMP9 in a CD151knockdown carcinoma line confirmed the involvement of CD151 in MMP expression, complex formation, and colocalization at the leading edge of lamellipodia (Shiomi et al 2005; Hasegawa et al. 2007). Transfection of FAK competent and deficient fibroblasts with CD151 cDNA provided evidence that FAK is needed for CD151-mediated increased migration, matrigel invasion, and metastasis formation (Kohno et al. 2002). Further studies confirmed that CD151 has no effect on steady state a3 and a6 integrin chain expression, but is important for proper localization of laminin5-binding integrins during tumor cell–stromal cell interactions, where under stimulatory conditions migration and invasion are strengthened by a3b1 cointernalization with CD151 (Hasegawa et al. 2007). Accordingly, CD151-knockdown cells display impaired motility, anomalously persistent adhesive contacts, and impaired a3b1 internalization (Winterwood et al. 2006). Notably, too, CD151 regulates glycosylation of a3b1 with reduced a3b1 glycosylation in CD151-knockdown cells showing strongly impaired migration towards laminin (Baldwin et al. 2008). Confirming the importance of CD151 for integrin traffic, mutating the CD151 sorting motif markedly attenuated a3b1, a5b1, and a6b1 endocytosis and accumulation in intracellular vesicular compartments (Liu et al. 2007). Thus, CD151 plays a critical role in integrin recycling as a mechanism to regulate tumor cell migration. CD151 is also an important regulator of collective tumor cell migration. Monolayers of CD151-knockdown cells display strikingly increased remodeling rates and junctional instability, which is caused by excessive RhoA activation and loss of actin organization at cell–cell junctions. There is evidence that CD151 regulates the stability of tumor cell–cell interaction through its association with a3b1 (Johnson et al. 2009). The importance of CD151 for metastasis formation has been confirmed in vivo (Zijlstra et al. 2008). A CD151 blocking antibody inhibits matrix-mediated migration, but has no impact on extravasation. Migration inhibition is due to a failure to detach at the rear end. As migration of CD151-knockout cells is unimpaired, the authors suggest that – if present – CD151 might recruit partner molecules that
572
M. Zöller
control deadhesion, but are hampered in their activity in the presence of the CD151 blocking antibody (Zijlstra et al. 2008). Taking into account that rear end adhesion was not restricted to selective components of the ECM, one could hypothesize that recruitment of MMPs may only be hampered by a CD151-specific antibody if CD151-associated integrins are involved in the adhesion process, which might be a consequence of a direct interaction with CD151, but not with other tetraspanins. The contribution of CD151 to cancer metastasis provides another convincing example of the facilitator role of tetraspanins. CD151 regulates cell migration, mostly through its association with a3b1, a6b4, and MMPs. The TEM location, which facilitates the recruitment of integrins, additional transmembrane and cytosolic proteins like E-cadherin, b-catenin, kinases, and phosphatases, in multimolecular complexes, contributes to this dominating theme (Hemler 2005). Tspan8 and Metastasis Expression of Tspan8, formerly CO-29 in human (Sela et al. 1989; Szala et al. 1990) and D6.1A in the rat (Matzku et al. 1989), has not been systemically evaluated in humans. It is relatively broad in rats, but it is not expressed in the epidermis, lymphocytes, and platelets (Claas et al. 1996). Human and rat Tspan8 associates with CD9, CD81, CD151, and several integrins including a3b1 and a6b4, but both a3b1 and a6b4 associations are weaker than described for CD151. Nonintegrin partners are EWI-F, EpCAM, CD13, CD44, PKC, and PI4KII (Claas et al 1998, 2005; Herlevsen et al. 2003; Schmidt et al. 2004; Gesierich et al. 2005; Jung et al. 2009). The Yxxf sorting motif of Tspan8 is located close to the TM region, which could hamper binding of the µ subunit of the AP-2 adaptor complex (Nakatsu and Ohno 2003; Berditchevski and Odintsova 2007). Overexpression of Tspan8 (CO-029) in colorectal cancer was described in 1989 (Sela et al. 1989) and was subsequently confirmed for colorectal, pancreatic, hepatocellular carcinoma, and esophageal cancer (Kanetaka et al. 2001; Gesierich et al. 2005; Kuhn et al. 2007; Zhou et al. 2008; reviewed in Zöller 2006, 2009). Tspan8 overexpression correlates with poor differentiation and intrahepatic spread of hepatoma rather than with the primary tumor growth. Only a Tspan8overexpressing hepatoma clone develops intrahepatic metastases (Kanetaka et al. 2003). Profiling the tetraspanin web of human colon cancer revealed Tspan8 expression in two metastatic, but not a nonmetastatic subline (Le Naour et al. 2006). Increased CO-029 expression in a metastasis versus the primary tumorderived colon carcinoma line supports a role in tumor progression (Huerta et al. 2003). First evidence that Tspan8 acts via associating molecules is derived from the observation that expression of an EpCAM–claudin-7 complex in human colorectal cancer promotes tumor progression only when Tspan8 is associated in TEM (Kuhn et al. 2007). Tspan8 can exert four tumor growth-promoting features, support of tumor cell proliferation, apoptosis protection, angiogenesis induction, and strengthening tumor cell motility. Only the latter is directly related to the metastatic process.
26 Tetraspanins and Cancer Metastasis
573
Increased Tspan8 expression in a dedifferentiated rat hepatoma cell line promoted cell proliferation (Tanaka et al. 2002). Furthermore, an interaction with platelets and leukocytes was suggested to provide tumor cells with a survival advantage in the hostile environment encountered during metastatic spread (Kanetaka et al. 2001, 2003). Several groups observed high Tspan8 expression to be associated with increased apoptosis resistance (Huerta et al. 2003; Kuhn et al. 2007), likely via a Tspan8-associated EpCAM–claudin-7 complex. This has been deduced from the striking decrease in drug resistance upon a knockdown of EpCAM or claudin-7 in human and rat cancer lines, which is accompanied by reduced PI3K activation and loss of Akt and downstream anti-apoptotic protein phosphorylation. Signals are initiated by the recruitment of the EpCAM-claudin-7 complex into TEM, which is accompanied by claudin-7 phosphorylation, possibly via Tspan8-associated PKC (Ladwein et al. 2005; Kuhn et al. 2007; Nübel et al. 2009). Tspan8-promoted tumor cell motility and liver metastasis formation may proceed via its association with a6b4, as it is seen in a Tspan8 and a6b4 overexpressing tumor line, but not a Tspan8 or a6b4, only, expressing line (Herlevsen et al. 2003; Gesierich et al. 2005). Tspan8 associates with a6b4 only after stimulation and disassembly of hemidesmosomes, which is accompanied by transient internalization of the Tspan8–a6b4 complex and increased motility (Huerta et al. 2003; Herlevsen et al. 2003). This kind of continuing internalization to the endosomal compartment and rapid recycling back to the cell surface via a short loop recycling machinery under the control of rab4 has been described for several integrins (Caswell and Norman 2008). An additional feature of pro-metastatic activity of Tspan8, angiogenesis induction, will be detailed in the following section. However, it should be mentioned that the Tspan8-mediated support of tumor cell migration or angiogenesis strikingly depends on the tumor cell’s integrin profile. Angiogenesis induction dominates in the absence of a6b4 and may be actively suppressed by a6b4 (Gesierich 2006). These comparably few studies point towards metastasis-promoting activities of Tspan8. In line with the general feature of tetraspanins, the TEM localization of Tspan8 is decisive and functional activity is determined by complex formation with associating molecules.
Tetraspanins, Premetastatic Niche, Angiogenesis, Thrombosis, and Exosomes Tetraspanins and Exosomes The MVB derive from vesicles sorted from the trans-Golgi network or from internalized membranes. MVB fuse with lysosomes for protein degradation (Stahl and Barbieri 2002) or their intraluminal vesicles fuse with the plasma membrane and are released as exosomes (Denzer et al. 2000; Stoorvogel et al. 2002; Lakkaraju and Rodriguez-Boulan 2008; Simpson et al. 2009). The latter process, first described for
574
M. Zöller
the release of the transferrin receptor during reticulocyte maturation, was considered as a mode to eliminate obsolete proteins (Pan et al. 1985). Exosome research became highly stimulated when it was noted that antigen presenting cells release exosomes derived from the MVB of the MHC class II compartment, which function similar to antigen presenting cells and stimulate T cells in vitro and in vivo (Denzer et al. 2000; André et al. 2002; Admyre et al. 2006; Iero et al. 2008; Schorey and Bhatnagar 2008). Exosomes are abundantly released by tumor cells (de Gassart et al. 2004; Keller et al. 2006; Wieckowski and Whiteside 2006; Lakkaraju and Rodriguez-Boulan 2008; Schorey and Bhatnagar 2008; van Niel et al. 2006). The molecular composition of exosomes reflects their origin from intraluminal vesicles. Besides a common set of membrane and cytosolic molecules, which includes several tetraspanins, like CD9, CD37, CD53, CD63, CD81, CD82, CD151, and Tspan8, exosomes harbor subsets of proteins that are linked to cell type-specific functions (Schorey and Bhatnagar 2008; Mathivanan et al. 2010) (Fig. 26.2). Exosomal proteins maintain their functional activity. Thus, microglial-derived exosomal CD13, a zink-dependent metalloprotease (Zhang and Xu 2008) is active in cleaving neuropeptides (Potolicchio et al. 2005); ADAM10 (a disintegrin and metalloprotease), a member of these multidomain membrane proteins, which share a disintegrin and a zinc metalloprotease domain (Edwards et al. 2008), cleaves proteins, like CD44 and L1/CD171 within exosomes (Stoeck et al. 2006). Another notable feature of exosomes is expression of phosphatidylserine at the outer membrane leaflet, which appears to be essential for exosome budding in the late endosomes (Fomina et al. 2003) and also can trigger exosome uptake by phosphatidylserinebinding proteins like scavenger receptors, integrins, and complement receptors (Zakharova et al. 2007). Most importantly, exosomes contain mRNA and microRNA, so-called shuttle RNA that is transferred to the target cell. The horizontal transfer of exosomal genetic material between cells can induce exogenous gene expression and mediate RNA silencing (Ratajczak et al. 2006; Deregibus et al. 2007; Valadi et al. 2007; Burghoff et al. 2008; Lakkaraju and Rodriguez-Boulan 2008). Exosomemediated horizontal gene transfer is specific with respect to the target cell, such that it is transcribed in one, but not another type of target cells (Valadi et al. 2007; Simons and Raposo 2009). The relative abundance of proteins, mRNA, and miRNAs differs between exosomes and donor cells, which implies active sorting into MVBs (Lakkaraju and Rodriguez-Boulan 2008). The mechanisms underlying selective sorting of mRNA and miRNA into exosomes are unknown (Subra et al. 2007). Instead, it is known that sorting of proteins into exosomes can rely on monoubiquitination (Gruenberg and Stenmark 2004; Hurley and Emr 2006; Smalheiser 2007), localization in cholesterol-rich membrane microdomains, or higher order oligomerization (Fang et al. 2007; Smalheiser 2007). Though many questions remain to be answered, it is well appreciated that exosomes may be one of the most important delivery systems (Johnstone 2006; Belting and Wittrup 2008; Simpson et al. 2008; Pap et al. 2009; Seow and Wood 2009; Xiao et al. 2009). Tetraspanins are a constitutive component of exosomes (Escola et al. 1998). Though targeting into exosomes apparently differs for individual tetraspanins and their associated molecules (Abache et al. 2007), tetraspanin enrichment in
26 Tetraspanins and Cancer Metastasis
575
exosomes is well in line with the preferential internalization of oligomeric proteins and/or proteins localized in cholesterol-rich membrane domains and may be supported by the internalization motif of some tetraspanins and/or their associated proteins. As exosomes could well contribute to premetastatic niche preparation and have been demonstrated to be most important in angiogenesis as well as in coagulation and homeostasis (Al-Nedawi et al. 2009), defining the function of tetraspanins in exosomes becomes demanding. Though not yet proven, tetraspanins have been implicated in the adhesion of exosomes to target cells (Pols and Klumperman 2009; Zöller 2009). Tetraspanins and the Premetastatic Niche Settlement of metastasizing tumor cells is facilitated by the establishment of special niches in (pre)metastatic organs (Bissell and Labarge 2005). Niche preparation involves stimulation of local fibroblasts by tumor-derived factors and chemokines that attract tumor cells and hematopoietic progenitors (Kaplan et al. 2005; Kaplan et al. 2006), lysyl oxidase being important for marrow cell recruitment (Erler et al. 2009). Nonetheless, information on long-distance communication between a tumor and host organs is still limited. We suggest that tumor cells avail special delivery systems and hypothesize that a concerted activity between tumor-derived factors and exosomes is required (Fevrier and Raposo 2004). An involvement of exosomes in metastasis was first described for platelet-derived exosomes. These transferred the aIIb integrin chain to lung cancer cells, stimulated the MAPK pathway and MT1-MMP expression, increased cyclinD2 expression, stimulated angiogenic factor expression as well as adhesion to fibrinogen and human umbilical vein ECs (Janowska-Wieczorek et al. 2005). A direct transfer of metastatic capacity by exosomes was demonstrated for B16 melanoma cells. Exosomes derived from a highly metastatic variant transferred metastatic capacity to low metastatic B16F1 cells. Lung metastasis formation by B16F1 was accompanied by protein uptake from exosomes of the metastasizing subclone (Hao et al. 2006). Our own findings suggest that Tspan8 and/or CD151-containing exosomes, depending on the exosomal tetraspanin web, also contribute to premetastatic niche formation (Jung et al. 2009). As the metastasizing capacity of a rat pancreatic adenocarcinoma, which metastasizes via the lymphatic system, but does not grow locally, is strikingly reduced by a knockdown of CD44v4-v7, this model allowed to define tumor cellderived components required for (pre)metastatic niche formation. Conditioned medium of the metastasizing subline strongly supports (pre)metastatic niche preparation, where fractionation of the conditioned medium revealed that the soluble matrix of the metastasizing subline can also cooperate with exosomes of the nonmetastatic subline. This implies that exosomes are the final actors, but require a soluble matrix, whose assembly depends on CD44v. A dominating component of the exosomes is the tetraspanins CD151 and Tspan8, which coimmunoprecipitate, besides others, with a6b4 (Jung et al. 2009). These exosomes preferentially bind to and are taken up by lymph node stroma cells and lung fibroblasts and binding is inhibited by anti-Tspan8 and even more efficiently by anti-CD151 and anti-a6b4 (unpublished finding).
576
M. Zöller
According to published evidence and our own findings, we suppose that exosomal tetraspanins and their associated molecules are most important for selective targeting and delivery of exosomes. Our work on the role of Tspan8 in angiogenesis strengthens this hypothesis. Tetraspanins and Angiogenesis CD151 has repeatedly been reported to be important in angiogenesis induction. This has been suggested to be due to CD151 being expressed abundantly on EC and particularly on EC–cell junctions (Fitter et al. 1995; Maecker et al. 1997). However, patients with mutations of the CD151 gene show skin and kidney disorders, but no obvious defects in vasculogenesis (Karamatic Crew et al. 2004) and in CD151knockout mice vasculogenesis is unimpaired. Instead, defects are seen in angiogenesis (Takeda et al. 2007a, b; Wright et al. 2004a, b; Sachs et al. 2006). Thus, CD151 expression of the tumor-bearing host facilitates tumor growth due to angiogenesis induction (Takeda et al. 2007a, b). CD151 supports EC invasiveness, migration, cable formation, matrigel contraction, tube formation and sprouting, which activities are all impaired in CD151-knockout mice (Takeda et al. 2007a, b). Selective defects in adhesion-dependent signaling activation of PKB/c-Akt, e-NOS, Rac, and Cdc42 on laminin substrates contribute to impaired angiogenesis induction (Takeda et al. 2007a, b; Zheng and Liu 2007). Also, overexpression of CD151 promotes neovascularization and improves blood perfusion in an ischemia model (Lan et al. 2005). Taken together, like in tumor cells, CD151 supports functional activity of ECs via the association with integrins, particularly laminin-binding integrins, that organization is distorted in CD151-/- mice. Instead, EC proliferation, ERK, p38 MAPK, FAK, src, and Raf activation are unimpaired in CD151-/- ECs (Takeda et al. 2007a, b). Besides the contribution of EC CD151 in angiogenesis, expression of tetraspanins in tumor cells and tumor-derived exosomes can be most important in angiogenesis. Rats receiving a Tspan8 overexpressing tumor line develop disseminated intravascular coagulation, which could be prevented by a Tspan8-specific antibody (Claas et al. 1998). A prothrombotic state that can culminate in disseminated intravascular coagulation is frequent in cancer patients. Though multifactorial, tumor-initiated angiogenesis and the leakiness of tumor vessels are considered to be important (De Cicco 2004). In fact, Tspan8 is a strong angiogenesis inducer that contributes to a systemic angiogenic switch by delivery of Tspan8 in tumor cell-derived exosomes (Gesierich et al. 2006). Based on these observations, we explored how Tspan8-containing, tumorderived exosomes interact with EC. Tspan8 contributes to a selective recruitment of proteins and mRNA into exosomes. Furthermore, possibly among others, Tspan8-associated CD49d is essential in exosome–EC binding. Transient recovery of selective Tspan8-exosomes enriched mRNA in EC revealed that Tspan8–CD49d complex-containing exosomes are internalized by EC. Exosome uptake induced VEGF-independent regulation of several angiogenesis-related genes, which was accompanied by enhanced proliferation, migration, and sprouting of EC.
26 Tetraspanins and Cancer Metastasis
577
Importantly, Tspan8 exosomes also bind to EC progenitors and suffice for EC progenitor maturation (Nazarenko et al. 2010). Both CD151 and Tspan32 (TSSC6) also are important for normal platelet functions. Though CD151-knockout mice also revealed only mild deficiencies (Karamatic Crew et al. 2004; Wright et al. 2004a, b; Sachs et al. 2006), CD151and Tspan32-knockout mice show defective platelet aggregation, impaired spreading on fibrinogen, and delayed clot retraction (Lau et al. 2004; Goschnick et al. 2006). Both tetraspanins associate with aIIbb3 and modulate outside-in signaling. Tspan32 additionally is required for stable platelet thrombus formation. CD63 may regulate platelet spreading by its association with aIIbb3 (Israels and McMillan-Ward 2005). Platelet-derived exosomes constitute about 70–90% of circulating exosomes in the plasma (Berckmans et al. 2001) with a life span of about 30 min (Flaumenhaft 2006). The procoagulant activity of platelet-derived exosomes is well known. It is suggested that exosomes provide negatively charged phospholipids, which are required for factor IXa and Xa activation (Kessels et al. 1994; Shet et al. 2003). Though still controversial, the therapeutic efficacy of antiglycoprotein IIb/IIIa is suggested to be a consequence of altered platelet exosome formation (Craft and Marsh 2003; Morel et al. 2004; Razmara et al. 2007). Taking into account the abundance of platelet-derived exosomes and their functional activity in coagulation, it will be important to analyze the contribution of exosomal CD151, Tspan32 and CD63. Taken together, by their localization in TEM supported by an internalization motif of some tetraspanins, tetraspanins are prone for internalization. Independent of the donor cell, tetraspanins are enriched in exosomes and the tetraspanin web is mostly maintained (Abache et al. 2007). It remains to be clarified, why some endosomes are delivered to the degradation machinery of the cell, whereas others are preferentially released as exosomes. It is also unanswered whether tetraspanins are involved in this vesicle sorting process (Gibbings et al. 2009; Simons and Raposo 2009). Instead, our data provide convincing evidence that tetraspanins are important in target cell selection, where in the two systems, which we analyzed, tetraspanin-associated integrins account for stable binding. Additional tetraspanin partners have not been excluded. It is also likely that other tetraspanin partners might account for tetraspanin uptake by different cells. Tetraspanins are involved in membrane fusion events (Hemler 2005; Levy and Shoham 2005a, b), which could facilitate the exosome uptake by target cells. However, further elaboration of the fusogenic activity of tetraspanins is required. So far, there is no evidence that the fate of the target cell and its modulation by the transferred proteins and/or the exosomal mRNA and/or microRNA is influenced by the tetraspanins contained in the exosome. Yet, it remains to be explored, whether tetraspanins contribute not only to the protein composition of exosomes, but also to the incorporation of mRNA and miRNA. In view of the exciting power of exosomes and our preliminary evidence for the engagement of tetraspanins in target cell selection, it is demanding to further elaborate the function of tetraspanins in exososmes as a possible tool to interfere with pathological angiogenesis and metastasis formation, two serious handicaps in cancer therapy, that rely on long-distance communication.
578
M. Zöller
Tetraspanins and Cancer Therapy Taking into account the importance of some tetraspanins in tumor progression and angiogenesis, it is demanding to consider these molecules as therapeutic targets. Due to their mode of activity as molecular facilitators this will not be an easy task. Indeed, the author is not aware of any standard therapy based on selectively attacking tetraspanins. As some tetraspanins function as metastasis suppressors, while others promote metastasis, the author will discuss these two aspects of tetraspanin-based therapies.
Rescuing the Metastasis Suppressor Gene CD82 Rescuing the metastasis suppressor gene CD82 offers itself for therapeutic interference, which requires an awareness of the regulation of CD82 gene transcription as well as of the mechanisms that downregulate CD82 expression in tumor cells (Tonoli and Barrett 2005; Liu and Zhang 2006). Regulation of CD82 transcription and silencing are complex processes (Gao et al. 2003; Tonoli and Barrett 2005; Liu and Zhang 2006). There is no evidence for gene mutation or loss of heterozygosity (Tagawa et al. 1999; Liu et al. 2000) and hypermethylation of CpG islands in the CD82 gene has only been seen in patients with multiple myeloma, where combined demethylation and deacetylation induced increased expression of CD82 mRNA (Jackson et al. 2000; Drucker et al. 2006). Histone deacetylase has also been described to target CD82 in glioma (Gensert et al. 2007). CD82 downregulation has also been related to the p53 status. Binding motifs for the transcription factor AP2 in the CD82 promoter function synergistically with p53 and junB such that the absence of wild-type p53 and/or loss of junB and AP2 protein expression correlate with CD82 mRNA downregulation (Marreiros et al. 2003, 2005). There have been some controversial results on the involvement of NFkB in CD82 transcription, which likely can be explained by the nature of the recruited cofactors. In nonmetastatic cells, IL-1b supports the recruitment of a Tip60 (HIV-1 TAT-interactive protein 60)/Fe65-Pontin complex, which acts as a coactivator together with NFkB p50 and accounts for the dismissal of the corepressor N-Cor/TAB2 (TAK1-binding AP)/HDAC3 (histone deacetylase 3) complex from NFkB p50, which would turn off CD82 transcription. In metastasizing tumor cells, Tip60 is downregulated and a b-catenin–reptin complex replaces the Tip60– Pontin complex and represses NFkB activity (Kim et al. 2005; Telese et al. 2005). An additional mode of CD82 downregulation could rely on alternative splicing. Expression of a splice variant lacking exon 7, frequently seen in metastatic tissue, confers increased metastatic potential (Lee et al. 2003). The potential therapeutic efficacy of CD82 was already demonstrated. CD82-transfected Lewis lung carcinoma cells loose the capacity to form lymph node metastasis (Takeda et al. 2007a, b). Furthermore, nerve growth factor has been shown to rescue CD82 expression and to abrogate tumorigenicity (Sigala et al. 1999). Besides reviving CD82 expression at the transcriptional level, CD82 expression may also become rescued by proteasome inhibitors or by targeting specific
26 Tetraspanins and Cancer Metastasis
579
components of the ubiquitin system, as ubiquitin ligase gp78 regulates CD82 expression (Tsai et al. 2007).
Interfering with Metastasis-Promoting Activities of Tetraspanins Therapeutic approaches concerned about interference with metastasis-promoting activities of tetraspanins are mostly based on antibodies, recombinant soluble ECL2, or posttranscriptional gene silencing via siRNA (Hemler 2008). Tetraspanin-specific antibodies, though not yet commercially available, have been shown in several instances to be of potential clinical relevance. Intratumoral application of anti-CD9 inhibited colon carcinoma growth (Ovalle et al. 2007), anti-CD37 improved the survival of B-CLL xenografted mice (Levy et al. 1998), and anti-CD151 interfered with metastasis formation (Testa et al. 1999; Kohno et al. 2002; Zijlstra et al. 2008). Though the underlying mechanisms have not been fully elucidated, it has been suggested that antibodies may interfere with the lateral associations of tetraspanins or promote clustering of tetraspanins and tetraspanin-associated molecules in TEM and thereby interfere with the activity not only of the targeted tetraspanin, but also of associated molecules including cytoplasmic partners. In line with this suggestion, tetraspanin antibodies have in some instances been shown to exert stronger effects than the knockout of an individual tetraspanin, e.g., anti-CD81 (Oren et al. 1990; Boismenu et al. 1996; Miyazaki et al. 1997; Tsitsikov et al. 1997; Levy et al. 1998). Taking this into account, one has to be aware that particularly the activity of tetraspanin-specific antibodies may vary depending on the recognized epitope (Serru et al. 1999; Yauch et al. 2000; Geary et al. 2001), which may enhance or block the effect of a tetraspanin as demonstrated for anti-CD151-promoting adhesion (Zijlstra et al. 2008) and for antiCD9 that can amplify the tumor suppressor function (Ovalle et al. 2007). Besides there blocking or enhancing activity, tetraspanin-specific antibodies, like therapeutic antibodies recognizing other molecules, have been described to induce apoptosis (Murayama et al. 2004) and to support complement and antibodydependent cellular cytotoxicity (Zhao et al. 2007). Finally, antibodies can be used as drug transporter as reported for 131I-labeled anti-CD37 (Press et al. 1989) or for transporting nanoparticles with siRNA (Peer et al. 2008). The latter has not yet been explored for tetraspanins. However, several approaches have been undertaken to silence tetraspanins via siRNA. CD9 silencing resulted in pronounced ovarian cancer dissemination (Furuya et al. 2005) and CD151 silencing interfered with integrin-dependent adhesion and migration (Winterwood et al. 2006). Though not metastasis-related, lentiviral CD81 shRNA delivery into the nucleus accumbens or the ventral tegmental resulted in a significant decrease in locomotor activity (Bahi et al. 2005). So far, the soluble form of the large extracellular domain (ECL2) of tetraspanins as a competitor has mainly been tested with respect to leukocyte–EC interaction via CD9 and CD151 (Barreiro et al. 2005), egg–sperm fusion (Zhu et al. 2002) and infectivity, where the ECL2 by exerting direct and indirect effects may be superior to antibodies (Molina et al. 2008).
580
M. Zöller
Other therapeutic setting based on tetraspanins is currently being discussed. These include modulation of AA important for TM folding (Tarasova et al. 1999), where perturbation of TM interactions of CD82 significantly affect the molecules intrinsic activities (Bari et al. 2009). The authors argue that a therapeutic perturbation of TM interactions may open a new avenue to prevent cancer invasion, which could be far easier approached than a blockade of individual signaling pathways. Modulation of the PDZ domain (Dev 2004; Latysheva et al. 2006), of key interaction sites in the ECL2 (Yauch et al. 2000; Seigneuret 2006), of palmitoylation sites (Berditchevski et al. 2002; Charrin et al. 2002; Yang et al. 2002, 2004; Kovalenko et al. 2005) including targeting of the responsible acyltransferase (Sharma et al. 2008) are additional therapeutic approaches currently being discussed. Last, not least, taking into account the increasingly appreciated role of exosomes as intercellular communicators and the strong presence of tetraspanins in exosome membranes, it is tempting to speculate that tetraspanins could be used as delivery system. This requires further elaboration of the engagement of tetraspanins and the associated molecule that together selectively bind to and become internalized by selective targets (Nazarenko et al. 2010). Knowledge on the exosomal binding to and uptake of tetraspanin complexes by selective target cells could well allow to generate competitive exosomes carrying desired siRNA or other drugs that interfere with exosome-initiated premetastatic niche preparation, angiogenesis, and thrombosis.
Conclusion Tetraspanins function as molecular facilitators that assemble a web including many distinct families of TM proteins in specialized membrane microdomains that serve as a scaffold for signal transducing and cytoskeletal proteins and are prone for internalization. The reversibility of TEM and their composition, which depends on the cell’s activation state, the abundance of associating molecules and their ligands adds a major constraint in defining tetraspanin functions. Nonetheless, modulation of cell motility, cell fusion and, as defined more recently, intercellular communication via exosomes may well cover the essential activities of tetraspanins. Taking a simplified view (exemplified in Fig. 26.3), the involvement of tetraspanins in these
Fig. 26.3 Modes of tetraspanin activity. (a) Tetraspanins may act as a receptor for defined ligands: CD82 binds DARC on endothelial cells (ECs). Thereby TBX2 becomes downregulated, ARF and p21 are released from repression and force tumor cell senescence. (b) Tetraspanins can directly influence adhesion, signal transduction, or gene transcription via associated molecules: CD151 can associate with proMMP7. This facilitates matrix degradation and strengthens invasiveness. (c) Tetraspanins only indirectly initiate activities via the recruitment of different molecules into TEM, a process frequently involving gangliosides. High level of a GM3–CD9 complex is accompanied by strong a3b1 ECM adhesion. High-level GM3 expression also initiates recruitment of Csk into TEM, where it inhibits src activation with the consequence of reduced rac and PI3K/Akt activation, which results in pronounced adhesion and increased apoptosis susceptibility. (d) Tetraspanins initiate internalization and relocation in distinct
26 Tetraspanins and Cancer Metastasis
a
581
b
c
TEM interactions
lateral tetraspanin interaction
tetraspanin-ligand interaction
ECM strong adhesion
endothelium ECM CD82
GM3high
DARC
α3β1high
CD9high
α3β1 degradation CD151
proMMP7
src ARF p21
nucleus TBX2
Csk rac PI3K
invasion
actin Akt
forced senescence
adhesion / survival
d
e
exosomal tetraspanin activity
tetraspanin internalization α3β1 α5β1 α6β1
CD151
non-metastatic, angiogenesis
non-metastatic α6β4
α4β1
α3β1
α6β1 CD151
CD151
α6β1
α4β1
α3β1 CD151
CD9
CD9
Tspan8
Tspan8 cDNA
Yxxφ
Yxxφ MMP9 transcription
internalization
motility
α4β1
exosome
exosome
mRNA
endothelium
endosome
enriched
recycling
VCAM exosome
Fig. 26.3 (continued) membrane regions of associated molecules: CD151 recruits the major laminin-binding integrins a3b1 and a6b4; via the CD151 Yxxf motif, the CD151–integrin complex becomes internalized. The integrin redistribution with re-expression in the leading lamella is associated with reduced adhesion and increased motility. (e) Tetraspanins initiate recruitment into MVB and release of TEMs in exosomes: Transfection of a nonmetastasizing tumor line with Tspan8 is accompanied by systemic angiogenesis and thrombosis as well as by upregulation of a4b1 and MMP9. Both a4b1 and MMP9 are enriched in exosomes and exosomes suffice for angiogenesis induction, where exosomes bind via Tspan8-associated a4b1 to endothelial VCAM as the initial step of exosome uptake and modulation of gene transcription in ECs. Thus, exosomal tetraspanins and tetraspanin-associated molecules are of major importance in target cell selection and target cell modulation
582
M. Zöller
actions can follow five routes: (a) tetraspanins may act as a receptor for defined ligands, this is rather exceptionally the case; (b) tetraspanins directly influence adhesion, signal transduction, or gene transcription via associated molecules; (c) tetraspanins only indirectly initiate activities via the recruitment of different molecules into TEM, a process that frequently involves gangliosides; (d) tetraspanins initiate internalization and relocation in distinct membrane regions of associated molecules; and (e) tetraspanins initiate recruitment into MVB and release of TEMs in exosomes. Exosomal tetraspanin-associated molecules may be of major importance in target cell selection and exosomal tetraspanins in exosome fusion with the target cell. All these activities have been demonstrated to contribute to tetraspaninmediated metastasis inhibition and promotion, to premetastatic niche formation, to angiogenesis and the tumor-associated prothrombotic state. Nonetheless, one of the key questions, why some tetraspanins rather consistently suppress (CD82, CD81, CD9, CD63) or promote (CD151, Tspan8) tumor progression remains unanswered. Neither the structure of the tetraspanins (CD82 and Tspan8 belong to the same subgroup) (Seigneuret et al. 2001) nor the assembly of the associating TM and signal transduction molecules in TEM differs fundamentally between CD9, CD63, CD81, CD82, CD151, and Tspan8 (Hemler 2005; Levy and Shoham 2005a, b). A well-positioned sorting motif is present in CD82 and CD151 (Berditchevski and Odintsova 2007) and, thus, cannot account for the distinct activity. Taking into account that the most consistent distinction between metastasis-suppressing and -promoting tetraspanins relates to the strong adhesion initiated by CD82 or CD9 versus CD151- and Tspan8-initiated transient internalization and motility, it is tempting to speculate that distinct vesicular journeys of internalized tetraspanins may be decisive for metastasis inhibition versus promotion. This hypothesis urgently requires to become answered, which may provide a solid ground for therapeutic interference with tetraspanin activities in tumor progression. Acknowledgments This work was supported by the Deutsche Forschungsgemeinschaft (grant ZO 40/12-1), the Deutsche Krebshilfe, and the Tumorzentrum Heidelberg/Mannheim.
References Abache T, Le Naour F, Planchon S et al (2007) The transferrin receptor and the tetraspanin web molecules CD9, CD81, and CD9P-1 are differentially sorted into exosomes after TPA treatment of K562 cells. J Cell Biochem 102:650–664 Adachi M, Taki T, Ieki Y et al (1996) Correlation of KAI1/CD82 gene expression with good prognosis in patients with non-small cell lung cancer. Cancer Res 56:1751–1755 Adachi M, Taki T, Huang C et al (1998) Reduced integrin alpha3 expression as a factor of poor prognosis of patients with adenocarcinoma of the lung. J Clin Oncol 16:1060–1067 Admyre C, Johansson SM, Paulie S et al (2006) Direct exosome stimulation of peripheral human T cells detected by ELISPOT. Eur J Immunol 36:1772–1781 Aharon A, Brenner B (2009) Microparticles, thrombosis and cancer. Best Pract Res Clin Haematol 22:61–69
26 Tetraspanins and Cancer Metastasis
583
Ahmad A, Hart IR (1997) Mechanisms of metastasis. Crit Rev Oncol Hematol 26:163–173 Albini A, Mirisola V, Pfeffer U (2008) Metastasis signatures: genes regulating tumor-microenvironment interactions predict metastatic behavior. Cancer Metastasis Rev 27:75–83 Al-Nedawi K, Meehan B, Rak J (2009) Microvesicles: messengers and mediators of tumor progression. Cell Cycle 8:2014–2018 André F, Schartz NE, Chaput N et al (2002) Tumor-derived exosomes: a new source of tumor rejection antigens. Vaccine 20 Suppl 4:A28–A31 André M, Le Caer JP, Greco C et al (2006) Proteomic analysis of the tetraspanin web using LC-ESI-MS/MS and MALDI-FTICR-MS. Proteomics 6:1437–1449 Ang J, Lijovic M, Ashman LK et al (2004) CD151 protein expression predicts the clinical outcome of low-grade primary prostate cancer better than histologic grading: a new prognostic indicator? Cancer Epidemiol Biomarkers Prev 13:1717–1721 Ardón-Alonso M, Yañez-Mó M, Barreiro O et al (2006) Tetraspanins CD9 and CD81 modulate HIV-1-induced membrane fusion. J Immunol 177:5129–5137 Arduise C, Abache T, Li L et al (2008) Tetraspanins regulate ADAM10-mediated cleavage of TNF-alpha and epidermal growth factor. J Immunol 181:7002–7013 Bahi A, Boyer F, Kolira M et al (2005) In vivo gene silencing of CD81 by lentiviral expression of small interference RNAs suppresses cocaine-induced behaviour. J Neurochem 92:1243–1255 Baldwin G, Novitskaya V, Sadej R et al (2008) Tetraspanin CD151 regulates glycosylation of (alpha)3(beta)1 integrin. J Biol Chem 283:35445–35454 Bandyopadhyay S, Zhan R, Chaudhuri A et al (2006) Interaction of KAI1 on tumor cells with DARC on vascular endothelium leads to metastasis suppression. Nat Med 12:933–938 Bari R, Zhang YH, Zhang F et al (2009) Transmembrane interactions are needed for KAI1/CD82mediated suppression of cancer invasion and metastasis. Am J Pathol 174:647–660 Barreiro O, Yáñez-Mó M, Sala-Valdés M et al (2005) Endothelial tetraspanin microdomains regulate leukocyte firm adhesion during extravasation. Blood 105:2852–2861 Bass R, Werner F, Odintsova E et al (2005) Regulation of urokinase receptor proteolytic function by the tetraspanin CD82. J Biol Chem 280:14811–14818 Belting M, Wittrup A (2008) J Nanotubes, exosomes, and nucleic acid-binding peptides provide novel mechanisms of intercellular communication in eukaryotic cells: implications in health and disease. Cell Biol 183:1187–1191 Berckmans RJ, Neiuwland R, Böing AN et al (2001) Cell-derived microparticles circulate in healthy humans and support low grade thrombin generation. Thromb Haemost 85:639–646 Berditchevski F (2001) Complexes of tetraspanins with integrins: more than meets the eye. J Cell Sci 114:4143–4151 Berditchevski F, Odintsova E (2007) Tetraspanins as regulators of protein trafficking. Traffic 8: 89–96 Berditchevski F, Bazzoni G, Hemler ME (1995) Specific association of CD63 with the VLA-3 and VLA-6 integrins. J Biol Chem 270:17784–17790 Berditchevski F, Tolias KF, Wong K et al (1997) A novel link between integrins, transmembrane-4 superfamily proteins (CD63 and CD81), and phosphatidylinositol 4-kinase. J Biol Chem 272:2595–2598 Berditchevski F, Odintsova E, Sawada S et al (2002) Expression of the palmitoylation-deficient CD151 weakens the association of alpha 3 beta 1 integrin with the tetraspanin-enriched microdomains and affects integrin-dependent signaling. J Biol Chem 277:36991–37000 Bérubé NG, Speevak MD, Chevrette M (1994) Suppression of tumorigenicity of human prostate cancer cells by introduction of human chromosome del(12)(q13). Cancer Res 54:3077–3081 Bienstock RJ, Barrett JC (2001) KAI1, a prostate metastasis suppressor: prediction of solvated structure and interactions with binding partners; integrins, cadherins, and cell-surface receptor proteins. Mol Carcinog 32:139–153 Bijlmakers MJ, Marsh M (2003) The on-off story of protein palmitoylation. Trends Cell Biol 13:32–42 Birchmeier C, Birchmeier W, Gherardi E et al (2003) Met, metastasis, motility and more. Nat Rev Mol Cell Biol 4:915–925
584
M. Zöller
Bissell MJ, Labarge MA (2005) Context, tissue plasticity, and cancer: are tumor stem cells also regulated by the microenvironment? Cancer Cell 7:17–23 Boismenu R, Rhein M, Fischer WH et al (1996) A role for CD81 in early T cell development. Science 271:198–200 Boucheix C, Rubinstein E (2001) Tetraspanins. Cell Mol Life Sci 58:1189–1205 Boukerche H, Su ZZ, Emdad L et al (2005) mda-9/Syntenin: a positive regulator of melanoma metastasis. Cancer Res 65:10901–10911 Brabletz T, Jung A, Spaderna S et al (2005) Opinion: migrating cancer stem cells – an integrated concept of malignant tumour progression. Nat Rev Cancer 5:744–749 Bredel M, Bredel C, Juric D et al (2005) Functional network analysis reveals extended gliomagenesis pathway maps and three novel MYC-interacting genes in human gliomas. Cancer Res 65:8679–8689 Briese J, Schulte HM, Sajin M et al (2008) Correlations between reduced expression of the metastasis suppressor gene KAI-1 and accumulation of p53 in uterine carcinomas and sarcomas. Virchows Arch 453:89–96 Burghoff S, Ding Z, Gödecke S et al (2008) Horizontal gene transfer from human endothelial cells to rat cardiomyocytes after intracoronary transplantation. Cardiovasc Res 77:534–543 Cannon KS, Cresswell P (2001) Quality control of transmembrane domain assembly in the tetraspanin CD82. EMBO J 20:2443–2453 Caswell P, Norman J (2008) Endocytic transport of integrins during cell migration and invasion. Trends Cell Biol 18:257–263 Charrin S, Le Naour F, Oualid M et al (2001) The major CD9 and CD81 molecular partner. Identification and characterization of the complexes. J Biol Chem 276:14329–14337 Charrin S, Manié S, Oualid M et al (2002) Differential stability of tetraspanin/tetraspanin interactions: role of palmitoylation. FEBS Lett 516:139–144 Charrin S, Manié S, Thiele C et al (2003) A physical and functional link between cholesterol and tetraspanins. Eur J Immunol 33:2479–2489 Chen Z, Mustafa T, Trojanowicz B et al (2004) CD82, and CD63 in thyroid cancer. Int J Mol Med 14:517–527 Cherukuri A, Carter RH, Brooks S et al (2004) B cell signaling is regulated by induced palmitoylation of CD81. J Biol Chem 279:31973–31982 Claas C, Herrmann K, Matzku S et al (1996) Developmentally regulated expression of metastasisassociated antigens in the rat. Cell Growth Differ 7:663–678 Claas C, Seiter S, Claas A et al (1998) Association between the rat homologue of CO-029, a metastasis-associated tetraspanin molecule and consumption coagulopathy. J Cell Biol 141: 267–280 Claas C, Stipp CS, Hemler ME (2001) Evaluation of prototype transmembrane 4 superfamily protein complexes and their relation to lipid rafts. J Biol Chem 276:7974–7984 Claas C, Wahl J, Orlicky DJ et al (2005) The tetraspanin D6.1A and its molecular partners on rat carcinoma cells. Biochem J 389:99–110 Clark KL, Oelke A, Johnson ME et al (2004) CD81 associates with 14–3–3 in a redox-regulated palmitoylation-dependent manner. J Biol Chem 279:19401–19406 Coombs GS, Covey TM, Virshup DM (2008) Wnt signaling in development, disease and translational medicine. Curr Drug Targets 9:513–531 Craft JA, Marsh NA (2003) Increased generation of platelet-derived microparticles following percutaneous transluminal coronary angioplasty. Blood Coagul Fibrinolysis 14:719–728 De Bruyne E, Andersen TL, De Raeve H et al (2006) Endothelial cell-driven regulation of CD9 or motility-related protein-1 expression in multiple myeloma cells within the murine 5T33MM model and myeloma patients. Leukemia 20:1870–1879 De Cicco M (2004) The prothrombotic state in cancer: pathogenic mechanisms. Crit Rev Oncol Hematol 50:187–196 Defilippi P, Di Stefano P, Cabodi S (2006) p130Cas: a versatile scaffold in signaling networks. Trends Cell Biol 16:257–263
26 Tetraspanins and Cancer Metastasis
585
de Gassart A, Géminard C, Hoekstra D et al (2004) Exosome secretion: the art of reutilizing nonrecycled proteins? Traffic 5:896–903 Denzer K, Kleijmeer MJ, Heijnen HF et al (2000) Exosome: from internal vesicle of the multivesicular body to intercellular signaling device. J Cell Sci 113:3365–3374 de Parseval A, Lerner DL, Borrow P et al (1997) Blocking of feline immunodeficiency virus infection by a monoclonal antibody to CD9 is via inhibition of virus release rather than interference with receptor binding. J Virol 71:5742–5749 Deregibus MC, Cantaluppi V, Calogero R et al (2007) Endothelial progenitor cell derived microvesicles activate an angiogenic program in endothelial cells by a horizontal transfer of mRNA. Blood 110:2440–2448 Deryugina EI, Quigley JP (2006) Matrix metalloproteinases and tumor metastasis. Cancer Metastasis Rev 25:9–34 Dev KK (2004) Making protein interactions druggable: targeting PDZ domains. Nat Rev Drug Discov 3:1047–1056 Devaux PF, Morris R (2004) Transmembrane asymmetry and lateral domains in biological membranes. Traffic 5:241–246 Dong JT, Lamb PW, Rinker-Schaeffer CW et al (1995) KAI1, a metastasis suppressor gene for prostate cancer on human chromosome 11p11.2. Science 268:884–886 Dong LM, Potter JD, White E et al (2008) Genetic susceptibility to cancer: the role of polymorphisms in candidate genes. JAMA 299:2423–2436 Drapkin R, Crum CP, Hecht JL (2004) Expression of candidate tumor markers in ovarian carcinoma and benign ovary: evidence for a link between epithelial phenotype and neoplasia. Hum Pathol 35:1014–1021 Drucker L, Tohami T, Tartakover-Matalon S et al (2006) Promoter hypermethylation of tetraspanin members contributes to their silencing in myeloma cell lines. Carcinogenesis 27: 197–204 Duffield A, Kamsteeg EJ, Brown AN et al (2003) The tetraspanin CD63 enhances the internalization of the H,K-ATPase beta-subunit. Proc Natl Acad Sci USA 100:15560–15565 Eble JA, Haier J (2006) Integrins in cancer treatment. Curr Cancer Drug Targets 6:89–105 Edwards DR, Handsley MM, Pennington CJ (2008) The ADAM metalloproteinases. Mol Aspects Med 29:258–289 Erler JT, Bennewith KL, Cox TR et al (2009) Hypoxia-induced lysyl oxidase is a critical mediator of bone marrow cell recruitment to form the premetastatic niche. Cancer Cell 15:35–44 Escola JM, Kleijmeer MJ, Stoorvogel W et al (1998) Selective enrichment of tetraspan proteins on the internal vesicles of multivesicular endosomes and on exosomes secreted by human B-lymphocytes. J Biol Chem 273:20121–20127 Fan H, Derynck R (1999) Ectodomain shedding of TGF-alpha and other transmembrane proteins is induced by receptor tyrosine kinase activation and MAP kinase signaling cascades. EMBO J 18:6962–6972 Fang Y, Wu N, Gan X et al (2007) Higher-order oligomerization targets plasma membrane proteins and HIV gag to exosomes. PLoS Biol 5:e158 Farhadieh RD, Smee R, Ow K et al (2004) Down-regulation of KAI1/CD82 protein expression in oral cancer correlates with reduced disease free survival and overall patient survival. Cancer Lett 213:91–98 Fevrier B, Raposo G (2004) Exosomes: endosomal-derived vesicles shipping extracellular messages. Curr Opin Cell Biol 16:415–421 Fitter S, Tetaz TJ, Berndt MC et al (1995) Molecular cloning of cDNA encoding a novel plateletendothelial cell tetra-span antigen, PETA-3. Blood 86:1348–1355 Flaumenhaft R (2006) Formation and fate of platelet microparticles. Blood Cells Mol Dis 36: 182–187 Fomina AF, Deerinck TJ, Ellisman MH (2003) Regulation of membrane trafficking and subcellular organization of endocytic compartments revealed with FM1–43 in resting and activated human T cells. Exp Cell Res 291:150–166
586
M. Zöller
Fraley TS, Tran TC, Corgan AM et al (2003) Phosphoinositide binding inhibits alpha-actinin bundling activity. J Biol Chem 278:24039–24045 Friedl P, Wolf K (2003) Tumour-cell invasion and migration: diversity and escape mechanisms. Nat Rev Cancer 3:362–374 Funakoshi T, Tachibana I, Hoshida Y et al (2003) Expression of tetraspanins in human lung cancer cells: frequent downregulation of CD9 and its contribution to cell motility in small cell lung cancer. Oncogene 22:674–687 Furuya M, Kato H, Nishimura N et al (2005) Down-regulation of CD9 in human ovarian carcinoma cell might contribute to peritoneal dissemination: morphologic alteration and reduced expression of beta1 integrin subsets. Cancer Res 65:2617–2625 Fuss H, Dubitzky W, Downes CS et al (2008) SRC family kinases and receptors: analysis of three activation mechanisms by dynamic systems modeling. Biophys J 94:1995–2006 Gao AC, Lou W, Dong JT et al (2003) Defining regulatory elements in the human KAI1 (CD 82) metastasis suppressor gene. Prostate 57:256–260 Garcia E, Nikolic DS, Piguet V (2008a) HIV-1 replication in dendritic cells occurs through a tetraspanin-containing compartment enriched in AP-3. Traffic 9:200–214 Garcia-España A, Chung PJ, Sarkar IN et al (2008) Appearance of new tetraspanin genes during vertebrate evolution. Genomics 91:326–334 Gassmann P, Enns A, Haier J (2004) Role of tumor cell adhesion and migration in organ-specific metastasis formation. Onkologie 27:577–582 Geary SM, Cambareri AC, Sincock PM et al (2001) Differential tissue expression of epitopes of the tetraspanin CD151 recognised by monoclonal antibodies. Tissue Antigens 58:141–153 Gensert JM, Baranova OV, Weinstein DE et al (2007) CD81, a cell cycle regulator, is a novel target for histone deacetylase inhibition in glioma cells. Neurobiol Dis 26:671–680 Gesierich S (2006) Das Tetraspanin CO-029/D6.1A in Membrankomplexen und Exosomen: Einfluss auf Tumorprogression und Angiogenese. Dissertation, University of Karlsruhe, Germany Gesierich S, Paret C, Hildebrand D et al (2005) Colocalization of the tetraspanins, CO-029 and CD151, with integrins in human pancreatic adenocarcinoma: impact on cell motility. Clin Cancer Res 11:2840–2852 Gesierich S, Berezovskiy I, Ryschich E et al (2006) Systemic induction of the angiogenesis switch by the tetraspanin D6.1A/CO-029. Cancer Res 66:7083–7094 Gibbings DJ, Ciaudo C, Erhardt M et al (2009) Multivesicular bodies associate with components of miRNA effector complexes and modulate miRNA activity. Nat Cell Biol 11:1143–1149 Goschnick MW, Lau LM, Wee JL et al (2006) Impaired “outside-in” integrin alphaIIbbeta3 signaling and thrombus stability in TSSC6-deficient mice. Blood 108:1911–1918 Gruenberg J, Stenmark H (2004) The biogenesis of multivesicular endosomes. Nat Rev Mol Cell Biol 5:317–323 Guo J, Wenk MR, Pellegrini L et al (2003) Phosphatidylinositol 4-kinase type IIalpha is responsible for the phosphatidylinositol 4-kinase activity associated with synaptic vesicles. Proc Natl Acad Sci USA 100:3995–4000 Guo XZ, Xu JH, Liu MP et al (2005) KAI1 inhibits anchorage-dependent and -independent pancreatic cancer cell growth. Oncol Rep 14:59–63 Hakomori SI (2010) Glycosynaptic microdomains controlling tumor cell phenotype through alteration of cell growth, adhesion, and motility. FEBS Lett 584:1901–1906 Hammond C, Denzin LK, Pan M et al (1998) The tetraspan protein CD82 is a resident of MHC class II compartments where it associates with HLA-DR, -DM, and -DO molecules. J Immunol 161:3282–3291 Hao S, Ye Z, Li F et al (2006) Epigenetic transfer of metastatic activity by uptake of highly metastatic B16 melanoma cell-released exosomes. Exp Oncol 28:126–131 Hasegawa H, Watanabe H, Nomura T et al (1997) Molecular cloning and expression of mouse homologue of SFA-1/PETA-3 (CD151), a member of the transmembrane 4 superfamily. Biochim Biophys Acta 1353:125–130 Hasegawa M, Furuya M, Kasuya Y et al (2007) CD151 dynamics in carcinoma-stroma interaction: integrin expression, adhesion strength and proteolytic activity. Lab Invest 87:882–892
26 Tetraspanins and Cancer Metastasis
587
Hashida H, Takabayashi A, Tokuhara T et al (2003) Clinical significance of transmembrane 4 superfamily in colon cancer. Br J Cancer 89:158–167 Hashimoto A, Tarner IH, Bohle RM et al (2007) Analysis of vascular gene expression in arthritic synovium by laser-mediated microdissection. Arthritis Rheum 56:1094–1105 He B, Liu L, Cook GA et al (2005) Tetraspanin CD82 attenuates cellular morphogenesis through down-regulating integrin alpha6-mediated cell adhesion. J Biol Chem 280:3346–3354 Helle F, Dubuisson J (2008) Hepatitis C virus entry into host cells. Cell Mol Life Sci 65:100–112 Hemler ME (2001) Specific tetraspanin functions. J Cell Biol 155:1103–1107 Hemler ME (2003) Tetraspanin proteins mediate cellular penetration, invasion, and fusion events and define a novel type of membrane microdomain. Annu Rev Cell Dev Biol 19:397–422 Hemler ME (2005) Tetraspanin functions and associated microdomains. Nat Rev Mol Cell Biol 6:801–811 Hemler ME (2008) Targeting of tetraspanin proteins – potential benefits and strategies. Nat Rev Drug Discov 7:747–758 Hemler ME, Mannion BA, Berditchevski F (1996) Association of TM4SF proteins with integrins: relevance to cancer. Biochim Biophys Acta 1287:67–71 Herlevsen M, Schmidt DS, Miyazaki K et al (2003) The association of the tetraspanin D6.1A with the alpha6beta4 integrin supports cell motility and liver metastasis formation. J Cell Sci 116:4373–4390 Herrlich A, Klinman E, Fu J et al (2008) Ectodomain cleavage of the EGF ligands HB-EGF, neuregulin1-{beta}, and TGF-{alpha} is specifically triggered by different stimuli and involves different PKC isoenzymes. FASEB J 22:4281–4295 Higashiyama M, Taki T, Ieki Y et al (1995) Reduced motility related protein-1 (MRP-1/CD9) gene expression as a factor of poor prognosis in non-small cell lung cancer. Cancer Res 55:6040–6044 Higashiyama M, Doi O, Kodama K et al (1997) Immunohistochemically detected expression of motility-related protein-1 (MRP-1/CD9) in lung adenocarcinoma and its relation to prognosis. Int J Cancer 74:205–211 Hirst J, Bright NA, Rous B et al (1999) Characterization of a fourth adaptor-related protein complex. Mol Biol Cell 10:2787–2802 Hlubek F, Spaderna S, Schmalhofer O et al (2007) Wnt/FZD signaling and colorectal cancer morphogenesis. Front Biosci 12:458–470 Hong IK, Kim YM, Jeoung DI et al (2005) Tetraspanin CD9 induces MMP-2 expression by activating p38 MAPK, JNK and c-Jun pathways in human melanoma cells. Exp Mol Med 37:230–239 Hong IK, Jin YJ, Byun HJ et al (2006) Homophilic interactions of Tetraspanin CD151 up-regulate motility and matrix metalloproteinase-9 expression of human melanoma cells through adhesion-dependent c-Jun activation signaling pathways. J Biol Chem 281:24279–24292 Horejsí V, Vlcek C (1991) Novel structurally distinct family of leucocyte surface glycoproteins including CD9, CD37, CD53 and CD63. FEBS Lett 288:1–4 Hori H, Yano S, Koufuji K et al (2004) CD9 expression in gastric cancer and its significance. J Surg Res 117:208–215 Horváth G, Serru V, Clay D et al (1998) CD19 is linked to the integrin-associated tetraspans CD9, CD81, and CD82. J Biol Chem 273:30537–30543 Hotta H, Ross AH, Huebner K et al (1998) Molecular cloning and characterization of an antigen associated with early stages of melanoma tumor progression. Cancer Res 48:2955–2962 Houle CD, Ding XY, Foley JF et al (2002) Loss of expression and altered localization of KAI1 and CD9 protein are associated with epithelial ovarian cancer progression. Gynecol Oncol 86:69–78 Huang CI, Kohno N, Ogawa E et al (1998) Correlation of reduction in MRP-1/CD9 and KAI1/ CD82 expression with recurrences in breast cancer patients. Am J Pathol 153:973–983 Huang CL, Liu D, Masuya D et al (2004) MRP-1/CD9 gene transduction downregulates Wnt signal pathways. Oncogene 23:7475–7483 Huang H, Groth J, Sossey-Alaoui K et al (2005) Aberrant expression of novel and previously described cell membrane markers in human breast cancer cell lines and tumors. Clin Cancer Res 11:4357–4364
588
M. Zöller
Huang CL, Ueno M, Liu D et al (2006) MRP-1/CD9 gene transduction regulates the actin cytoskeleton through the downregulation of WAVE2. Oncogene 25:6480–6488 Huang H, Sossey-Alaoui K, Beachy SH et al (2007) The tetraspanin superfamily member NET-6 is a new tumor suppressor gene. J Cancer Res Clin Oncol 133:761–769 Huerta S, Harris DM, Jazirehi A et al (2003) Gene expression profile of metastatic colon cancer cells resistant to cisplatin-induced apoptosis. Int J Oncol 22:663–670 Hurley JH, Emr SD (2006) The ESCRT complexes: structure and mechanism of a membranetrafficking network. Annu Rev Biophys Biomol Struct 35:277–298 Ichikawa T, Ichikawa Y, Isaacs JT (1991) Genetic factors and suppression of metastatic ability of prostatic cancer. Cancer Res 51:3788–3792 Iero M, Valenti R, Huber V et al (2008) Tumour-released exosomes and their implications in cancer immunity. Cell Death Differ 15:80–88 Ikeyama S, Koyama M, Yamaoko M et al (1993) Suppression of cell motility and metastasis by transfection with human motility-related protein (MRP-1/CD9) DNA. J Exp Med 177: 1231–1237 Imai T, Yoshie O (1993) C33 antigen and M38 antigen recognized by monoclonal antibodies inhibitory to syncytium formation by human T cell leukemia virus type 1 are both members of the transmembrane 4 superfamily and associate with each other and with CD4 or CD8 in T cells. J Immunol 151:6470–6481 Imhof I, Gasper WJ, Derynck R (2008) Association of tetraspanin CD9 with transmembrane TGF{alpha} confers alterations in cell-surface presentation of TGF{alpha} and cytoskeletal organization. J Cell Sci 121:2265–2274 Inoue G, Horiike N, Michitaka K et al (2001) The CD81 expression in liver in hepatocellular carcinoma. Int J Mol Med 7:67–71 Ishitani T, Kishida S, Hyodo-Miura J et al (2003) The TAK1-NLK mitogen-activated protein kinase cascade functions in the Wnt-5a/Ca(2+) pathway to antagonize Wnt/beta-catenin signaling. Mol Cell Biol 23:131–139 Israels SJ, McMillan-Ward EM (2005) CD63 modulates spreading and tyrosine phosphorylation of platelets on immobilized fibrinogen. Thromb Haemost 93:311–318 Jackson P, Millar D, Kingsley E et al (2000) Methylation of a CpG island within the promoter region of the KAI1 metastasis suppressor gene is not responsible for down-regulation of KAI1 expression in invasive cancers or cancer cell lines. Cancer Lett 157:169–176 Jackson P, Marreiros A, Russell PJ (2005) KAI1 tetraspanin and metastasis suppressor. Int J Biochem Cell Biol 37:530–534 Jackson P, Rowe A, Grimm MO (2007) An alternatively spliced KAI1 mRNA is expressed at low levels in human bladder cancers and bladder cancer cell lines and is not associated with invasive behaviour. Oncol Rep 18:1357–1363 Jang HI, Lee H (2003) A decrease in the expression of CD63 tetraspanin protein elevates invasive potential of human melanoma cells. Exp Mol Med 35:317–323 Janmey PA, Lindberg U (2004) Cytoskeletal regulation: rich in lipids. Nat Rev Mol Cell Biol 5:658–666 Janowska-Wieczorek A, Wysoczynski M, Kijowski J et al (2005) Microvesicles derived from activated platelets induce metastasis and angiogenesis in lung cancer. Int J Cancer 113: 752–760 Janvier K, Bonifacino JS (2005) Role of the endocytic machinery in the sorting of lysosomeassociated membrane proteins. Mol Biol Cell 16:4231–4242 Johnson JL, Winterwood N, DeMali KA et al (2009) Tetraspanin CD151 regulates RhoA activation and the dynamic stability of carcinoma cell-cell contacts. J Cell Sci 122:2263–2273 Johnstone RM (2006) Exosomes biological significance: A concise review. Blood Cells Mol Dis 36:315–321 Jung KK, Liu XW, Chirco R et al (2006) Identification of CD63 as a tissue inhibitor of metalloproteinase-1 interacting cell surface protein. EMBO J 25:3934–3942 Jung T, Castellana D, Klingbeil P et al (2009) CD44v6 dependence of premetastatic niche preparation by exosomes. Neoplasia 11:1093–1105
26 Tetraspanins and Cancer Metastasis
589
Kaji K, Oda S, Shikano T et al (2000) The gamete fusion process is defective in eggs of Cd9deficient mice. Nat Genet 24:279–282 Kanetaka K, Sakamoto M, Yamamoto Y et al (2001) Overexpression of tetraspanin CO-029 in hepatocellular carcinoma. J Hepatol 35:637–642 Kanetaka K, Sakamoto M, Yamamoto Y et al (2003) Possible involvement of tetraspanin CO-029 in hematogenous intrahepatic metastasis of liver cancer cells. J Gastroenterol Hepatol 18:1309–1314 Kaplan RN, Riba RD, Zacharoulis S et al (2005) VEGFR1-positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438:820–827 Kaplan RN, Rafii S, Lyden D (2006) Preparing the “soil”: the premetastatic niche. Cancer Res 66:11089–11093 Karamatic Crew V, Burton N, Kagan A et al (2004) CD151, the first member of the tetraspanin (TM4) superfamily detected on erythrocytes, is essential for the correct assembly of human basement membranes in kidney and skin. Blood 104:2217–2223 Kawashima M, Doh-ura K, Mekada E et al (2002) CD9 expression in solid non-neuroepithelial tumors and infiltrative astrocytic tumors. J Histochem Cytochem 50:1195–1203 Keller S, Sanderson MP, Stoeck A et al (2006) Exosomes: from biogenesis and secretion to biological function. Immunol Lett 107:102–108 Kerbel R, Folkman J (2002) Clinical translation of angiogenesis inhibitors. Nat Rev Cancer 2:727–739 Kessels H, Béguin S, Andree H et al (1994) Measurement of thrombin generation in whole blood – the effect of heparin and aspirin. Thromb Haemost 72:78–83 Kim JH, Kim B, Cai L et al (2005) Transcriptional regulation of a metastasis suppressor gene by Tip60 and beta-catenin complexes. Nature 434:921–926 Kitani S, Berenstein E, Mergenhagen S et al (1991) A cell surface glycoprotein of rat basophilic leukemia cells close to the high affinity IgE receptor (Fc epsilon RI). Similarity to human melanoma differentiation antigen ME491. J Biol Chem 266:1903–1909 Kohno M, Hasegawa H, Miyake M et al (2002) CD151 enhances cell motility and metastasis of cancer cells in the presence of focal adhesion kinase. Int J Cancer 97:336–343 Kolesnikova TV, Kazarov AR, Lemieux ME et al (2009) Glioblastoma inhibition by cell surface immunoglobulin protein EWI-2, in vitro and in vivo. Neoplasia 11:77–86 Koo TH, Lee JJ, Kim EM et al (2002) Syntenin is overexpressed and promotes cell migration in metastatic human breast and gastric cancer cell lines. Oncogene 21:4080–4088 Kovalenko OV, Yang X, Kolesnikova TV et al (2004) Evidence for specific tetraspanin homodimers: inhibition of palmitoylation makes cysteine residues available for cross-linking. Biochem J 377:407–417 Kovalenko OV, Metcalf DG, DeGrado WF et al (2005) Structural organization and interactions of transmembrane domains in tetraspanin proteins. BMC Struct Biol 5:11 Kovalenko OV, Yang XH, Hemler ME (2007) A novel cysteine cross-linking method reveals a direct association between claudin-1 and tetraspanin CD9. Mol Cell Proteomics 6: 1855–1867 Kuhn S, Koch M, Nübel T et al (2007) A complex of EpCAM, claudin-7, CD44 variant isoforms, and tetraspanins promotes colorectal cancer progression. Mol Cancer Res 5:553–567 Kusukawa J, Ryu F, Kameyama T et al (2001) Reduced expression of CD9 in oral squamous cell carcinoma: CD9 expression inversely related to high prevalence of lymph node metastasis. J Oral Pathol 30:73–79 Kwon MS, Shin SH, Yim SH et al (2007) CD63 as a biomarker for predicting the clinical outcomes in adenocarcinoma of lung. Lung Cancer 57:46–53 Ladwein M, Pape UF, Schmidt DS et al (2005) The cell-cell adhesion molecule EpCAM interacts directly with the tight junction protein claudin-7. Exp Cell Res 309:345–357 Lagaudrière-Gesbert C, Lebel-Binay S, Wiertz E et al (1997) The tetraspanin protein CD82 associates with both free HLA class I heavy chain and heterodimeric beta 2-microglobulin complexes. J Immunol 158:2790–2797 Lakkaraju A, Rodriguez-Boulan E (2008) Itinerant exosomes: emerging roles in cell and tissue polarity. Trends Cell Biol 18:199–209
590
M. Zöller
Lan RF, Liu ZX, Liu XC et al (2005) CD151 promotes neovascularization and improves blood perfusion in a rat hind-limb ischemia model. J Endovasc Ther 12:469–478 Latysheva N, Muratov G, Rajesh S et al (2006) Syntenin-1 is a new component of tetraspaninenriched microdomains: mechanisms and consequences of the interaction of syntenin-1 with CD63. Mol Cell Biol 26:7707–7718 Lau LM, Wee JL, Wright MD et al (2004) The tetraspanin superfamily member CD151 regulates outside-in integrin alphaIIbbeta3 signaling and platelet function. Blood 104:2368–2375 Lazo PA (2007) Functional implications of tetraspanin proteins in cancer biology. Cancer Sci 98:1666–1677 Leavey PJ, Timmons C, Frawley W et al (2006) KAI-1 expression in pediatric high-grade osteosarcoma. Pediatr Dev Pathol 9:219–224 Lee JH, Seo YW, Park SR et al (2003) Expression of a splice variant of KAI1, a tumor metastasis suppressor gene, influences tumor invasion and progression. Cancer Res 63:7247–7255 Lee JH, Park SR, Chay KO et al (2004) KAI1 COOH-terminal interacting tetraspanin (KITENIN), a member of the tetraspanin family, interacts with KAI1, a tumor metastasis suppressor, and enhances metastasis of cancer. Cancer Res 64:4235–4243 Lekishvili T, Fromm E, Mujoomdar M et al (2008) The tumour-associated antigen L6 (L6-Ag) is recruited to the tetraspanin-enriched microdomains: implication for tumour cell motility. J Cell Sci 121:685–694 Le Naour F, Zöller M (2008) The tumor antigen EpCAM: tetraspanins and the tight junction protein claudin-7, new partners, new functions. Front Biosci 13:5847–5865 Le Naour F, André M, Boucheix C, Rubinstein E (2006) Membrane microdomains and proteomics: lessons from tetraspanin microdomains and comparison with lipid rafts. Proteomics 6: 6447–6454 Levy S, Shoham T (2005a) Protein-protein interactions in the tetraspanin web. Physiology (Bethesda) 20:218–224 Levy S, Shoham T (2005b) The tetraspanin web modulates immune-signalling complexes. Nat Rev Immunol 5:136–148 Levy S, Todd SC, Maecker HT (1998) CD81 (TAPA-1): a molecule involved in signal transduction and cell adhesion in the immune system. Annu Rev Immunol 16:89–109 Lewis TB, Robison JE, Bastien R et al (2005) Molecular classification of melanoma using realtime quantitative reverse transcriptase-polymerase chain reaction. Cancer 104:1678–1686 Li Q, Li L, Shi W et al (2006) Mechanism of action differences in the antitumor effects of transmembrane and secretory tumor necrosis factor-alpha in vitro and in vivo. Cancer Immunol Immunother 55:1470–1479 Linder ME, Deschenes RJ (2007) Palmitoylation: policing protein stability and traffic. Nat Rev Mol Cell Biol 8:74–84 Lishner M, Zismanov V, Tohami T et al (2008) Tetraspanins affect myeloma cell fate via Akt signaling and FoxO activation. Cell Signal 20:2309–2316 Little KD, Hemler ME, Stipp CS (2004) Dynamic regulation of a GPCR-tetraspanin-G protein complex on intact cells: central role of CD81 in facilitating GPR56-Galpha q/11 association. Mol Biol Cell 15:2375–2387 Liu WM, Zhang XA (2006) KAI1/CD82, a tumor metastasis suppressor. Cancer Lett 240:183–194 Liu FS, Dong JT, Chen JT et al (2000) Frequent down-regulation and lack of mutation of the KAI1 metastasis suppressor gene in epithelial ovarian carcinoma. Gynecol Oncol 78:10–15 Liu L, He B, Liu WM et al (2007) Tetraspanin CD151 promotes cell migration by regulating integrin trafficking. J Biol Chem 282:31631–31642 Loewen CJ, Moritz OL, Tam BM et al (2003) The role of subunit assembly in peripherin-2 targeting to rod photoreceptor disk membranes and retinitis pigmentosa. Mol Biol Cell 14:3400–3413 Lombardi DP, Geradts J, Foley JF et al (1999) Loss of KAI1 expression in the progression of colorectal cancer. Cancer Res 59:5724–5731 Longo N, Yáñez-Mó M, Mittelbrunn M et al (2001) Regulatory role of tetraspanin CD9 in tumorendothelial cell interaction during transendothelial invasion of melanoma cells. Blood 98:3717–3726
26 Tetraspanins and Cancer Metastasis
591
Maecker HT, Todd SC, Levy S (1997) The tetraspanin superfamily: molecular facilitators. FASEB J 11:428–442 Malik FA, Sanders AJ, Jiang WG (2009) KAI-1/CD82, the molecule and clinical implication in cancer and cancer metastasis. Histol Histopathol 24:519–530 Malinda KM (2009) In vivo matrigel migration and angiogenesis assay. Methods Mol Biol 467:287–294 Mannion BA, Berditchevski F, Kraeft SK et al (1996) Transmembrane-4 superfamily proteins CD81 (TAPA-1), CD82, CD63, and CD53 specifically associated with integrin alpha 4 beta 1 (CD49d/CD29). J Immunol 157:2039–2047 Marhaba R, Zöller M (2004) CD44 in cancer progression: adhesion, migration and growth regulation. J Mol Histol 35:211–231 Marken JS, Schieven GL, Hellström I et al (1992) Cloning and expression of the tumor-associated antigen L6. Proc Natl Acad Sci USA 89:3503–3507 Marks MS, Ohno H, Kirchnausen T et al (1997) Protein sorting by tyrosine-based signals: adapting to the Ys and wherefores. Trends Cell Biol 7:124–128 Marreiros A, Czolij R, Yardley G et al (2003) Identification of regulatory regions within the KAI1 promoter: a role for binding of AP1, AP2 and p53. Gene 302:155–164 Marreiros A, Dudgeon K, Dao V et al (2005) KAI1 promoter activity is dependent on p53, junB and AP2: evidence for a possible mechanism underlying loss of KAI1 expression in cancer cells. Oncogene 24:637–649 Martin F, Roth DM, Jans DA et al (2005) Tetraspanins in viral infections: a fundamental role in viral biology? J Virol 79:10839–10851 Mathivanan S, Lim JW, Tauro BJ et al (2010) Proteomic analysis of A33-immunoaffinity-purified exosomes released from the human colon tumor cell line LIM1215 reveals a tissue-specific protein signature. Mol Cell Proteomics 9:197–208 Matzku S, Wenzel A, Liu S et al (1989) Antigenic differences between metastatic and nonmetastatic BSp73 rat tumor variants characterized by monoclonal antibodies. Cancer Res 49:1294–1299 Mazzocca A, Liotta F, Carloni V (2008) Tetraspanin CD81-regulated cell motility plays a critical role in intrahepatic metastasis of hepatocellular carcinoma. Gastroenterology 135: 244–256 Mhawech P, Herrmann F, Coassin M et al (2003) Motility-related protein 1 (MRP-1/CD9) expression in urothelial bladder carcinoma and its relation to tumor recurrence and progression. Cancer 98:1649–1657 Mhawech P, Dulguerov P, Tschanz E et al (2004) Motility-related protein-1 (MRP-1/CD9) expression can predict disease-free survival in patients with squamous cell carcinoma of the head and neck. Br J Cancer 90:471–475 Mimori K, Kataoka A, Yoshinaga K et al (2005) Identification of molecular markers for metastasis-related genes in primary breast cancer cells. Clin Exp Metastasis 22:59–67 Mitsuzuka K, Handa K, Satoh M et al (2005) A specific microdomain (“glycosynapse 3”) controls phenotypic conversion and reversion of bladder cancer cells through GM3-mediated interaction of alpha3beta1 integrin with CD9. J Biol Chem 280:35545–35553 Miura Y, Kainuma M, Jiang H et al (2004) Reversion of the Jun-induced oncogenic phenotype by enhanced synthesis of sialosyllactosylceramide (GM3 ganglioside). Proc Natl Acad Sci USA 101:16204–16209 Miyake M, Adachi M, Huang C et al (1999) A novel molecular staging protocol for non-small cell lung cancer. Oncogene 18:2397–2404 Miyamoto S, Maruyama A, Okugawa K et al (2001) Loss of motility-related protein 1 (MRP1/ CD9) and integrin alpha3 expression in endometrial cancers. Cancer 92:542–548 Miyazaki T, Müller U, Campbell KS (1997) Normal development but differentially altered proliferative responses of lymphocytes in mice lacking CD81. EMBO J 16:4217–4225 Miyazaki T, Kato H, Shitara Y et al (2000) Mutation and expression of the metastasis suppressor gene KAI1 in esophageal squamous cell carcinoma. Cancer 89:955–962 Mohan A, Nalini V, Mallikarjuna K et al (2007) Expression of motility-related protein MRP1/ CD9, N-cadherin, E-cadherin, alpha-catenin and beta-catenin in retinoblastoma. Exp Eye Res 84:781–789
592
M. Zöller
Molina S, Castet V, Pichard-Garcia L et al (2008) Serum-derived hepatitis C virus infection of primary human hepatocytes is tetraspanin CD81 dependent. J Virol 82:569–574 Morel O, Hugel B, Jesel L et al (2004) Circulating procoagulant microparticles and soluble GPV in myocardial infarction treated by primary percutaneous transluminal coronary angioplasty. A possible role for GPIIb-IIIa antagonists. J Thromb Haemost 2:1118–1126 Mori M, Mimori K, Shiraishi T et al (1998) Motility related protein 1 (MRP1/CD9) expression in colon cancer. Clin Cancer Res 4:1507–1510 Moss ML, Bartsch JW (2004) Therapeutic benefits from targeting of ADAM family members. Biochemistry 43:7227–7235 Mott JD, Werb Z (2004) Regulation of matrix biology by matrix metalloproteinases. Curr Opin Cell Biol 16:558–564 Murayama Y, Miyagawa J, Oritani K et al (2004) CD9-mediated activation of the p46 Shc isoform leads to apoptosis in cancer cells. J Cell Sci 117:3379–3388 Murayama Y, Shinomura Y, Oritani K et al (2008) The tetraspanin CD9 modulates epidermal growth factor receptor signaling in cancer cells. J Cell Physiol 216:135–143 Nakamura K, Mitamura T, Takahashi T et al (2000) Importance of the major extracellular domain of CD9 and the epidermal growth factor (EGF)-like domain of heparin-binding EGF-like growth factor for up-regulation of binding and activity. J Biol Chem 275:18284–18290 Nakatsu F, Ohno H (2003) Adaptor protein complexes as the key regulators of protein sorting in the post-Golgi network. Cell Struct Funct 28:419–429 Nakazawa Y, Sato S, Naito M et al (2008) Tetraspanin family member CD9 inhibits Aggrus/ podoplanin-induced platelet aggregation and suppresses pulmonary metastasis. Blood 112: 1730–1739 Nazarenko I, Rana S, Baumann A et al (2010) Cell surface tetraspanin Tspan8 contributes to molecular pathways of exosome-induced endothelial cell activation. Cancer Res 70: 1668–1678 Ng T, Shima D, Squire A et al (1999) PKCalpha regulates beta1 integrin-dependent cell motility through association and control of integrin traffic. EMBO J 18:3909–3923 Nichols TC, Guthridge JM, Karp DR et al (1998) Gamma-glutamyl transpeptidase, an ectoenzyme regulator of intracellular redox potential, is a component of TM4 signal transduction complexes. Eur J Immunol 28:4123–4129 Nübel T, Preobraschenski J, Tuncay H et al (2009) Claudin-7 regulates EpCAM-mediated functions in tumor progression. Mol Cancer Res 7:285–299 Odintsova E, Sugiura T, Berditchevski F (2000) Attenuation of EGF receptor signaling by a metastasis suppressor, the tetraspanin CD82/KAI-1. Curr Biol 10:1009–1012 Odintsova E, Voortman J, Gilbert E et al (2003) Tetraspanin CD82 regulates compartmentalisation and ligand-induced dimerization of EGFR. J Cell Sci 116:4557–4566 Odintsova E, Butters TD, Monti E et al (2006) Gangliosides play an important role in the organization of CD82-enriched microdomains. Biochem J 400:315–325 Ono M, Handa K, Withers DA et al (1999) Motility inhibition and apoptosis are induced by metastasis-suppressing gene product CD82 and its analogue CD9, with concurrent glycosylation. Cancer Res 59:2335–2339 Oren R, Takahashi S, Doss C et al (1990) TAPA-1, the target of an antiproliferative antibody, defines a new family of transmembrane proteins. Mol Cell Biol 10:4007–4015 Ovalle S, Gutiérrez-López MD, Olmo N et al (2007) The tetraspanin CD9 inhibits the proliferation and tumorigenicity of human colon carcinoma cells. Int J Cancer 121:2140–2152 Pan BT, Teng K, Wu C et al (1985) Electron microscopic evidence for externalization of the transferrin receptor in vesicular form in sheep reticulocytes. J Cell Biol 101:942–948 Pap E, Pállinger E, Pásztói M, Falus A (2009) Highlights of a new type of intercellular communication: microvesicle-based information transfer. Inflamm Res 58:1–8 Peer D, Park EJ, Morishita Y et al (2008) Systemic leukocyte-directed siRNA delivery revealing cyclin D1 as an anti-inflammatory target. Science 319:627–630 Percherancier Y, Planchenault T, Valenzuela-Fernandez A et al (2001) Palmitoylation-dependent control of degradation, life span, and membrane expression of the CCR5 receptor. J Biol Chem 276:31936–31944
26 Tetraspanins and Cancer Metastasis
593
Phillips KK, White AE, Hicks DJ et al (1998) Correlation between reduction of metastasis in the MDA-MB-435 model system and increased expression of the Kai-1 protein. Mol. Carcinog 21:111–120 Pileri P, Uematsu Y, Campagnoli S et al (1998) Binding of hepatitis C virus to CD81. Science 282:938–941 Pols MS, Klumperman J (2009) Trafficking and function of the tetraspanin CD63. Exp Cell Res 315:1584–1592 Potolicchio I, Carven GJ, Xu X et al (2005) Proteomic analysis of microglia-derived exosomes: metabolic role of the aminopeptidase CD13 in neuropeptide catabolism. J Immunol 175:2237–2243 Press OW, Eary JF, Badger CC et al (1989) Treatment of refractory non-Hodgkin’s lymphoma with radiolabeled MB-1 (anti-CD37) antibody. J Clin Oncol 7:1027–1038 Prince S, Carreira S, Vance KW et al (2004) Tbx2 directly represses the expression of the p21(WAF1) cyclin-dependent kinase inhibitor. Cancer Res 64:1669–1674 Protzel C, Kakies C, Kleist B et al (2008) Down-regulation of the metastasis suppressor protein KAI1/CD82 correlates with occurrence of metastasis, prognosis and presence of HPV DNA in human penile squamous cell carcinoma. Virchows Arch 452:369–375 Radford KJ, Mallesch J, Hersey P (1995) Suppression of human melanoma cell growth and metastasis by the melanoma-associated antigen CD63 (ME491). Int J Cancer 62:631–635 Radford KJ, Thorne RF, Hersey P (1996) CD63 associates with transmembrane 4 superfamily members, CD9 and CD81, and with beta 1 integrins in human melanoma. Biochem Biophys Res Commun 222:13–18 Ratajczak J, Miekus K, Kucia M et al (2006) Embryonic stem cell-derived microvesicles reprogram hematopoietic progenitors: evidence for horizontal transfer of mRNA and protein delivery. Leukemia 20:847–856 Razmara M, Hu H, Masquelier M et al (2007) Glycoprotein IIb/IIIa blockade inhibits platelet aminophospholipid exposure by potentiating translocase and attenuating scramblase activity. Cell Mol Life Sci 64:999–1008 Ribatti D, Nico B, Crivellato E et al (2007) The history of the angiogenic switch concept. Leukemia 21:44–52 Rocha-Perugini V, Lavie M, Delgrange D et al (2009) The association of CD81 with tetraspaninenriched microdomains is not essential for Hepatitis C virus entry. BMC Microbiol 9:111 Rous BA, Reaves BJ, Ihrke G et al (2002) Role of adaptor complex AP-3 in targeting wild-type and mutated CD63 to lysosomes. Mol Biol Cell 13:1071–1082 Rowe A, Jackson P (2006) Expression of KITENIN, a KAI1/CD82 binding protein and metastasis enhancer, in bladder cancer cell lines: relationship to KAI1/CD82 levels and invasive behaviour. Oncol Rep 16:1267–1272 Rubinstein E, Le Naour F, Lagaudrière-Gesbert C et al (1996) CD9, CD63, CD81, and CD82 are components of a surface tetraspan network connected to HLA-DR and VLA integrins. Eur J Immunol 26:2657–2665 Rubinstein E, Ziyyat A, Wolf JP et al (2006) The molecular players of sperm-egg fusion in mammals. Semin Cell Dev Biol 17:254–263 Sachs N, Kreft M, van den Bergh Weerman MA et al (2006) Kidney failure in mice lacking the tetraspanin CD151. J Cell Biol 175:33–39 Saito Y, Tachibana I, Takeda Y et al (2006) Absence of CD9 enhances adhesion-dependent morphologic differentiation, survival, and matrix metalloproteinase-2 production in small cell lung cancer cells. Cancer Res 66:9557–9565 Sala-Valdés M, Ursa A, Charrin S et al (2006) EWI-2 and EWI-F link the tetraspanin web to the actin cytoskeleton through their direct association with ezrin-radixin-moesin proteins. J Biol Chem 281:19665–19675 Sauer G, Kurzeder C, Grundmann R et al (2003a) Expression of tetraspanin adaptor proteins below defined threshold values is associated with in vitro invasiveness of mammary carcinoma cells. Oncol Rep 10:405–410 Sauer G, Windisch J, Kurzeder C et al (2003b) Progression of cervical carcinomas is associated with down-regulation of CD9 but strong local re-expression at sites of transendothelial invasion. Clin Cancer Res 9:6426–6431
594
M. Zöller
Sawada S, Yoshimoto M, Odintsova E et al (2003) The tetraspanin CD151 functions as a negative regulator in the adhesion-dependent activation of Ras. J Biol Chem 278:26323–26326 Schindl M, Birner P, Bachtiary B et al (2000) The impact of expression of the metastasis suppressor protein KAI1 on prognosis in invasive squamous cell cervical cancer. Anticancer Res 20: 4551–4555 Schmidt DS, Klingbeil P, Schnölzer M et al (2004) CD44 variant isoforms associate with tetraspanins and EpCAM. Exp Cell Res 297:329–347 Schöniger-Hekele M, Hänel S, Wrba F et al (2005) Hepatocellular carcinoma – survival and clinical characteristics in relation to various histologic molecular markers in Western patients. Liver Int 25:62–69 Schorey JS, Bhatnagar S (2008) Exosome function: from tumor immunology to pathogen biology. Traffic 9:871–881 Seigneuret M (2006) Complete predicted three-dimensional structure of the facilitator transmembrane protein and hepatitis C virus receptor CD81: conserved and variable structural domains in the tetraspanin superfamily. Biophys J 90:212–227 Seigneuret M, Delaguillaumie A, Lagaudrière-Gesbert C et al (2001) Structure of the tetraspanin main extracellular domain. A partially conserved fold with a structurally variable domain insertion. J Biol Chem 276:40055–40064 Sela BA, Steplewski Z, Koprowski H (1989) Colon carcinoma-associated glycoproteins recognized by monoclonal antibodies CO-029 and GA22–2. Hybridoma 8:481–491 Seow Y, Wood MJ (2009) Biological gene delivery vehicles: beyond viral vectors. Mol Ther 17:767–777 Serru V, Le Naour F, Billard M et al (1999) Selective tetraspan-integrin complexes (CD81/alpha4beta1, CD151/alpha3beta1, CD151/alpha6beta1) under conditions disrupting tetraspan interactions. Biochem J 340:103–111 Sharma C, Yang XH, Hemler ME (2008) DHHC2 affects palmitoylation, stability, and functions of tetraspanins CD9 and CD151. Mol Biol Cell 19:3415–3425 Shet AS, Aras O, Gupta K et al (2003) Sickle blood contains tissue factor-positive microparticles derived from endothelial cells and monocytes. Blood 102:2678–2683 Shi W, Fan H, Shum L et al (2000) The tetraspanin CD9 associates with transmembrane TGFalpha and regulates TGF-alpha-induced EGF receptor activation and cell proliferation. J Cell Biol 148:591–602 Shinohara T, Nishimura N, Hanibuchi M et al (2001) Transduction of KAI1/CD82 cDNA promotes hematogenous spread of human lung-cancer cells in natural killer cell-depleted SCID mice. Int J Cancer 94:16–23 Shiomi T, Inoki I, Kataoka F et al (2005) Pericellular activation of proMMP-7 (promatrilysin-1) through interaction with CD151. Lab Invest 85:1489–1506 Sierko E, Wojtukiewicz MZ (2007) Inhibition of platelet function: does it offer a chance of better cancer progression control? Semin Thromb Hemost 33:712–721 Sigala S, Faraoni I, Botticini D et al (1999) Suppression of telomerase, reexpression of KAI1, and abrogation of tumorigenicity by nerve growth factor in prostate cancer cell lines. Clin Cancer Res 5:1211–1218 Silvie O, Rubinstein E, Franetich JF et al (2003) Hepatocyte CD81 is required for Plasmodium falciparum and Plasmodium yoelii sporozoite infectivity. Nat Med 9:93–96 Simons M, Raposo G (2009) Exosomes – vesicular carriers for intercellular communication. Curr Opin Cell Biol 21:575–581 Simpson RJ, Jensen SS, Lim JW (2008) Proteomic profiling of exosomes: current perspectives. Proteomics 8:4083–4099 Simpson RJ, Lim JW, Moritz RL et al (2009) Exosomes: proteomic insights and diagnostic potential. Expert Rev Proteomics 6:267–283 Sincock PM, Mayrhofer G, Ashman LK (1997) Localization of the transmembrane 4 superfamily (TM4SF) member PETA-3 (CD151) in normal human tissues: comparison with CD9, CD63, and alpha5beta1 integrin. J Histochem Cytochem 45:515–525
26 Tetraspanins and Cancer Metastasis
595
Sincock PM, Fitter S, Parton RG et al (1999) PETA-3/CD151, a member of the transmembrane 4 superfamily, is localised to the plasma membrane and endocytic system of endothelial cells, associates with multiple integrins and modulates cell function. J Cell Sci 112:833–844 Singethan K, Müller N, Schubert S et al (2008) CD9 clustering and formation of microvilli zippers between contacting cells regulates virus-induced cell fusion. Traffic 9:924–235 Skubitz KM, Campbell KD, Iida J et al (1996) CD63 associates with tyrosine kinase activity and CD11/CD18, and transmits an activation signal in neutrophils. J Immunol 157: 3617–3626 Smalheiser NR (2007) Exosomal transfer of proteins and RNAs at synapses in the nervous system. Biol Direct 2:35 Son BH, Choi JS, Lee JH (2005) Prognostic values of KAI1 and survivin expression in an infiltrating ductal carcinoma of the breast. Pathology 37:131–136 Sood SL (2009) Cancer-associated thrombosis. Curr Opin Hematol 16:378–385 Sordat I, Decraene C, Silvestre T et al (2002) Complementary DNA arrays identify CD63 tetraspanin and alpha3 integrin chain as differentially expressed in low and high metastatic human colon carcinoma cells. Lab Invest 82:1715–1724 Spoden G, Freitag K, Husmann M et al (2008) Clathrin- and caveolin-independent entry of human papillomavirus type 16 – involvement of tetraspanin-enriched microdomains (TEMs). PLoS One 3:e3313 Sridhar SC, Miranti CK (2006) Tetraspanin KAI1/CD82 suppresses invasion by inhibiting integrin-dependent crosstalk with c-Met receptor and Src kinases. Oncogene 25:2367–2378 Stafford LJ, Vaidya KS, Welch DR (2008) Metastasis suppressors genes in cancer. Int J Biochem Cell Biol 40:874–891 Staflin K, Zuchner T, Honeth G et al (2009) Identification of proteins involved in neural progenitor cell targeting of gliomas. BMC Cancer 9:206 Stahl PD, Barbieri MA (2002) Multivesicular bodies and multivesicular endosomes: the “ins and outs” of endosomal traffic. Sci STKE 2002(141):PE32 Steeg PS (2003) Metastasis suppressors alter the signal transduction of cancer cells. Nat Rev Cancer 3:55–63 Sterk LM, Geuijen CA, Oomen LC et al (2000) The tetraspan molecule CD151, a novel constituent of hemidesmosomes, associates with the integrin alpha6beta4 and may regulate the spatial organization of hemidesmosomes. J Cell Biol 149:969–982 Stipp CS, Kolesnikova TV, Hemler ME (2001a) EWI-2 is a major CD9 and CD81 partner and member of a novel Ig protein subfamily. J Biol Chem 276:40545–40554 Stipp CS, Orlicky D, Hemler ME (2001b) FPRP, a major, highly stoichiometric, highly specific CD81- and CD9-associated protein. J Biol Chem 276:4853–4862 Stipp CS, Kolesnikova TV, Hemler ME (2003) Functional domains in tetraspanin proteins. Trends Biochem Sci 28:106–112 Stoeck A, Keller S, Riedle S et al (2006) A role for exosomes in the constitutive and stimulusinduced ectodomain cleavage of L1 and CD44. Biochem J 393:609–618 Stoorvogel W, Kleijmeer MJ, Geuze HJ et al (2002) The biogenesis and functions of exosomes. Traffic 3:321–330 Stracke ML, Liotta LA (1992) Multi-step cascade of tumor cell metastasis. In Vivo 6:309–316 Subra C, Laulagnier K, Perret B et al (2007) Exosome lipidomics unravels lipid sorting at the level of multivesicular bodies. Biochimie 89:205–212 Suzuki-Inoue K, Fuller GL, García A et al (2006) A novel Syk-dependent mechanism of platelet activation by the C-type lectin receptor CLEC-2. Blood 107:542–549 Szala S, Kasai Y, Steplewski Z et al (1990) Molecular cloning of cDNA for the human tumorassociated antigen CO-029 and identification of related transmembrane antigens. Proc Natl Acad Sci USA 87:6833–6837 Szöllósi J, Horejsí V, Bene L et al (1996) Supramolecular complexes of MHC class I, MHC class II, CD20, and tetraspan molecules (CD53, CD81, and CD82) at the surface of a B cell line JY. J Immunol 157:2939–2946
596
M. Zöller
Tagawa K, Arihiro K, Takeshima Y et al (1999) Down-regulation of KAI1 messenger RNA expression is not associated with loss of heterozygosity of the KAI1 gene region in lung adenocarcinoma. Jpn J Cancer Res 90:970–976 Takahashi M, Sugiura T, Abe M et al (2007) Regulation of c-Met signaling by the tetraspanin KAI-1/CD82 affects cancer cell migration. Int J Cancer 121:1919–1929 Takaoka A, Hinoda Y, Satoh S et al (1998) Suppression of invasive properties of colon cancer cells by a metastasis suppressor KAI1 gene. Oncogene 16:1443–1453 Takeda T, Hattori N, Tokuhara T et al (2007a) Adenoviral transduction of MRP-1/CD9 and KAI1/ CD82 inhibits lymph node metastasis in orthotopic lung cancer model. Cancer Res 67: 1744–1749 Takeda Y, Kazarov AR, Butterfield CE et al (2007b) Deletion of tetraspanin Cd151 results in decreased pathologic angiogenesis in vivo and in vitro. Blood 109:1524–1532 Takenawa T, Suetsugu S (2007) The WASP-WAVE protein network: connecting the membrane to the cytoskeleton. Nat Rev Mol Cell Biol 8:37–48 Takino T, Miyamori H, Kawaguchi N et al (2003) Tetraspanin CD63 promotes targeting and lysosomal proteolysis of membrane-type 1 matrix metalloproteinase. Biochem Biophys Res Commun 304:160–166 Tanaka F, Hori N, Sato K (2002) Identification of differentially expressed genes in rat hepatoma cell lines using subtraction and microarray. J Biochem 131:39–44 Tarasova NI, Rice WG, Michejda CJ (1999) Inhibition of G-protein-coupled receptor function by disruption of transmembrane domain interactions. J Biol Chem 274:34911–34915 Telese F, Bruni P, Donizetti A et al (2005) Transcription regulation by the adaptor protein Fe65 and the nucleosome assembly factor SET. EMBO Rep 6:77–82 Testa JE, Brooks PC, Lin JM et al (1999) Eukaryotic expression cloning with an antimetastatic monoclonal antibody identifies a tetraspanin (PETA-3/CD151) as an effector of human tumor cell migration and metastasis. Cancer Res 59:3812–3820 Todeschini RA, Hakomori SI (2008) Functional role of glycosphingolipids and gangliosides in control of cell adhesion, motility, and growth, through glycosynaptic microdomains. Biochim Biophys Acta 1780:421–433 Todeschini AR, Dos Santos JN, Handa K et al (2008) Ganglioside GM2/GM3 complex affixed on silica nanospheres strongly inhibits cell motility through CD82/cMet-mediated pathway. Proc Natl Acad Sci USA 105:1925–1930 Tohami T, Drucker L, Shapiro H et al (2007) Overexpression of tetraspanins affects multiple myeloma cell survival and invasive potential. FASEB J 21:691–699 Tokuhara T, Hasegawa H, Hattori N et al (2001) Clinical significance of CD151 gene expression in non-small cell lung cancer. Clin Cancer Res 7:4109–4114 Tonoli H, Barrett JC (2005) CD82 metastasis suppressor gene: a potential target for new therapeutics? Trends Mol Med 11:563–570 Tsai YC, Mendoza A, Mariano JM et al (2007) The ubiquitin ligase gp78 promotes sarcoma metastasis by targeting KAI1 for degradation. Nat Med 13:1504–1509 Tsitsikov EN, Gutierrez-Ramos JC, Geha RS (1997) Impaired CD19 expression and signaling, enhanced antibody response to type II T independent antigen and reduction of B-1 cells in CD81-deficient mice. Proc Natl Acad Sci USA 94:10844–10849 Tsukita S, Furuse M (2000) The structure and function of claudins, cell adhesion molecules at tight junctions. Ann N Y Acad Sci 915:129–135 Tsukita S, Yonemura S (1999) Cortical actin organization: lessons from ERM (ezrin/radixin/ moesin) proteins. J Biol Chem 274:34507–34510 Tsutsumi S, Shimura T, Morinaga N et al (2005) Loss of KAI1 expression in gastric cancer. Hepatogastroenterology 52:281–284 Ueda T, Ichikawa T, Tamaru J et al (1996) Expression of the KAI1 protein in benign prostatic hyperplasia and prostate cancer. Am J Pathol 149:1435–1440 Valadi H, Ekström K, Bossios A et al (2007) Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol 9:654–659
26 Tetraspanins and Cancer Metastasis
597
van Niel G, Porto-Carreiro I, Simoes S, Raposo G (2006) Exosomes: a common pathway for a specialized function. J Biochem 140:13–21 Walker JW (2008) Protein scaffolds, lipid domains and substrate recognition in protein kinase C function: implications for rational drug design. Handb Exp Pharmacol 186:185–203 Wang JC, Bégin LR, Bérubé NG et al (2007a) Down-regulation of CD9 expression during prostate carcinoma progression is associated with CD9 mRNA modifications. Clin Cancer Res 13:2354–2361 Wang XQ, Yan Q, Sun P et al (2007b) Suppression of epidermal growth factor receptor signaling by protein kinase C-alpha activation requires CD82, caveolin-1, and ganglioside. Cancer Res 67:9986–9995 Wells A, Grandis JR (2003) Phospholipase C-gamma1 in tumor progression. Clin Exp Metastasis 20:285–290 White A, Lamb PW, Barrett JC (1998) Frequent downregulation of the KAI1(CD82) metastasis suppressor protein in human cancer cell lines. Oncogene 16:3143–3149 Wieckowski E, Whiteside TL (2006) Human tumor-derived vs dendritic cell-derived exosomes have distinct biologic roles and molecular profiles. Immunol Res 36:247–254 Winterwood NE, Varzavand A, Meland MN et al (2006) A critical role for tetraspanin CD151 in alpha3beta1 and alpha6beta4 integrin-dependent tumor cell functions on laminin-5. Mol Biol Cell 17:2707–2721 Wright MD, Geary SM, Fitter S et al (2004a) Characterization of mice lacking the tetraspanin superfamily member CD151. Mol Cell Biol 24:5978–5988 Wright MD, Moseley GW, van Spriel AB (2004b) Tetraspanin microdomains in immune cell signalling and malignant disease. Tissue Antigens 64:533–542 Wu Q, Ji Y, Zhang MQ, Chen YQ et al (2003) Role of tumor metastasis suppressor gene KAI1 in digestive tract carcinomas and cancer cells. Cell Tissue Res 314:237–249 Xiao Z, Blonder J, Zhou M et al (2009) Proteomic analysis of extracellular matrix and vesicles. J Proteomics 72:34–45 Xu L, Hynes RO (2007) GPR56 and TG2: possible roles in suppression of tumor growth by the microenvironment. Cell Cycle 6:160–165 Yalaoui S, Zougbédé S, Charrin S et al (2008) Hepatocyte permissiveness to Plasmodium infection is conveyed by a short and structurally conserved region of the CD81 large extracellular domain. PLoS Pathog 4:e1000010 Yamamoto H, Vibitketkumnuen A, Adachi Y et al (2004) Association of matrilysin-2 (MMP-26) expression with tumor progression and activation of MMP-9 in esophageal squamous cell carcinoma. Carcinogenesis 25:2353–2360 Yamamoto Y, Grubisic K, Oelgeschläger M (2007) Xenopus Tetraspanin-1 regulates gastrulation movements and neural differentiation in the early Xenopus embryo. Differentiation 75:235–245 Yamane H, Tachibana I, Takeda Y et al (2005) Propionibacterium acnes-induced hepatic granuloma formation is impaired in mice lacking tetraspanin CD9. J Pathol 206:486–492 Yanez-Mo M, Barreiro O, Gonzalo P et al (2008) MT1-MMP collagenolytic activity is regulated through association with tetraspanin CD151 in primary endothelial cells. Blood 112:3217–3226 Yang J, Weinberg RA (2008) Epithelial-mesenchymal transition: at the crossroads of development and tumor metastasis. Dev Cell 14:818–829 Yang X, Welch DR, Phillips KK et al (1997) KAI1, a putative marker for metastatic potential in human breast cancer. Cancer Lett 119:149–155 Yang X, Claas C, Kraeft SK et al (2002) Palmitoylation of tetraspanin proteins: modulation of CD151 lateral interactions, subcellular distribution, and integrin-dependent cell morphology. Mol Biol Cell 13:767–781 Yang X, Kovalenko OV, Tang W et al (2004) Palmitoylation supports assembly and function of integrin-tetraspanin complexes. J Cell Biol 167:1231–1240 Yang XH, Kovalenko OV, Kolesnikova TV et al (2006) Contrasting effects of EWI proteins, integrins, and protein palmitoylation on cell surface CD9 organization. J Biol Chem 281:12976–12985
598
M. Zöller
Yang XH, Richardson AL, Torres-Arzayus MI et al (2008) CD151 accelerates breast cancer by regulating alpha 6 integrin function, signaling, and molecular organization. Cancer Res 68: 3204–3213 Yates AJ, Rampersaud A (1998) Sphingolipids as receptor modulators. An overview. Ann N Y Acad Sci 845:57–71 Yauch RL, Hemler ME (2000) Specific interactions among transmembrane 4 superfamily (TM4SF) proteins and phosphoinositide 4-kinase. Biochem J 351:629–637 Yauch RL, Berditchevski F, Harler MB et al (1998) Highly stoichiometric, stable, and specific association of integrin alpha3beta1 with CD151 provides a major link to phosphatidylinositol 4-kinase, and may regulate cell migration. Mol Biol Cell 9:2751–2765 Yauch RL, Kazarov AR, Desai B et al (2000) Direct extracellular contact between integrin alpha(3)beta(1) and TM4SF protein CD151. J Biol Chem 275:9230–9238 Yoshida T, Kawano Y, Sato K et al (2008) A CD63 mutant inhibits T-cell tropic human immunodeficiency virus type 1 entry by disrupting CXCR4 trafficking to the plasma membrane. Traffic 9:540–558 You Z, Saims D, Chen S et al (2002) Wnt signaling promotes oncogenic transformation by inhibiting c-Myc-induced apoptosis. J Cell Biol 157:429–440 Yunta M, Lazo PA (2003) Tetraspanin proteins as organisers of membrane microdomains and signalling complexes. Cell Signal 15:559–564 Zakharova L, Svetlova M, Fomina AF (2007) T cell exosomes induce cholesterol accumulation in human monocytes via phosphatidylserine receptor. J Cell Physiol 212:174–181 Zhang X, Xu W (2008) Aminopeptidase N (APN/CD13) as a target for anti-cancer agent design. Curr Med Chem 15:2850–2865 Zhang XA, Bontrager AL, Hemler ME (2001) Transmembrane-4 superfamily proteins associate with activated protein kinase C (PKC) and link PKC to specific beta(1) integrins. J Biol Chem 276:25005–25013 Zhang XA, Kazarov AR, Yang X et al (2002) Function of the tetraspanin CD151-alpha6beta1 integrin complex during cellular morphogenesis. Mol Biol Cell 13:1–11 Zhang XA, He B, Zhou B et al (2003a) Requirement of the p130CAS-Crk coupling for metastasis suppressor KAI1/CD82-mediated inhibition of cell migration. J Biol Chem 278:27319–27328 Zhang XA, Lane WS, Charrin S et al (2003b) EWI2/PGRL associates with the metastasis suppressor KAI1/CD82 and inhibits the migration of prostate cancer cells. Cancer Res 63:2665–2674 Zhao X, Lapalombella R, Joshi T et al (2007) Targeting CD37-positive lymphoid malignancies with a novel engineered small modular immunopharmaceutical. Blood 110:2569–2577 Zheng ZZ, Liu ZX (2007) Activation of the phosphatidylinositol 3-kinase/protein kinase Akt pathway mediates CD151-induced endothelial cell proliferation and cell migration. Int J Biochem Cell Biol 39:340–348 Zhijun X, Shulan Z, Zhuo Z (2007) Expression and significance of the protein and mRNA of metastasis suppressor gene ME491/CD63 and integrin alpha5 in ovarian cancer tissues. Eur J Gynaecol Oncol 28:179–183 Zhou B, Liu L, Reddivari M et al (2004) The palmitoylation of metastasis suppressor KAI1/CD82 is important for its motility- and invasiveness-inhibitory activity. Cancer Res 64:7455–7463 Zhou Z, Ran YL, Hu H et al (2008) TM4SF3 promotes esophageal carcinoma metastasis via upregulating ADAM12m expression. Clin Exp Metastasis 25:537–548 Zhu GZ, Miller BJ, Boucheix C et al (2002) Residues SFQ (173–175) in the large extracellular loop of CD9 are required for gamete fusion. Development 129:1995–2002 Zijlstra A, Lewis J, Degryse B et al (2008) The inhibition of tumor cell intravasation and subsequent metastasis via regulation of in vivo tumor cell motility by the tetraspanin CD151. Cancer Cell 13:221–234 Zöller M (2006) Gastrointestinal tumors: metastasis and tetraspanins. Z Gastroenterol 44:573–586 Zöller M (2009) Tetraspanins: push and pull in suppressing and promoting metastasis. Nat Rev Cancer 9:40–55 Zvieriev V, Wang JC, Chevrette M (2005) Over-expression of CD9 does not affect in vivo tumorigenic or metastatic properties of human prostate cancer cells. Biochem Biophys Res Commun 337:498–504
Part VI
Secreted Proteins
Chapter 27
Chemokines and Metastasis Kalyan C. Nannuru, Seema Singh, and Rakesh K. Singh
Abstract Chemokines are a large group of low-molecular-weight cytokines that are known to direct migration of leukocytes in response to inflammation and pathologic stimuli. Although the primary function of chemokines is well recognized as leukocyte attractants, recent evidences indicate that chemokines and their receptors influence tumor development, growth, angiogenesis, invasion, and metastasis. Chemokines activate cells through cell surface seven trans-membrane, G-protein-coupled receptors (GPCR), resulting in cell invasion, motility, and survival. The role played by chemokines and their receptors in tumor pathophysiology is complex as some chemokines favor tumor growth and metastasis, whereas others may enhance anti-tumor immunity. These diverse functions of chemokines establish them as key mediators between the tumor cells and their microenvironment and play a critical role during tumor progression and metastasis. Manipulating chemokine–chemokine receptor network is an emerging novel targeted therapeutic strategy for various malignancies. In this chapter, we will review recent advances in chemokine research with special emphasis on their role in host–tumor interaction during tumor progression, angiogenesis, and metastasis.
Introduction Despite advances in the use of aggressive adjuvant chemotherapy and radiotherapy, which in combination with surgery, are often successful in the eradication of the primary tumor, most deaths in cancer patients result from metastasis. The process of cancer progression and metastasis comprises a series of sequential interrelated steps, each of which can be rate-limiting (Fidler 2003). The major steps in the formation of a metastasis are as follows: (1) following initial transforming event,
R.K. Singh (*) Department of Pathology and Microbiology, University of Nebraska Medical Center, 985900 Nebraska Medical Center, Omaha, NE 68198-5900, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_27, © Springer Science+Business Media, LLC 2010
601
602
K.C. Nannuru et al.
growth of neoplastic cells must be progressive which requires extensive vascularization if a tumor mass is to exceed 2-mm in diameter; (2) local invasion of the host stroma by some tumor cells occurs as thin-walled venules, like lymphatic channels, offer very little resistance to penetration by tumor cells and provide the most common pathway for tumor-cell entry into the circulation; (3) detachment and embolization of tumor-cell aggregates occurs with subsequent arrest in the capillary beds of organs; (4) extravasation to secondary organ; and (5) proliferation and neovascularization within the distant organ parenchyma to produce detectable metastatic lesions. Intrinsic properties of the tumor cells, as well as their surrounding microenvironment, are crucial in defining the progression of cancer and the fate of metastases. Several factors have been identified that facilitate the interplay between tumor cells and their microenvironment. Proliferation, neovascularization, invasion, and migration of malignant cells to distinct organs are crucial steps for tumor progression and metastasis that can be regulated by chemokines (Fig. 27.1). Over the last 25 years, it has been increasingly recognized that chemokines play an important part in regulation of the metastatic cascade, as they are known to be expressed by tumor cells, as well as by host cells in their proximity and at metastatic Proliferation/Vascularization Migration
Inflammation
Circulation Metastasis Chemokine receptor
Chemokines
Intravasation
Extravasation Chemokines
Primary Tumor Chemokine Receptors
Leukocytes
Chemokines
Angiogenesis
Invasion and motility
Cell growth/survival Immunosuppression
Tumor Progression and Metastasis Fig. 27.1 The multifaceted role of chemokines in tumor growth, invasion, and metastasis. Chemokines produced by tumor cells attract infiltrating leukocyte and/or promote proliferation and can also affect the microenvironment by promoting vascularization. Chemokines can stimulate their specific receptors that alter the adhesive capacity of tumor cells and their migration/invasion into circulation, and extravasation toward distant organs
27 Chemokines and Metastasis
603
sites (Singh et al. 2007; Balkwill 2004). The interactions between chemokine receptors and their ligands coordinate cellular trafficking and responses within different microenvironments. Recent evidences demonstrate the multifunctional role of chemokines and their receptors in regulating cellular growth, survival, angiogenesis, motility and metastasis, and controlling leukocyte infiltration to affect anti-tumor immunity (Ruffini et al. 2007; Balkwill 2004; Vandercappellen et al. 2008; Singh et al. 2007; Ransohoff 2009). A clear understanding of emerging roles of chemokines and the mechanisms of their actions in the processes of malignant progression and metastasis will open new doors for therapeutic interventions. This chapter highlights the functional role of chemokines and their receptors, with focus on angiogenesis, tumor-stromal interaction, and distant metastasis.
Chemokines and Their Receptors The chemokine superfamily includes a large number of low-molecular-weight chemotactic proteins that regulate the trafficking of leukocytes to inflammatory sites (Locati et al. 2002; Murphy et al. 2000; Zlotnik and Yoshie 2000). Chemokines are generally 8–15kDa in size and contain four conserved cysteine amino acid residues linked by disulfide bonds. Structurally, chemokines are classified into four (CXC, CC, C, and CX3C) subfamilies (Murphy et al. 2000; Zlotnik and Yoshie 2000). There are more than 50 chemokines, the majority of which belong to the major CC and CXC chemokine subfamilies (Tables 27.1 and 27.2). According to a new classification, chemokine ligands/receptors are named ‘L’ or ‘R’, respectively (Zlotnik and Yoshie 2000). Receptors are also grouped into four subfamilies, as each receptor binds to one of the four chemokine subfamilies (Tables 27.1 and 27.2). To avoid confusion, we have used the new designations and have listed their old and new names along with their respective receptors in Tables 27.1 and 27.2. Members of the CXC (or a-chemokine) subfamily contain one non-conserved amino acid (X) between the first and second cysteine residues. On the basis of the presence or absence of a Glu-Leu-Arg (ELR) motif, the CXC chemokines can be further subdivided into two groups (ELR+and ELR¯) (Murphy et al. 2000; Murphy 2002; Zlotnik and Yoshie 2000). The ELR motif is located at the N-terminus immediately before the first cysteine amino acid residue (Baggiolini et al. 1997). Extensive investigations regarding the functions of the CXC subfamily have revealed that the presence/absence of the ELR motif determines whether the chemokine is angiogenic or angiostatic (Strieter et al. 1995b, 2006). The ELR+chemokines are primarily chemotactic for endothelial cells and neutrophils. These chemokines are potent promoters of angiogenesis, as the recruited neutrophils are known to synthesize and store angiogenic molecules like vascular endothelial growth factors (VEGF)-A (Scapini et al. 2004; Belperio et al. 2000; Luster 1998; Strieter et al. 1995b, 2006). On the other hand, the main targets for ELR¯ members are T cells and B cells and are potent inhibitors of angiogenesis (Strieter et al. 1995a). The angiogenic CXC chemokine family members include CXCL1-3 and CXCL5-8
604
K.C. Nannuru et al.
Table 27.1 Human CXC, C, and CX3 C chemokines in cancer Chemokines Alternative name Known receptor(s) CXC (a-chemokines) CXCL1 GRO-a /SCYB-1/MGSA/ CXCR1, CXCR2 GRO-1/NAP-3
CXCL2
CXCR2
CXCL9
GRO-b/SCYB-2/GRO-2/ MIp-2a GRO-g/SCYB-3/GRO-3/ MIp-2b PF-4/SCYB-4 ENA-78/SCYB-5 GCP-2/SCYB-6 NAP-2/(SCYB-7/PBP/ CTAP-III/b-TG SCYB-8/GCP-1/NAP-1/ MDNCF MIG/SCYB-9
CXCL10
IP-10/SCYB-10
CXCL11
I-TAC/SCYB-11/b-R1/ H174/IP-9 SDF-1/SCYB-12/PBSF BCA-1/SCYB-13
CXCR3, KSHVGPCR CXCR3
CXCL3 CXCL4 CXCL5 CXCL6 CXCL7 CXCL8
CXCL12 CXCL13
CXCL14 CXCL16
BRAK/SCYB-14/ Bolekine Small inducible cytokine B6
C (g-chemokines) XCL1 Lymphotactin/ SCYC1/SCM-1a/ Lymphotactin a XCL2 SCM-1b/SCYC2/ Lymphotactin b CX3C (DELTA chemokines) Fractalkine/SCYD1 CX3CL1
CXCR2 Unknown CXCR2 CXCR1, CXCR2 CXCR1, CXCR2 CXCR1, CXCR2 CXCR3
CXCR4 CXCR5
Unknown CXCR6
Tumor type Melanoma, breast cancer, myeloma, colon cancer, ovarian cancer Melanoma, breast cancer Melanoma, colon cancer Lung cancer Lung cancer Lung cancer Myelodysplastic syndrome Various malignancies Breast cancer, melanoma Colon cancer, breast cancer Colon cancer, breast cancer Various malignancies Prostate cancer, melanoma, breast cancer Breast cancer, oral carcinoma Prostate cancer, pancreatic cancer
XCR1
Anti-tumor activity
XCR1
Anti-tumor activity
CX3CR1
Various malignancies
Cancer cells express different chemokine and chemokine receptors and modulate tumor growth angiogenesis and metastasis
(Strieter et al. 1995b), whereas angiostatic CXC chemokine members include CXCL4 (Maione et al. 1990) and CXCL9-11 (Maione et al. 1990; Nagpal et al. 2004; Strieter et al. 1995a, 2006). Although CXCL12, an ELR¯chemokine, was originally described as a pre-B-cell growth-stimulating factor (Nagasawa et al. 1996), it has been shown to exhibit angiogenic activity (Mirshahi et al. 2000; Orimo et al. 2005; Salcedo et al. 1999; Tachibana et al. 1998). Furthermore,
27 Chemokines and Metastasis
605
Table 27.2 Human CC chemokines in cancer Chemokines Alternative name Known receptor(s) CC (b-chemokines) CCL1 I-309 /SCYA1 CCR8 CCL2 MCP-1/SCYA2/ CCR2, CCR5,CCR10 MCAF/HC11
Tumor types
MIP-1a/SCYA 3/ LD78a/SIS- a MIP-1b/SCYA4/ACT2/G-26/HC21/ LAG-1/SIS-g RANTES/SCYA5/ SIS-d
CCR1,CCR5
Colon cancer Prostate cancer, melanoma, colon cancer, breast cancer Lung cancer
CCR5, CCR10
Lung cancer
CCR1, CCR3, CCR5, CCR10
CCL7
MCP-3/SCYA7
CCL8
MCP-2/SCYA8/HC14
CCL11
Eotaxin/SCYA11
CCR1, CCR2, CCR3 CCR5 CCR2, CCR3, CCR5 CCR1 CCR3
Prostate cancer, melanoma, colon cancer, breast cancer, renal cell carcinoma Anti-tumor activity
CCL13
MCP-4/SCYA13/Ck b10/NCC-1 HCC-1/SCYA14/Ck b1/MCIF/NCC-2/ CC-1 MIP-1 d/SCYA15/ Lkn-1/HCC-2/ MIP-5/NCC-3/CC-2 HCC-4/SCYA16/Ck b12/LEC/LCC-1/ NCC-4/ILINCK/LMC/ Mtn-1 TARC/SCYA17
CCR1, CCR2, CCR3, CCR5 CCR1
PARC/SCYA18/Ckb7/ DC-CK1/ AMAC-1/MIP-4/ DCtactin MIP-3b/SCYA19/ Ckb11/ELC/ Exodus-3 MIP-3a/SCYA20/ LARC/Exodus-1 6Ckine/SCYA21/ Ckb9/SLC/ Exodus-2 MDC/SCYA22
Unknown
CCL3 CCL4
CCL5
CCL14
CCL15
CCL16
CCL17 CCL18
CCL19
CCL20 CCL21
CCL22
Anti-tumor activity Renal cell carcinoma, Colon cancer Unknown Ovarian cancer
CCR1, CCR3
Melanoma
CCR1
Anti-tumor activity
CCR4
Non-Hodgkin’s lymphoma Melanoma
CCR7
Breast cancer, colon cancer
CCR6
Colon cancer
CCR7
Breast cancer, colon cancer, prostate cancer Breast cancer
CCR4
(continued)
606
K.C. Nannuru et al.
Table 27.2 (continued) Chemokines Alternative name
Known receptor(s)
Tumor types
CCL23
MPIF/SCYA23/Ckb8/ Ckb8-1/MIP-3/ MPIP-1 Eotaxin-2/SCYA24/ Ckb6/MPIF-2 TECK/SCYA25/ Ckb15
CCR1
Unknown
CCR3
Colon cancer
CCR9
Eotaxin-3/SCYA26/ MIP-4a/TSC-1/ IMA CTACK/SCYA27/ ESkine/Skinkine CCL28/SCYA28/MEC
CCR3
Prostate cancer, melanoma, colon cancer, breast cancer Unknown
CCR3, CCR2, CCR10
Melanoma
CCR10, CCR3
Unknown
CCL24 CCL25
CCL26
CCL27 CCL28
Cancer cells express different chemokine and chemokine receptors and modulate tumor growth angiogenesis and metastasis
ELR¯CXC chemokines have also been shown to inhibit neovascularization induced by classical angiogenic factors, such as basic fibroblast growth factor (FGF-2) and vascular endothelial cell growth factor (VEGF) (Strieter et al. 1995a). The CC chemokines (or b-chemokines) represent the largest family of chemokines and have adjacent cysteine residues. The known members of this family are CCL1-5, CCL7-8, CCL11, and CCL13-28, which are listed in Table 27.2. Members of this family exhibit the most diverse range of target cell specificities. They generally attract leukocytes, including monocytes, macrophages, T cells, B cells, basophils, eosinophils, dendritic cells, mast cells, and natural killer cells (Baggiolini et al. 1994; Bischoff et al. 1993; Dahinden et al. 1994; Garcia-Zepeda et al. 1996; Imai et al. 1996; Jose et al. 1994; Kameyoshi et al. 1992; Ponath et al. 1996; Rot et al. 1992). So far, neutrophils have not been shown to respond to chemotactic stimuli from any of the CC chemokines. Chemokines of the C subfamily (g-chemokines) have only one of the cysteine residues, and XCL1 (lymphotactin-a) and XCL2 (lymphotactin-b; Kelner et al. 1994) are the two members of this subfamily, with a molecular size of 16kDa. These chemokines seem to be lymphocyte specific (Kelner et al. 1994). The CX3C chemokine (d-chemokines) has three non-conserved amino acids between the first two cysteines (Bazan et al. 1997). This family also has only one known member, CX3CL1 (Fractalkine), which has been shown to induce both the migration and the adhesion of leukocytes (Segerer et al. 2002; Umehara and Imai 2001). In general, chemokines are secretory proteins, and CX3CL1 is the only exception being a membrane-bound chemokine (Pan et al. 1997). All chemokines exert their biological function by binding to G-protein-coupled receptors (GPCR; Murphy et al. 2000; Murphy 2002; Thelen 2001). Chemokine receptors have an N-terminus outside the cell, three extracellular and three intracellular
27 Chemokines and Metastasis
607
loops, and a C-terminus containing serine and threonine phosphorylation sites in the cytoplasm. In addition, they also possess seven hydrophobic trans-membrane domains. Currently, 6 CXC receptors, 11 CC receptors, and 1 receptor each for C and CX3C have been identified (Tables 27.1and 27.2). Of the six CXC receptors, CXCR1 binds both CXCL8 and CXCL6 with high affinity (Lee et al. 1992; Wuyts et al. 1998). CXCR2, another receptor for CXCL8, has 78% identity with CXCR1 at the amino acid level. It has also been reported that CXCR2 can bind to all of the ELR+CXC chemokines with high affinity (Baggiolini and Loetscher 2000; Locati and Murphy 1999; Wuyts et al. 1998). The third receptor, CXCR3, has been shown to bind the ELRˉ CXC chemokines (CXCL9-11; Cole et al. 1998; Farber 1997; Loetscher et al. 1998a). CXCR4 has been shown to be a cofactor for human immunodeficiency virus (HIV) infection of T lymphocytes (Feng et al. 1996). For CXCR4, the only known ligand is CXCL12, and this ELR¯ CXC chemokine can inhibit HIV infection by competing with lymphotropic HIV virus for binding of CXCR4 (Oberlin et al. 1996). CXCR5, identified as a receptor on B lymphocytes, is responsible for B cell chemotaxis mediated by CXCL13 (Legler et al. 1998). CXCR6 is a receptor for CXCL16, and was described previously as a fusion cofactor for HIV-1 and simian immunodeficiency virus (SIV) (Deng et al. 1997; Matloubian et al. 2000). Among the CC chemokine receptors, CCR1 is a receptor for CCL3, CCL5, CCL7-8, CCL13-16, and CCL23 (Berkhout et al.2000; Gong et al.1996; Hwang et al.2005; Neote et al.1993; Tsou et al.1998).CCR2 exists as two splice variants, CCR2a and CCR2b, among them CCR2b is the most studied and appears to bind at least CCL2, CCL7-8, CCL13, and CCL27 (Charo et al. 1994; Moore et al. 1997; Stellato et al. 1997). CCR3 binds to CCL5, along with CCL7-8, CCL11, CCL13, NNY-CCL14 (CCL14 analogue), CCL15, CCL24, and CCL26-28 (Forssmann et al. 1997, 2004; Ponath et al. 1996; Uguccioni et al. 1997; Youn et al. 1997). CCR4 is a receptor for CCL17 and CCL22 (Imai et al. 1997, 1998), whereas CCR5 is a major co-receptor for macrophage (M)-tropic HIV-1, HIV-2, and SIV strains (Littman 1998). CCR5 can bind to other CC chemokines (CCL2, CCL7-8, and CCL13) with decreased affinity (Blanpain et al. 1999). CCR6 has been shown to be a receptor for CCL20 (Baba et al. 1997). CCL19 and CCL21 bind to CCR7 (Yoshida et al. 1997a), while CCR8 appears to bind CCL1 specifically (Tiffany et al. 1997). CCR9 is a receptor for CCL25 (Zaballos et al. 1999). Functional studies conducted on CCR10 indicate that it has high binding affinity for CCL2, CCL4, CCL27, and CCL28 (Bonini and Steiner 1997; Wang et al. 2000). So far, one receptor each for C (XCR1) and CX3C (CX3CR1) chemokines have been identified (Table 27.1). One of the remarkable features of the chemokine receptor superfamily is their promiscuity in ligand binding (Mantovani 1999). This suggests that the regulation of chemokine activities and the response in target cells are complex events. The intricate functional and regulatory natures of chemokine activities give rise to diverse responses in normal homeostasis and pathological conditions, including tumor growth, invasion, angiogenesis, and metastasis.
608
K.C. Nannuru et al.
Chemokines on Leukocyte Recruitment and Activation in Malignant Tumors Chemokines and their receptor expression, which is dependent on intrinsic properties of the individual cells and the microenvironment, regulate their recruitment and activation within specific microenvironment. The pattern of chemokine receptors’ and their ligands’ expression in a tissue generally correlates with the frequency and type of leukocyte infiltrates. The chemokine gradient that regulates the recruitment can be created by different cell population in response to pathological and inflammatory stimuli. The leukocyte infiltration within tumor microenvironment is regulated by a complex network of chemokines, which influences frequency and phenotype of the immune cells, critical for modulating tumor growth, progression, and metastasis. Most malignant tumors contain tumor cells and stromal cells (macrophages, T and B lymphocytes, eosinophils, granulocytes, natural killer cells, endothelial cells, and fibroblasts; Brigati et al. 2002; Coussens and Werb 2002; Balkwill 2004; Pollard 2009). The number and type of leukocyte infiltration depend on the type of chemokine present in the microenvironment and the specific receptors expressed on the infiltrating cells. The presence of CC chemokines is an important determinant for macrophage and lymphocytic infiltration in a variety of human malignancies (Balkwill and Mantovani 2001; Zlotnik 2006). The extent of macrophage and lymphocyte infiltration into tumors of the same histological origin can vary widely. However, the percentage of tumor-associated macrophages (TAMs) and lymphocytes (TILs) for each tumor is usually maintained as a relatively stable property of a particular tumor during progression (Talmadge et al. 1981; Whiteside et al. 1986). The functional role that these macrophages play in tumor growth is controversial. These cells are located predominantly at the tumor and host cellular interface and represent a potential target for therapy based on immune manipulation. Infiltrating lymphocytes are multifunctional and capable of producing cytokines, enzymes, and different growth inhibitory/stimulatory factors to regulate the initiation, maintenance, and termination of tumor (Perussia 1992; Mantovani et al. 2008). The mechanism of recruitment and the significance of macrophages and lymphocytes in the growth and metastasis of breast cancer have not been studied. In addition, a better understanding of “cross-talk” between these cells and the malignant cell population is essential before the potential of macrophages and lymphocytes as antitumor effector cells can be realized and successful therapeutic strategies implemented. Accumulation of macrophages and lymphocytes at tumor sites has been shown to be mediated by tumor-derived CC-chemokines (Balkwill and Mantovani 2001; Zlotnik 2006). By virtue of receptors on the target cells, CC-chemokines have the potential to preferentially recruit macrophages and T lymphocytes, NK cells, and dendritic cells into the tumors (Balkwill 2004; Ruffini et al. 2007). However, the relationship between CC-chemokines and macrophage- and lymphocyte-mediated immune response in different malignancies remains poorly understood.
27 Chemokines and Metastasis
609
CCL2 was initially considered a monocyte-specific chemoattractant, until in vitro studies showed that CCL2 was the prime agent in phytohemagglutinin (PHA)-stimulated leukocyte culture responsible for T cell chemotaxis (Carr et al. 1994). Importantly, flow cytometric analysis of responding population indicated that only T cells with memory phenotype (CD45RA+, CD45RO+, CD29+, L-selectin+) were involved, because naïve T cells (CD45RA+, CD45RO+CD29 weak, L-selectin+) were unresponsive. In addition, although monocytes began to migrate in response to CCL2 within 1h, a significant T cell response to CCL2 was not evident until 4h. These data were of great significance because although CCL2 had been previously demonstrated within inflammatory sites characterized by monocytes and lymphocytes infiltration, there was no concept of a casual link. This theme of CCL2 involvement in T cell recruitment was elaborated in subsequent studies that confirmed a T cell chemotactic response to CCL2 and extended the data by showing analogous response to CCL2 and CCL7 (Loetscher et al. 1998b, 2000; Loetscher and Clark-Lewis 2001; Youngs et al. 1997). Recent reports demonstrate that differential expression of chemokine receptors and their responsiveness may dictate, to a large extent, the migration and homing of T helper cell type 1 (Th1) and Th2s (Negus et al. 1995, 1998; Milliken et al. 2002; Gu et al. 2000). A two way interaction (negative, positive, or both) occurs between tumor cells and infiltrating leukocytes at certain stage of tumor progression. The role of infiltrating macrophages and lymphocytes in regulating tumor growth can best be viewed as a balance between stimulatory and inhibitory activities. The host response against tumors can be exerted in situ by interaction of different subsets of leukocytes, which results into either positive (by production of immunosuppressive factors, tumor growth, or angiogenic factors) or negative (potent cellular immune response) effects on tumor growth (Locati et al. 2002; Sica et al. 2002). Functional data demonstrate that tumor-infiltrating macrophages and lymphocytes are usually unable to react against the tumors and display a depressed function. There is evidence that macrophage and lymphocyte recruitment and activation is systemically impaired in some cancer patients (Dinapoli et al. 1996; Young and Wright 1992; Varney et al. 2005a, b). If this is a cytokine defect, and not an inherent macrophage and/or lymphocyte defect, manipulation of CC-chemokines may enhance antitumor responses (Huang et al. 1994, 1995). Aberrant levels of chemokines in tumor microenvironment can influence immunosuppressive type 2 macrophage (M2), which release immunosuppressive cytokines interleukin 10 and transforming growth factor (TGF)b (Sica et al. 2000). M2 macrophages also produce higher amounts of CCL2, which can contribute to T-helper 2 (Th2) polarized immunity (Gu et al. 2000). In addition to being immunosuppressive, infiltrating leukocytes contribute to tumor progression by producing matrix metalloproteinases (MMPs), growth, and angiogenic factors (Varney et al. 2005a, b; Lu et al. 2007a, b; Kuroda et al. 2005; Nakashima et al. 1995). Moreover, chronic expression of CCL2 in tumor microenvironment directly or indirectly enhance tumorigenecity and metastasis (Lu et al. 2007a, b; Kuroda et al. 2005; Nakashima et al. 1995; Sica et al. 2002; Varney et al. 2009).
610
K.C. Nannuru et al.
The relationship between patient prognosis, the extent of lymphoreticular infiltration and chemokine, and their receptor expression remains poorly understood. The mechanism(s) of leukocytic recruitment and activation and the significance of this process in tumor growth and metastasis are intensely investigated. A detailed model is required to determine the interrelations (whether positive, negative, or both) between leukocytic infiltration, chemokine and their receptor expression, prognosis, and therapeutic effectiveness, which may offer strategies for the development of novel adjuvant therapies.
Chemokines in Tumor Angiogenesis Angiogenesis is a biological process of new blood vessel formation from preexisting blood vessels and hence is also referred to as neovascularization. It is fundamental to many physiological as well as pathological processes in living organisms (Folkman and Cotran 1976; Folkman 1985; Folkman and Klagsbrun 1987; Leibovich and Wiseman 1988). The process of angiogenesis is regulated by many angiogenic growth factors, enzymes, lipids, and carbohydrates, including the members of the chemokine superfamily (Strieter et al. 2004, 2006). Specific members of the chemokine superfamily can act as pro-angiogenic molecules and support the formation of new blood vessels, while others can antagonize these activities and therefore are angiostatic (Belperio et al. 2000; Strieter et al. 2004). Among all the chemokines, CXCL8 is extensively studied as a potent mediator of angiogenesis. The pro-angiogenic activity of CXCL8 in vivo was confirmed through the use of the rat mesenteric window assay, the rat and rabbit corneal assay, and a subcutaneous sponge model (Hu et al. 1993; Koch et al. 1992; Norrby 1996; Strieter et al. 1992). Human recombinant CXCL8 was shown to be angiogenic when implanted in the rat cornea and induced proliferation and chemotaxis of human umbilical vein endothelial cells (HUVEC) (Koch et al. 1992). In addition, the angiogenic properties of conditioned media from activated monocytes and macrophages were attenuated by CXCL8 anti-sense oligonucleotides (Koch et al. 1992). Furthermore, it was shown that CXCL8 can act directly on vascular endothelial cells by promoting their survival (Yoshida et al. 1997b). Studies from our lab and other groups suggest that CXCL8 stimulates both endothelial cell proliferation and capillary tube formation in vitro in a dose-dependent manner, and both of these effects can be blocked by monoclonal antibodies to CXCL8 (Li et al. 2003; Shono et al. 1996). In addition, CXCL8 was shown to inhibit apoptosis of endothelial cells (Li et al. 2005). CXCL8 exerts its angiogenic activity by up-regulating MMP-2 and MMP-9 in tumor and endothelial cells (Inoue et al. 2000; Li et al. 2005; Luca et al. 1997). MMP-mediated degradation of the extracellular matrix is required for endothelial cell migration, organization, and hence, angiogenesis (McCawley and Matrisian 2000). It has been demonstrated by our group that CXCL8 directly enhances endothelial cell proliferation, survival, and MMP expression in
27 Chemokines and Metastasis
611
CXCR1- and CXCR2-expressing endothelial cells, thus may be an important player in the process of angiogenesis (Li et al. 2003). Several investigators have suggested the angiogenic effect of CXCL8 is independent of its chemotactic and pro-inflammatory effects, since CXCL8 promotes angiogenesis in the absence of inflammatory cells (Hu et al. 1993; Strieter et al. 1992). In addition, it has been reported that there is a direct correlation between high levels of CXCL8 and tumor angiogenesis, progression, and metastasis in nude xenograft models of human cancer cells (Luca et al. 1997; Xie 2001). In an experimental model of ovarian cancer, the expression of CXCL8 was directly correlated with neovascularization and poor survival (Yoneda et al. 1998). CXCL8 may also play an important role in angiogenesis in prostate and breast cancers as elevated serum levels of CXCL8 in the patients with these cancers correlate with disease stage (Benoy et al. 2004; Veltri et al. 1999; Aalinkeel et al. 2004; Lehrer et al. 2004; Uehara et al. 2005). The ability of CXCL8 to elicit angiogenic activity depends on the expression of its receptor by endothelial cells. CXCL8 and its receptors, CXCR1 and CXCR2, have been observed on endothelial cells and have been shown to play a role in endothelial cell proliferation (Koch et al. 1992; Murdoch et al. 1999; Salcedo et al. 2000b). Recent studies indicate that CXCR1 is highly and CXCR2 is moderately expressed on human microvascular endothelial cells (HMEC), whereas HUVEC show low levels of CXCR1 and CXCR2 expressions (Salcedo et al. 2000b). Neutralizing antibodies to CXCR1 and CXCR2 abrogated CXCL8-induced migration of endothelial cells, indicating that these two receptors are critical for the CXCL8 angiogenic response (Li et al. 2005; Salcedo et al. 2000b). Of these two high affinity receptors for CXCL8, the importance of CXCR2 in mediating chemokine-induced angiogenesis was demonstrated to be fundamental to CXCL8-induced neovascularization (Addison et al. 2000; Strieter et al. 2004). The role of CXCR2 in promoting tumor-associated angiogenesis has been confirmed in other tumor systems (Keane et al. 2004; Mestas et al. 2005). In addition to CXCL8, other members of the chemokine family have been shown to play important roles in angiogenesis. Elevated levels of CXCL5 and CXCL8 correlated with the vascularity of non-small cell lung cancer (NSCLC) (Numasaki et al. 2005; Strieter et al. 2004). In a severe, combined immune-deficient (SCID) mouse model system, depletion of CXCL5 resulted in the attenuation of tumor growth, angiogenesis, and spontaneous metastasis (Arenberg et al. 1998). In renal cell carcinoma, elevated levels of CXCL1, CXCL3, CXCL5, and CXCL8 were found to be expressed in the tumor tissue and detected in the plasma, and CXCR2 was found to be expressed on endothelial cells within the tumor biopsies (Mestas et al. 2005). Thus, multiple studies mentioned above demonstrate the important role of ELR+CXC chemokines in tumor angiogenesis. Despite of being a non-ELR, CXC chemokine, CXCL12 is of major interest in tumor angiogenesis. Several lines of evidence indicate that CXCL12 induces endothelial cell migration, proliferation, and tube formation, and increases VEGF release by endothelial cells (Kanda et al. 2003; Neuhaus et al. 2003; Salcedo and Oppenheim 2003). Blockade of CXCL12/CXCR4 signaling results in decreased
612
K.C. Nannuru et al.
tumor growth in vivo due to inhibition of angiogenesis in a VEGF-independent manner (Guleng et al. 2005). The source of CXCL12 that drives angiogenesis is likely to be derived from specialized stromal cells and tumor cells (Barbero et al. 2003). In a recent study, it has been suggested that CXCL12 is partly responsible for the ability of breast carcinoma-associated fibroblasts to promote angiogenesis (Orimo et al. 2005). However, the effects of CXCL12 on angiogenesis cannot be generalized to all tumor systems, as inhibition of metastasis in a model of NSCLC by CXCL12 neutralization did not show reduction of tumor angiogenesis (Phillips et al. 2003). These observations suggest that the action of CXCL12 may be tumor specific and/or may act in cooperation with other angiogenic proteins. As mentioned above, the ELRˉ CXC chemokines include angiostatic members that are known to inhibit neovascularization (Belperio et al. 2000; Salcedo and Oppenheim 2003; Strieter et al. 2004). The angiostatic role of these chemokines has been demonstrated in several studies. For example, CXCL10 potently inhibited CXCL8- and FGF-2-induced angiogenesis (Strieter et al. 1995a). Delivery of CXCL9 or CXCL10 into tumors by injection or by genetic manipulation has been shown to suppress tumor angiogenesis (Arenberg et al. 1996b; Sgadari et al. 1996, 1997). In murine cancer models, intratumoral delivery of immunotherapeutic agents correlates with increased expression of CXCL9 and/or CXCL10 (Dorsey et al. 2002; Ruehlmann et al. 2001). Clinical studies of renal carcinoma patients have also indicated that the intratumoral expression of CXCL9 and CXCL10 results in a decrease in the tumor size (Kondo et al. 2004). It has been shown that CXCR3 is expressed on endothelial cells in a cell cycle-dependent manner, and its expression mediates the angiostatic activity of CXCL9-11 (Romagnani et al. 2001). Recently, it has been suggested that overexpression of CXCL10 in human prostate LNCaP cells activates CXCR3 expression and inhibits cell proliferation (Nagpal et al. 2006). These findings provide definitive evidence of CXCR3-mediated angiostatic activity by angiostatic ELRˉ CXC chemokines. The presence of angiogenic and angiostatic regulators in the CXC chemokine family suggests that tumor angiogenesis may also be affected by the relative expression/activities of the different chemokines in the tumor microenvironment. Recent reports demonstrate that CC chemokines can also participate in angiogenic activity in addition to members of CXC chemokine (Salcedo et al. 2000a). CCL2 has been added to the growing list of angiogenic modulators (Hong et al. 2005; Salcedo et al. 2000a). Previous studies suggest that CCL2 indirectly stimulates angiogenesis (Goede et al. 1999; Leung et al. 1997); however, it has recently been shown that CCL2 may also mediate angiogenic effects by acting directly on endothelial cells and increasing vascularity (Hong et al. 2005). Fractalkine (FKN, CX3CL1), a member of the CX3C chemokine family, also belongs to the list of angiogenesis regulators (Blaschke et al. 2003; Nanki et al. 2004; Volin et al. 2001). Recent studies have suggested that the interaction of FKN and CX3CR1 contributes to the pathogenesis of atherosclerosis (Combadiere et al. 2003; Eriksson 2004) and kidney diseases through the firm adhesion of leukocytes to endothelial cells (Segerer et al. 2002; Umehara et al. 2001). FKN has also been shown to participate in the pathogenesis of rheumatoid arthritis, probably by
27 Chemokines and Metastasis
613
increasing the angiogenic process through endothelial cell activation (Blaschke et al. 2003; Nanki et al. 2004; Volin et al. 2001). The in vivo effect of FKN on angiogenesis has clearly shown that FKN plays a significant role in facilitating inflammatory angiogenesis by activating the GPCR (Lee et al. 2006). Several lines of evidence suggest that a biological imbalance in the production of angiogenic and angiostatic factors, such as chemokines, contributes to the pathogenesis of several angiogenesis-dependent disorders, including cancer, rheumatoid arthritis, and psoriasis (Singh and Fidler 1996; Folkman 1995; Keane et al. 1997; Arenberg et al. 1997; Strieter et al. 1995a). The relative levels of these chemokines and their role in regulating tumor angiogenesis are not clear. More studies are needed to provide evidence, if any, that the imbalance in the expression of angiogenic or angiostatic chemokines regulates tumor angiogenesis. On the basis of the published reports, one can predict that a shift in the balance of expression of these angiogenic and angiostatic chemokines dictates whether the tumor grows and develops metastasis or regresses. If this is correct, it will provide an opportunity to shift this imbalance in favor of angiostasis by modulating the expression of the specific chemokine by pharmacological intervention, which will inhibit tumor growth and metastasis.
Chemokines in Tumor Growth and Metastasis Cancer metastasis consists of multiple, complex interacting and interdependent steps (Fidler 1995; Fidler and Ellis 1994; Nicolson 1991; Singh and Fidler 1996; Fidler et al. 2000) where tumor cells exploit the host responses, which are a part of normal, physiological processes, in order to grow and metastasize, and chemokines play important roles in tumor–host interaction. To begin with, chemokines can provide chemo-attractive signaling that can be critical for cellular trafficking to distant organ sites. Infiltrating leukocytes are not the only cells that express chemokine receptors and respond to chemokine gradient in cancer, many cancer cells themselves express chemokine, chemokine receptors and respond to chemokine gradient (Singh et al. 2007; Balkwill 2003; Muller et al. 2001; Murphy 2001). It has been shown that the site of metastasis depends on the characteristics of neoplastic cells and the specific microenvironment of the secondary organ (Fidler 2002). Similar to the process of leukocyte trafficking, constitutive expression of chemokines by the secondary organ of metastasis can provide signaling cues for malignant cell homing. The role of chemokines in organ-specific metastasis was initially suggested in breast cancer. It was demonstrated that CCR7 and CXCR4 that are highly expressed in breast cancer cells determine the invasion and organ specificity of breast cancer metastasis (Muller et al. 2001). The ligands CCL21 and CXCL12 for these receptors are highly expressed in the organs (lung, liver, and bone) that are preferred sites for breast cancer metastasis (Muller et al. 2001). Recently, it has been suggested that osteoclasts may promote metastasis of lung cancer cells expressing CCR4 to the bone marrow by producing its ligand CCL22 (Nakamura et al. 2006).
614
K.C. Nannuru et al.
Stimulation of osteoclast-like cells with CCR4 results in up-regulation of its ligand, CCL22. In addition, it has been demonstrated that a human lung cancer cell line which expresses CCR4 metastasizes to bone when injected intravenously into NK cell-depleted SCID mice, (Nakamura et al. 2006). Together, the above data supports the idea that the expression of chemokine ligands or their receptors in the organ environment or by malignant cells plays a major role in organ-specific metastasis. CXCR4 appears to be the major chemokine receptor expressed on cancer cells. The expression of CXCR4 has been reported in malignant cells from more than 23 different types of cancers. The accumulating evidence suggests that CXCR4 is an important regulator of breast cancer metastasis and can be predictive of lymph node metastasis (Cabioglu et al. ; Hao et al. 2007; Smith et al. 2004). In experimental studies, treatment of CXCR4-expressing breast cancer cells with neutralizing anti-CXCR4 antibody reduced metastasis to lung in a mouse model (Muller et al. 2001). Gene array analysis of human breast cancer myoepithelial cells and myofibroblasts demonstrated up-regulation of both CXCR4 and CXCL12 which serve to enhance migration and invasion (Allinen et al. 2004). CXCR4 expression also mediates organ-specific metastasis of pancreatic cancer cells and a strong association of CXCR4 with advanced pancreatic cancer has also been suggested (Saur et al. 2005; Wehler et al. 2006). In vitro stimulation of CXCR4-positive cancer cells with CXCL12 resulted in directed migration/ invasion of ovarian cancer cells (Scotton et al. 2002). Stimulation with CXCL12 also up regulated integrin expression and facilitated adhesion in lung cancer cells (Burger et al. 2003). Furthermore, it has been shown that the expression of CXCR4 on human renal cell carcinoma (RCC) correlates with their metastatic ability in both heterotopic and orthotopic SCID mouse models (Pan et al. 2006). In an orthotopic SCID mouse model of RCC, treatment with specific antiCXCL12 antibodies markedly abrogated metastasis of RCC to target organs expressing high levels of CXCL12, without significant changes in tumor cell proliferation, apoptosis, or tumor-associated angiogenesis (Pan et al. 2006). The above findings support the notion that the CXCL12/CXCR4 axis plays a critical role in regulating tumor growth and metastasis. In addition, the clinical relevance of CXCR4 expression has been demonstrated in various types of cancers. In esophageal cancer and melanoma, CXCR4 expression is associated with poor clinical outcome (Kaifi et al. 2005; Scala et al. 2005). In breast cancer, CXCR4 expression predicted lymph node metastasis (Cabioglu et al. 2005). Patients with high CXCR4 expression in NSCLC, colorectal or prostate cancer was more prone to metastasis (Arya et al. 2004; Schimanski et al. 2005; Su et al. 2005). The expression of CXCR4 in nasopharyngeal carcinoma and osteosarcoma is also associated with metastasis (Hu et al. 2005; Laverdiere et al. 2005). In colorectal cancer, CXCR4 expression in primary tumors demonstrated significant association with recurrence, survival, and liver metastasis (Kim et al. 2005). In contrast, a study on neuroblastoma patients found that, although neuroblastoma cells expressed CXCR4, but was not functional (Airoldi et al. 2006), and a study on approximately 300 breast cancer patients found that expression of CXCR4 was not
27 Chemokines and Metastasis
615
associated with metastasis (Weigelt et al. 2005). Above data strongly suggest the role of CXCR4 in regulating the progression of tumor cells to metastasize; however, additional investigations are needed to resolve the ambiguities in different studies. CXCL8, a potent chemoattractant, has been demonstrated to contribute to human cancer progression through its potential function as a mitogenic and angiogenic factor. Elevated levels of CXCL8 have been detected in variety of tumors, such as ovarian carcinoma (Ivarsson et al. 2000), NSCLC (Arenberg et al. 1996a), metastatic melanoma (Varney et al. 2006), and colon carcinoma (Grimm et al. 1996). Studies from our laboratory and others suggest that the expression of CXCL8 correlates positively with disease progression (Nurnberg et al. 1999; Scheibenbogen et al. 1995; Singh et al. 1994; Ugurel et al. 2001). A concomitant up-regulation of one of the two putative CXCL8 receptors has been reported in human melanoma specimens. Analysis of CXCR1 in human melanoma specimens from different Clark levels demonstrated that it is expressed ubiquitously in all Clark levels. In contrast, CXCR2 is expressed predominantly by higher grade melanoma tumors and metastases, suggesting an association between expression of CXCL8 and CXCR2 with vessel density in advanced lesions and metastasis (Varney et al. 2006). More specifically, the effect of CXCL8 can be mediated by CXCR1 and CXCR2, with CXCR1 being a selective receptor for CXCL8 (Addison et al. 2000). Together, these data implicate that CXCL8 can directly modulate growth and the metastatic phenotype of cancer cells. Antibodies to chemokines have shown some promise as a therapeutic modality for treatment of malignant melanoma. Earlier studies have also demonstrated that neutralizing antibodies to CXCR1 and CXCR2 inhibits melanoma cell proliferation and their invasive potential (Varney et al. 2003). Humanized antibodies to CXCL8 have also been shown to inhibit melanoma growth, angiogenesis, and metastasis (Villares et al. 2008; Melnikova and Bar-Eli 2006). Neutralizing antibodies against other chemokines also shows similar results as shown by antibodies against CXCL8 suggesting that melanoma may utilize more than one chemokine ligand pathway to support its growth (Fujisawa et al. 1999). All these evidences emphasize on targeting CXCL8 receptors rather than CXCL8 alone. Also, it has been reported that 17beta-estradiol, progesterone, and dihydrotestosterone suppress the growth of melanoma by inhibiting CXCL8 production in a receptordependent manner (Bendrik and Dabrosin 2009). Antagonist for CXCL8 receptors is also under consideration for melanoma therapy. Small molecule inhibitors with affinity for CXCR1 such as repertaxin or affinity for CXCR2 such as SB-225002 or SB-332235 have been used against inflammatory diseases (Bertini et al. 2004; Thatcher et al. 2005). A recent study has shown the potential of the CXCR2/1specific inhibitors, SCH-479833 and SCH-527123, in inhibiting human melanoma growth by decreasing tumor cell proliferation, survival, and invasion (Singh et al. 2009a). Treatment of melanoma cells with SCH-479833 or SCH-527123 also inhibited tumor growth. Histological and histochemical analyses showed significant (p<0.05) decreases in tumor cell proliferation and microvessel density in tumors. Moreover, we observed a significant increase in melanoma cell apoptosis
616
K.C. Nannuru et al.
in SCH-479833- or SCH-527123-treated animals as compared to controls (Singh et al. 2009a, b). These observations provide basis for an exciting opportunity to utilize these antagonists for future melanoma therapy. CCR7 is another chemokine receptor highly expressed by several tumor types. CCR7 plays a critical role in lymphocyte and dendritic cell trafficking into and within lymph nodes. Human cancer cells from malignant breast, gastric, oesophageal, NSCLC, squamous cell carcinoma of the head and neck, and colorectal carcinomas express functional CCR7 (Ding et al. 2003; Gunther et al. 2005; Mashino et al. 2002; Muller et al. 2001; Takanami 2003; Wang et al. 2005). Cancer cells that overexpress CCR7 can migrate toward its ligand CCL21, which is strongly expressed in lymph nodes (Mashino et al. 2002; Takeuchi et al. 2004). It has been reported that when CCR7 was overexpressed in a murine melanoma cell line, metastasis to lymph nodes was increased in vivo (Wiley et al. 2001). Recently, it has been reported that CCR7 activation on thyroid carcinoma cell by CCL21 favors tissue invasion and cell proliferation, and therefore may promote thyroid carcinoma growth and lymph node metastasis (Sancho et al. 2006). CCR7 is a potential marker to predict metastasis in breast and colorectal cancer and is correlated with disease progression (Cabioglu et al. 2005; Gunther et al. 2005). In addition, patients with oral and oropharyngeal squamous cell carcinomas that express CCR7 have lower survival rates than those with CCR7-negative tumors (Tsuzuki et al. 2006). In addition, there are many more chemokines that have been detected in various cancers and are known to play a role in metastasis. Recently, CCL2 was also shown to be extremely important in tumorigenesis and bone metastasis of several solid tumors (Craig and Loberg 2006). The expression of CCR2 in prostate cancer and CCR3 in human renal carcinoma correlates with disease progression (Lu et al. 2006; Johrer et al. 2005). Co-expression of CCR4 and CCR10, the known pair of skin-homing chemokine receptors, may play an important role in adult T-cell leukemia/lymphoma (ATLL) invasion into the skin (Harasawa et al. 2006). CCL5 and its receptor CCR5 have been detected on prostate cancer cell lines (Vaday et al. 2006). CCR6 and its ligand have been shown to be involved in colorectal cancer and its metastasis (Ghadjar et al. 2006); CCR9 on prostate and melanoma cells (Letsch et al. 2004; Singh et al. 2004); and CCR10 on melanoma cells (Muller et al. 2001). The expression of CCR10 and CCL27 in human melanomas increases the ability of neoplastic cells to grow, invade, and disseminate to lymph node (Simonetti et al. 2006). Expression of CXCR5 in carcinomas seemed highly unlikely, but recently it has been shown that CXCR5 is expressed by colon carcinoma cells and promotes tumor growth and metastasis to liver (Meijer et al. 2006). In addition, prostate cancer cells express CX3CR1 and stimulation with its ligand CX3CL1 resulted in increased adhesion and migration (Shulby et al. 2004). It is possible that the cancer cell can exploit interferon-inducible chemokines for their own propagation, survival, and metastasis. For example, CXCL9 and CXCL10 induce the migration and invasion of melanoma cell and multiple myeloma cells (Kawada et al. 2004; Pellegrino et al. 2004; Soejima and Rollins 2001). CXCR3 the receptor for CXCL9 and CXCL10 was shown to be expressed by cancer cells of solid tumors, such as breast tumor cells, renal carcinoma cells,
27 Chemokines and Metastasis
617
and melanoma cells (Goldberg-Bittman et al. 2004; Kawada et al. 2004; LongoImedio et al. 2005; Suyama et al. 2005). A direct role for CXCR3 in inducing tumor cell migration to metastatic sites was suggested by observations showing that the expression of this chemokine receptor by melanoma cells is causally involved in metastasis to lymph nodes (Soejima and Rollins 2001). Recently, it has also been demonstrated that CXCR3 plays a critical role in colon cancer cell metastasis to lymph nodes by inducing diverse cellular effects (Kawada et al. 2007). These data suggest that anti- or pro-malignant behavior of these chemokines (such as CXCL9 and CXCL10) would be in part dictated by their receptor expression by the cancer cell. Overall, the aberrant expression of chemokines and their receptors may be a common feature of various epithelial cancers. Although, the functional significance of some of these chemokines is yet to be elucidated, we can contemplate that the expression of certain chemokines and their receptors plays a key role in dictating the fate of developing tumors and their ability to metastasize to certain preferred organ sites.
Chemokines Targeting and Chemotherapy Current strategies focused on systemic therapy for the treatment of metastatic solid tumors have shown no impact on improving patient survival. Different therapeutic approaches have been evaluated including chemotherapy and biological therapy, both as single agent or in combination. Conventional systemic chemotherapy is still considered the mainstay of treatment for higher stage tumors. Recent reports suggest that the expression of CXCR2 ligands and the activation of CXCR2-dependent pathways might provide survival signal for therapy-resistant malignant tumor cells (De Larco et al. 2001a, b; Acosta and Gil 2009). An increase in the expression of CXCR1 and CXCR2 and their ligands in response to chemotherapy in various cancers has been observed (Lev et al. 2003; De Larco et al. 2001a, b; Maxwell et al. 2007). Increase in the level of CXCL8 and CXCL1 after chemotherapy suggests that cancerous cells may use this phenomenon either as an escape mechanism from the drug action or it is the druginduced pathway to kill cancerous cells. CXCL-8 level increases in vitro and in patients with ovarian cancer after treatment with paclitaxel suggesting that it kills cancerous cells through CXCL-8-mediated mechanism (Uslu et al. 2005), whereas the survivors of sequential treatment with Adriamycin and/or 5¢-fluro2¢-deoxyuridine have higher CXCL-8 levels and display the greatest phenotypic variance from parental cells in case of MCF-7 tumors and these tumors have more rapid initial growth phase in situ and give rise to spontaneous lung metastases within 10 weeks in breast cancer (De Larco et al. 2001b). Treatment of malignant melanoma cells with Dacarbazine transcriptionally up regulates CXCL-8 expression, which might render them resistant to the cytotoxic effect of drugs (Lev et al. 2003).
618
K.C. Nannuru et al.
Conclusion and Future Perspective Overexpression of chemokines and chemokine receptors appears to be a hallmark of cancer. Identifying the specific networks of chemokines and their receptors and understanding the effect of their expression on local immune response, angiogenesis, and their interaction with other factors at the tumor milieu will reveal novel opportunities for restricting tumor growth and metastasis. The accumulating evidence resulting from the experimental studies points toward a critical role of chemokines and their receptors in cancer progression and metastasis. Many retrospective studies have now demonstrated that the expression of various chemokines and their receptors is associated with poor prognosis. This advocates that the measurement of chemokines and their receptor expression in tumor samples has the potential of becoming a biomarker in assessing tumor aggressiveness. This will be helpful in selecting appropriate treatments for certain cancer patients. From the scientific perspective, we foresee that a better understanding of the biological and molecular mechanisms of chemokine-dependent regulation of cellular phenotypes associated with tumor growth, angiogenesis, and metastasis will identify the cellular and molecular targets that regulate distinct functional interactions of chemokines with their receptors on malignant cells and endothelial cells. An understanding of the mechanism(s) regulating modulation of chemokine-receptor pathways will be critical in prognosis and designing effective strategies for the development of novel targeted molecular therapeutics. Acknowledgements This work was supported in part by grants CA72781 (R.K.S.) and Cancer Center Support Grant (P30CA036727) from National Cancer Institute, National Institutes of Health and Nebraska Research Initiative Cancer Glycobiology Program (R.K.S.).
References Aalinkeel R, Nair MP, Sufrin G, Mahajan SD, Chadha KC, Chawda RP, Schwartz SA (2004) Gene expression of angiogenic factors correlates with metastatic potential of prostate cancer cells. Cancer Res 64:5311–5321 Acosta JC, Gil J (2009) A role for CXCR2 in senescence, but what about in cancer? Cancer Res 69:2167–2170 Addison CL, Daniel TO, Burdick MD, Liu H, Ehlert JE, Xue YY, Buechi L, Walz A, Richmond A, Strieter RM (2000) The CXC chemokine receptor 2, CXCR2, is the putative receptor for ELR+ CXC chemokine-induced angiogenic activity. J Immunol 165:5269–5277 Airoldi I, Raffaghello L, Piovan E, Cocco C, Carlini B, Amadori A, Corrias MV, Pistoia V (2006) CXCL12 does not attract CXCR4+ human metastatic neuroblastoma cells: clinical implications. Clin Cancer Res 12:77–82 Allinen M, Beroukhim R, Cai L, Brennan C, Lahti-Domenici J, Huang H, Porter D, Hu M, Chin L, Richardson A, Schnitt S, Sellers WR, Polyak K (2004) Molecular characterization of the tumor microenvironment in breast cancer. Cancer Cell 6:17–32 Arenberg DA, Kunkel SL, Polverini PJ, Glass M, Burdick MD, Strieter RM (1996a) Inhibition of interleukin-8 reduces tumorigenesis of human non-small cell lung cancer in SCID mice. J Clin Invest 97:2792–2802
27 Chemokines and Metastasis
619
Arenberg DA, Kunkel SL, Polverini PJ, Morris SB, Burdick MD, Glass MC, Taub DT, Iannettoni MD, Whyte RI, Strieter RM (1996b) Interferon-gamma-inducible protein 10 (IP-10) is an angiostatic factor that inhibits human non-small cell lung cancer (NSCLC) tumorigenesis and spontaneous metastases. J Exp Med 184:981–992 Arenberg DA, Polverini PJ, Kunkel SL, Shanafelt A, Hesselgesser J, Horuk R, Strieter RM (1997) The role of CXC chemokines in the regulation of angiogenesis in non-small cell lung cancer. J Leukoc Biol 62:554–562 Arenberg DA, Keane MP, DiGiovine B, Kunkel SL, Morris SB, Xue YY, Burdick MD, Glass MC, Iannettoni MD, Strieter RM (1998) Epithelial-neutrophil activating peptide (ENA-78) is an important angiogenic factor in non-small cell lung cancer. J Clin Invest 102: 465–472 Arya M, Patel HR, McGurk C, Tatoud R, Klocker H, Masters J, Williamson M (2004) The importance of the CXCL12-CXCR4 chemokine ligand-receptor interaction in prostate cancer metastasis. J Exp Ther Oncol 4:291–303 Baba M, Imai T, Nishimura M, Kakizaki M, Takagi S, Hieshima K, Nomiyama H, Yoshie O (1997) Identification of CCR6, the specific receptor for a novel lymphocyte-directed CC chemokine LARC. J Biol Chem 272:14893–14898 Baggiolini M, Loetscher P (2000) Chemokines in inflammation and immunity. Immunol Today 21:418–420 Baggiolini M, Dewald B, Moser B (1994) Interleukin-8 and related chemotactic cytokines – CXC and CC chemokines. Adv Immunol 55:97–179 Baggiolini M, Dewald B, Moser B (1997) Human chemokines: an update. Annu Rev Immunol 15:675–705 Balkwill F (2003) Chemokine biology in cancer. Semin Immunol 15:49–55 Balkwill F (2004) Cancer and the chemokine network. Nat Rev Cancer 4:540–550 Balkwill F, Mantovani A (2001) Inflammation and cancer: back to Virchow? Lancet 357:539–545 Barbero S, Bonavia R, Bajetto A, Porcile C, Pirani P, Ravetti JL, Zona GL, Spaziante R, Florio T, Schettini G (2003) Stromal cell-derived factor 1alpha stimulates human glioblastoma cell growth through the activation of both extracellular signal-regulated kinases 1/2 and Akt. Cancer Res 63:1969–1974 Bazan JF, Bacon KB, Hardiman G, Wang W, Soo K, Rossi D, Greaves DR, Zlotnik A, Schall TJ (1997) A new class of membrane-bound chemokine with a CX3C motif. Nature 385:640–644 Belperio JA, Keane MP, Arenberg DA, Addison CL, Ehlert JE, Burdick MD, Strieter RM (2000) CXC chemokines in angiogenesis. J Leukoc Biol 68:1–8 Bendrik C, Dabrosin C (2009) Estradiol increases IL-8 secretion of normal human breast tissue and breast cancer in vivo. J Immunol 182:371–378 Benoy IH, Salgado R, Van DP, Geboers K, Van ME, Scharpe S, Vermeulen PB, Dirix LY (2004) Increased serum interleukin-8 in patients with early and metastatic breast cancer correlates with early dissemination and survival. Clin Cancer Res 10:7157–7162 Berkhout TA, Gohil J, Gonzalez P, Nicols CL, Moores KE, Macphee CH, White JR, Groot PH (2000) Selective binding of the truncated form of the chemokine CKbeta8 (25–99) to CC chemokine receptor 1(CCR1). Biochem Pharmacol 59:591–596 Bertini R, Allegretti M, Bizzarri C, Moriconi A, Locati M, Zampella G, Cervellera MN, Di Cioccio V, Cesta MC, Galliera E, Martinez FO, Di Bitondo R, Troiani G, Sabbatini V, D’Anniballe G, Anacardio R, Cutrin JC, Cavalieri B, Mainiero F, Strippoli R, Villa P, Di GM, Martin F, Gentile M, Santoni A, Corda D, Poli G, Mantovani A, Ghezzi P, Colotta F (2004) Noncompetitive allosteric inhibitors of the inflammatory chemokine receptors CXCR1 and CXCR2: prevention of reperfusion injury. Proc Natl Acad Sci U S A 101:11791–11796 Bischoff SC, Krieger M, Brunner T, Rot A, von Tscharner V, Baggiolini M, Dahinden CA (1993) RANTES and related chemokines activate human basophil granulocytes through different G protein-coupled receptors. Eur J Immunol 23:761–767
620
K.C. Nannuru et al.
Blanpain C, Migeotte I, Lee B, Vakili J, Doranz BJ, Govaerts C, Vassart G, Doms RW, Parmentier M (1999) CCR5 binds multiple CC-chemokines: MCP-3 acts as a natural antagonist. Blood 94:1899–1905 Blaschke S, Koziolek M, Schwarz A, Benohr P, Middel P, Schwarz G, Hummel KM, Muller GA (2003) Proinflammatory role of fractalkine (CX3CL1) in rheumatoid arthritis. J Rheumatol 30:1918–1927 Bonini JA, Steiner DF (1997) Molecular cloning and expression of a novel rat CC-chemokine receptor (rCCR10rR) that binds MCP-1 and MIP-1beta with high affinity. DNA Cell Biol 16:1023–1030 Brigati C, Noonan DM, Albini A, Benelli R (2002) Tumors and inflammatory infiltrates: friends or foes? Clin Exp Metastasis 19:247–258 Burger M, Glodek A, Hartmann T, Schmitt-Graff A, Silberstein LE, Fujii N, Kipps TJ, Burger JA (2003) Functional expression of CXCR4 (CD184) on small-cell lung cancer cells mediates migration, integrin activation, and adhesion to stromal cells. Oncogene 22:8093–8101 Cabioglu N, Yazici MS, Arun B, Broglio KR, Hortobagyi GN, Price JE, Sahin A (2005) CCR7 and CXCR4 as novel biomarkers predicting axillary lymph node metastasis in T1 breast cancer. Clin Cancer Res 11:5686–5693 Carr MW, Roth SJ, Luther E, Rose SS, Springer TA (1994) Monocyte chemoattractant protein 1 acts as a T-lymphocyte chemoattractant. Proc Natl Acad Sci USA 91:3652–3656 Charo IF, Myers SJ, Herman A, Franci C, Connolly AJ, Coughlin SR (1994) Molecular cloning and functional expression of two monocyte chemoattractant protein 1 receptors reveals alternative splicing of the carboxyl-terminal tails. Proc Natl Acad Sci U S A 91:2752–2756 Cole KE, Strick CA, Paradis TJ, Ogborne KT, Loetscher M, Gladue RP, Lin W, Boyd JG, Moser B, Wood DE, Sahagan BG, Neote K (1998) Interferon-inducible T cell alpha chemoattractant (I-TAC): a novel non-ELR CXC chemokine with potent activity on activated T cells through selective high affinity binding to CXCR3. J Exp Med 187:2009–2021 Combadiere C, Potteaux S, Gao JL, Esposito B, Casanova S, Lee EJ, Debre P, Tedgui A, Murphy PM, Mallat Z (2003) Decreased atherosclerotic lesion formation in CX3CR1/apolipoprotein E double knockout mice. Circulation 107:1009–1016 Coussens LM, Werb Z (2002) Inflammation and cancer. Nature 420:860–867 Craig MJ, Loberg RD (2006) CCL2 (monocyte chemoattractant protein-1) in cancer bone metastases. Cancer Metastasis Rev 25:611–619 Dahinden CA, Geiser T, Brunner T, von Tscharner V, Caput D, Ferrara P, Minty A, Baggiolini M (1994) Monocyte chemotactic protein 3 is a most effective basophil- and eosinophil-activating chemokine. J Exp Med 179:751–756 De Larco JE, Wuertz BRK, Manivel JC, Furcht LT (2001a) Progression and enhancement of metastatic potential after exposure of tumor cells to chemotherapeutic agents. Cancer Res 61: 2857–2861 De Larco JE, Wuertz BRK, Rosner KA, Erickson SA, Gamache DE, Manivel JC, Furcht LT (2001b) A potential role for interleukin-8 in the metastatic phenotype of breast carcinoma cells. Am J Pathol 158:639–646 Deng HK, Unutmaz D, KewalRamani VN, Littman DR (1997) Expression cloning of new receptors used by simian and human immunodeficiency viruses. Nature 388:296–300 Dinapoli MR, Calderon CL, Lopez DM (1996) The altered tumoricidal capacity of macrophages isolated from tumor-bearing mice is related to reduce expression of the inducible nitric oxide synthase gene. J Exp Med 183:1323–1329 Ding Y, Shimada Y, Maeda M, Kawabe A, Kaganoi J, Komoto I, Hashimoto Y, Miyake M, Hashida H, Imamura M (2003) Association of CC chemokine receptor 7 with lymph node metastasis of esophageal squamous cell carcinoma. Clin Cancer Res 9:3406–3412 Dorsey R, Kundu N, Yang Q, Tannenbaum CS, Sun H, Hamilton TA, Fulton AM (2002) Immunotherapy with interleukin-10 depends on the CXC chemokines inducible protein-10 and monokine induced by IFN-gamma. Cancer Res 62:2606–2610 Eriksson EE (2004) Mechanisms of leukocyte recruitment to atherosclerotic lesions: future prospects. Curr Opin Lipidol 15:553–558
27 Chemokines and Metastasis
621
Farber JM (1997) Mig and IP-10: CXC chemokines that target lymphocytes. J Leukoc Biol 61:246–257 Feng Y, Broder CC, Kennedy PE, Berger EA (1996) HIV-1 entry cofactor: functional cDNA cloning of a seven-transmembrane, G protein-coupled receptor. Science 272:872–877 Fidler IJ (1995) Cancer biology: invasion and metastasis. Churchill Livingstone, New York Fidler IJ (2002) The organ microenvironment and cancer metastasis. Differentiation 70:498–505 Fidler IJ (2003) The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nat Rev Cancer 3:453–458 Fidler IJ, Ellis LM (1994) The implications of angiogenesis for the biology and therapy of cancer metastasis. Cell 79:185–188 Fidler IJ, Singh RK, Yoneda J, Kumar R, Xu L, Dong Z, Bielenberg DR, McCarty M, Ellis LM (2000) Critical determinants of neoplastic angiogenesis. Cancer J Sci Am 6(Suppl 3): S225–S236 Folkman J (1985) Tumor angiogenesis. Adv Cancer Res 43:175–203 Folkman J (1995) Angiogenesis in cancer, vascular, rheumatoid and other disease. Nat Med 1:27–31 Folkman J, Cotran R (1976) Relation of vascular proliferation to tumor growth. Int Rev Exp Pathol 16:207–248 Folkman J, Klagsbrun M (1987) Vascular physiology. A family of angiogenic peptides. Nature 329:671–672 Forssmann U, Uguccioni M, Loetscher P, Dahinden CA, Langen H, Thelen M, Baggiolini M (1997) Eotaxin-2, a novel CC chemokine that is selective for the chemokine receptor CCR3, and acts like eotaxin on human eosinophil and basophil leukocytes. J Exp Med 185:2171–2176 Forssmann U, Hartung I, Balder R, Fuchs B, Escher SE, Spodsberg N, Dulkys Y, Walden M, Heitland A, Braun A, Forssmann WG, Elsner J (2004) n-Nonanoyl-CC chemokine ligand 14, a potent CC chemokine ligand 14 analogue that prevents the recruitment of eosinophils in allergic airway inflammation. J Immunol 173:3456–3466 Fujisawa N, Hayashi S, Miller EJ (1999) A synthetic peptide inhibitor for alpha-chemokines inhibits the tumour growth and pulmonary metastasis of human melanoma cells in nude mice. Melanoma Res 9:105–114 Garcia-Zepeda EA, Rothenberg ME, Ownbey RT, Celestin J, Leder P, Luster AD (1996) Human eotaxin is a specific chemoattractant for eosinophil cells and provides a new mechanism to explain tissue eosinophilia. Nat Med 2:449–456 Ghadjar P, Coupland SE, Na IK, Noutsias M, Letsch A, Stroux A, Bauer S, Buhr HJ, Thiel E, Scheibenbogen C, Keilholz U (2006) Chemokine receptor CCR6 expression level and liver metastases in colorectal cancer. J Clin Oncol 24:1910–1916 Goede V, Brogelli L, Ziche M, Augustin HG (1999) Induction of inflammatory angiogenesis by monocyte chemoattractant protein-1. Int J Cancer 82:765–770 Goldberg-Bittman L, Neumark E, Sagi-Assif O, Azenshtein E, Meshel T, Witz IP, Ben-Baruch A (2004) The expression of the chemokine receptor CXCR3 and its ligand, CXCL10, in human breast adenocarcinoma cell lines. Immunol Lett 92:171–178 Gong JH, Uguccioni M, Dewald B, Baggiolini M, Clark-Lewis I (1996) RANTES and MCP-3 antagonists bind multiple chemokine receptors. J Biol Chem 271:10521–10527 Grimm MC, Elsbury SK, Pavli P, Doe WF (1996) Interleukin 8: cells of origin in inflammatory bowel disease. Gut 38:90–98 Gu L, Tseng S, Horner RM, Tam C, Loda M, Rollins BJ (2000) Control of TH2 polarization by the chemokine monocyte chemoattractant protein-1. Nature 404:407–411 Guleng B, Tateishi K, Ohta M, Kanai F, Jazag A, Ijichi H, Tanaka Y, Washida M, Morikane K, Fukushima Y, Yamori T, Tsuruo T, Kawabe T, Miyagishi M, Taira K, Sata M, Omata M (2005) Blockade of the stromal cell-derived factor-1/CXCR4 axis attenuates in vivo tumor growth by inhibiting angiogenesis in a vascular endothelial growth factor-independent manner. Cancer Res 65:5864–5871 Gunther K, Leier J, Henning G, Dimmler A, Weissbach R, Hohenberger W, Forster R (2005) Prediction of lymph node metastasis in colorectal carcinoma by expressionof chemokine receptor CCR7. Int J Cancer 116:726–733
622
K.C. Nannuru et al.
Hao L, Zhang C, Qiu Y, Wang L, Luo Y, Jin M, Zhang Y, Guo TB, Matsushima K, Zhang Y (2007) Recombination of CXCR4, VEGF, and MMP-9 predicting lymph node metastasis in human breast cancer. Cancer Lett 253:34–42 Harasawa H, Yamada Y, Hieshima K, Jin Z, Nakayama T, Yoshie O, Shimizu K, Hasegawa H, Hayashi T, Imaizumi Y, Ikeda S, Soda H, Soda H, Atogami S, Takasaki Y, Tsukasaki K, Tomonaga M, Murata K, Sugahara K, Tsuruda K, Kamihira S (2006) Survey of chemokine receptor expression reveals frequent co-expression of skin-homing CCR4 and CCR10 in adult T-cell leukemia/lymphoma. Leuk Lymphoma 47:2163–2173 Hong KH, Ryu J, Han KH (2005) Monocyte chemoattractant protein-1-induced angiogenesis is mediated by vascular endothelial growth factor-A. Blood 105:1405–1407 Hu DE, Hori Y, Fan TP (1993) Interleukin-8 stimulates angiogenesis in rats. Inflammation 17:135–143 Hu J, Deng X, Bian X, Li G, Tong Y, Li Y, Wang Q, Xin R, He X, Zhou G, Xie P, Li Y, Wang JM, Cao Y (2005) The expression of functional chemokine receptor CXCR4 is associated with the metastatic potential of human nasopharyngeal carcinoma. Clin Cancer Res 11:4658–4665 Huang S, Singh RK, Xie K, Gutman M, Berry KK, Bucana CD, Fidler IJ, Bar-Eli M (1994) Expression of the JE/MCP-1 gene suppresses metastatic potential in murine colon carcinoma cells. Cancer Immunol Immunother 39:231–238 Huang S, Xie K, Singh RK, Gutman M, Bar-Eli M (1995) Suppression of tumor growth and metastasis of murine renal adenocarcinoma by syngeneic fibroblasts genetically engineered to secrete the JE/MCP-1 cytokine. J Interferon Cytokine Res 15:655–665 Hwang J, Son KN, Kim CW, Ko J, Na DS, Kwon BS, Gho YS, Kim J (2005) Human CC chemokine CCL23, a ligand for CCR1, induces endothelial cell migration and promotes angiogenesis. Cytokine 30:254–263 Imai T, Yoshida T, Baba M, Nishimura M, Kakizaki M, Yoshie O (1996) Molecular cloning of a novel T cell-directed CC chemokine expressed in thymus by signal sequence trap using Epstein-Barr virus vector. J Biol Chem 271:21514–21521 Imai T, Baba M, Nishimura M, Kakizaki M, Takagi S, Yoshie O (1997) The T cell-directed CC chemokine TARC is a highly specific biological ligand for CC chemokine receptor 4. J Biol Chem 272:15036–15042 Imai T, Chantry D, Raport CJ, Wood CL, Nishimura M, Godiska R, Yoshie O, Gray PW (1998) Macrophage-derived chemokine is a functional ligand for the CC chemokine receptor 4. J Biol Chem 273:1764–1768 Inoue K, Slaton JW, Eve BY, Kim SJ, Perrotte P, Balbay MD, Yano S, Bar-Eli M, Radinsky R, Pettaway CA, Dinney CP (2000) Interleukin 8 expression regulates tumorigenicity and metastases in androgen-independent prostate cancer. Clin Cancer Res 6:2104–2119 Ivarsson K, Ekerydh A, Fyhr IM, Janson PO, Brannstrom M (2000) Upregulation of interleukin-8 and polarized epithelial expression of interleukin-8 receptor A in ovarian carcinomas. Acta Obstet Gynecol Scand 79:777–784 Johrer K, Zelle-Rieser C, Perathoner A, Moser P, Hager M, Ramoner R, Gander H, Holtl L, Bartsch G, Greil R, Thurnher M (2005) Up-regulation of functional chemokine receptor CCR3 in human renal cell carcinoma. Clin Cancer Res 11:2459–2465 Jose PJ, Griffiths-Johnson DA, Collins PD, Walsh DT, Moqbel R, Totty NF, Truong O, Hsuan JJ, Williams TJ (1994) Eotaxin: a potent eosinophil chemoattractant cytokine detected in a guinea pig model of allergic airways inflammation. J Exp Med 179:881–887 Kaifi JT, Yekebas EF, Schurr P, Obonyo D, Wachowiak R, Busch P, Heinecke A, Pantel K, Izbicki JR (2005) Tumor-cell homing to lymph nodes and bone marrow and CXCR4 expression in esophageal cancer. J Natl Cancer Inst 97:1840–1847 Kameyoshi Y, Dorschner A, Mallet AI, Christophers E, Schroder JM (1992) Cytokine RANTES released by thrombin-stimulated platelets is a potent attractant for human eosinophils. J Exp Med 176:587–592 Kanda S, Mochizuki Y, Kanetake H (2003) Stromal cell-derived factor-1alpha induces tube-like structure formation of endothelial cells through phosphoinositide 3-kinase. J Biol Chem 278:257–262
27 Chemokines and Metastasis
623
Kawada K, Sonoshita M, Sakashita H, Takabayashi A, Yamaoka Y, Manabe T, Inaba K, Minato N, Oshima M, Taketo MM (2004) Pivotal role of CXCR3 in melanoma cell metastasis to lymph nodes. Cancer Res 64:4010–4017 Kawada K, Hosogi H, Sonoshita M, Sakashita H, Manabe T, Shimahara Y, Sakai Y, Takabayashi A, Oshima M, Taketo MM (2007) Chemokine receptor CXCR3 promotes colon cancer metastasis to lymph nodes. Oncogene 26:4679–4688 Keane MP, Arenberg DA, Lynch JP III, Whyte RI, Iannettoni MD, Burdick MD, Wilke CA, Morris SB, Glass MC, DiGiovine B, Kunkel SL, Strieter RM (1997) The CXC chemokines, IL-8 and IP-10, regulate angiogenic activity in idiopathic pulmonary fibrosis. J Immunol 159:1437–1443 Keane MP, Belperio JA, Xue YY, Burdick MD, Strieter RM (2004) Depletion of CXCR2 inhibits tumor growth and angiogenesis in a murine model of lung cancer. J Immunol 172:2853–2860 Kelner GS, Kennedy J, Bacon KB, Kleyensteuber S, Largaespada DA, Jenkins NA, Copeland NG, Bazan JF, Moore KW, Schall TJ (1994) Lymphotactin: a cytokine that represents a new class of chemokine. Science 266:1395–1399 Kim J, Takeuchi H, Lam ST, Turner RR, Wang HJ, Kuo C, Foshag L, Bilchik AJ, Hoon DS (2005) Chemokine receptor CXCR4 expression in colorectal cancer patients increases the risk for recurrence and for poor survival. J Clin Oncol 23:2744–2753 Koch AE, Polverini PJ, Kunkel SL, Harlow LA, DiPietro LA, Elner VM, Elner SG, Strieter RM (1992) Interleukin-8 as a macrophage-derived mediator of angiogenesis. Science 258:1798–1801 Kondo T, Ito F, Nakazawa H, Horita S, Osaka Y, Toma H (2004) High expression of chemokine gene as a favorable prognostic factor in renal cell carcinoma. J Urol 171:2171–2175 Kuroda T, Kitadai Y, Tanaka S, Yang X, Mukaida N, Yoshihara M, Chayama K (2005) Monocyte chemoattractant protein-1 transfection induces angiogenesis and tumorigenesis of gastric carcinoma in nude mice via macrophage recruitment. Clin Cancer Res 11:7629–7636 Laverdiere C, Hoang BH, Yang R, Sowers R, Qin J, Meyers PA, Huvos AG, Healey JH, Gorlick R (2005) Messenger RNA expression levels of CXCR4 correlate with metastatic behavior and outcome in patients with osteosarcoma. Clin Cancer Res 11:2561–2567 Lee J, Horuk R, Rice GC, Bennett GL, Camerato T, Wood WI (1992) Characterization of two high affinity human interleukin-8 receptors. J Biol Chem 267:16283–16287 Lee SJ, Namkoong S, Kim YM, Kim CK, Lee H, Ha KS, Chung HT, Kwon YG, Kim YM (2006) Fractalkine stimulates angiogenesis by activating the Raf-1/MEK/ERK- and PI3K/Akt/eNOSdependent signal pathways. Am J Physiol Heart Circ Physiol 291:H2836–H2846 Legler DF, Loetscher M, Roos RS, Clark-Lewis I, Baggiolini M, Moser B (1998) B cell-attracting chemokine 1, a human CXC chemokine expressed in lymphoid tissues, selectively attracts B lymphocytes via BLR1/CXCR5. J Exp Med 187:655–660 Lehrer S, Diamond EJ, Mamkine B, Stone NN, Stock RG (2004) Serum interleukin-8 is elevated in men with prostate cancer and bone metastases. Technol Cancer Res Treat 3:411 Leibovich SJ, Wiseman DM (1988) Macrophages, wound repair and angiogenesis. Prog Clin Biol Res 266:131–145 Letsch A, Keilholz U, Schadendorf D, Assfalg G, Asemissen AM, Thiel E, Scheibenbogen C (2004) Functional CCR9 expression is associated with small intestinal metastasis. J Invest Dermatol 122:685–690 Leung SY, Wong MP, Chung LP, Chan AS, Yuen ST (1997) Monocyte chemoattractant protein-1 expression and macrophage infiltration in gliomas. Acta Neuropathol (Berl) 93:518–527 Lev DC, Ruiz M, Mills L, McGary EC, Price JE, Bar-Eli M (2003) Dacarbazine causes transcriptional up-regulation of interleukin 8 and vascular endothelial growth factor in melanoma cells: a possible escape mechanism from chemotherapy. Mol Cancer Ther 2:753–763 Li A, Dubey S, Varney ML, Dave BJ, Singh RK (2003) IL-8 directly enhanced endothelial cell survival, proliferation, and matrix metalloproteinases production and regulated angiogenesis. J Immunol 170:3369–3376
624
K.C. Nannuru et al.
Li A, Varney ML, Valasek J, Godfrey M, Dave BJ, Singh RK (2005) Autocrine role of interleukin-8 in induction of endothelial cell proliferation, survival, migration and MMP-2 production and angiogenesis. Angiogenesis 8:63–71 Littman DR (1998) Chemokine receptors: keys to AIDS pathogenesis? Cell 93:677–680 Locati M, Murphy PM (1999) Chemokines and chemokine receptors: biology and clinical relevance in inflammation and AIDS. Annu Rev Med 50:425–440 Locati M, Otero K, Schioppa T, Signorelli P, Perrier P, Baviera S, Sozzani S, Mantovani A (2002) The chemokine system: tuning and shaping by regulation of receptor expression and coupling in polarized responses. Allergy 57:972–982 Loetscher P, Clark-Lewis I (2001) Agonistic and antagonistic activities of chemokines. J Leukoc Biol 69:881–884 Loetscher M, Loetscher P, Brass N, Meese E, Moser B (1998a) Lymphocyte-specific chemokine receptor CXCR3: regulation, chemokine binding and gene localization. Eur J Immunol 28:3696–3705 Loetscher P, Uguccioni M, Bordoli L, Baggiolini M, Moser B, Chizzolini C, Dayer JM (1998b) CCR5 is characteristic of Th1 lymphocytes. Nature 391:344–345 Loetscher P, Moser B, Baggiolini M (2000) Chemokines and their receptors in lymphocyte traffic and HIV infection. Adv Immunol 74:127–180 Longo-Imedio MI, Longo N, Trevino I, Lazaro P, Sanchez-Mateos P (2005) Clinical significance of CXCR3 and CXCR4 expression in primary melanoma. Int J Cancer 117:861–865 Lu Y, Cai Z, Galson DL, Xiao G, Liu Y, George DE, Melhem MF, Yao Z, Zhang J (2006) Monocyte chemotactic protein-1 (MCP-1) acts as a paracrine and autocrine factor for prostate cancer growth and invasion. Prostate 66:1311–1318 Lu Y, Cai Z, Xiao G, Keller ET, Mizokami A, Yao Z, Roodman GD, Zhang J (2007a) Monocyte chemotactic protein-1 mediates prostate cancer-induced bone resorption. Cancer Res 67:3646–3653 Lu Y, Cai Z, Xiao G, Liu Y, Keller ET, Yao Z, Zhang J (2007b) CCR2 expression correlates with prostate cancer progression. J Cell Biochem 101:676–685 Luca M, Huang S, Gershenwald JE, Singh RK, Reich R, Bar-Eli M (1997) Expression of interleukin-8 by human melanoma cells up-regulates MMP-2 activity and increases tumor growth and metastasis. Am J Pathol 151:1105–1113 Luster AD (1998) Chemokines–chemotactic cytokines that mediate inflammation. N Engl J Med 338:436–445 Maione TE, Gray GS, Petro J, Hunt AJ, Donner AL, Bauer SI, Carson HF, Sharpe RJ (1990) Inhibition of angiogenesis by recombinant human platelet factor-4 and related peptides. Science 247:77–79 Mantovani A (1999) The chemokine system: redundancy for robust outputs. Immunol Today 20:254–257 Mantovani A, Allavena P, Sica A, Balkwill F (2008) Cancer-related inflammation. Nature 454:436–444 Mashino K, Sadanaga N, Yamaguchi H, Tanaka F, Ohta M, Shibuta K, Inoue H, Mori M (2002) Expression of chemokine receptor CCR7 is associated with lymph node metastasis of gastric carcinoma. Cancer Res 62:2937–2941 Matloubian M, David A, Engel S, Ryan JE, Cyster JG (2000) A transmembrane CXC chemokine is a ligand for HIV-coreceptor Bonzo. Nat Immunol 1:298–304 Maxwell PJ, Gallagher R, Seaton A, Wilson C, Scullin P, Pettigrew J, Stratford IJ, Williams KJ, Johnston PG, Waugh DJJ (2007) HIF-1 and NF-[kappa]B-mediated upregulation of CXCR1 and CXCR2 expression promotes cell survival in hypoxic prostate cancer cells. Oncogene 26:7333–7345 McCawley LJ, Matrisian LM (2000) Matrix metalloproteinases: multifunctional contributors to tumor progression. Mol Med Today 6:149–156 Meijer J, Zeelenberg IS, Sipos B, Roos E (2006) The CXCR5 chemokine receptor is expressed by carcinoma cells and promotes growth of colon carcinoma in the liver. Cancer Res 66:9576–9582
27 Chemokines and Metastasis
625
Melnikova VO, Bar-Eli M (2006) Bioimmunotherapy for melanoma using fully human antibodies targeting MCAM/MUC18 and IL-8. Pigment Cell Res 19:395–405 Mestas J, Burdick MD, Reckamp K, Pantuck A, Figlin RA, Strieter RM (2005) The role of CXCR2/CXCR2 ligand biological axis in renal cell carcinoma. J Immunol 175: 5351–5357 Milliken D, Scotton C, Raju S, Balkwill F, Wilson J (2002) Analysis of chemokines and chemokine receptor expression in ovarian cancer ascites. Clin Cancer Res 8:1108–1114 Mirshahi F, Pourtau J, Li H, Muraine M, Trochon V, Legrand E, Vannier J, Soria J, Vasse M, Soria C (2000) SDF-1 activity on microvascular endothelial cells: consequences on angiogenesis in in vitro and in vivo models. Thromb Res 99:587–594 Moore UM, Kaplow JM, Pleass RD, Castro SW, Naik K, Lynch CN, Daly S, Roach AG, Jaye M, Williams RJ (1997) Monocyte chemoattractant protein-2 is a potent agonist of CCR2B. J Leukoc Biol 62:911–915 Muller A, Homey B, Soto H, Ge N, Catron D, Buchanan ME, McClanahan T, Murphy E, Yuan W, Wagner SN, Barrera JL, Mohar A, Verastegui E, Zlotnik A (2001) Involvement of chemokine receptors in breast cancer metastasis. Nature 410:50–56 Murdoch C, Monk PN, Finn A (1999) Cxc chemokine receptor expression on human endothelial cells. Cytokine 11:704–712 Murphy PM (2001) Chemokines and the molecular basis of cancer metastasis. N Engl J Med 345:833–835 Murphy PM (2002) International union of pharmacology. XXX. Update on chemokine receptor nomenclature. Pharmacol Rev 54:227–229 Murphy PM, Baggiolini M, Charo IF, Hebert CA, Horuk R, Matsushima K, Miller LH, Oppenheim JJ, Power CA (2000) International union of pharmacology. XXII. Nomenclature for chemokine receptors. Pharmacol Rev 52:145–176 Nagasawa T, Nakajima T, Tachibana K, Iizasa H, Bleul CC, Yoshie O, Matsushima K, Yoshida N, Springer TA, Kishimoto T (1996) Molecular cloning and characterization of a murine pre-B-cell growth-stimulating factor/stromal cell-derived factor 1 receptor, a murine homolog of the human immunodeficiency virus 1 entry coreceptor fusin. Proc Natl Acad Sci U S A 93: 14726–14729 Nagpal ML, Chen Y, Lin T (2004) Effects of overexpression of CXCL10 (cytokine-responsive gene-2) on MA-10 mouse Leydig tumor cell steroidogenesis and proliferation. J Endocrinol 183:585–594 Nagpal ML, Davis J, Lin T (2006) Overexpression of CXCL10 in human prostate LNCaP cells activates its receptor (CXCR3) expression and inhibits cell proliferation. Biochim Biophys Acta 1762:811–818 Nakamura ES, Koizumi K, Kobayashi M, Saitoh Y, Arita Y, Nakayama T, Sakurai H, Yoshie O, Saiki I (2006) RANKL-induced CCL22/macrophage-derived chemokine produced from osteoclasts potentially promotes the bone metastasis of lung cancer expressing its receptor CCR4. Clin Exp Metastasis 23:9–18 Nakashima E, Mukaida N, Kubota Y, Kuno K, Yasumoto K, Ichimura F, Nakanishi I, Miyasaka M, Matsushima K (1995) Human MCAF gene transfer enhances the metastatic capacity of a mouse cachectic adenocarcinoma cell line in vivo. Pharm Res 12:1598–1604 Nanki T, Urasaki Y, Imai T, Nishimura M, Muramoto K, Kubota T, Miyasaka N (2004) Inhibition of fractalkine ameliorates murine collagen-induced arthritis. J Immunol 173:7010–7016 Negus RP, Stamp GW, Relf MG, Burke F, Malik ST, Bernasconi S, Allavena P, Sozzani S, Mantovani A, Balkwill FR (1995) The detection and localization of monocyte chemoattractant protein-1 (MCP-1) in human ovarian cancer. J Clin Invest 95:2391–2396 Negus RP, Turner L, Burke F, Balkwill FR (1998) Hypoxia down-regulates MCP-1 expression: implications for macrophage distribution in tumors. J Leukoc Biol 63:758–765 Neote K, Darbonne W, Ogez J, Horuk R, Schall TJ (1993) Identification of a promiscuous inflammatory peptide receptor on the surface of red blood cells. J Biol Chem 268:12247–12249 Neuhaus T, Stier S, Totzke G, Gruenewald E, Fronhoffs S, Sachinidis A, Vetter H, Ko YD (2003) Stromal cell-derived factor 1alpha (SDF-1alpha) induces gene-expression of early growth
626
K.C. Nannuru et al.
response-1 (Egr-1) and VEGF in human arterial endothelial cells and enhances VEGF induced cell proliferation. Cell Prolif 36:75–86 Nicolson GL (1991) Tumor and host molecules important in the organ preference of metastasis. Semin Cancer Biol 2:143–154 Norrby K (1996) Interleukin-8 and de novo mammalian angiogenesis. Cell Prolif 29:315–323 Numasaki M, Watanabe M, Suzuki T, Takahashi H, Nakamura A, McAllister F, Hishinuma T, Goto J, Lotze MT, Kolls JK, Sasaki H (2005) IL-17 enhances the net angiogenic activity and in vivo growth of human non-small cell lung cancer in SCID mice through promoting CXCR2-dependent angiogenesis. J Immunol 175:6177–6189 Nurnberg W, Tobias D, Otto F, Henz BM, Schadendorf D (1999) Expression of interleukin-8 detected by in situ hybridization correlates with worse prognosis in primary cutaneous melanoma. J Pathol 189:546–551 Oberlin E, Amara A, Bachelerie F, Bessia C, Virelizier JL, Renzana-Seisdedos F, Schwartz O, Heard JM, Clark-Lewis I, Legler DF, Loetscher M, Baggiolini M, Moser B (1996) The CXC chemokine SDF-1 is the ligand for LESTR/fusin and prevents infection by T-cell-line-adapted HIV-1. Nature 382:833–835 Orimo A, Gupta PB, Sgroi DC, Renzana-Seisdedos F, Delaunay T, Naeem R, Carey VJ, Richardson AL, Weinberg RA (2005) Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell 121:335–348 Pan J, Mestas J, Burdick MD, Phillips RJ, Thomas GV, Reckamp K, Belperio JA, Strieter RM (2006) Stromal derived factor-1 (SDF-1/CXCL12) and CXCR4 in renal cell carcinoma metastasis. Mol Cancer 5:56 Pan Y, Lloyd C, Zhou H, Dolich S, Deeds J, Gonzalo JA, Vath J, Gosselin M, Ma J, Dussault B, Woolf E, Alperin G, Culpepper J, Gutierrez-Ramos JC, Gearing D (1997) Neurotactin, a membrane-anchored chemokine upregulated in brain inflammation. Nature 387:611–617 Pellegrino A, Antonaci F, Russo F, Merchionne F, Ribatti D, Vacca A, Dammacco F (2004) CXCR3-binding chemokines in multiple myeloma. Cancer Lett 207:221–227 Perussia B (1992) Tumor infiltrating cells. Lab Invest 67:155–157 Phillips RJ, Burdick MD, Lutz M, Belperio JA, Keane MP, Strieter RM (2003) The stromal derived factor-1/CXCL12-CXC chemokine receptor 4 biological axis in non-small cell lung cancer metastases. Am J Respir Crit Care Med 167:1676–1686 Pollard JW (2009) Trophic macrophages in development and disease. Nat Rev Immunol 9:259–270 Ponath PD, Qin S, Ringler DJ, Clark-Lewis I, Wang J, Kassam N, Smith H, Shi X, Gonzalo JA, Newman W, Gutierrez-Ramos JC, Mackay CR (1996) Cloning of the human eosinophil chemoattractant, eotaxin. Expression, receptor binding, and functional properties suggest a mechanism for the selective recruitment of eosinophils. J Clin Invest 97:604–612 Ransohoff RM (2009) Chemokines and chemokine receptors: standing at the crossroads of immunobiology and neurobiology. Immunity 31:711–721 Romagnani P, Annunziato F, Lasagni L, Lazzeri E, Beltrame C, Francalanci M, Uguccioni M, Galli G, Cosmi L, Maurenzig L, Baggiolini M, Maggi E, Romagnani S, Serio M (2001) Cell cycle-dependent expression of CXC chemokine receptor 3 by endothelial cells mediates angiostatic activity. J Clin Invest 107:53–63 Rot A, Krieger M, Brunner T, Bischoff SC, Schall TJ, Dahinden CA (1992) RANTES and macrophage inflammatory protein 1 alpha induce the migration and activation of normal human eosinophil granulocytes. J Exp Med 176:1489–1495 Ruehlmann JM, Xiang R, Niethammer AG, Ba Y, Pertl U, Dolman CS, Gillies SD, Reisfeld RA (2001) MIG (CXCL9) chemokine gene therapy combines with antibody-cytokine fusion protein to suppress growth and dissemination of murine colon carcinoma. Cancer Res 61:8498–8503 Ruffini PA, Morandi P, Cabioglu N, Altundag K, Cristofanilli M (2007) Manipulating the chemokine–chemokine receptor network to treat cancer. Cancer 109:2392–2404 Salcedo R, Oppenheim JJ (2003) Role of chemokines in angiogenesis: CXCL12/SDF-1 and CXCR4 interaction, a key regulator of endothelial cell responses. Microcirculation 10:359–370
27 Chemokines and Metastasis
627
Salcedo R, Wasserman K, Young HA, Grimm MC, Howard OM, Anver MR, Kleinman HK, Murphy WJ, Oppenheim JJ (1999) Vascular endothelial growth factor and basic fibroblast growth factor induce expression of CXCR4 on human endothelial cells: in vivo neovascularization induced by stromal-derived factor-1alpha. Am J Pathol 154:1125–1135 Salcedo R, Ponce ML, Young HA, Wasserman K, Ward JM, Kleinman HK, Oppenheim JJ, Murphy WJ (2000a) Human endothelial cells express CCR2 and respond to MCP-1: direct role of MCP-1 in angiogenesis and tumor progression. Blood 96:34–40 Salcedo R, Resau JH, Halverson D, Hudson EA, Dambach M, Powell D, Wasserman K, Oppenheim JJ (2000b) Differential expression and responsiveness of chemokine receptors (CXCR1-3) by human microvascular endothelial cells and umbilical vein endothelial cells. FASEB J 14:2055–2064 Sancho M, Vieira JM, Casalou C, Mesquita M, Pereira T, Cavaco BM, Dias S, Leite V (2006) Expression and function of the chemokine receptor CCR7 in thyroid carcinomas. J Endocrinol 191:229–238 Saur D, Seidler B, Schneider G, Algul H, Beck R, Senekowitsch-Schmidtke R, Schwaiger M, Schmid RM (2005) CXCR4 expression increases liver and lung metastasis in a mouse model of pancreatic cancer. Gastroenterology 129:1237–1250 Scala S, Ottaiano A, Ascierto PA, Cavalli M, Simeone E, Giuliano P, Napolitano M, Franco R, Botti G, Castello G (2005) Expression of CXCR4 predicts poor prognosis in patients with malignant melanoma. Clin Cancer Res 11:1835–1841 Scapini P, Morini M, Tecchio C, Minghelli S, Di CE, Tanghetti E, Albini A, Lowell C, Berton G, Noonan DM, Cassatella MA (2004) CXCL1/macrophage inflammatory protein-2-induced angiogenesis in vivo is mediated by neutrophil-derived vascular endothelial growth factor-A. J Immunol 172:5034–5040 Scheibenbogen C, Mohler T, Haefele J, Hunstein W, Keilholz U (1995) Serum interleukin-8 (IL-8) is elevated in patients with metastatic melanoma and correlates with tumour load. Melanoma Res 5:179–181 Schimanski CC, Schwald S, Simiantonaki N, Jayasinghe C, Gonner U, Wilsberg V, Junginger T, Berger MR, Galle PR, Moehler M (2005) Effect of chemokine receptors CXCR4 and CCR7 on the metastatic behavior of human colorectal cancer. Clin Cancer Res 11:1743–1750 Scotton CJ, Wilson JL, Scott K, Stamp G, Wilbanks GD, Fricker S, Bridger G, Balkwill FR (2002) Multiple actions of the chemokine CXCL12 on epithelial tumor cells in human ovarian cancer. Cancer Res 62:5930–5938 Segerer S, Hughes E, Hudkins KL, Mack M, Goodpaster T, Alpers CE (2002) Expression of the fractalkine receptor (CX3CR1) in human kidney diseases. Kidney Int 62:488–495 Sgadari C, Angiolillo AL, Cherney BW, Pike SE, Farber JM, Koniaris LG, Vanguri P, Burd PR, Sheikh N, Gupta G, Teruya-Feldstein J, Tosato G (1996) Interferon-inducible protein-10 identified as a mediator of tumor necrosis in vivo. Proc Natl Acad Sci U S A 93:13791–13796 Sgadari C, Farber JM, Angiolillo AL, Liao F, Teruya-Feldstein J, Burd PR, Yao L, Gupta G, Kanegane C, Tosato G (1997) Mig, the monokine induced by interferon-gamma, promotes tumor necrosis in vivo. Blood 89:2635–2643 Shono T, Ono M, Izumi H, Jimi SI, Matsushima K, Okamoto T, Kohno K, Kuwano M (1996) Involvement of the transcription factor NF-kappaB in tubular morphogenesis of human microvascular endothelial cells by oxidative stress. Mol Cell Biol 16:4231–4239 Shulby SA, Dolloff NG, Stearns ME, Meucci O, Fatatis A (2004) CX3CR1-fractalkine expression regulates cellular mechanisms involved in adhesion, migration, and survival of human prostate cancer cells. Cancer Res 64:4693–4698 Sica A, Saccani A, Bottazzi B, Bernasconi S, Allavena P, Gaetano B, Fei F, LaRosa G, Scotton C, Balkwill F, Mantovani A (2000) Defective expression of the monocyte chemotactic protein-1 receptor CCR2 in macrophages associated with human ovarian carcinoma. J Immunol 164:733–738 Sica A, Saccani A, Mantovani A (2002) Tumor-associated macrophages: a molecular perspective. Int Immunopharmacol 2:1045–1054
628
K.C. Nannuru et al.
Simonetti O, Goteri G, Lucarini G, Filosa A, Pieramici T, Rubini C, Biagini G, Offidani A (2006) Potential role of CCL27 and CCR10 expression in melanoma progression and immune escape. Eur J Cancer 42:1181–1187 Singh RK, Fidler IJ (1996) Regulation of tumor angiogenesis by organ-specific cytokines. In:Gunthert U, Birchmeier W (eds) Attempts to understand metastasis formation II. SpringerVerlag, New York, pp 1–11 Singh RK, Gutman M, Radinsky R, Bucana CD, Fidler IJ (1994) Expression of interleukin 8 correlates with the metastatic potential of human melanoma cells in nude mice. Cancer Res 54:3242–3247 Singh S, Singh UP, Stiles JK, Grizzle WE, Lillard JW Jr (2004) Expression and functional role of CCR9 in prostate cancer cell migration and invasion. Clin Cancer Res 10:8743–8750 Singh S, Sadanandam A, Singh RK (2007) Chemokines in tumor angiogenesis and metastasis. Cancer Metastasis Rev 26:453–467 Singh S, Nannuru KC, Sadanandam A, Varney ML, Singh RK (2009a) CXCR1 and CXCR2 enhances human melanoma tumourigenesis, growth and invasion. Br J Cancer 100:1638–1646 Singh S, Sadanandam A, Nannuru KC, Varney ML, Mayer-Ezell R, Bond R, Singh RK (2009b) Small-molecule antagonists for CXCR2 and CXCR1 inhibit human melanoma growth by decreasing tumor cell proliferation, survival, and angiogenesis. Clin Cancer Res 15:2380–2386 Smith MC, Luker KE, Garbow JR, Prior JL, Jackson E, Piwnica-Worms D, Luker GD (2004) CXCR4 regulates growth of both primary and metastatic breast cancer. Cancer Res 64:8604–8612 Soejima K, Rollins BJ (2001) A functional IFN-gamma-inducible protein-10/CXCL10-specific receptor expressed by epithelial and endothelial cells that is neither CXCR3 nor glycosaminoglycan. J Immunol 167:6576–6582 Stellato C, Collins P, Ponath PD, Soler D, Newman W, La RG, Li H, White J, Schwiebert LM, Bickel C, Liu M, Bochner BS, Williams T, Schleimer RP (1997) Production of the novel C-C chemokine MCP-4 by airway cells and comparison of its biological activity to other C-C chemokines. J Clin Invest 99:926–936 Strieter RM, Kunkel SL, Elner VM, Martonyi CL, Koch AE, Polverini PJ, Elner SG (1992) Interleukin-8. A corneal factor that induces neovascularization. Am J Pathol 141:1279–1284 Strieter RM, Kunkel SL, Arenberg DA, Burdick MD, Polverini PJ (1995a) Interferon gammainducible protein 10 (IP-10), a member of the C-X-C chemokine family, is an inhibitor of angiogenesis. Biochem Biophys Res Commun 210:51–57 Strieter RM, Polverini PJ, Kunkel SL, Arenberg DA, Burdick MD, Kasper J, Dzuiba J, Van DJ, Walz A, Marriott D (1995b) The functional role of the ELR motif in CXC chemokinemediated angiogenesis. J Biol Chem 270:27348–27357 Strieter RM, Belperio JA, Phillips RJ, Keane MP (2004) CXC chemokines in angiogenesis of cancer. Semin Cancer Biol 14:195–200 Strieter RM, Burdick MD, Mestas J, Gomperts B, Keane MP, Belperio JA (2006) Cancer CXC chemokine networks and tumour angiogenesis. Eur J Cancer 42:768–778 Su L, Zhang J, Xu H, Wang Y, Chu Y, Liu R, Xiong S (2005) Differential expression of CXCR4 is associated with the metastatic potential of human non-small cell lung cancer cells. Clin Cancer Res 11:8273–8280 Suyama T, Furuya M, Nishiyama M, Kasuya Y, Kimura S, Ichikawa T, Ueda T, Nikaido T, Ito H, Ishikura H (2005) Up-regulation of the interferon gamma (IFN-gamma)-inducible chemokines IFN-inducible T-cell alpha chemoattractant and monokine induced by IFN-gamma and of their receptor CXC receptor 3 in human renal cell carcinoma. Cancer 103:258–267 Tachibana K, Hirota S, Iizasa H, Yoshida H, Kawabata K, Kataoka Y, Kitamura Y, Matsushima K, Yoshida N, Nishikawa S, Kishimoto T, Nagasawa T (1998) The chemokine receptor CXCR4 is essential for vascularization of the gastrointestinal tract. Nature 393:591–594 Takanami I (2003) Overexpression of CCR7 mRNA in nonsmall cell lung cancer: correlation with lymph node metastasis. Int J Cancer 105:186–189
27 Chemokines and Metastasis
629
Takeuchi H, Fujimoto A, Tanaka M, Yamano T, Hsueh E, Hoon DS (2004) CCL21 chemokine regulates chemokine receptor CCR7 bearing malignant melanoma cells. Clin Cancer Res 10:2351–2358 Talmadge JE, Key M, Fidler IJ (1981) Macrophage content of metastatic and nonmetastatic rodent neoplasms. J Immunol 126:2245–2248 Thatcher TH, McHugh NA, Egan RW, Chapman RW, Hey JA, Turner CK, Redonnet MR, Seweryniak KE, Sime PJ, Phipps RP (2005) Role of CXCR2 in cigarette smoke-induced lung inflammation. Am J Physiol Lung Cell Mol Physiol 289:L322–L328 Thelen M (2001) Dancing to the tune of chemokines. Nat Immunol 2:129–134 Tiffany HL, Lautens LL, Gao JL, Pease J, Locati M, Combadiere C, Modi W, Bonner TI, Murphy PM (1997) Identification of CCR8: a human monocyte and thymus receptor for the CC chemokine I-309. J Exp Med 186:165–170 Tsou CL, Gladue RP, Carroll LA, Paradis T, Boyd JG, Nelson RT, Neote K, Charo IF (1998) Identification of C-C chemokine receptor 1 (CCR1) as the monocyte hemofiltrate C-C chemokine (HCC)-1 receptor. J Exp Med 188:603–608 Tsuzuki H, Takahashi N, Kojima A, Narita N, Sunaga H, Takabayashi T, Fujieda S (2006) Oral and oropharyngeal squamous cell carcinomas expressing CCR7 have poor prognoses. Auris Nasus Larynx 33:37–42 Uehara H, Troncoso P, Johnston D, Bucana CD, Dinney C, Dong Z, Fidler IJ, Pettaway, CA (2005) Expression of interleukin-8 gene in radical prostatectomy specimens is associated with advanced pathologic stage. Prostate 64:40–49 Uguccioni M, Mackay CR, Ochensberger B, Loetscher P, Rhis S, LaRosa GJ, Rao P, Ponath PD, Baggiolini M, Dahinden CA (1997) High expression of the chemokine receptor CCR3 in human blood basophils. Role in activation by eotaxin, MCP-4, and other chemokines. J Clin Invest 100:1137–1143 Ugurel S, Rappl G, Tilgen W, Reinhold U (2001) Increased serum concentration of angiogenic factors in malignant melanoma patients correlates with tumor progression and survival. J Clin Oncol 19:577–583 Umehara H, Imai T (2001) Role of fractalkine in leukocyte adhesion and migration and in vascular injury. Drug News Perspect 14:460–464 Umehara H, Goda S, Imai T, Nagano Y, Minami Y, Tanaka Y, Okazaki T, Bloom ET, Domae N (2001) Fractalkine, a CX3C-chemokine, functions predominantly as an adhesion molecule in monocytic cell line THP-1. Immunol Cell Biol 79:298–302 Uslu R, Sanli UA, Dikmen Y, Karabulut B, Ozsaran A, Sezgin C, Muezzinoglu GG, Omay SB, Goker E (2005) Predictive value of serum interleukin-8 levels in ovarian cancer patients treated with paclitaxel-containing regimens. Int J Gynecol Cancer 15:240–245 Vaday GG, Peehl DM, Kadam PA, Lawrence DM (2006) Expression of CCL5 (RANTES) and CCR5 in prostate cancer. Prostate 66:124–134 Vandercappellen J, Van Damme J, Struyf S (2008) The role of CXC chemokines and their receptors in cancer. Cancer Lett 267:226–244 Varney ML, Li A, Dave BJ, Bucana CD, Johansson SL, Singh RK (2003) Expression of CXCR1 and CXCR2 receptors in malignant melanoma with different metastatic potential and their role in interleukin-8 (CXCL-8)-mediated modulation of metastatic phenotype. Clin Exp Metastasis 20:723–731 Varney ML, Johansson SL, Singh RK (2005a) Tumour-associated macrophage infiltration, neovascularization and aggressiveness in malignant melanoma: role of monocyte chemotactic protein-1 and vascular endothelial growth factor-A. Melanoma Res 15:417–425 Varney ML, Olsen KJ, Mosley RL, Singh RK (2005b) Paracrine regulation of vascular endothelial growth factor – a expression during macrophage–melanoma cell interaction: role of monocyte chemotactic protein-1 and macrophage colony-stimulating factor. J Interferon Cytokine Res 25:674–683 Varney, ML, Johansson SL, Singh RK (2006) Distinct expression of CXCL8 and its receptors CXCR1 and CXCR2 and their association with vessel density and aggressiveness in malignant melanoma. Am J Clin Pathol 125:209–216
630
K.C. Nannuru et al.
Varney ML, Singh S, Backora M, Chen Z, Singh RK (2009) Lymphangiogenesis and anti-tumor immune responses. Curr Mol Med 9:694–701 Veltri RW, Miller MC, Zhao G, Ng A, Marley GM, Wright GL Jr, Vessella RL, Ralph D (1999) Interleukin-8 serum levels in patients with benign prostatic hyperplasia and prostate cancer. Urology 53:139–147 Villares GJ, Zigler M, Wang H, Melnikova VO, Wu H, Friedman R, Leslie MC, Vivas-Mejia PE, Lopez-Berestein G, Sood AK, Bar-Eli M (2008) Targeting melanoma growth and metastasis with systemic delivery of liposome-incorporated protease-activated receptor-1 small interfering RNA. Cancer Res 68:9078–9086 Volin MV, Woods JM, Amin MA, Connors MA, Harlow LA, Koch AE (2001) Fractalkine: a novel angiogenic chemokine in rheumatoid arthritis. Am J Pathol 159:1521–1530 Wang J, Xi L, Gooding W, Godfrey TE, Ferris RL (2005) Chemokine receptors 6 and 7 identify a metastatic expression pattern in squamous cell carcinoma of the head and neck. Adv Otorhinolaryngol 62:121–133 Wang W, Soto H, Oldham ER, Buchanan ME, Homey B, Catron D, Jenkins N, Copeland NG, Gilbert DJ, Nguyen N, Abrams J, Kershenovich D, Smith K, McClanahan T, Vicari AP, Zlotnik A (2000) Identification of a novel chemokine (CCL28), which binds CCR10 (GPR2). J Biol Chem 275:22313–22323 Wehler T, Wolfert F, Schimanski CC, Gockel I, Herr W, Biesterfeld S, Seifert JK, Adwan H, Berger MR, Junginger T, Galle PR, Moehler M (2006) Strong expression of chemokine receptor CXCR4 by pancreatic cancer correlates with advanced disease. Oncol Rep 16:1159–1164 Weigelt B, Wessels LF, Bosma AJ, Glas AM, Nuyten DS, He YD, Dai H, Peterse JL, van’t Veer LJ (2005) No common denominator for breast cancer lymph node metastasis. Br J Cancer 93:924–932 Whiteside TL, Miescher S, Hulimann J, Moretta L, Von Hiedner U (1986) Clonal analysis and in situcharacterization of lymphocytes infiltrating human breast carcinomas. Cancer Immunol Immunother 23:169–178 Wiley HE, Gonzalez EB, Maki W, Wu MT, Hwang ST (2001) Expression of CC chemokine receptor-7 and regional lymph node metastasis of B16 murine melanoma. J Natl Cancer Inst 93:1638–1643 Wuyts A, Proost P, Lenaerts JP, Ben-Baruch A, Van DJ, Wang JM (1998) Differential usage of the CXC chemokine receptors 1 and 2 by interleukin-8, granulocyte chemotactic protein-2 and epithelial-cell-derived neutrophil attractant-78. Eur J Biochem 255:67–73 Xie K (2001) Interleukin-8 and human cancer biology. Cytokine Growth Factor Rev 12:375–391 Yoneda J, Kuniyasu H, Crispens MA, Price JE, Bucana CD, Fidler IJ (1998) Expression of angiogenesis-related genes and progression of human ovarian carcinomas in nude mice. J Natl Cancer Inst 90:447–454 Yoshida R, Imai T, Hieshima K, Kusuda J, Baba M, Kitaura M, Nishimura M, Kakizaki M, Nomiyama H, Yoshie O (1997a) Molecular cloning of a novel human CC chemokine EBI1ligand chemokine that is a specific functional ligand for EBI1, CCR7. J Biol Chem 272:13803–13809 Yoshida S, Ono M, Shono T, Izumi H, Ishibashi T, Suzuki H, Kuwano M (1997b) Involvement of interleukin-8, vascular endothelial growth factor, and basic fibroblast growth factor in tumor necrosis factor alpha-dependent angiogenesis. Mol Cell Biol 17:4015–4023 Youn BS, Zhang SM, Lee EK, Park DH, Broxmeyer HE, Murphy PM, Locati M, Pease JE, Kim KK, Antol K, Kwon BS (1997) Molecular cloning of leukotactin-1: a novel human betachemokine, a chemoattractant for neutrophils, monocytes, and lymphocytes, and a potent agonist at CC chemokine receptors 1 and 3. J Immunol 159:5201–5205 Young MRI, Wright MA (1992) Myelopoiesis-associated immune suppressor cells in mice bearing metastatic Lewis lung carcinoma tumors: Gamma-interferon plus tumor necrosis factor-alpha
27 Chemokines and Metastasis
631
synergistically reduces immune suppressor and tumor growth-promoting activities of bone marrow cells and diminishes tumor recurrence and metastasis. Cancer Res 52:6335–6340 Youngs SJ, Ali SA, Taub DD, Rees RC (1997) Chemokines induce migrational responses in human breast carcinoma cell lines. Int J Cancer 71:257–266 Zaballos A, Gutierrez J, Varona R, Ardavin C, Marquez G (1999) Cutting edge: identification of the orphan chemokine receptor GPR-9-6 as CCR9, the receptor for the chemokine TECK. J Immunol 162:5671–5675 Zlotnik A (2006) Chemokines and cancer. Int J Cancer 119:2026–2029 Zlotnik A, Yoshie O (2000) Chemokines: a new classification system and their role in immunity. Immunity 12:121–127
Chapter 28
Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis Sonia B. Jakowlew
Abstract Transforming growth factor-beta (TGF-b) is a multifunctional regulatory polypeptide that is the prototypical member of a large family of cytokines that controls many aspects of cellular function, including cellular proliferation, differentiation, migration, apoptosis, adhesion, angiogenesis, immune surveillance, and survival. Contrary to the initial concept that these proteins may be down-regulated in cancer cells to promote their growth, a marked increase in the expression of TGF-b has often been found in human cancers in vivo, including cancer of the lung. Moreover, in lung cancer, increased expression of TGF-b correlates with more advanced stages of malignancy and metastasis and with decreased survival. Increased expression of TGF-b is usually accompanied by a loss in the growth inhibitory response to TGF-b. Indeed, some lung cancer cells in culture demonstrate a progressive loss of the growth inhibitory response to TGF-b that varies directly with the malignant stage of the original tumor. The study of the molecular events associated with the escape of cancer cells from growth regulation by TGF-b has provided insight into mechanisms underlying carcinogenesis. Specific defects in TGF-b receptors, TGF-b-related signal transduction/gene activation, and TGF-b-regulated cell cycle proteins, have been implicated in the oncogenesis of human lung cancer and metastasis. This review provides background information on TGF-b and updates the status of our knowledge of the role of TGF-b in lung cancer, carcinogenesis, and metastasis.
Introduction Metastasis is a leading cause of mortality for cancer patients, including lung cancer patients. Despite this, the molecular mechanisms underlying metastasis are still poorly understood. It was established more than 100 years ago that some types of S.B. Jakowlew (*) National Cancer Institute, Center for Cancer Training, Cancer Training Branch, Bethesda, MD 20892, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_28, © Springer Science+Business Media, LLC 2010
633
634
S.B. Jakowlew
tumors have a propensity to metastasize to certain organs (Paget 1889). For example, breast tumors tend to metastasize to lung, bones, brain, and liver. This non-random distribution of metastases seems to be independent of vascular anatomy, rate of blood flow, and the number of tumor cells delivered to each organ (Fidler 2003). Instead, metastases may be due to expression of particular genes within the tumor cells that drive their invasion of a particular organ (Kang et al. 2003; Minn et al. 2007). Cancer cells need to overcome multiple challenges in order to successfully seed a tissue site. First, these cells need to escape from the primary tumor and invade local tissue. Second, these cells need to enter the bloodstream or lymphatic system where they must survive. Third, these cells need to move from the blood stream into tissues by extravasation at sites that are far from the primary tumor site. Finally, these cells need to adapt to or modify the metastatic niche to establish a new tumor. Lung cancer is one of the most prevalent and lethal cancers. In 2009, it is estimated that there will be more than 384,600 new cases of lung cancer among both men and women in the United States and more than 159,300 deaths (Jemel et al. 2009), representing 28% of all cancer deaths. Lung cancer has become the leading cause of cancer mortality among both men and women in the United States and is also the leading cause of cancer death worldwide, with more than 1.1 million annual deaths (Jemal et al. 2005). There are four pathological types of lung cancer, including lung adenocarcinoma (about 35% of cases), lung squamous cell carcinoma (about 30% of cases), large cell lung carcinoma (about 15% of cases) and small cell lung carcinoma (SCLC about 20% of cases), with adenocarcinoma, squamous cell carcinoma, and large cell carcinoma, commonly grouped as nonsmall cell lung cancer (NSCLC). While most of the common neoplasms in breast, colon, and prostate are predominantly adenocarcinomas, only 40% of lung cancers are adenocarcinomas. Within lung adenocarcinoma, histology is heterogeneous and associated with tissue invasion and clinical outcomes. The World Health Organization has subclassified adenocarcinoma on the basis of predominant cell morphology and growth pattern (Brambilla et al. 2001), such as bronchioloalveolar carcinoma (BAC), adenocarcinoma with mixed subtypes, and homogenously invasive tumors with a variety of histologic patterns. BAC is a type of well-differentiated adenocarcinoma in the periphery of the lung that is composed of Clara cells or alveolar type II cells that has a tendency to spread chiefly within the confines of the lung by aerogenous and lymphatic routes (Liebow 1960). BAC cells are cuboidal to columnar, with or without mucin, and grow in a non-invasive fashion along alveolar walls. The histologic distinction between BAC and other adenocarcinoma subclassifications is tissue invasion, the first step of the metastasis process, in which epithelial cells lose cell–cell adhesion, gain motility, and invade adjacent stroma (Fidler 2003). Adenocarcinomas with mixed subtypes frequently contain regions of non-invasive tumor at the periphery of invasive tumor. Pure invasive adenocarcinomas are devoid of bronchioloalveolar morphology. The spectrum of intratumoral histologic heterogeneity in adenocarcinoma suggests invasiveness represents a continuum of disease, from non-invasive BAC to adenocarcinoma-mixed subtype with BAC components, to pure invasive adenocarcinoma. The molecular events essential
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
635
to this transition in the lung are largely unknown. Recent survival data show that only 10–12% of all patients are long-term survivors for NSCLC and 3% for SCLC (Skarkin 1994; Skarki 1993). The poor outcome of lung cancer compared with other common cancers (12–15% vs. 62–97% average 5-year survival) is partially attributable to the current limited ability to distinguish fundamental differences in tumor biological predisposition to metastasis that may be associated with histologic heterogeneity. Lung cancer metastasis is frequent and approximately 40% of patients have distant metastases at the time of diagnosis. Furthermore, among patients who present with localized resectable disease, approximately 30% will develop metastases and succumb to their disease within 5 years. Lung cancer metastasis to lymphatics and visceral organ beds via the systemic circulation is the result of several well-characterized, sequentially acquired properties of tumor cells. These steps include enzyme-mediated invasion of organ stroma, circulation in lymphatic or vascular channels, and extravasation and proliferation in distant organ beds (Fidler 2003). Transforming growth factor-beta (TGF-b) is one of several polypeptide growth factors that have been identified in the lung and regulates many aspects of cellular function, including proliferation, differentiation, adhesion, and migration (Heldin et al. 2009; Massagué 2008). TGF-b initiates these responses by interacting with cell surface receptors, three types of which have been identified and characterized and shown to be widely distributed in most cells that are responsive to TGF-b (Massagué et al. 1994). These receptors include TGF-b type I (RI), type II (RII), and type III (RIII). Molecular cloning and functional analyses have demonstrated that TGF-b RI and TGF-b RII are serine/threonine kinases, which through the formation of a heteroduplex with TGF-b are essential components for TGF-b signal transduction (Franzen et al. 1993; Lin et al. 1992; Laiho et al. 1990). The direct involvement of TGF-b RI and TGF-b RII in TGF-b signal transduction implies that loss of functional TGF-b RI and/or TGF-b RII expression could contribute to loss of the growth inhibitory activity of TGF-b. Close association between inactivation of TGF-bRII and escape from TGF-b-mediated growth inhibition has been demonstrated in several cancer cell lines (Geiser et al. 1992; Park et al. 1994; Sun et al. 1994). A dysfunctional TGF-b-mediated signal transduction pathway due to loss of TGF-bRII function has been connected to the development of several human malignancies, including breast, prostate, colorectal, colon, and hepatocellular cancer (Grady et al. 1998; Furta et al. 1999; Markowitz et al. 1995; Hahn et al. 1996; De Souza et al. 1992). On the other hand, the TGF-b signaling system has also been implicated to be a tumor suppressor pathway in several tumor systems (Chakravarthy et al. 1999; Guo et al. 1997; Matsushita et al. 1999). Thus, TGF-b is a multifunctional growth factor with a complicated dual role in tumorigenesis and metastasis (Leivonen and Kähäri 2007; Derynk et al. 2001; Wakefield and Roberts 2002; Roberts and Wakefield 2003). During the early stages of tumor formation, TGF-b acts as a tumor suppressor, inhibiting proliferation and inducing apoptosis of tumor cells. During later stages of tumorigenesis, however, many tumor cells become unresponsive to the growth inhibitory functions of TGF-b, and instead, respond to
636
S.B. Jakowlew
this growth factor by becoming more motile, more invasive, and more resistant to apoptosis. Furthermore, TGF-b is induced in response to hypoxia and inflammation and can have a protective effect on tumor cells. TGF-b has also been observed to drive epithelial-to-mesenchymal transition (EMT) in cancer cells, which increases their metastatic capability. The 100-year-old “seed and soil” theory explains key features of cancer metastasis, including early initiation, late appearance, and organ-specificity (Fidler 2003). The “seed” is the cancer cell; it undergoes genetic alterations, disturbing the cellular activities that maintain normal tissue organization, and in so doing, initiating the formation of invasive and metastatic tumors. The “soil” consists of tumorassociated host cells, including endothelial cells and pericytes forming blood and lymph vessels attracted to the cancer cells by vascular endothelial growth factor (VEGF), nerve cells, fibroblasts converted into myofibroblasts by cancer cellreleased TGF-b, inflammatory cells attracted by cancer chemokines, and osteoclasts activated by metastatic cancer cells in the bone marrow. All these host cells engage in continuous molecular crosstalk with the cancer cells, and influence invasion and metastasis. Tumor-associated host cells are themselves invasive and some of them arrive at the site of metastasis ahead of cancer cells. The high radiosensitivity of tumor-associated host cells leads to speculation that radiotherapy may affect invasion and metastasis. For example, analgesic effects on bone metastasis and prevention of lung metastasis to the brain by ionizing radiation might be possibly due to alterations of host cells. Thus, tumor-associated host cells might be considered when new strategies are developed for cancer radiotherapy, chemotherapy, and surgery (Mareel and Madani 2006).
Transforming Growth Factor-b Signaling Over 60 different TGF-b family members have been identified in nature, with nearly 30 of these proteins being found in the human. The proteins that compose the TGF-b superfamily include four TGF-b ligands, five activins, eight bone morphogenetic proteins (BMP), and 15 growth and differentiation factors (GDF). Three TGF-b isoforms have been identified in humans, including TGF-b1, TGF-b2, and TGF-b3; these ligands are homodimeric polypeptides with a molecular weight of 25 kDa. Crystallographic studies show that TGF-b2 is composed of two monomers, with each monomer consisting of two antiparallel pairs of b-strands that form a flat surface and a separate a-helix (Schlunegger and Grutter 1992). In addition, two intrachain disulfide bonds form a ring that is threaded by a third intrachain disulfide bond; this arrangement has become to be referred to as the “cysteine knot.” Active TGF-b functions through specific high affinity receptors. Five TGF-b superfamily type I receptors and seven type II receptors have been identified in mammals (Derynck et al. 2001; Massagué 2000). The type I receptor family includes activin-like kinases (ALKs) 1 through 7. The type II receptors include TGF-b RII, BMP RII, activin RIIA, and activin RIIB (Attisano and Wrana 2002;
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
637
Derynck and Zhang 2003; Glick 2004). The type I and type II receptors are structurally-related transmembrane glycoproteins that consist of an extracellular N-terminal ligand-binding domain with more than ten cysteine residues that regulate the dimeric structure, a transmembrane region, and a C-terminal serine/threonine kinase domain. Unlike the type II receptors, the type I receptors have a highly conserved region that is rich in glycine and serine residues, referred to as the GS domain, in the juxtamembrane domain next to the N-terminus of the kinase domain. The GS domain is a target for the type II receptor kinase, and after phosphorylation of specific serine and threonine residues, the type I receptor becomes activated (Shi and Massagué 2003; Heldin et al. 1997). Located downstream of the type II receptor, the type I receptor functions to determine the specificity of intracellular signals. The type I and II receptors exist as homodimers at the cell surface in the absence of ligands, but have an inherent heteromeric affinity for each other. Only select combinations of type I and II receptors act as ligand-binding signaling complexes. The molecular basis of the selectivity of the type I–type II receptor interactions remains unknown for the most part; but, it is thought that the structural complement at the interface may help define the selectivity of the receptor combinations. Most of the TGF-b ligands bind with high affinity to the type I receptor or to the type II receptor, while others bind efficiently only to heteromeric receptor combinations. The binding of the ligand to the extracellular domain of the type II receptors induces a conformational change resulting in the phosphorylation and activation of type I receptors. The signaling initiated by the kinase activity of TGF-b involves the phosphorylation of multiple mothers against decapentaplegic homolog 1 (Smad) family proteins; this activity in turn brings numerous changes in the transcriptional regulation of various response genes. The Smad family proteins include Smads 1, 2, 3, 4, 5, 7, and 8. The Smads are divided into three subclasses that are defined by their structure and function and consist of receptor-regulated Smads (R-Smads), commonmediator Smad (Co-Smad), and inhibitory Smads (I-Smads). The R-Smads, Smads 2 and 3, function downstream of the TGF-b/nodal/activin ligands, while Smads 1, 5, and 8 function downstream of the BMP and GDF subfamilies of ligands. Smads 1, 2, 3, 5, and 8 are direct substrates for the TGF-b type I receptor kinase. Co-Smad Smad 4 functions in Smad complex formation. The I-Smads, Smad6, and Smad7 interfere with TGF-b-induced Smad-dependent signal transduction (Park 2005). Activation of cell surface receptors by ligands leads to phosphorylation of the R-Smads. This phosphorylation then permits the R-Smads to form both homomeric and heteromeric complexes with Smad4 in the nucleus that then permits transcriptional regulation of target genes. Smad6 and Smad7 are natural inhibitors of TGFb-induced Smad-mediated signaling. The inhibitory Smads antagonize TGF-b signaling by binding to the TGF-b type I receptor and interfering with the activation of Smad2 and Smad3 by preventing their interaction with the TGF-b type I receptor and phosphorylation (Röijer et al. 1998; Hayashi et al. 1997; Nakao et al. 1997; Hanyu et al. 2001; Mochizuki et al. 2004). The I-Smads interfere with R-Smad recruitment and phosphorylation. Smad1 and Smad5 induce Smad6 phosphorylation, whereas Smad3 induces Smad7 expression. BMP signaling induces an inhibitory
638
S.B. Jakowlew
feedback loop through Smad6 expression, while TGF-b induces an inhibitory feedback loop through Smad7 expression, although BMPs and TGF-b can also induce Smad7 and Smad6 expression, respectively. Smad6 inhibits BMP and TGF-b signaling with similar potency, while Smad7 inhibits TGF-b signaling more efficiently than Smad6 (Miyazono et al. 2003; Canalis et al. 2003).
Transforming Growth Factor-b Isoforms Immunohistochemical staining for TGF-b1, TGF-b RI, and TGF-b RII using specific antibodies in normal human lung shows expression of all three proteins in the epithelium of bronchi and bronchioles, as well as in alveoli (Kang et al. 2000). Differential expression of the proteins and mRNAs for TGF-b1, TGF-b RI, and TGF-b RII can also be found in various well-characterized NSCLC cell lines and primary NSCLC specimens at variable levels (Kang et al. 2000; Jakowlew et al. 1995, 1997c). Reduced TGF-b RII was detected in poorly differentiated adenocarcinomas and squamous cell carcinomas and some moderately differentiated adenocarcinomas. In situ hybridization studies conducted with specific riboprobes for TGF-b1, TGF-b RI, and TGF-b RII showed corresponding localization of expression of the mRNAs in the specimens that showed positive immunostaining for the proteins. In patients diagnosed with NSCLC carcinoma, tumors expressing high levels of TGF-b1 and TGF-b receptors were shown to have higher proliferation and metastasis capability (Woszczyk et al. 2004). Tumors with lower levels of TGF expression were less metastatic. NSCLC patients with low-grade or high-grade lymphoma showed expression of TGF-b1 and TGF-b RI, TGF-b RII, and TGF-b RIII that was twice as high in patients with a diagnosis of high-grade lymphomas as in the group of patients diagnosed with low-grade lymphomas by quantitative reverse transcription polymerase chain reaction (QRT-PCR) amplification. Results also showed a clear difference in TGF-b1 expression in patients with non-Hodgkin’s lymphoma depending on the subtype of the lymphoma, suggesting its significant role in the patho-mechanism of this group of malignant diseases, as well as its potential value as a prognostic factor. Immunohistochemical analysis was also used to examine the TGF-b isoforms in osteosarcoma tissue and to investigate its relationship to subsequent lung metastasis. The results showed the presence of one or more TGF-b isoforms in the majority (81%) of tumor cells in osteosarcoma tissues in all of the subtypes (Yang et al. 1998). The expression of either TGF-b1 or TGF-b3 isoforms was also associated with a higher rate of subsequent lung metastasis. Interestingly, only minimal amounts of any of the TGF-b isoforms were found in the tumor bone matrix. The plasma levels of TGF-b1, TGF-b3, and endoglin–TGF-b3 complexes were also measured in patients with untreated earlystage breast cancer, including some patients of which were histologically confirmed as having axillary lymph node metastases (Li et al. 1998). The levels of both TGF-b3 and endoglin–TGF-b3 complex were significantly elevated in patients with positive lymph nodes compared to those without node metastasis and the
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
639
levels of both TGF-b3 and endoglin–TGF-b3 complex correlated with lymph node status. Interestingly, the only patient who died of the disease had very high plasma levels of TGF-b3 and endoglin–TGF-b3 complex and positive lymph nodes; this patient developed lung metastases within two years of diagnosis. The findings suggest that plasma levels of TGF-b3 and endoglin–TGF-b3 complex may be of prognostic value in the early detection of metastasis of breast cancer. To characterize the effects of TGF-b1 on the adhesion, motility, and invasiveness of a metastatic human pulmonary carcinoma cell line in vitro, lung cancer A549 cells were used to demonstrate that TGF-b1 can stimulate the invasion of cells into type I collagen gels in a dose-dependent manner, with both the number of cells entering the gel and the depth of invasion of cells into the gel increasing (Mooradian et al. 1992). The TGF-b1-mediated increase in invasion and motility was accompanied by a fourfold increase in A549 cell adhesion to type I collagen and suggests that TGF-b1 can influence cellular recognition of extracellular matrix (ECM) components and can modulate cellular adhesion and migration on these components, leading to increased invasive potential. A549 cells were also used to examine the role of homeobox genes that regulate sets of genes that determine cellular fates in embryonic morphogenesis and maintenance of adult tissue architecture by regulating cellular motility and cell–cell interactions that interact with the TGF-b1 signaling pathway. Studies have shown that a specific homeobox member, homeobox D3 (HOXD3), when overexpressed, upregulates integrin b3 expression in A549 cells and human erythroleukemia HEL cells, and enhances their motility and invasiveness (Taniguchi et al. 1995; Hamada et al. 2001). A microarray study of over 7075 genes conducted to determine the mechanisms underlying the HOXD3-enhanced motility and invasiveness highlighted a set of TGF-b-up-regulated genes, which included matrix metalloproteinase-2 (MMP-2), syndecan-1, cluster of differentiation 44 (CD44), and TGF-b-induced 68 kDa protein (Miyazaki et al. 2002). Exogenous TGF-b also caused this pattern of up-regulation in A549 cells and enhanced their migratory and invasive activity, confirming the involvement of TGF-b signaling. HOXD3 also reduced the expression of TGF-b-independent genes coding for desmosomal components such as desmoglein, desmoplakin, and plakoglobin, which are known to suppress tumor invasion and metastasis, suggesting that HOXD3 enhances the invasive and metastatic potential of cancer cells through TGF-b-dependent and -independent pathways. The androgen-independent R3327-MATLyLu Dunning rat prostate cancer epithelial cell line, which produces metastatic anaplastic lung tumors when reinoculated in vivo, was also evaluated for TGF-b expression. When R3327-MATLyLu cells were stably transfected with an expression vector that codes for latent TGF-b1 and subclones of cells that overexpressed TGF-b1 mRNA were isolated, the TGFb1-overproducing MATLyLu tumors were 50% larger, markedly less necrotic, and produced more extensive metastatic disease (lung metastases in 73% of all lobes and lymph node metastases in 88% of animals) compared to control MATLyLu tumors (lung metastases, 21%; lymph node metastases, 7%) (Steiner and Barrack 1992). TGF-b1-overproducing cells became transiently growth inhibited, and then resumed proliferation. In addition, racemic gossypol [(+/–)-GP], a naturally occurring
640
S.B. Jakowlew
polyphenolic yellow pigment present in cottonseed products that has been shown to inhibit in vitro proliferation of MAT-LyLu cells, human prostate cancer PC3 cells derived from a bone marrow metastasis and primary cultured human prostate cells, also has the ability to inhibit the metastasis of lung and lymph nodes of MAT-LyLu cells after implantation into Copenhagen rats (Jiang et al. 2004). RT-PCR and ELISA analyses showed that (+/–)-GP elevates the mRNA expression and protein secretion of TGF-b1 and caused reductions in DNA synthesis and prolonged the doubling times in PC3 cells. Furthermore, the growth inhibition of PC3 cells by conditioned media collected from the (+/–)-GP-treated-PC3 cells was completely reversed by addition of mouse monoclonal TGF-b1, TGF-b2, and TGF-b3 antibodies, suggesting the involvement of TGF-b1 in (+/–)-GP-induced growth inhibition of PC3 cells. These results suggest the inhibitory effects of (+/–)-GP on the proliferation of PC3 cells are associated with induction of TGF-b1. In addition, Chinese hamster ovary (CHO) cells transfected with the TGF-b1 gene showed large numbers of metastatic colonies in the lungs of nude mice inoculated with transfected CHO cells with excessive secretion of TGF-b1 and showed that tumors derived from these transfected cells demonstrated marked angiogenesis (Ueki et al. 1993). This suggests that overproduction of TGF-b1 by tumors may participate in the metastatic progression following establishment of angiogenesis at the primary tumor site. Syngeneic F344 rats were also used to evaluate the experimental metastatic potential of the 13762NF mammary adenocarcinoma clone MTLn3 after treatment with TGF-b1 and appearance of lung tumors after inoculation. A bell-shaped dose– response curve with two- to threefold increase in number of surface lung metastases was seen (Welch et al. 1990). Increased metastatic potential appears to occur from an increased propensity of cells to extravasate as tested in the membrane invasion culture system, where MTLn3 cells penetrated reconstituted basement-membrane barriers two- to threefold more than did untreated control cells, depending upon length of TGF-b1 exposure. Increased invasive potential was apparently due, in part, to an increase in type IV collagenolytic (gelatinolytic) and heparanase activity. TGF-b1 treatment of MTLn3 cells did not alter their growth rate or morphology in the presence of serum; however, growth was inhibited in serum-free medium. This suggests that TGF-b1 may modulate metastatic potential of mammary tumor cells by controlling their ability to break down and penetrate basement-membrane barriers.
Transforming Growth Factor-b Receptors The role of TGF-b signaling in cancer is complex, with tumor suppressor and pro-oncogenic activities depending on the particular tumor cell and its stage in malignant progression (Leivonen and Kähäri 2007; Derynk et al. 2001; Wakefield and Roberts 2002; Roberts and Wakefield 2003). TGF-b RI interacts with TGF-b RII to make this signaling pathway functional. Loss or impairment of either TGF-b receptor can make this signaling pathway non-functional. It has also been demonstrated in breast cancer cell lines that Smad2/3 signaling plays a dominant role in mediating tumor suppressor effects on well-differentiated breast cancer cell lines
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
641
grown as xenografts and prometastatic effects on a more invasive, metastatic cell line (Tian et al. 2003). Selective interference with activation of endogenous Smad2 and Smad3 by stable expression of a mutant form of the TGF-b RI (RImL45) unable to bind Smad2/3 but with a functional kinase also showed that reduction in Smad2/3 signaling by expression of RImL45 enhanced the malignancy of xenografted tumors of the well-differentiated MCF10A-derived tumor cell line MCF10CA1h, resulting in formation of larger tumors with a higher proliferative index and more malignant histologic features (Tian et al. 2004). In contrast, expression of RImL45 in the more aggressive MCF10CA1a cell line strongly suppressed formation of lung metastases following tail vein injection. These findings suggest a causal, dominant role for the endogenous Smad2/3 signaling pathway in the tumor suppressor and prometastatic activities of TGF-b. In addition, non-Smad signaling pathways, including p38 mitigen-activated protein kinase (MAPK) and c-Jun NH(2)-terminal kinase (JNK), cooperate with TGF-b/Smads in enhancing migration of metastatic MCF10CA1a cells. Mutations of the TGF-b RI gene have been reported to occur at high frequency in breast cancer metastases to lung (42%) and much less frequently in primary breast cancer tumors (6%), with all mutations being an identical C to A transversion at nucleotide 1160 of the gene (Chen et al. 1998). Such a mutation would result in a serine to tyrosine substitution at codon 387 (S387Y) that would disrupt receptor function and represent a pivotal genetic alteration in breast cancer progression. However, using both single-strand conformation polymorphism screening and sequencing of lung adenocarcinoma metastases, breast cancer metastases and colorectal cancer metastases for possible mutations at this site, no mutations of the TGF-b RI gene were found in any of the samples. These results suggest the S387Y mutation of the TGF-b RI gene is not common in these types of human cancers (Anbazhagan et al. 1999). Additional studies will be needed to resolve this discrepancy. Use of inhibitors of TGF-b signaling that disrupt the TGF-b receptors may be a novel strategy for the treatment of patients with metastatic cancer. Using a small molecule inhibitor, A-83-01, which is structurally similar to reported ALK-5 inhibitors previously developed (Sawyer et al. 2004) that block signaling of type I serine/threonine kinase receptors for cytokines of the TGF-b superfamily with a TGF-b-responsive reporter construct in lung cells, transcriptional activity induced by TGF-b type I receptor ALK-5 and that by activin type IB receptor ALK-4 and nodal type I receptor ALK-7 was inhibited (Tojo et al. 2005). A-83-01 was found to be more potent in the inhibition of ALK5 than another ALK-5 inhibitor, SB-431542, and also to prevent phosphorylation of Smad2/3 and the growth inhibition induced by TGF-b. In contrast, A-83-01 had little or no effect on bone morphogenetic protein type I receptors, p38 MAPK, or extracellular-regulated kinase (ERK). Consistent with these findings, A-83-01 inhibited EMT induced by TGF-b, suggesting that A-83-01 and related molecules may be useful for preventing the progression of advanced cancers. The efficacy of another TGF-b RI kinase inhibitor (TGF-b RI-I) to limit early systemic metastases in an orthotopic xenograft model of lung metastasis and in an intracardiac injection model of experimental bone and lung metastasis was also examined using human breast carcinoma MDA-MB-435-F-L cells, a highly
642
S.B. Jakowlew
metastatic variant of human breast cancer MDA-MB-435 cells, expressing the enhanced green fluorescent protein. Treatment of cells with the TGF-b RI-I had no effect on their growth, but blocked TGF-b-stimulated expression of integrin a(v)b(3) and cell migration in vitro (Bandyopadhyay et al. 2006). Members of the integrin family recognize a variety of spatially-restricted extracellular ligands. Classically, ligation of integrins activates cytoplasmic signals in the integrin-expressing cell and contributes to cell adhesion, migration, proliferation, and survival. Integrin-mediated TGF-b activation has been shown to play important roles in modulating tissue fibrosis, acute lung injury, and pulmonary emphysema. Systemic administration of TGF-b RI-I effectively reduced the number and size of the lung metastasis in both orthotopic xenograft and experimental metastasis models with no effects on primary tumor growth rate compared with controls. TGF-b RI-I treatment also reduced the incidence of widespread early skeletal metastases in the femur, tibia, mandible, and spine detected by whole-body fluorescence imaging. Tumor burden in femora and tibiae was also reduced after TGF-b RI-I treatment as detected by histomorphometry analysis compared with the placebo controls. These findings show that abrogation of TGF-b signaling by systemic administration of the TGF-b RI-I can inhibit both early lung and bone metastasis in animal model systems and suggest antimetastatic therapeutic potential of the TGF-b RI-I. At least two other members of the integrin family, a(v)b(6) and a(v)b(8), perform an additional function, activation of latent complexes of TGF-b. This process allows integrins on one cell to activate signals on adjacent [in the case of integrin a(v)b(6)] or nearby cells [in the case of integrin a(v)b(8)]. Given the important roles that TGF-b plays in modulating epithelial cell growth, EMT and tumor invasion and metastasis, integrin-mediated TGF-b activation is likely to play important roles in tumor growth and metastasis (Sheppard 2005). Expression of TGF-b RII is necessary for TGF-b to inhibit the growth of most epithelial cells. Microarray analysis of microdissected non-invasive BAC and invasive adenocarcinoma and adenocarcinoma-mixed type with BAC features identified transcriptional profiles of lung adenocarcinoma invasiveness (Borczuk et al. 2005). Among the signature set that was reduced in adenocarcinoma-mixed compared with BAC was TGF-b RII, suggesting down-regulation of TGF-b RII is an early event in lung adenocarcinoma metastasis. Immunostaining in independently acquired specimens demonstrated a correlation between TGF-b RII expression and length of tumor invasion. Repression of TGF-b RII in lung cancer cells increased tumor cell invasiveness and activated p38 MAPK. Microarray analysis of invasive cells identified potential downstream mediators of TGF-b RII with differential expression in lung adenocarcinomas. This suggests repression of TGF-b RII may act as a significant determinant of lung adenocarcinoma invasiveness, an early step in tumor progression toward metastasis. Reduced TGF-b RII expression has been found in various other cancers, including human thyroid differentiated and undifferentiated carcinomas. In rat thyroid transformed cells, the resistance to TGF-b is associated with a decreased expression of TGF-b RII mRNA and protein. After transfection of K-ras-transformed thyroid
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
643
cells with an expression vector carrying the human TGF-b RII gene regulated by an inducible promoter, the resulting isolated clones, overexpressing TGF-b RII, showed a reduction in the anchorage-dependent and -independent cell growth, compared with control K-ras-transformed cells (Turco et al. 1999). When transplanted into athymic nude mice, the transfected clones presented a decrease in tumorigenicity with respect to the highly malignant parental cells. Moreover, the diminished tumorigenic ability of the clones studied was accompanied by a statistically significant reduction in spontaneous and lung metastases. These findings demonstrate that TGF-b RII acts as a potent tumor suppressor gene when overexpressed in malignant thyroid cells. To characterize the impact of increased production of TGF-b in a xenograft model of human breast cancer, TGF-b-responsive MDA-231 cells were genetically modified by stable transfection so as to increase their production of active TGF-b1. Compared with control cells, cells that produced increased amounts of TGF-b proliferated in vitro more slowly (Tobin et al. 2002). In vivo, however, tumors derived from these cells exhibited increased proliferation and grew at an accelerated pace. To evaluate the role of autocrine TGF-b signaling, cells were also transfected with a dominant-negative truncated TGF-b RII. Disruption of autocrine TGF-b signaling in the TGF-b-overexpressing cells reduced their in vivo growth rate. Tissue invasion by the tumor was a distinctive feature of the TGF-b-overexpressing cells, whether or not the autocrine loop was intact. Furthermore, tumors derived from TGF-b-overexpressing cells, irrespective of the status of the autocrine TGF-bsignaling pathway, had a higher incidence of lung metastasis. Thus, in this experimental model system, in vitro assays of cell proliferation and invasion did not accurately reflect in vivo observations, perhaps due to autocrine and paracrine effects of TGF-b that influence the important in vivo-based phenomena of tumor growth, invasion, and metastasis. TGF-b is a potent immunosuppressive cytokine that is frequently associated with mechanisms of tumor escape from immunosurveillance (Massagué 2008; Leivonen and Kähäri 2007; Miyazono et al. 2003). It was reported that transplantation of murine bone marrow expressing a dominant-negative TGF-b RII (TGF-b RIIDN) leads to the generation of mature leukocytes capable of a potent antitumor response in vivo (Shah et al. 2002). Hematopoietic precursors in murine bone marrow from donor mice were rendered insensitive to TGF-b via retroviral expression of the TGF-b RIIDN construct and were transplanted in C57BL/6 mice before tumor challenge. After administration of B16-F10 murine melanoma cells into TGF-b RIIDN-bone marrow transplanted recipients, survival of challenged mice was 70% (7 of 10) versus 0% (0 of 10) for vector-control treated mice, and surviving TGF-b RIIDN-bone marrow mice showed a virtual absence of metastatic lesions in the lung. The effect of stable transfection of TGF-b RIIDN cDNA was also examined in a human oral carcinoma cell line that contained normal Ras and was growth-inhibited by TGF-b1. Treatment of cells with exogenous TGF-b1 resulted in a decrease in ligand-induced growth inhibition and loss of c-myc downregulation in test cells compared to controls (Huntley et al. 2004). Cells containing TGF-b RIIDN grew faster in monolayer culture, expressed less keratin-10, and
644
S.B. Jakowlew
exhibited increased motility and invasion in vitro compared to control cell lines. After orthotopic transplantation in athymic mice, cells containing TGF-b RIIDN formed comparable numbers of primary tumors at the site of inoculation as controls but the tumors were less differentiated as demonstrated by the absence of keratin-10 immunostaining. Further, metastatic dissemination to the lungs and lymphatics was more evident in grafts of cells containing TGF-b RIIDN than controls. Taken together, the results demonstrate that attenuation of TGF-b signaling through transfection of TGF-b RIIDN cDNA leads to an enhanced growth rate, a loss of tumor cell differentiation and an increase in migration and invasion; these characteristics correspond to the development of the metastatic phenotype. To elucidate the role of TGF-b signaling in mammary gland development, tumorigenesis, and metastasis, the gene encoding TGF-b RII (Tgfbr2) was conditionally deleted in the mammary epithelium (Tgfbr2MGKO), and Tgfbr2MGKO mice were mated to the mouse mammary tumor virus-polyomavirus middle T antigen (PyVmT) transgenic mouse model of metastatic breast cancer. Loss of Tgfbr2 in the context of PyVmT expression resulted in a shortened median tumor latency and an increased formation of pulmonary metastases (Forrester et al. 2005). These findings support a tumor-suppressive role for epithelial TGF-b signaling in mammary gland tumorigenesis and show that pulmonary metastases can occur and are even enhanced in the absence of TGF-b signaling in the carcinoma cells. A soluble Fc:TGF-b RII fusion protein (Fc:TGF-b RII) was tested in transgenic and transplantable models of breast cancer metastases. Systemic administration of Fc:TGF-b RII did not alter primary mammary tumor latency in mouse mammary tumor virus (MMTV)-Polyomavirus middle T antigen transgenic mice (Muraoka et al. 2002). However, Fc:TGF-b RII increased apoptosis in primary tumors, while reducing tumor cell motility, intravasation, and lung metastases. These effects correlated with inhibition of Akt/protein kinase B activity and forkhead box O3 related-1 (FKHRL1) phosphorylation. Fc:TGF-b RII also inhibited metastases from transplanted 4T1 and EMT-6 mammary tumors in syngeneic BALB/c mice. This method of interfering with TGF-b signaling may reduce tumor cell viability and migratory potential and represents a testable therapeutic approach against metastatic carcinomas. In another report, Cre/LoxP technology was used with the whey acidic protein promoter driving transgenic expression of Cre recombinase (WAP-Cre) and TGF-b RII expression was ablated specifically within mouse mammary alveolar progenitors. Transgenic expression of the polyoma virus middle T antigen, under control of the MMTV enhancer/promoter, was used to produce mammary tumors in the absence or presence of Cre (TGF-b RII((fl/fl);PY) and TGF-b RII((fl/fl);PY;WC), respectively). The loss of TGF-b signaling significantly decreased tumor latency and increased the rate of pulmonary metastasis and correlated with increased tumor size and enhanced carcinoma cell survival (Bierie et al. 2008). In addition, significant differences were observed in stromal fibrovascular abundance and composition accompanied by increased recruitment of F4/80(+) cell populations in TGF-b RII((fl/fl);PY;WC) mice when compared with TGF-b RII((fl/fl);PY) controls. The recruitment of F4/80(+) cells correlated with increased expression of known
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
645
inflammatory genes including chemokine (C-X-C motif) ligand 11 (Cxc11), Cxcl5, and prostaglandin-endoperoxide synthase 2 (Ptgs2). An enriched K5(+) dNp63(+) cell population was identified in primary TGF-b RII((fl/fl);PY;WC) tumors and corresponding pulmonary metastases, suggesting that loss of TGF-b signaling in this subset of carcinoma cells can contribute to metastasis. Together, this shows that loss of TGF-b signaling in mammary alveolar progenitors may affect tumor initiation, progression, and metastasis through regulation of both intrinsic cell signaling and adjacent stromal–epithelial interactions in vivo. Expression of a truncated soluble extracellular domain of TGF-b RIII (TGF-b sRIII) in human breast cancer MDA-MB-231 cells was also used to antagonize the tumor-promoting activity of TGF-b by sequestering active TGF-b isoforms that are produced by the cancer cells. The secretion of TGF-b sRIII reduced the amount of active TGF-b1 and TGF-b2 in the conditioned medium (Bandyopadhyay et al. 1999). This led to a significant reduction of the growth-inhibitory activity of the medium conditioned by TGF-b sRIII-expressing cells on the growth of mink lung epithelial CCL64 cells in comparison with the medium conditioned by the control cells. The tumor incidence and growth rate of all of the three TGF-b sRIII-expressing clones studied were significantly lower than those of the control cells in athymic nude mice. Four of five control cell-inoculated mice showed spontaneous metastasis in the lung, whereas none of the TGF-b sRIII-expressing cell-inoculated mice had any lung metastasis. Thus, these findings suggest that the TGF-b sRIII may be used to antagonize the tumor-promoting activity of TGF-b.
Smads TGF-b facilitates progression to metastasis during the advanced stages of cancer (Leivonen and Kähäri 2007; Derynk et al. 2001; Wakefield and Roberts 2002; Roberts and Wakefield 2003). Inhibitory Smad6 and Smad7 block signaling by TGF-b. Adenovirus-mediated gene transfer of Smad6 and Smad7 natural inhibitors in a mouse model of breast cancer with mouse mammary carcinoma JygMC(A) cells, which spontaneously metastasize to lung and liver, was used to examine proliferation, migration, and invasion. High-throughput western blotting analysis was used to examine the expression levels for 47 signal transduction proteins in JygMC(A) cells and primary tumors. Control mice bearing tumors derived from JygMC(A) cells showed many metastases to the lung and liver and all animals died by 50 days after cell inoculation (Azuma et al. 2005). By contrast, mice treated with AdCMV-Smad6 or AdCMVSmad7 demonstrated a dramatic decrease in metastasis and statistically significantly longer survival than control mice. Expression of Smad7 in JygMC(A) cells was associated with increased expression of major components of adherens and tight junctions, including E-cadherin, decreased expression of N-cadherin, and decreases in the migratory and invasive abilities of the JygMC(A) cells. Smad7 inhibits metastasis, possibly by regulating cell–cell adhesion and systemic expression of Smad7 may be a novel strategy for the prevention of metastasis of advanced cancers.
646
S.B. Jakowlew
Ski-related novel protein N (SnoN) is another important negative regulator of TGF-b signaling through its ability to interact with and repress the activity of Smad proteins. SnoN expression is highly elevated in many human cancer cell lines, and this high level of SnoN promotes mitogenic transformation of breast and lung cancer cell lines in vitro and tumor growth in vivo, consistent with its proposed prooncogenic role (Zhu et al. 2007). However, a high level of SnoN expression also inhibits EMT. Supporting this observation, in an in vivo breast cancer metastasis model, reducing SnoN expression was found to moderately enhance metastasis of human breast cancer cells to bone and lung. Thus, SnoN plays both pro-tumorigenic and antitumorigenic roles at different stages of mammalian malignant progression. These findings establish the importance of SnoN in mammalian epithelial carcinogenesis and reveal a novel aspect of SnoN function in malignant progression. In addition to SnoN, Sloan Kettering infected protein (Ski) is an important corepressor of TGF-b signaling through its ability to bind to and repress the activity of the Smad proteins. It was reported that reducing Ski expression in breast and lung cancer cells does not affect tumor growth but enhances tumor metastasis in vivo (Le Scolan et al. 2008). Thus, in these cells, Ski plays an antitumorigenic role. In addition, it was also shown that TGF-b induces Ski degradation through the ubiquitin-dependent proteasome in malignant human cancer cells. Upon TGF-b treatment, the E3 ubiquitin ligase, Arkadia, mediates degradation of Ski in a Smad-dependent manner (Le Scolan et al. 2008). Although Arkadia interacts with Ski in the absence of TGF-b, binding of phosphorylated Smad2 or Smad3 to Ski is required to induce efficient degradation of Ski by Arkadia. These studies suggest that the ability of TGF-b to induce degradation of Ski could be an additional mechanism contributing to its protumorigenic activity. In another study, it was shown that down-regulation of Ski through lentivirus-mediated RNA interference decreases tumor growth both in vitro and in vivo, butt promotes cell invasiveness in vitro, and lung metastasis in vivo in the pancreatic cancer cell line SW1990, which contain wild-type Smad4 expression, and the BxPC3 line, which is Smad4 deficient (Wang et al. 2009). It was also shown that the down-regulation of Ski increases TGF-b-induced transcriptional activity, which is associated with increased TGF-b-dependent Smad2/3 phosphorylation, and results in an altered expression profile of TGF-b-inducible genes involved in lung metastasis, angiogenesis and cell proliferation and EMT. These results suggest that Ski may act as a tumor proliferation promoting factor or as a metastatic suppressor in human pancreatic cancer.
Microenvironment Although many studies exploring the molecular mechanisms of metastasis have focused on features intrinsic to tumor cells, the tumor microenvironment also plays an important role (McSherry et al. 2007). Primary carcinomas, as well as metastases, are comprised of both tumor cells and cells of the stroma, including fibroblasts, endothelial cells, and inflammatory cells. There is strong evidence that interactions between the tumor cells and the stroma influence cancer growth and metastasis
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
647
(Karnoub et al. 2007; Lin et al. 2001; Olumi et al. 1999). One accepted role for stromal cells during metastasis is to enhance the ability of tumor cells to leave the primary tumor and to invade local tissue. However, extravasation and tumor establishment at sites distant from the primary tumor are limiting steps of metastasis, and the ways in which the primary tumor microenvironment may promote these later stages of metastasis is still unclear. One polypeptide growth factor that is produced abundantly by stromal cells in the tumor microenvironment is TGF-b. TGF-b is produced by stromal cells of the tumor microenvironment in response to such activities as hypoxia or inflammation, or by carcinoma-associated fibroblasts (CAFs). Cancer cells in advanced tumors are unresponsive to the growth inhibitory effects of TGF-b, and instead, respond to this growth factor by increasing their growth, survival, motility, and invasion, and by undergoing EMT. In addition to these effects, TGF-b stimulates expression of the adipokine angiopoietin-like 4 (ANGPTL4) by activating SMAD transcription factors. Secretion of ANGPTL4 enables tumor cells to extravasate into lung tissue and to seed micrometastase (Padua et al. 2008). The role of TGF-b produced by stromal cells in the metastasis of breast tumor cells to the lung was examined and a gene expression microarray approach was used to define a group of target genes expressed by human breast tumor cells in response to TGF-b exposure. A TGF-b-response signature (TBRS) was found to correlate with metastasis to lung tissue in patients with estrogen receptor (ER)-negative breast cancer. Comparison of gene-expression profiles between the TBRS and a lung metastasis signature (LMS) (Minn et al. 2007) revealed that ER-negative breast tumors scoring positive for both the TBRS and the LMS were uniquely associated with a high risk of lung metastasis. No link was identified between the TBRS and metastasis to bone, or metastasis of ER-positive breast tumor cells. To test the requirement for TGF-b in metastasis of ER-negative breast cancer cells, the authors turned to an ER-negative human breast cancer cell line, LM2, which had been selected experimentally for the capability to metastasize to lung (Minn et al. 2007). Blocking TGF-b signaling led to reduced metastasis to the lungs of recipient mice, but did not affect tumor growth or intravasation. These findings suggest that TGF-b is critical for ER-negative breast tumor cells to complete one or more of the later stages of metastasis, including survival in the bloodstream, extravasation, and colonization of the lungs. Treatment of the tumor cells with TGF-b for a short time prior to intravenous injection into mice was found to significantly increase the number of cells detected in the lung in the first 24h after inoculation (Padua et al. 2008). Because this effect was not sustained through outgrowth of metastatic colonies, this suggests that TGF-b selectively primes tumor cells for extravasation to lung. Analysis of two genes whose expression overlap between the TBRS and the LMS profiles led to the identification of ANGPTL4 as a possible mediator of TGF-b-mediated lung colonization. TGF-b signaling induces expression of ANGPTL4 in a number of cultured tumor cell lines, and also in primary malignant pleural cell effusion samples from breast cancer patients. Interestingly, blocking ANGPTL4 expression using short hairpin RNAs inhibited lung metastases from LM2 breast cancer
648
S.B. Jakowlew
cell xenografts in mouse mammary glands. This effect was due to decreased extravasation of the tumor cells into the lung. Secretion of ANGPTL4 by the breast cancer cell line led to disruption of junctions between vascular endothelial cells, causing increased permeability of lung capillaries. Together, these results suggest that exposure to TGF-b produced by the stroma enables ER-negative breast cancer cells to form metastatic lesions when they reach the lung. Interestingly, exposure of breast tumor cells to TGF-b had no effect on their ability to colonize bone following injection into the arterial circulation of recipient mice, presumably because of the absence of vascular endothelial junctions in the bone marrow vasculature. By focusing on critical downstream elements of TGF-b signaling during metastasis, a functional role for ANGPTL4 was identified in the extravasation step that enables breast tumor cells to colonize lung tissue. Using high-throughput genomic strategies, the molecular programs driving the tumor–stromal interactions that lead to metastases are becoming well characterized, with delineation of roles for transcription factors (Yang et al. 2004; Liotta and Kohn 2001), proteinases, such as MMP-11 and cathepsin L2 (Paik et al. 2004), and chemokines, such as chemokine (C-X-C motif ligand 12) (CXCL12) and CXCL14 (Allinen et al. 2004). Molecular signatures that are discriminative of NSCLC differentiation and prognosis have been reported (Bhattacharjee et al. 2001; Shah et al. 2004; Beer et al. 2002; Borczuk et al. 2003). Gene signatures of lung carcinoma prognosis often contain gene classifiers of metastasis in other tumor systems. A potential limitation of molecular signatures of prognosis derived from resected tumors is that associations of gene expression with survival may be confounded by tumor biological heterogeneity and non-tumor-related properties, such as patient performance status, co-morbid disease, and non-cancer-related causes of death. Many proteins are involved in metastasis, including MMPs which form a family of zinc-dependent matrix-degrading enzymes secreted by mesenchymal cells and tumor cells that have been implicated in invasion and metastasis (StetlerStevenson et al. 1993; Kahari and Saarialho-Kere 1999). Urokinase-type plasminogen activator (uPA) plays a key role in tumor-associated proteolysis resulting in the invasion and dissemination of tumor cells (Dano et al. 1994). uPA-catalyzed proteolysis involves the urokinase-type plasminogen activator receptor (uPAR) (Pepper et al. 1993). The activity of the receptor-bound uPA is regulated by inhibitors of plasminogen activator, PAI-1 and PAI-2 (Dano et al. 1994). Integrins comprise a supergene family of transmembrane proteins composed of non-covalently-associated a and b subunits (Hynes 1992). Integrins are involved in tumor cell adhesion, spreading and migration, and play important roles in tumor cell growth and metastasis (Liotta 1986). The proteins involved in cell growth and apoptosis also influence the process of metastasis and participate in one of four main activities, including ECM adhesion disruption, ECM degradation, cell cycle deregulation, and escape from apoptosis (Owen-Schaub et al. 1986; Holmgren et al. 1995).
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
649
Epithelial-to-Mesenchymal Transition Epithelial–mesenchymal transition (EMT) is a normal physiological process that regulates tissue development, remodeling, and repair (Thiery and Sleeman 2006). However, aberrant EMT also elicits disease development in humans, including lung fibrosis, rheumatoid arthritis, and cancer cell metastasis (Heldin et al. 2009; Yang and Weinberg 2008). EMT transition is also a fundamental biological process whereby epithelial cells lose their polarity and undergo a transition to a mesenchymal phenotype (Thiery 2003; Wheelock and Johnson 2003). When cancer cells invade adjacent tissues, they seem to use a mechanism akin to EMT (Yang and Weinberg 2008), and understanding the molecular mechanisms that drive this transition will facilitate studies into new targets for prevention of metastasis. Extracellular stimuli, such as growth factors, and their cytosolic effectors cooperate to promote EMT. TGF-b is a master regulator of EMT in normal mammary epithelial cells (MECs), wherein this pleiotropic cytokine also functions as a potent suppressor of mammary tumorigenesis (Massagué 2008; Leivonen and Kähäri 2007; Derynk et al. 2001; Wakefield and Roberts 2002; Roberts and Wakefield 2003). In contrast, malignant MECs typically evolve resistance to TGF-b-mediated cytostasis and develop the ability to proliferate, invade, and metastasize when stimulated by TGF-b. Establishing how TGF-b promotes EMT may offer new insights into targeting the oncogenic activities of TGF-b in human breast cancers. MEC cells were also used to show here that the TGF-b gene target, fibulin-5 (FBLN5), initiates EMT and enhances that induced by TGF-b by monitoring alterations in the actin cytoskeleton and various markers of EMT (Lee et al. 2008). It was also shown that the TGF-b gene target, fibulin-5 (FBLN5), initiates EMT and enhances that induced by TGF-b. Whereas normal MECs contain few FBLN5 transcripts, those induced to undergo EMT by TGF-b show significant up-regulation of FBLN5 messenger RNA, suggesting that EMT and the dedifferentiation of MECs override the repression of FBLN5 expression in polarized MECs. It was also shown that FBLN5 stimulated MMP expression and activity, leading to MEC invasion and EMT, to elevated Twist expression and to reduced E-cadherin expression. Finally, FBLN5 promoted anchorage-independent growth in normal and malignant MECs, as well as enhanced the growth of 4T1 tumors in mice. Taken together, these findings identify a novel EMT and tumor-promoting function for FBLN5 in developing and progressing breast cancers. In highly fibrotic cancers like lung cancer, it is thought that ECM, including collagen, can initiate signals that promote EMT. It was reported that collagen I induces EMT in NSCLC cell lines, which is prevented by blocking TGF-b3 signaling (Shintani et al. 2008). In addition, it was shown in the same report that collagen I-induced EMT is prevented by inhibitors of phosphoinositide-3 kinase (PI3K) and ERK signaling, which promotes transcription of TGF-b3 mRNA in these cells. These studies are consistent with the hypothesis that collagen I induces EMT in lung cancer cells by activating autocrine TGF-b3 signaling.
650
S.B. Jakowlew
The metastatic spread of epithelial cancer cells from the primary tumor to distant organs mimics the cell migrations that occur during embryogenesis with participation of EMT. Using gene expression profiling, it was found that the forkhead box C2 (FOXC2) transcription factor, which is involved in specifying mesenchymal cell fate during embryogenesis, is associated with the metastatic capabilities of cancer cells (Mani et al. 2007). FOXC2 expression is required for the ability of murine mammary carcinoma cells to metastasize to the lung, and overexpression of FOXC2 enhances the metastatic ability of mouse mammary carcinoma cells. FOXC2 expression was shown to be induced in cells undergoing EMT triggered by a number of signals, including TGF-b1 and several EMT-inducing transcription factors, such as Snail, Twist, and Goosecoid. FOXC2 specifically promotes mesenchymal differentiation during an EMT and may serve as a key mediator to orchestrate the mesenchymal component of the EMT program. Expression of FOXC2 is significantly correlated with the highly aggressive basal-like subtype of human breast cancers. These observations indicate that FOXC2 plays a central role in promoting invasion and metastasis and that it may prove to be a highly specific molecular marker for human basal-like breast cancers. It was reported that TGF-b1 increases expression of high-molecular weight tropomyosins (HMW-tropomyosins) and formation of actin stress fibers in normal epithelial cells (Bakin et al. 2004). The role of HMW-tropomyosins was investigated in TGF-b1-mediated cell motility and invasion and it was shown that TGF-b1 restricts motility of normal epithelial cells, although it promotes EMT and formation of actin stress fibers and focal adhesions (Zheng et al. 2008). Cell motility was enhanced by siRNA-mediated suppression of HMW-tropomyosins. TGF-b1 stimulated migration and matrix proteolysis in breast cancer MDA-MB-231 cells that express low levels of HMW-tropomyosins. Tet-Off-regulated expression of HMWtropomyosin inhibited cell migration and matrix proteolysis without affecting expression of MMPs. HMW-tropomyosins increased cell adhesion to matrix by enhancing actin fibers and focal adhesions. Finally, HMW-tropomyosins impaired the ability of tumor cells to form lung metastases in SCID mice. These studies suggest that HMW-tropomyosins are important for TGF-b-mediated control of cell motility and acquisition of the metastatic potential. Loss of epithelial morphology and the acquisition of mesenchymal characteristics may contribute to metastasis formation during colorectal tumorigenesis and metastasis with the Wnt, Notch, and TGF-b signaling pathways controlling tissue homeostasis and tumor development. The relationship between the activity of these pathways and the expression of epithelial and mesenchymal markers was investigated in a series of primary colorectal tumors and their corresponding metastases. When compared with normal mucosa, primary colorectal tumors showed a marked increase in the levels of cytoplasmic mesenchymal vimentin and nuclear epithelial b-catenin, phospho-Smad2 and hairy and enhancer of split homology 1 (HES1) in a tissue microarray format (Veenendaal et al. 2008). Increased vimentin expression correlated with the presence of oncogenic K-ras and with nuclear b-catenin. The corresponding liver, lymph node, brain and lung metastases did not express vimentin and displayed significantly lower levels of nuclear phospho-Smad2 and HES1,
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
651
while retaining nuclear b-catenin. The expression of mesenchymal (vimentin, fibronectin) and epithelial (E-cadherin) markers was related to markers of Wnt (b-catenin), Notch (HES1), and TGF-b (phospho-Smad2) signaling. Primary colorectal carcinomas display aberrant expression of vimentin, and have activated Notch and TGF-b signaling pathways. Surprisingly, many regional and distant metastases have lost nuclear HES1 and pSmad2, suggesting that the activity of the Notch and TGF-b pathways is reduced in secondary colorectal tumors. The role of E-cadherin and the retinoblastoma tumor suppressor protein (Rb) was also examined in metastasis. The Rb protein is mutated or expressed at very low levels in several tumor types, including retinoblastoma and osteosarcoma, as well as small cell lung, colon, prostate, bladder, and breast carcinomas and loss or reduction of Rb expression is seen most commonly in high-grade breast adenocarcinomas. This suggests that a relationship may exist between loss of Rb function and a less-differentiated state, increased proliferation, and high metastatic potential. It was reported that knockdown of Rb by small interfering RNA in MCF7 breast cancer cells disrupts cell–cell adhesion and induces a mesenchymal-like phenotype (Arima et al. 2008). Additionally, Rb is decreased during growth factor- and cytokine-induced EMT and overexpression of Rb inhibits the EMT in MCF10A human mammary epithelial cells. Ectopic expression and knockdown of Rb resulted in increased or reduced expression of E-cadherin, which is specifically involved in epithelial cell–cell adhesion, along with other EMT-related transcriptional factors, including Slug and zinc finger E-box binding homeobox-1 (Zeb-1). In addition, concurrent down-regulation of Rb and E-cadherin expression was observed in mesenchymal-like invasive cancers. These findings suggest that Rb inactivation contributes to tumor progression due to not only loss of cell proliferation control but also conversion to an invasive phenotype and that the inhibition of EMT is a novel tumor suppressor function of Rb. In most human breast cancers, lowering of TGF-b receptor or Smad gene expression combined with increased levels of TGF-b in the tumor microenvironment is sufficient to abrogate TGF-b’s tumor suppressive effects and to induce a mesenchymal, motile and invasive phenotype. In genetic mouse models, TGF-b signaling suppresses de novo mammary cancer formation, but promotes metastasis of tumors that have broken through TGF-b tumor suppression. In mouse models of triple-negative or basal-like breast cancer, treatment with TGF-b neutralizing antibodies or receptor kinase inhibitors strongly inhibits development of lung and bone metastases (Tan et al. 2009). These TGF-b antagonists do not significantly affect tumor cell proliferation or apoptosis. Rather, they de-repress anti-tumor immunity, inhibit angiogenesis and reverse the mesenchymal, motile, invasive phenotype characteristic of basal-like and HER2-positive breast cancer cells. Patterns of TGF-b target genes up-regulation in human breast cancers suggest that TGF-b may drive tumor progression in estrogen-independent cancer, while it mediates a suppressive host cell response in estrogen-dependent luminal cancers. In addition, TGF-b appears to play a key role in maintaining the mammary epithelial cancer stem cell pool, in part by inducing a mesenchymal phenotype, while differentiated, estrogen receptor-positive, luminal cells are unresponsive to TGF-b because the
652
S.B. Jakowlew
TGF-b RII gene is transcriptionally silent. These same cells respond to estrogen by down-regulating TGF-b, while antiestrogens act by up-regulating TGF-b. This model predicts that inhibiting TGF-b signaling should drive the differentiation of mammary stem cells into ductal cells. Consequently, TGF-b antagonists may convert basal-like or human epidermal growth factor receptor 2 (Her2)-positive cancers to a more epithelioid, non-proliferating, and perhaps, non-metastatic phenotype. Conversely, these agents might antagonize the therapeutic effects of anti-estrogens in estrogen-dependent luminal cancers. These predictions need to be addressed prospectively in clinical trials and should inform the selection of patient populations most likely to benefit from this novel anti-metastatic therapeutic approach. In human hepatocellular carcinoma, EMT correlates with aggressiveness of tumors and poor survival. A model of EMT based on immortalized p19 alternate reading frame (ARF) null hepatocytes, which display tumor growth upon expression of oncogenic Ras and undergo EMT through the synergism of Ras and TGF-b was used to show that the interleukin-related protein interleukin-like EMT inducer (ILEI), a novel EMT tumor- and metastasis-inducing protein, cooperates with oncogenic Ras to cause TGF-b-independent EMT (Lahsnig et al. 2009). Ras-transformed hepatocytes overexpressing ILEI showed cytoplasmic E-cadherin, loss of tight junction protein zona occludes-1 (ZO-1), and induction of a-smooth muscle actin as well as platelet-derived growth factor (PDGF)/PDGF receptor (PDGF-R) isoforms. As shown by dominant-negative PDGF-R expression in these cells, ILEI-induced PDGF signaling was required for enhanced cell migration, nuclear accumulation of b-catenin, nuclear pY-signal transducer of activator of transcription 3 (Stat3) and accelerated growth of lung metastases. In hepatocytes expressing the Ras mutant V12-C40, ILEI collaborated with PI3 kinase signaling resulting in tumor formation without EMT. Clinically, human hepatocellular carcinoma samples showed granular or cytoplasmic localization of ILEI correlating with well and poorly differentiated tumors, respectively. In conclusion, these data indicate that ILEI requires cooperation with oncogenic Ras to govern hepatocellular EMT through mechanisms involving PDGF-R/b-catenin and PDGF-R/Stat3 signaling. EMT has been considered essential for metastasis, including local invasion, intravasation, extravasation, and proliferation at distant sites (Yang and Weinberg 2008). However, controversy remains as to whether EMT truly happens and how important it is to metastasis. The involvement of EMT in individual steps of metastasis was studied and it was found that p12 cyclin-dependent kinase-2-activator protein-1 (CDK2-AP1), a downstream effector of TGF-b, induced EMT of hamster cheek pouch carcinoma-1 cells by promoting the expression of Twist2 (Tsuji et al. 2008). EMT cells have an increased invasive but decreased metastatic phenotype. When inoculated subcutaneously, both EMT and non-EMT cells established primary tumors, but only EMT cells invaded into the adjacent tissues and blood vessels; however, neither cells formed lung metastases. However, when inoculated intravenously, only non-EMT cells established lung metastases. Moreover, subcutaneous inoculation of a mixture of the two cell types resulted in intravasation of both cell types and formation of lung metastasis from non-EMT cells. These findings show that EMT and non-EMT cells cooperate to complete the spontaneous metastasis process.
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
653
Immune System TGF-b is a potent immunosuppressant and overproduction of TGF-b by tumor cells may lead to tumor evasion from the host immune surveillance and tumor progression (Massagué 2008; Leivonen and Kähäri 2007; Derynk et al. 2001; Wakefield and Roberts 2002; Roberts and Wakefield 2003). Transgenic mouse model for prostate cancer-C2 (TRAMP-C2) cells that produce large amounts of TGF-b1 were used as an experimental model to develop a treatment strategy through adoptive transfer of tumor-reactive TGF-b-insensitive CD8+ T cells. C57BL/6 mice were primed with irradiated TRAMP-C2 cells. CD8+ T cells were isolated from the spleens of C57BL/6 mice primed with irradiated TRAMP-C2 cells, expanded ex vivo, and rendered TGF-b insensitive by infecting with a retrovirus containing dominantnegative TGF-b RII. Results of in vitro cytotoxic assay revealed that these CD8+ T cells showed a specific and robust tumor-killing activity against TRAMP-C2 cells but were ineffective against an irrelevant tumor line, B16-F10 (Zhang et al. 2005). After recipient mice were challenged with a single injection of TRAMP-C2 cells before adoptive transfer of CD8+ T cells was done, pulmonary metastasis was either eliminated or significantly reduced in the group receiving adoptive transfer of tumor-reactive TGF-b-insensitive CD8+ T cells. Immunofluorescent studies showed that only tumor-reactive TGF-b-insensitive CD8+ T cells were able to infiltrate into the tumor and mediate apoptosis in tumor cells. Furthermore, transferred tumor-reactive TGF-b-insensitive CD8+ T cells were able to persist in tumor-bearing hosts but declined in tumor-free animals. In another study, tumorreactive CD8+ T cells from C57BL/6 mice were isolated, expanded ex vivo, and rendered insensitive to TGF-b by introducing a dominant-negative TGF-b RII vector. Following subcutaneous injection of TRAMP-C2 cells into the flank of BALB/ c-Rag1(–/–) mice, tumor-reactive, TGF-b-insensitive CD8+ T cells were transferred with and without the cotransfer of an equal number of CD8+-depleted splenocytes from C57BL/6 donors. Forty days following the transfer, the average tumor weight in animals that received cotransfer of tumor-reactive, TGF-b-insensitive CD8+ T cells and CD8-depleted splenocytes was at least 50% less than that in animals of all other groups (Zhang et al. 2006). Tumors in animals of the former group showed a massive infiltration of CD8+ T cells. This was associated with secretion of relevant cytokines, decreased tumor proliferation, reduced angiogenesis, and increased tumor apoptosis. The leukocyte activation marker cluster of differentiation 69 (CD69) is a novel regulator of the immune response that modulates the production of cytokines, including TGF-b. An antimurine CD69 monoclonal antibody, CD69.2.2, was generated which down-regulates CD69 expression in vivo but does not deplete CD69expressing cells. Therapeutic administration of CD69.2.2 to wild-type mice was shown to induces significant natural killer (NK) cell-dependent antitumor responses to major histocompatibility complex class I low RMA-S lymphomas and to RM-1 prostatic carcinoma lung metastases (Esplugues et al. 2005). These in vivo antitumor responses are comparable to those seen in CD69(–/–) mice. In vitro studies
654
S.B. Jakowlew
demonstrated the novel ability of anti-CD69 mAbs to activate resting NK cells in a Fc receptor-independent manner, resulting in a substantial increase in both NK-cell cytolytic activity and interferon gamma production. Modulation of the innate immune system with monoclonal antibodies to host CD69 thus provides a novel means to antagonize tumor growth and metastasis. Dendritic cell (DC)-based immunotherapy has not been as effective as expected in most solid tumors even in the murine model, particularly in renal cell carcinoma and pulmonary metastasis- and tumor recurrence-inhibitory effects of DC-vaccination was investigated in solid tumor-bearing mice. In these experiments, it was found that the limitations of DC-based immunotherapy to solid renal cell carcinoma likely result from tumor-mediated TGF-b hindrance of immune attack rather than insufficient immune induction by DC therapy (Lim et al. 2007). In fact, the circulating T lymphocyte (CTL) response induced by DC therapy was quite sufficient and functional for the inhibition of tumor recurrence after surgery or of tumor metastasis induced by additional tumor-challenge to the tumor-bearing mice. Taken together, these findings suggest the potential of DC immunotherapy in tumor patients for hindering or blocking disease progression by inhibition of tumor metastasis and/or tumor recurrence after surgery. Patients with idiopathic pulmonary fibrosis have a high incidence of lung cancer and a worse prognosis for clinical treatment. A few molecules with antifibrosis properties have been shown to promote cancer progression in clinical trials. Bleomycin was used to induce pulmonary fibrosis in mice with or without naringenin treatment in both passive and spontaneous metastatic C57BL/6 and BALB/c mouse models to test the hypothesis that mice with pulmonary fibrosis could have an increased risk of lung cancer and associated cancer progression. It was shown that mice with lung fibrosis challenged using tumors show an increased incidence of lung metastasis and shorter life spans compared with the mice without lung fibrosis (Du et al. 2009). A fibrotic environment in the lung results in increased abundance of TGF-b1 and CD4(+)CD25(+)Foxp3(+) regulatory T cells and a decreased proportion of activated effector T cells. This grave immunosuppressive environment favors tumor localization and growth. Naringenin significantly reduces lung metastases in mice with pulmonary fibrosis and increases their survival by improving the immunosuppressive environment through down-regulating TGF-b1 and reducing regulatory T cells. Naringenin could be an ideal therapeutic agent in the treatment of both cancer and fibrosis.
Drugs, Treatments, and Therapies Despite major improvements in patient management, the prognosis for patients with lung cancer remains dismal (Jemel et al. 2009). As our knowledge of the molecular biology of cancers has increased, new targets for therapeutic interventions have been identified. Established concepts, such as retinoid metabolism and the inhibition of cyclooxygenase-2 metabolism undergo new testing. In addition,
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
655
newer targets, such as TGF-b signaling, Janus-activated kinase (JNK)/signal transducers and activators of transcription pathway, and cell invasion are being investigated. Studies demonstrate that multiple, often overlapping, mechanisms of disruption are present in lung cancer cells, presenting a plethora of molecular targets (Gazdar et al. 2004). Tranilast (N-[3,4-dimethoxycinnamonyl]-anthranilic acid) is a drug of low toxicity that is orally administered, and has been used clinically in Japan as an antiallergic and antifibrotic agent. Its antifibrotic effect is thought to depend on the inhibition of TGF-b. Tranilast inhibited the proliferation of several tumor cell lines including mouse mammary carcinoma (4T1), rat mammary carcinoma stem cell (LA7), and human breast carcinoma (MDA-MB-231 and MCF-7) (Chakrabarti et al. 2009). In the highly metastatic 4T1 cell line, tranilast inhibited phospho-Smad2 generation, consistent with a blockade of TGF-b signaling. It also inhibited the activation of MAP kinases (ERK1 and 2 and JNK), which have been linked to TGF-b-dependent EMT and, indeed, it blocked EMT. Tranilast reduced (>50%) the growth of the primary tumor. However, its effects on metastasis were more striking, with more than 90% reduction of metastases in lungs and no metastasis in the liver. BALB/c mice or severe combined immunodeficient-beige mice were treated with the immunosuppressive drug, tacrolimus, also known as Fujimycin, and the effect of treatment on mouse renal cancer cell pulmonary metastasis was investigated. Treatment with tacrolimus resulted in a dose-dependent increase in the number of pulmonary metastases in the BALB/c mice and in the severe combined immunodeficient-beige mice (Maluccio et al. 2003). Tacrolimusinduced TGF-b1 overexpression may be a pathogenetic mechanism in tumor progression. Bryostatin 1 and phorbol-12-myristate-13-acetate (PMA) have been used to reduce the intracellular melanin level in high metastatic overexpressing nPKCdelta BL6 (BL6T) cells, thereby inducing white experimental metastasis in syngeneic mice. The possible differences between white and black metastases induced by both treatments on the proliferative and metastatic potential as well as on the expression of some cytokines involved in the metastatic process such as TGF-b, interleukin-10 (IL-10), and interferon (IFN)-gamma were evaluated after the injection of bryostatin 1- or PMA-treated cells into the tail vein of syngenic mice. The results showed only one significant difference between bryostatin I and phorbol ester, namely the cells obtained from white bryostatin 1-treated cells return to a black phenotype after a few passages in culture (La Porta and Comolli 2000). In vivo white metastases showed higher levels of TGF-b and IFN-gamma and when re-injected into syngeneic mice, gave big black metastases. Therefore, in murine melanoma cells, treatment with bryostatin 1 induces the appearance of a white population expressing different levels of TGF-b and IFN-gamma. The effect of TGF-b and PMA alone and in combination was examined in human NSCLC and normal human bronchial epithelial (NHBE) cells and shown to increase expression of TGF-b1 mRNA 24h after their addition to both cell lines (Jakowlew et al. 1997a, b). TGF-b1 and PMA both caused a persistent increase in expression of the mRNAs for both PAI-1 and PA up to 24h in most NSCLC cells, with the increase in PAI-1 mRNA beginning several hours before that of PA mRNA. In contrast, while
656
S.B. Jakowlew
TGF-b1 also increased expression of PAI-1 mRNA in NHBE cells, the expression of PA mRNA decreased simultaneously. The effect of PMA on PAI-1 and PA mRNAs was opposite of TGF-b1 in these cells, with expression of PAI-1 mRNA decreasing and PA mRNA increasing after addition of PMA. The responses of the mRNAs and proteins of TGF-b1, PAI-1 and PA to TGF-b1 and PMA were inhibited by the serine/threonine kinase inhibitor H7 in NSCLC cells. These studies show that there is parallel regulation of the genes for TGF-b1, PAI-1 and PA by TGF-b1 and PMA in NSCLC, but differential regulation of the genes for PAI-1 and PA by these agents in NHBE cells and because the effects of TGF-b and PMA on the different TGF-b isoforms, PA, and PAI, as well as on and fibronectin in NHBE and NSCLC cells are complex, these studies suggest that there are distinct mechanisms for controlling the different TGF-b isoforms, PA, PAI and extracellular matrix proteins in normal lung and lung cancer cells. Both PMA and TGF-b1 can also induce early growth response gene-1 (Egr) mRNA expression in NSCLC cells and NHBE cells with PMA inducing Egr-1 mRNA similarly in both cell types, while TGF-b1 induces Egr-1 mRNA expression more rapidly and more transiently in NSCLC cells than in NHBE cells (You and Jakowlew 1997). 13-cis-Retinoic acid can mediate differentiation of transformed cells and slow the proliferation of malignant cells, suggesting its use as a potential intervention tool. Specific cDNA probes for retinoic acid receptors demonstrated the expression of mRNAs for the different retinoic acid receptor isoforms in SCLC cell lines (Avis et al. 1995). Addition of 13-cis-retinoic acid to SCLC cells resulted in an increase in the level of the retinoic acid receptor-beta mRNAs and in a significant dosedependent, growth-inhibitory effect using serum-free conditions that decreased when cells were cultured in medium containing serum or serum components. Preincubating serum with triglycerides restored the inhibitory effects of 13-cisretinoic acid demonstrated in serum-free systems and suggests that 13-cis-retinoic acid preferentially binds to serum albumin, restricting its inhibitory effects on epithelial cell receptors. Blocking retinoic acid–albumin interactions with a fatty acid source may improve the bioavailability of 13-cis-retinoic acid and significantly enhance the inhibitory effect in vivo. In addition, TGF-b2 transcripts increased while TGF-b3 transcripts decreased in multiple NSCLC cells upon treatment with 13-cis-retinoic acid (Jakowlew et al. 2000). Rapamycin is an effective inhibitor of human renal cancer metastasis. A human renal cell cancer pulmonary metastasis model was developed using human RCC 786-O as the tumor challenge and the severe combined immunodeficient (SCID) beige mouse as the host. Rapamycin reduced the number of pulmonary metastases, whereas interestingly, cyclosporine increased pulmonary metastases (Luan et al. 2003). Rapamycin was effective in cyclosporine-treated mice, and rapamycin or rapamycin plus cyclosporine prolonged survival. The modifying effects of a Kunitz trypsin inhibitor (KTI) and a Bowman-Birk trypsin inhibitor (BBI), purified from soybean trypsin inhibitor, as dietary supplements on experimental and spontaneous pulmonary metastasis of murine Lewis lung carcinoma 3LL cells as well as peritoneal disseminated metastasis model in human ovarian cancer HRA cells were also investigated in mouse models. It was shown in an in vivo spontaneous metastasis
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
657
assay that diet supplementation with KTI, but not with BBI, for 28days immediately after tumor cell inoculation significantly inhibited the formation of lung metastasis in C57BL/6 mice in a dose-dependent manner (Kobayashi et al. 2004). The inhibition of lung metastasis was not due to direct antitumor effects of KTI. In an in vivo experimental metastasis assay, the diet supplementation with KTI or BBI for 21days after tumor cell inoculation did not reduce the number of lung tumor colonies. In addition, KTI treatment in a peritoneal disseminated metastasis model of HRA cells resulted in a 40% reduction in total tumor burden when compared with control animals. These results suggest that dietary supplementation of KTI more efficiently regulates the mechanism involved in the entry into vascular circulation of tumor cells in intravasation than in extravasation during the metastatic process. KTI treatment may also be beneficial for ovarian cancer patients with or at risk for peritoneal disseminated metastasis; it greatly reduces tumor burden in part by inhibiting phosphorylation of MAP kinase and PI3 kinase, leading to suppression of uPA expression. Calcium homeostasis is another area of drug study. Calcium homeostasis is a tightly regulated process that involves the co-ordinated efforts of the skeleton, kidney, parathyroid glands, and intestine. Neoplasms can alter this homeostasis indirectly through the production of endocrine factors resulting in humoral hypercalcemia of malignancy. Relatively common with breast and lung cancer, this paraneoplastic condition is most often due to tumor production of parathyroid hormone-related protein (PTHrP) and ensuing increased osteoclastic bone resorption. The metastasis of tumor cells to bone represents another skeletal complication of malignancy (Clines and Guise 2005). As explained in the “seed and soil” hypothesis (Fidler 2003), bone represents a fertile ground for cancer cells to flourish. In the case of osteolytic bone disease, tumor-produced PTHrP stimulates osteoclasts that in turn secrete tumor-activating TGF-b that further stimulates local cancer cells. This vicious cycle of bone metastases represents reciprocal bone/cancer cellular signals that likely modulate osteoblastic bone metastatic lesions as well. The development of targeted therapies to block initial cancer cell chemotaxis, invasion and adhesion or to break the vicious cycle is dependent on a more complete understanding of bone metastases. Although bisphosphonates delay progression of skeletal metastases, it is clear that more effective therapies are needed. Cancer-associated bone morbidity remains a major public health problem, and to improve therapy and prevention it is important to understand the pathophysiology of the effects of cancer on bone. Decorin is a major extracellular matrix protein which has become the focus of various cancer studies. The potential of decorin as a novel biological target for the treatment of osteosarcoma has been studied, with lung metastases being the most crucial event affecting the therapeutic outcome of osteosarcoma. In one study, the LM8 murine osteosarcoma cell line with high metastatic potential to the lung was used and two cell lines were established, including LM8-DCN which stably expressed human decorin and LM8-mock control. When the LM8-DCN cell line was subcutaneously injected into the backs of mice, significantly fewer pulmonary metastases were observed in mice with LM8-DCN compared to mice inoculated
658
S.B. Jakowlew
with LM8 and LM8-mock (Shintani et al. 2008). In addition, the mice in the LM8DCN inoculated group survived significantly longer than those in the LM8 and LM8-mock inoculated group. There was no difference in the morphology and growth rates, but the motility and invasion of LM8 were inhibited by decorin. These results suggest that decorin has the therapeutic potential to prevent lung metastasis in osteosarcoma. Doxorubicin is a DNA-damaging drug, commonly used for treatment of cancer patients. Doxorubicin causes not only cytotoxic and cytostatic effects, but also inhibits metastasis formation. The influence of doxorubicin on TGF-b signaling in tumor cells was examined and demonstrated that Doxorubicin inhibited TGF-b signaling in human lung adenocarcinoma A549 cells; namely, it blocked TGF-b1induced activation of Smad3-responsive CAGA(12)-Luc reporter, but did not affect c-myc-Luc reporter (Filyak et al. 2008). That effect was observed as early as after 1–3h of treating these cells with doxorubicin, while the other drugs cisplatin or methotrexate did not alter activation of CAGA(12)-Luc reporter under the same conditions. Also, after 1h, doxorubicin abrogated TGF-b-induced translocation of Smad3-protein from the cytoplasm to the nucleus. Down-regulation of expression of Smad2, Smad3, and Smad4 proteins, and up-regulation of inhibitory Smad7 protein upon doxorubicin treatment, were found after 12–24h of doxorubicin treatment. Phosphorylation of Smad2/3 proteins was also affected by doxorubicin.
Genomics Lung cancer often metastasizes to bone in patients with advanced disease. The bone is a rich source of many chemokines and growth factors, including: insulin-like growth factor (IGF) I and II, TGF-b, interleukins, and tumor necrosis factor-alpha (TNF-a). Identification of other factors involved in the interactions between lung cancer cells and bone will improve the prevention and treatment of bone metastases. Changes in metastasis-related gene expression of human HARA lung squamous carcinoma cells co-cultured with neonatal mouse calvariae using a pathwayspecific microarray analysis. Nine genes were up-regulated and two genes downregulated in HARA cells co-cultured with mouse calvariae (Deng et al. 2007). Five of the nine up-regulated genes, including caveolin 1, CD44, ephrin-B2 (EphB2), ezrin, and PTHrP, and one down-regulated gene, secretory leukocyte protreinase inhibitor I (SLPI), were further confirmed by RT-PCR amplification. A mouse model was subsequently used to study the role of PTHrP and ezrin in bone metastasis in vivo. PTHrP (all three isoforms) and ezrin were up-regulated in HARA cells at sites of bone metastasis as detected by RT-PCR amplification and immunohistochemical analysis. The PTHrP-141 mRNA isoform was increased by the greatest extent in bone metastases compared to PTHrP-139 and PTHrP-173 mRNA. A HARA cell line was generated in which PTHrP expression was inducibly silenced by RNA interference. Silencing of PTHrP expression caused significant reduction of submembranous F-actin and decreased HARA cell invasion. Ezrin up-regulation
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
659
was confirmed by western blots on HARA cells co-cultured with adult mouse long bones. Further, TGF-b was identified as one of the factors in the bone microenvironment that was responsible for the up-regulation of ezrin. The identification of PTHrP and ezrin as important regulators of lung cancer bone metastasis offers new mechanistic insights into the metastasis of lung cancer and provides potential targets for the prevention and treatment of lung cancer metastasis. It has also been proposed that exposure of breast cancer cells to the bone microenvironment results in alterations in gene expression that favor the growth and proliferation of tumor cells in the bone. To investigate this, MDA-MB-231 breast carcinoma cells were exposed to bone-derived conditioned media generated by culturing fetal rat calvaria for 24h under serum free conditions. Using cDNA microarray technology, the insulin-like growth factor family of binding proteins (IGFBPs) was identified as genes whose expression profiles are consistently and significantly altered with exposure to this simulated bone environment in vitro, when compared with untreated controls (Giles and Singh 2003). These findings suggest that the up-regulation of IGFBP-3 seen with exposure to the bone microenvironment is directly linked to an increase in TGF-b-mediated cell proliferation and appears to be functioning through an IGF-independent mechanism. Gene expression profiling of metastatic brain tumors from primary lung adenocarcinoma, using a 17K-expression array, also revealed that 1561 genes were consistently altered (Zohrabian et al. 2007). Further functional classification placed the genes into seven categories: cell cycle and DNA damage repair, apoptosis, signal transduction molecules, transcription factors, invasion and metastasis, adhesion, and angiogenesis. Genes involved in apoptosis, such as caspase 2 (CASP2), transforming growth factor-b-inducible early gene (TIEG), and neuroprotective heat shock protein 70 (Hsp70) were underexpressed in metastatic brain tumors. Alterations in Rho GTPases (ARHGAP26, ARHGAP1), as well as down-regulation of the metastasis suppressor gene kisspeptin-1 (KiSS-1) were noted, which may contribute to tumor aggression. Overexpression of the invasion-related gene neurofibromatosis 1 (NF1), and angiogenesis-related genes VEGF-b and placental growth factor (PGF) was also observed, while brain-specific angiogenesis inhibitors 1 and 3 (BAI1 and BAI3) were underexpressed as well. Examination of cell-adhesion and migration-related genes revealed an increased expression of integrins and extracellular matrices collagen and laminin. The study also showed alterations in p53 protein-associated genes, among these increased gene expression of p53, up-regulation of Reprimo or candidate mediator of the p53-dependent G2-arrest, down-regulation of p53-regulated apoptosis-inducing protein 1 (p53 AIP1), p53 decreased expression of tumor protein inducible nuclear protein 1 (p53 DINP1), and down-regulation of transformed 3T3 cell double minute 4 (Mdm4). The results demonstrated that genes involved in adhesion, motility, and angiogenesis were consistently up-regulated in metastatic brain tumors, while genes involved in apoptosis, neuroprotection, and suppression of angiogenesis were markedly down-regulated, collectively making these cancer cells prone to metastasis. To identify and functionally characterize genes involved in the mechanisms of osseous metastasis, a murine lung cancer model was developed. Comparative transcriptomic
660
S.B. Jakowlew
analysis identified genes encoding signaling molecules (such as transcription factor 4 [TCF4] and protein kinase D [PRKD3]) and cell anchorage-related proteins (melanoma cell adhesion molecule [MCAM] and sushi domain-containing protein 5 precursor [SUSD5]), some of which were basally modulated by TGF-b in tumor cells and in conditions mimicking tumor–stromal interactions (Vicent et al. 2008). Triple gene combinations induced not only high osteoclastogenic activity but also a marked enhancement of global metalloproteolytic activities in vitro. These effects were strongly associated with robust bone colonization in vivo, whereas this gene subset was ineffective in promoting local tumor growth and cell homing activity to bone. Interestingly, global inhibition of metalloproteolytic activities and simultaneous TGF-b blockade in vivo led to increased survival and a remarkable attenuation of bone tumor burden and osteolytic metastasis. Thus, this metastatic gene signature mediates bone matrix degradation by a dual mechanism of induction of TGFb-dependent osteoclastogenic bone resorption and enhancement of stroma-dependent metalloproteolytic activities. These findings suggest the cooperative contribution of host-derived and cell autonomous effects directed by a small subset of genes in mediating aggressive osseous colonization.
Animal Models Various mouse animal model systems have been used to examine the role of TGF-b in lung cancer metastasis. The roles of TGF-b1, TGF-b2, and TGF-b3 were determined in the in-bred A/J mouse model challenged with the carcinogen ethyl carbamate to induce lung adenomas. Immunostaining for the TGF-b ligands and receptors was detected in the epithelia of the bronchioles of untreated and treated A/J mice at similar levels and in adenomas by two months (Jakowlew et al. 1998a, b). While immunostaining for TGF-b1, TGF-b2, and TGF-b3 and TGF-b RI in adenomas was detected at levels comparable to those in bronchioles, immunostaining for TGF-b RII was less intense in adenomas than in bronchioles, and decreased immunostaining for TGF-b RII in adenomas persisted for at least 8 months after exposure to ethyl carbamate. In situ hybridization studies conducted with TGF-b receptor riboprobes showed a corresponding reduction in expression of TGF-b RII mRNA, but not of TGF-b RI mRNA in adenomas compared with expression in normal bronchioles. Expression of TGF-b RII mRNA was also examined in nontumorigenic and tumorigenic mouse lung cells and found to be lower in tumorigenic cells derived from ethyl carbamate-induced lung tumors. These findings suggest that a decrease in expression of TGF-b RII may contribute to autonomous cell growth and may play an important role in mouse lung carcinogenesis induced by ethyl carbamate. Whereas expression of the proteins and mRNAs for TGF-b1 and TGF-b RI was found to be comparable in lung adenomas and bronchioles of in-bred A/J mice treated with benzo(a)pyrene in chemically-induced mouse lung tumorigenesis, decreased immunostaining and hybridization for TGF-b RII protein and mRNA was detected in 50% of lung adenomas in these mice (Kang et al. 2000)
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
661
along with increased levels of surfactant protein D (Zhang et al. 2003). The non-steroidal anti-inflammatory drug indomethacine was shown to be able to reduce lung tumor number in these mice (Moody et al. 2001). Interestingly, expression of TGF-b1 and the TGF-b receptor proteins in A/J mice was similar to that of bronchioles in C57B1/6 mice and their littermates heterozygous for deletion of the TGF-b1 gene treated with diethylnitrosamine (Tang et al. 1998). Such findings show that reduced levels of expression of TGF-b RII occur in some, but not all, human and mouse lung tumors. This suggests that different mechanisms of action, some of which may involve the TGF-b signaling pathway, may contribute to the progression of mouse lung tumorigenesis. To elucidate the role of TGF-b1 and TGF-b RII as tumor-suppressor genes in lung carcinogenesis, C57BL/6 mice heterozygous (HT) for deletion of the TGF-b1 gene were mated with A/J mice to produce AJBL6 TGF-b1 HT progeny and their wild-type (WT) littermates. Immunohistochemical staining, in situ hybridization, and northern blot analyses showed decreased staining and hybridization for TGF-b1 protein and mRNA, respectively, in the lungs of normal HT mice versus WT mice and competitive reverse transcription polymerase chain reaction (CRT-PCR) amplification showed the level of TGF-b1 mRNA in the lungs of HT mice to be fourfold lower than the level in WT lung (Kang et al. 2000). When challenged with ethyl carbamate, lung adenomas were detected in 55% of HT mice by 4 months but only in 25% of WT littermates at this time. Whereas all HT mice had adenomas by 6 months, it was not until 10 months before all WT mice had adenomas. After 12 months, the average number of adenomas was fivefold higher in HT lungs than in WT lungs. Most dramatic was the appearance of lung carcinomas in HT mice 8months before they were visible in WT mice. Increased susceptibility of HT mice to chemically-induced lung tumorigenesis was also shown to occur independently of K-ras (McKenna et al. 2001) and to result in reduced levels of expression of Nkx2 homeobox 1 transcription factor (Nkx2.1), also known as thyroid transcription factor 1 (TTF-1) (Kang et al. 2004). In addition, weak immunostaining for TGF-b RII was detected in 67% of HT carcinomas at 12 months, whereas only 22% of WT carcinomas showed weak staining for this protein. Individual lung carcinomas showing reduced TGF-b RII expression and adjacent normal bronchioles were excised from HT lungs using laser capture microdissection, and CRT-PCR amplification of the extracted RNA showed 12-fold less TGF-b RII mRNA in these carcinomas compared with bronchioles. Decreasing TGF-b RII mRNA levels occurred with increasing tumorigenesis in lung hyperplasias, adenomas, and carcinomas, with carcinomas having four- and sevenfold lower levels of TGF-b RII mRNA than adenomas and hyperplasias, respectively. These findings show enhanced ethyl carbamate-induced lung tumorigenesis in AJBL6 HT mice compared with WT mice, suggesting that both TGF-b1 alleles are necessary for tumor-suppressor activity. Reduction of TGF-b RII mRNA expression in progressive stages of lung tumorigenesis in HT mice suggests that loss of TGF-b RII may play an important role in the promotion of lung carcinogenesis in mice with reduced TGF-b1 gene dosage when challenged with carcinogen. Multiple components of the cyclin D/cyclin-dependent kinase-4 (CDK-4)/p16(ink4a)pRb signaling pathway were
662
S.B. Jakowlew
demonstrated to be frequently altered early in lung tumors of HT mice induced by ethyl carbamate (Kang et al. 2002). Heterozygous TGF-b1 mice were also mated with latent activatable (LA) mutated K-ras mice to generate TGF-b1(+/+), K-ras LA (wild-type (WT)/LA) and TGF-b1(+/–), K-ras LA (HT/LA) mice. While both HT/LA and WT/LA mice developed spontaneous lung tumors, HT/LA mice progressed to adenocarcinomas significantly earlier compared with WT/LA mice and HT/LA adenocarcinomas had significantly lower angiogenic activity compared with WT/LA adenocarcinomas (Pandey et al. 2007). Thus, while oncogenic K-ras mutation and insensitivity to the growth regulatory effects of TGF-b1 is essential for initiation and progression of mouse lung tumors to adenocarcinoma, a full gene dosage of TGF-b1 is required for tumor-induced angiogenesis and invasive potential. The influence of TGF-b signaling on Neu-induced mammary tumorigenesis and metastasis was examined with transgenic mouse models using mice expressing an activated TGF-b RI or dominant negative TGF-b RII under control of the mouse mammary tumor virus promoter. When crossed with mice expressing activated forms of the Neu receptor tyrosine kinase that selectively couple to the growth factor receptor-bound protein 2 (Grb2) or Src homology 2 domain-containing protein (Shc) signaling pathways, the activated TGF-b increased the latency of mammary tumor formation and also enhanced the frequency of extravascular lung metastasis (Siegel et al. 2003). Conversely, the expression of dominant-negative TGF-b RII decreased the latency of Neu-induced mammary tumor formation, while significantly reducing the incidence of extravascular lung metastases. These observations argue that TGF-b can promote the formation of lung metastases while impairing Neu-induced tumor growth and suggest that extravasation of breast cancer cells from pulmonary vessels is a point of action of TGF-b in the metastatic process. To address the role of TGF-b in the progression of established tumors while avoiding the confounding inhibitory effects of TGF-b on early transformation, doxycyclineinducible triple transgenic mice were generated in which active TGF-b1 expression could be conditionally regulated in mouse mammary tumor cells transformed by the polyomavirus middle T antigen. Doxycycline-mediated induction of TGF-b1 for as little as 2weeks increased lung metastases more than ten-fold without a detectable effect on primary tumor cell proliferation or tumor size (Muraoka-Cook et al. 2004). Doxycycline-induced active TGF-b1 protein and nuclear Smad2 were restricted to cancer cells, suggesting a causal association between autocrine TGF-b and increased metastases. Antisense-mediated inhibition of TGF-b1 in polyomavirus middle T antigen-expressing tumor cells also reduced basal cell motility, survival, anchorage-independent growth, tumorigenicity, and metastases. Thus, induction and/or activation of TGF-b in hosts with established TGF-b-responsive cancers can rapidly accelerate metastatic progression. Using the MMTV/PyVmT transgenic model of metastatic breast cancer, it was shown that administration of ionizing radiation or doxorubicin caused increased circulating levels of TGF-b1 as well as increased circulating tumor cells and lung metastases (Biswas et al. 2007). Circulating polyomavirus middle T antigen-expressing tumor cells did not grow ex vivo in the presence of the
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
663
TGF-b antibody, suggesting autocrine TGF-b is a survival signal in these cells. Radiation failed to enhance lung metastases in mice bearing tumors that lack TGF-b RII, suggesting that the increase in metastases was due, at least in part, to a direct effect of TGF-b on the cancer cells. Such findings implicate TGF-b induced by anticancer therapy as a pro-metastatic signal in tumor cells and provide a rationale for the simultaneous use of these therapies in combination with TGF-b inhibitors. Use of small interfering RNA has been perfected in the past few years to demonstrate functional roles of gene products. Inhibition of TGF-b1 expression with small interfering RNA constructs targeting TGF-b1 in metastatic breast cancer MDA-MB-435 cells showed a 35% decrease in migration and a 55% decrease in invasion in vitro, with a 50% increase in proliferation and no effect on apoptosis (Moore et al. 2008). While in vivo analysis indicated a 90% decrease in the number of mice bearing macroscopic lung metastases, primary tumors did not show any difference in the growth kinetics when compared with the parental MDA-MB-435 cells. Interestingly, analysis of TGF-b signaling pathways in the clonal derivatives showed a decrease in Smad2 activation and an increase in Akt/protein kinase B and ERK activation and decreased TGF-b RI and TGF-b RII expression in TGF-b1 silenced cells. These findings suggest that inhibition of TGF-b1 ligand may act as a negative feedback loop to disrupt the function of all TGF-b isoforms. Furthermore, therapies targeting the TGF-b signaling pathway may be more effective in latestage disease to prevent organ metastasis but not primary tumor formation and may be combined with other tumor-targeted therapies normally limited by increased circulating TGF-b levels. Clonal populations of 4NQO-induced rat malignant oral keratinocytes were also examined for metastatic capacity following orthotopic transplantation into athymic mice for effects on TGF-b. Polygonal and spindle cells formed well-differentiated squamous cell carcinomas and undifferentiated spindle cell tumors in almost 100% of animals at the site of inoculation (Davies et al. 1999). Transplantation of cells of either cell type at high cell density resulted in approximately 50% of animals forming pulmonary metastases, while inoculation of low density differentiated polygonal cells resulted in the formation of significantly fewer pulmonary metastases than the undifferentiated spindle cells. A single well-differentiated clone of polygonal cells and three of four of the undifferentiated spindle cell lines produced comparable levels of TGF-b1. One undifferentiated spindle cell line expressed significantly more TGF-b1 and following transplantation orthotopically, fewer animals formed pulmonary metastases despite the formation of primary tumors in almost all grafted animals, suggesting that TGF-b1 can act as a tumor suppressor in this cell type. The clones of polygonal cells were markedly inhibited and the spindle cells were only partially inhibited by exogenous TGF-b1. Both cell types expressed high-affinity TGF-b cell surface receptors. The results suggest that differentiated rat malignant oral keratinocytes are less aggressive and have a decreased potential to metastasise than their undifferentiated spindle cell counterparts and may be attributable, in part, to a change in TGF-b receptor profile leading to the partial loss of response to exogenous TGF-b1.
664
S.B. Jakowlew
The effects of TGF-b1 on the proliferation and experimental pulmonary metastasis of MCS-1 cells, undifferentiated type cloned tumor cells established from a mesenchymal chondrosarcoma which spontaneously occurred in the soft tissue of a female Chinese hamster were also examined. Treatment of MCS-1 cells with TGF-b1 was tested for in vitro growth using the 3-(4,5-dimethylthiazol-2-yl)2,5-diphenyltetrazolium bromide (MTT) method, in vivo growth by subcutaneous inoculation into athymic nude mice and experimental pulmonary metastasis by injection into the tail vein of athymic nude mice. TGF-b1 significantly inhibited in vitro growth of MCS-1, depending on concentration, and also experimental metastasis (Fujisawa et al. 1999). TGF-b1, however, was ineffective for in vivo subcutaneous growth of MCS-1 cells. These results indicate that TGF-b1 might be an inhibitor of metastasis of mesenchymal chondrosarcomas including other types of non-epithelial cartilage or bone formation tumors.
Conclusions The acquisition of metastatic characteristics that are associated with lung cancer often predicts significant cancer-associated mortality and morbidity as lung tumorigenesis progresses. Although drug treatments and therapies are being developed that target primary lung tumors, additional new drugs and therapies are required to combat the challenges of metastatic tumors that occur in the lung as well as in other organs. Because there is ample clinical and experimental evidence that shows that TGF-b and its downstream signaling components have an important role in the metastatic process, TGF-b has become an attractive candidate for use in anti-metastasis treatments and therapies. Several small molecule inhibitors targeting the TGF-b receptor kinases have been developed recently along with specific neutralizing antibodies and nucleic acid-based therapies to inhibit various components of the TGF-b signaling pathway. Some of these drug treatments and therapies have shown promise in mouse model and cell culture systems that include lung. Developing these drug therapies has not been easy, partially because of the complex role of TGF-b in tumor progression to metastasis. This is because TGF-b can act as a tumor suppressor or a tumor promoter depending on the tumor type and the stage of tumor progression. Although many investigators have already devoted several years to understanding the role of TGF-b in tumorigenesis, there is still a lot of research that needs to be performed in order to develop new strategies that can utilize the complex nature of TGF-b to make significant inroads into the prevention and elimination of lung cancer, carcinogenesis, and metastasis. Acknowledgments It is not possible to include every important contribution that has been made to our understanding of TGF-b in lung cancer and metastasis. Apologies are extended to those investigators whose contributions could not be included.
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
665
References Allinen M, Beroukhim R, Cai L et al (2004) Molecular characterization of the tumor microenvironment in breast cancer. Cancer Cell 6:17–32 Anbazhagan R, Bornman DM, Johnston JC et al (1999) The S387Y mutations of the transforming growth factor-b receptor type I gene is uncommon in metastases of breast cancer and other common types of adenocarcinoma. Cancer Res 59:3363–3364 Arima Y, Inoue Y, Shibata T et al (2008) Rb depletion results in deregulation of E-cadherin and induction of cellular phenotypic changes that are characteristic of the epithelial-to-mesenchymal transition. Cancer Res 68:5104–5112 Attisano L, Wrana JL (2002) Signal transduction by the TGF-b superfamily. Science 296:1646–1647 Avis I, Mathias A, Unsworth EJ et al (1995) Analysis of small cell lung cancer cell growth inhibition by 13-cis-retinoic acid: importance of bioavailability. Cell Growth Differ 6:485–492 Azuma H, Ehata S, Miyazaki H et al (2005) Effect of Smad7 expression on metastasis of mouse mammary carcinoma JygMC(A) cells. J Natl Cancer Inst 97:1734–1746 Bakin AV, Safina A, Rinehart C et al (2004) A critical role of tropomyosins in TGF-b regulation of the actin cytoskeleton and cell motility in epithelial cells. Mol Biol Cell 15:4682–4694 Bandyopadhyay A, Zhu Y, Cibull ML et al (1999) A soluble transforming growth factor-b type III receptor suppresses tumorigenicity and metastasis of human breast cancer MDA-MB-231 cells. Cancer Res 59:5041–5046 Bandyopadhyay A, Agyin JK, Wang L et al (2006) Inhibition of pulmonary and skeletal metastasis by a transforming growth factor-beta type I receptor kinase inhibitor. Cancer Res 66:6714–6721 Beer DG, Kardia SL, Huang CC et al (2002) Gene-expression profiles predict survival of patients with lung adenocarcinoma. Nat Med 8:816–824 Bhattacharjee A, Richards WG, Staunton J et al (2001) Classification of human lung carcinomas by mRNA expression profiling reveals distinct adenocarcinoma subclasses. Proc Natl Acad Sci U S A 98:13790–13795 Bierie B, Stover DG, Abel TW et al (2008) Transforming growth factor-b regulates mammary carcinoma cell survival and interaction with the adjacent microenvironment. Cancer Res 68:1809–1819 Biswas S, Guix M, Rinehart C et al (2007) Inhibition of TGF-b with neutralizing antibodies prevents radiation-induced acceleration of metastatic cancer progression. J Clin Invest 117:1305–1313 Borczuk AC, Gorenstein L, Walter KL et al (2003) Non-small-cell lung cancer molecular signatures recapitulate lung developmental pathways. Am J Pathol 163:1949–1960 Borczuk AC, Kim HK, Yegen HA et al (2005) Lung adenocarcinoma global profiling identifies type II transforming growth factor-b receptor as a repressor of invasiveness. Am J Respir Crit Care Med 172:729–737 Brambilla E, Travis WD, Colby TV et al (2001) The new World Health Organization classification of lung tumours. Eur Respir J 18:1059–1068 Canalis E, Economides AN, Gazzerro E (2003) Bone morphogenetic proteins, their antagonists, and the skeleton. Endocr Rev 24:218–235 Chakrabarti R, Subramaniam V, Abdalla S et al (2009) Tranilast inhibits the growth and metastasis of mammary carcinoma. Anticancer Drugs 20:334–345 Chakravarthy D, Green AR, Green VL et al (1999) Expression and secretion of TGF-b isoforms and expression of TGF-b receptors I, II and III in normal and neoplastic human breast. Int J Oncol 15:187–194 Chen T, Carter D, Garrigue-Antar L, Reiss M (1998) Transforming growth factor b type I receptor kinase mutant associated with metastatic breast cancer. Cancer Res 58:4805–4810
666
S.B. Jakowlew
Clines GA, Guise TA (2005) Hypercalcaemia of malignancy and basic research on mechanisms responsible for osteolytic and osteoblastic metastasis to bone. Endocr Relat Cancer 12:549–583 Dano K, Behrendt N, Brunner N, et al (1994) The urokinase receptor. Protein structure and role in plasminogen activation and cancer invasion. Fibrinolysis 8:189–203 Davies M, Paterson IC, Stone A et al (1999) Loss of differentiation of 4NQO-induced rat malignant oral keratinocytes correlates with metastatic dissemination and is associated with a reduced cellular response to TGF-b1 and an altered receptor profile. J Oral Pathol Med 28:397–405 Deng X, Tannehill-Gregg SH, Nadella MV et al (2007) Parathyroid hormone-related protein and ezrin are up-regulated in human lung cancer bone metastases. Clin Exp Metastasis 24:107–119 Derynck R, Zhang YE (2003) Smad-independent and smad-dependent pathways in TGF-b family signaling. Nature 425:577–584 Derynck R, Akhurst RJ, Balmain A (2001) TGF-b signaling in cancer: a double-edged sword. Trends Cell Biol 29:117–129 Derynk R, Akhurst RJ, Balmain A (2001) TGF-b signaling in tumor suppression and cancer progression. Nat Genet 29:117–129 De Souza AT, Hankins GR, Washington MK (1992) M6P/IGF2R gene is mutated in human hepatocellular carcinomas with loss of heterozygosity. Nat Genet 11:447–449 Du G, Jin L, Han X et al (2009) Naringenin: a potential immunomodulator for inhibiting lung fibrosis and metastasis. Cancer Res 69:3205–3212 Esplugues E, Vega-Ramos J, Cartoixà D et al (2005) Induction of tumor NK-cell immunity by anti-CD69 antibody therapy. Blood 105:4399–4406 Fidler IJ (2003) The pathogenesis of cancer metastasis: the ‘seed and soil’ hypothesis revisited. Nat Rev Cancer 3:453–458 Filyak Y, Filyak O, Souchelnytskyi S, Stoika R (2008) Doxorubicin inhibits TGF-b signaling in human lung carcinoma A549 cells. Eur J Pharmacol 590:67–73 Forrester E, Chytil A, Bierie B et al (2005) Effect of conditional knockout of the TGF-b RII gene in mammary epithelia on mammary gland development and polyomavirus middle T antigen induced tumor formation and metastasis. Cancer Res 65:2296–2302 Franzen P, ten Dijke P, Ichijo H et al (1993) Cloning of a receptor that forms a heteromeric complex with the TGF-b type II receptor. Cell 75:681–692 Fujisawa N, Sato NL, Motoyama TI (1999) Inhibitory effects of transforming growth factor-b1 pretreatment on experimental pulmonary metastasis of MCS-1 Chinese hamster mesenchymal chondrosarcoma cells. Tohoku J Exp Med 187:203–213 Furta K, Misao S, Takahashi K et al (1999) Gene mutation of transforming growth factor b1 type II receptor in hepatocellular carcinoma. Int J Cancer 81:851–853 Gazdar AF, Miyajima K, Reddy J et al (2004) Molecular targets for cancer therapy and prevention. Chest 125:97S–101S Geiser AG, Burmester JK, Webbink R et al (1992) Inhibition of growth by transforming growth factor-ß following fusion of two non-responsive human carcinoma cell lines. J Biol Chem 267:2588–2593 Giles ED, Singh G (2003) Role of insulin-like growth factor binding proteins (IGFBPs) in breast cancer proliferation and metastasis. Clin Exp Metastasis 20:481–487 Glick A (2004) TGF-b1, back to the future. Cancer Biol Ther 3:276–283 Grady WM, Rajput A, Myerhof L et al (1998) Mutation of the type II transforming growth factorb receptor is coincident with the transformation of human colon adenomas to malignant carcinomas. Cancer Res 58:3101–3104 Guo Y, Jacobs SC, Kyprianou N (1997) Down-regulation of protein and mRNA expression for transforming growth factor-b (TGF-b1) type I and type II receptors in human prostate cancer. Int J Cancer 71:573–579 Hahn SA, Schutte M, Hoque AT et al (1996) DPC4, a candidate tumor suppressor gene at human chromosome 18q21.1. Science 271:350–353
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
667
Hamada J, Omatsu T, Okada F et al (2001) Overexpression of homeobox gene HOXD3 induces coordinate expression of metastasis-related genes in human lung cancer cells. Int J Cancer 93:516–525 Hanyu A, Ishidou Y, Ebisawa T et al (2001) The N domain of Smad7 is essential for specific inhibition of transforming growth factor-b signaling. J Cell Biol 155:1017–1027 Hayashi H, Abdollah S, Qiu Y et al (1997) The MAD-related protein Smad7 associates with the TGFbeta receptor and functions as an antagonist of TGF-beta signaling. Cell 89:1165–1173 Heldin CH, Miyazono K, ten Dijke P (1997) TGF-b signaling from cell membrane to nucleus through SMAD proteins. Nature 390:465–471 Heldin CH, Landström M, Moustakas A (2009) Mechanism of TGF-b signaling to growth arrest, apoptosis, and epithelial–mesenchymal transition. Curr Opin Cell Biol 21:166–176 Holmgren L, O’Reilly MS, Folkman J (1995) Dormancy of micrometastases: balanced proliferation and apoptosis in the presence of angiogenesis suppression. Nat Med 1:149–153 Huntley SP, Davies M, Matthews JB et al (2004) Attenuated type II TGF-b receptor signalling in human malignant oral keratinocytes induces a less differentiated and more aggressive phenotype that is associated with metastatic dissemination. Int J Cancer 110:170–176 Hynes RO (1992) Integrins: versatility, modulation, and signaling in cell adhesion. Cell 69:11–25 Jakowlew SB, Mathias A, Chung P et al (1995) Expression of transforming growth factor-b ligand and receptor messenger RNAs in lung cancer cell lines. Cell Growth Differ 6:465–476 Jakowlew SB, Mariano JM, You L (1997a) Effects of transforming growth factor-b1 and phorbol ester on PAI-1 and PA genes in human lung cells. Growth Factors 14:177–197 Jakowlew SB, Mariano JM, You L et al (1997b) Differential regulation of protease and extracellular matrix protein expression by transforming growth factor-beta 1 in non-small cell lung cancer cells and normal human bronchial epithelial cells. Biochim Biophys Acta 1353:157–170 Jakowlew SB, Moody TW, Mariano JM (1997c) Transforming growth factor-b receptors in human cancer cell lines: analysis of transcript, protein and proliferation. Anticancer Res 17:1849–1860 Jakowlew SB, Moody TW, You L et al (1998a) Reduction in transforming growth factor-b type II receptor in mouse lung carcinogenesis. Mol Carcinog 22:46–56 Jakowlew SB, Moody TW, You L et al (1998b) Transforming growth factor-beta expression in mouse lung carcinogenesis. Exp Lung Res 24:579–593 Jakowlew SB, Zakowicz H, Moody TW (2000) Retinoic acid down-regulates VPAC(1) receptors and TGF-b3 but up-regulates TGF-b2 in lung cancer cells. Peptides 21:1831–1837 Jemal A, Murray T, Ward E et al (2005) Cancer statistics, 2005. CA Cancer J Clin 55:10–30 Jemel A, Siegel R, Ward E et al (2009) Cancer statistics, 2009. CA Cancer J Clin 59:225–249 Jiang J, Sugimoto Y, Liu S et al (2004) The inhibitory effects of gossypol on human prostate cancer cells-PC3 are associated with transforming growth factor beta1 (TGFb) signal transduction pathway. Anticancer Res 24:91–100 Kahari VM, Saarialho-Kere U (1999) Matrix metalloproteinases and their inhibitors in tumour growth and invasion. Ann Med 31:34–45 Kang Y, Mariano JM, Angdisen J, Moody TW, Diwan BA, Wakefield LM, Jakowlew SB (2000) Enhanced tumorigenesis and reduced transforming growth factor-b type II receptor in lung tumors from mice with reduced gene dosage of transforming growth factor-b1. Mol Carcinog 29:112–126 Kang Y, Prentice MA, Mariano JM et al (2000) Transforming growth factor-b1 and its receptors in human lung cancer and mouse lung carcinogenesis. Exp Lung Res 26:685–707 Kang Y, Ozbun LL, Angdisen J et al (2002) Altered expression of G1/S regulatory genes occurs early and frequently in lung carcinogenesis in transforming growth factor-b1 heterozygous mice. Carcinogenesis 23:1217–1227 Kang Y, Siegel PM, Shu W et al (2003) A multigenic program mediating breast cancer metastasis to bone. Cancer Cell 3:537–549 Kang Y, Hebron H, Ozbun L et al (2004) Nkx2.1 transcription factor in lung cells and a transforming growth factor-b1 heterozygous mouse model of lung carcinogenesis. Mol Carcinog 40:212–231
668
S.B. Jakowlew
Karnoub AE, Dash AB, Vo AP et al (2007) Mesenchymal stem cells within tumour stroma promote breast cancer metastasis. Nature 449:557–563 Kobayashi H, Fukuda Y, Yoshida R et al (2004) Suppressing effects of dietary supplementation of soybean trypsin inhibitor on spontaneous, experimental and peritoneal disseminated metastasis in mouse model. Int J Cancer 112:519–524 Lahsnig C, Mikula M, Petz M et al (2009) ILEI requires oncogenic Ras for the epithelial to mesenchymal transition of hepatocytes and liver carcinoma progression. Oncogene 28:638–650 Laiho M, Weiss FMB, Massagué J (1990) Concomitant loss of transforming growth factor-b receptors type I and type II in TGF-b-resistant cell mutants implicates both receptor types in signal transduction. J Biol Chem 265:18518–18524 La Porta CA, Comolli R (2000) Different levels of TGF-b, IL-10, IFNgamma and gelatinase A occur in experimental white and black metastases induced by bryostatin 1 or by phorbol estertreated BL6T murine melanoma cells. Clin Exp Metastasis 18:361–369 Lee YH, Albig AR, Regner M et al (2008) Fibulin-5 initiates epithelial-mesenchymal transition (EMT) and enhances EMT induced by TGF-b in mammary epithelial cells via a MMPdependent mechanism. Carcinogenesis 29:2243–2251 Leivonen SK, Kähäri YM (2007) Transforming growth factor-b signaling in cancer invasion and metastasis. Int J Cancer 121:2119–2124 Le Scolan E, Zhu Q, Wang L et al (2008) Transforming growth factor-b suppresses the ability of Ski to inhibit tumor metastasis by inducing its degradation. Cancer Res 68:3277–3285 Li C, Wang J, Wilson PB et al (1998). Role of transforming growth factor-b3 in lymphatic metastasis in breast cancer. Int J Cancer 79:455–459 Liebow AA (1960) Bronchioalveolar carcinoma. Adv Intern Med 10:329–335 Lim DS, Kim JH, Lee DS et al (2007) DC immunotherapy is highly effective for the inhibition of tumor metastasis or recurrence, although it is not efficient for the eradication of established solid tumors. Cancer Immunol Immunother 56:1817–1829 Lin EY, Nguyen AV, Russell RG, Pollard JW (2001) Colony-stimulating factor 1 promotes progression of mammary tumors to malignancy. J Exp Med 193:727–740 Lin HY, Wang X-F, Ng-Eaton E et al (1992) Expression cloning of the TGF-b type II receptor, a functional transmembrane serine/threonine kinase. Cell 68:775–785 Liotta LA (1986) Tumor invasion and metastases – role of the extracellular matrix: Rhoads memorial award lecture. Cancer Res 46:1–7 Liotta LA, Kohn EC (2001) The microenvironment of the tumour-host interface. Nature 411:375–379 Luan FL, Ding R, Sharma VK et al (2003) Rapamycin is an effective inhibitor of human renal cancer metastasis. Kidney Int 63:917–926 Maluccio M, Sharma V, Lagman M et al (2003) Tacrolimus enhances transforming growth factorb1 expression and promotes tumor progression. Transplantation 76:597–602 Mani SA, Yang J, Brooks M et al (2007) Mesenchyme Forkhead 1 (FOXC2) plays a key role in metastasis and is associated with aggressive basal-like breast cancers. Proc Natl Acad Sci U S A 104:10069–10074 Mareel M, Madani I (2006) Tumour-associated host cells participating at invasion and metastasis: targets for therapy? Acta Chir Belg 106:635–40 Markowitz S, Wang J, Myerhoff L (1995) Inactivation of the type II TGF-b receptor in colon cancer cells with microsatellite instability. Science 268:1336–1338 Massagué J (2000) How cells read TGF-b signals. Nat Rev Mol Cell Biol 1:169–178 Massagué J (2008) TGF-b in cancer. Cell 134:215–230 Massagué J, Attisano L, Wrana J (1994) The TGF-b family and its composite receptors. Trends Cell Biol 4:172–178 Matsushita M, Matsuzaki K, Date M et al (1999) Down-regulation of TGF-b receptors in human colorectal cancer: implications for cancer development. Br J Cancer 80:194–205 McKenna IM, Ramakrishna G, Diwan BA et al (2001) Heterozygous inactivation of TGF-b1 increases the susceptibility to chemically induced mouse lung tumorigenesis independently of mutational activation of K-ras. Toxicol Lett 123:151–158
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
669
McSherry EA, Donatello S, Hopkins AM, McDonnell S (2007) Molecular basis of invasion in breast cancer. Cell Mol Life Sci 64:3201–3218 Minn AJ, Gupta GP, Padua D et al (2007) Lung metastasis genes couple breast tumor size and metastatic spread. Proc Natl Acad Sci U S A 104:6740–6745 Miyazaki YJ, Hamada J, Tada M et al (2002) HOXD3 enhances motility and invasiveness through the TGF-b-dependent and -independent pathways in A549 cells. Oncogene 21:798–808 Miyazono K, Suzuki H, Imamura T (2003) Regulation of TGF-b signaling and its roles in progression of tumors. Cancer Sci 94:230–234 Mochizuki T, Miyazaki H, Hara T et al (2004) Roles for the MH2 domain of Smad7 in the specific inhibition of transforming growth factor-b superfamily signaling. J Biol Chem 279:31568–31574 Moody TW, Leyton J, Zakowicz H et al (2001) Indomethacin reduces lung adenoma number in A/J mice. Anticancer Res 21:1749–1755 Mooradian DL, McCarthy JB, Komanduri KV, Furcht LT (1992) Effects of transforming growth factor-beta 1 on human pulmonary adenocarcinoma cell adhesion, motility, and invasion in vitro. J Natl Cancer Inst 84:523–527 Moore LD, Isayeva T, Siegal GP, Ponnazhagan S (2008) Silencing of transforming growth factorb1 in situ by RNA interference for breast cancer: implications for proliferation and migration in vitro and metastasis in vivo. Clin Cancer Res 14:4961–4970 Muraoka-Cook RS, Kurokawa H, Koh Y et al (2004) Conditional overexpression of active transforming growth factor b1 in vivo accelerates metastases of transgenic mammary tumors. Cancer Res 64:9002–9011 Muraoka RS, Dumont N, Ritter CA et al (2002) Blockade of TGF-b inhibits mammary tumor cell viability, migration, and metastases. J Clin Invest 109:1551–1559 Nakao A, Afrakhte M, Mören A, et al (1997) Identification of Smad7, a TGF-b-inducible antagonist of TGF-b signaling. Nature 389:631–635 Olumi AF, Grossfeld GD, Hayward SW et al (1999) Carcinoma-associated fibroblasts direct tumor progression of initiated human prostatic epithelium. Cancer Res 59:5002–5011 Owen-Schaub LB, van Golen KL, Hill LL et al (1986) Fas and Fas ligand interactions suppress melanoma lung metastasis. J Exp Med 188:1717–1723 Padua D, Zhang X H-F, Wang Q et al (2008) TGFb primes breast tumors for lung metastasis seeding through angiopoietin-like 4. Cell 133:66–77 Paget S (1889) The distribution of secondary growths in cancer of the breast. Lancet 1:571–573 Paik S, Shak S, Tang G et al (2004) A multigene assay to predict recurrence of tamoxifen-treated, node-negative breast cancer. N Engl J Med 351:2817–2826 Pandey J, Umphress SM, Kang Y et al (2007) Modulation of tumor induction and progression of oncogenic K-ras-positive tumors in the presence of TGF-b1 haploinsufficiency. Carcinogenesis 28:2589–2596 Park K, Kim S-J, Bang Y-J et al (1994) Genetic changes in the transforming b type II receptor gene in human gastric cells: correlation with sensitivity to TGF-b. Proc Natl Acad Sci U S A 81:8772–8776 Park SH (2005) Fine tuning and cross-talking of TGF-b signal by inhibitory Smads. J Biochem Mol Biol 38:9–16 Pepper MS, Sappino AP, Stocklin R et al (1993) Upregulation of urokinase receptor expression on migrating endothelial cells. J Cell Biol 122:673–684 Roberts AB, Wakefield LM (2003) The two faces of transforming growth factor-b in carcinogenesis. Proc Natl Acad Sci U S A 100:8621–8623 Röijer E, Mören A, ten Dijke P, Stenman G (1998) Assignment of the Smad7 gene (MADH7) to human chromosome 18q21.1 by fluorescence in situ hybridization. Cytogenet Cell Genet 81:189–190. Sawyer JS, Beight DW, Britt KS et al (2004) Synthesis and activity of new aryl- and heteroarylsubstituted 5,6-dihydro-4H-pyrrolo[1,2-b]pyrazole inhibitors of the transforming growth factor-beta type I receptor kinase domain. Bioorg Med Chem Lett 14:3581–3584
670
S.B. Jakowlew
Schlunegger MP, Grutter MG (1992) An unusual feature revealed by the crystal structure at 2.2 Å resolution of human transforming growth factor-b2. Nature 358:430–434 Shah AH, Tabayoyong WB, Kundu SD et al (2002) Suppression of tumor metastasis by blockade of transforming growth factor b signaling in bone marrow cells through a retroviral-mediated gene therapy in mice. Cancer Res 62:7135–7138 Shah L, Walter KL, Borczuk AC et al (2004) Expression of syndecan-1 and expression of epidermal growth factor receptor are associated with survival in patients with non-small cell lung carcinoma. Cancer 101:1632–1638 Sheppard D (2005) Integrin-mediated activation of latent transforming growth factor b. Cancer Metastasis Rev 24:395–402 Shintani K, Matsumine A, Kusuzaki K et al (2008) Decorin suppresses lung metastases of murine osteosarcoma. Oncol Rep 19:1533–1539 Shintani Y, Maeda M, Chaika N et al (2008) Collagen I promotes epithelial-to-mesenchymal transition in lung cancer cells via transforming growth factor-b signaling. Am J Respir Cell Mol Biol 38:95–104 Shi Y, Massagué J (2003) Mechanisms of TGF-b signaling from cell membrane to the nucleus. Cell 113:685–700 Siegel PM, Shu W, Cardiff RD et al (2003) Transforming growth factor b signaling impairs Neuinduced mammary tumorigenesis while promoting pulmonary metastasis. Proc Natl Acad Sci U S A 100:8430–8435 Skarki A (1993) Analysis of long-term survivors with small-cell lung cancer. Chest 103:440S–444S Skarkin A (1994) Respiratory tract and head and neck cancer. In: Rubinstein E, Federman D (eds) Scientific American. Scientific American Medicine, New York, pp 1–35 Steiner MS, Barrack ER (1992) Transforming growth factor-beta 1 overproduction in prostate cancer: effects on growth in vivo and in vitro. Mol Endocrinol 6:15–25 Stetler-Stevenson WG, Aznavoorian S, Liotta LA (1993) Tumor cell interactions with the extracellular matrix during invasion and metastasis. Annu Rev Cell Biol 9:541–573 Sun L, Wu G, Wilson JK et al (1994) Expression of transforming growth factor b type II receptor leads to reduced malignancy in human breast cancer MCF-7 cells. J Biol Chem 269:26449–26455 Tan AR, Alexe G, Reiss M (2009) Transforming growth factor-b signaling: emerging stem cell target in metastatic breast cancer? Breast Cancer Res Treat 115:453–495 Tang B, Böttinger EP, Jakowlew SB et al (1998) Transforming growth factor-b1 is a new form of tumor suppressor with true haploid insufficiency. Nat Med 4:802–807 Taniguchi Y, Komatsu N, Moriuchi T (1995) Overexpression of the HOX4A (HOXD3) homeobox gene in human erythroleukemia HEL cells results in altered adhesive properties. Blood 85:2786–2794 Thiery JP (2003) Epithelial–mesenchymal transition in development and pathologies. Curr Opin Cell Biol 15:740–746 Thiery JP, Sleeman JP (2006) Complex networks orchestrate epithelial-mesenchymal transitions. Nat Rev Mol Cell Biol 7:131–142 Tian F, DaCosta Byfield S, Parks WT et al (2003) Reduction in Smad2/3 signaling enhances tumorigenesis but suppresses metastasis of breast cancer cell lines. Cancer Res 63:8284–8292 Tian F, Byfield SD, Parks WT et al (2004) Smad-binding defective mutant of transforming growth factor b type I receptor enhances tumorigenesis but suppresses metastasis of breast cancer cell lines. Cancer Res 64:4523–5230 Tobin SW, Douville K, Benbow U et al (2002) Consequences of altered TGF-b expression and responsiveness in breast cancer: evidence for autocrine and paracrine effects. Oncogene 21:108–118 Tojo M, Hamashima Y, Hanyu A et al (2005) The ALK-5 inhibitor A-83-01 inhibits Smad signaling and epithelial-to-mesenchymal transition by transforming growth factor-b. Cancer Sci 96:791–800
28 Transforming Growth Factor-b in Lung Cancer, Carcinogenesis, and Metastasis
671
Tsuji T, Ibaragi S, Shima K et al (2008) Epithelial-mesenchymal transition induced by growth suppressor p12CDK2-AP1 promotes tumor cell local invasion but suppresses distant colony growth. Cancer Res 68:10377–120386 Turco A, Coppa A, Aloe S et al (1999) Overexpression of transforming growth factor beta-type II receptor reduces tumorigenicity and metastatic potential of K-ras-transformed thyroid cells. Int J Cancer 80:85–91 Ueki N, Ohkawa T, Yokoyama Y et al (1993) Potentiation of metastatic capacity by transforming growth factor-beta 1 gene transfection. Jpn J Cancer Res 84:589–593 Veenendaal LM, Kranenburg O, Smakman N et al (2008) Differential Notch and TGF-b signaling in primary colorectal tumors and their corresponding metastases. Cell Oncol 30:1–11 Vicent S, Luis-Ravelo D, Antón I et al (2008) A novel lung cancer signature mediates metastatic bone colonization by a dual mechanism. Cancer Res 68:2275–2280 Wakefield LM, Roberts AB (2002) TGF-b signaling: positive and negative effects on tumorigenesis. Curr Opin Genet Dev 123:22–29 Wang P, Chen Z, Meng Z et al (2009) Dual role of Ski in pancreatic cancer cells: tumor-promoting versus metastasis-suppressive function. Carcinogenesis 30:1497–1506 Welch DR, Fabra A, Nakajima M (1990) Transforming growth factor b stimulates mammary adenocarcinoma cell invasion and metastatic potential. Proc Natl Acad Sci U S A 87: 7678–7682 Wheelock MJ, Johnson KR (2003) Cadherins as modulators of cellular phenotype. Annu Rev Cell Dev Biol 19:207–235 Woszczyk D, Gola J, Jurzak M et al (2004) Expression of TGF-b1 genes and their receptor types I, II, and III in low- and high-grade malignancy non-Hodgkin’s lymphomas. Med Sci Monit 75(10):CR33–37 Yang J, Weinberg RA (2008) Epithelial-mesenchymal transition: at the crossroads of development and tumor metastasis. Dev Cell 14:818–829 Yang J, Mani SA, Donaher JL (2004) Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 117:927–939 Yang RS, Wu CT, Lin KH et al (1998) Relation between histological intensity of transforming growth factor-beta isoforms in human osteosarcoma and the rate of lung metastasis. Tohoku J Exp Med 184:133–142 You L, Jakowlew SB (1997) Identification of early growth response gene-1 (Egr-1) as a phorbol myristate acetate-induced gene in lung cancer cells by differential mRNA display. Am J Respir Cell Mol Biol 17:617–624 Zhang F, Pao W, Umphress SM et al (2003) Serum levels of surfactant protein D are increased in mice with lung tumors. Cancer Res 63:5889–5894 Zhang Q, Yang X, Pins M et al (2005) Adoptive transfer of tumor-reactive transforming growth factor-b-insensitive CD8+ T cells: eradication of autologous mouse prostate cancer. Cancer Res 65:1761–1769 Zhang Q, Yang XJ, Kundu SD et al (2006) Blockade of transforming growth factor-b signaling in tumor-reactive CD8(+) T cells activates the antitumor immune response cycle. Mol Cancer Ther 5:1733–1743 Zheng Q, Safina A, Bakin AV (2008) Role of high-molecular weight tropomyosins in TGF-bmediated control of cell motility. Int J Cancer 122:78–90 Zhu Q, Krakowski AR, Dunham EE et al (2007) Dual role of SnoN in mammalian tumorigenesis. Mol Cell Biol 27:324–339 Zohrabian VM, Nandu H, Gulati N et al (2007) Gene expression profiling of metastatic brain cancer. Oncol Rep 18: 321–328
Chapter 29
Cooperative Interactions Between Integrins and Growth Factor Signaling in Pathological Angiogenesis Jennifer Roth, Eric Tweedie, and Peter C. Brooks
Abstract As with most complex biological processes, angiogenesis requires the integration of a number of molecular signaling networks to coordinate multiple cues from both the extracellular tissue microenvironment as well as the cell’s interior. Thus an important area of angiogenesis investigation involves understanding the mechanisms that facilitate cooperation between multiple receptor–ligand signaling pathways. Two crucial networks that play active role in angiogenesis include growth factor/growth factor receptors and extracellular matrix/integrin receptor signaling systems. Emerging evidence suggests that these two important signaling systems depend in large part on each other, and function cooperatively to control new blood vessel development. Given the tissue-specific variations in the expression of components within each of these systems, significant challenges exist in order to exploit these signaling pathways for clinical intervention. A more detailed understanding of how the molecular components of these two signaling systems communicate with each other to direct and coordinate downstream effector functions may lead to optimized anti-angiogenic strategies to control malignant tumor progression. In this regard, we will discuss the multiple ways by which growth factor and integrin signaling pathways function cooperatively to regulate pathological angiogenesis within the context of the tissue microenvironment.
Introduction Both physiological and pathological blood vessel formation occurs within a complex tissue microenvironment composed of interconnected networks of extracellular matrix (ECM) molecules and a variety of different cell types. It has been appreciated for decades that cellular behavior is regulated by a highly orchestrated P.C. Brooks (*) Maine Medical Center Research Institute, Center for Molecular Medicine, 81 Research Drive, Scarborough, ME, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_29, © Springer Science+Business Media, LLC 2010
673
674
J. Roth et al.
system of communication links between different cell types, secreted growth factors, cytokines, chemokines, and a variety of insoluble solid-state ECM molecules that constitute the tissue microenvironment (Contois et al. 2009; Rapisarda and Melillo 2009). However, how cells are able to assimilate and interpret this vast amount of information in a meaningful way to regulate processes such as new blood vessel development or angiogenesis remains in many respects unknown. Given the diverse interwoven signaling networks that are required to complete the angiogenic cascade, it is critical that we begin to understand in more detail the impact that disrupting one communication pathway may have on another, especially as it relates to development of anti-angiogenic strategies for controlling tumor progression. This concept is of particular importance given the highly integrated nature of feed back systems that exist between signaling pathways such as PI3K/Akt and MAPK/Erk cascades which have been implicated in angiogenesis (Huang and Fraenkel 2009; Wagner and Nebreda 2009; Engelman 2009). Viewed from an integrated systems perspective, disruption of one signaling node or hub may have unexpected consequences on a second distinct pathway. Thus, studying angiogenesis from a more global network systems approach may offer unique molecular insight into the coordinated events necessary for new vessel formation. This molecular insight may facilitate the development of more effective therapeutic strategies to control pathological blood vessel growth. As the complexities of angiogenesis continue to be unraveled, new studies are rapidly adding to the growing list of cell types that play active roles in controlling neovascularization. In the past, the majority of studies on angiogenic mechanisms focused largely on signaling within endothelial cells. Now, evidences indicates that pericytes, smooth muscle cells, fibroblasts, inflammatory infiltrates including monocytes, macrophages, mast cells, as well as platelets, bone marrow-derived progenitor cells, and malignant tumor cells all contribute to new blood vessel formation (Murdoch et al. 2008; Colmone and Sipkins 2008; Greenberg et al. 2008; Klement et al. 2008). These distinct cell types have been suggested to contribute to angiogenesis by mechanisms as diverse as physical incorporation into the vascular structure, to modulating the local concentrations of secreted pro- and anti-angiogenic factors, and to mechanically altering the local insoluble matrix surrounding vessels (Murdoch et al. 2008; Colmone and Sipkins 2008; Greenberg et al. 2008; Klement et al. 2008). Intriguing new data now suggests that cell-mediated biomechanical tissue contraction may contribute to the movement of pre-existing vessels into areas of hypoxia, thereby enhancing local tissue perfusion (Kilarski et al. 2009). Given the increasing numbers of distinct cell types and mechanisms thought to regulate angiogenesis, coupled with the diversity of soluble factors, understanding how a specific tissue microenvironment modulates a cells capacity to produce as well as response to both pro- and anti-angiogenic molecules will likely lead to more effective therapeutic strategies to control abnormal vessel growth. It is well accepted that the ability of a cell to respond to external cues such as pro- and anti-angiogenic regulatory molecules depends on bi-directional communication with secreted factors, insoluble ECM elements, and between different cell types. Integrins represent a major family of cell surface receptors known to play
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
675
important roles in mediating cell-matrix communication. This multifunctional family of receptors not only binds to insoluble ECM, but also bind a variety of soluble molecules (Contois et al. 2009; Avraamides et al. 2008). Given their capacity to sense and transmit signals derived from biochemical and mechanical alteration originating from inside and outside the cell, integrins are well suited to function as microenvironmental sensors, which integrate diverse signaling cascades. Thus, integrins may serve as “functional hubs” in the interconnected network of angiogenesis signaling (Contois et al.2009). Many growth factors and cytokines are known to regulate angiogenesis. Thus, distinct molecular mechanisms have evolved to allow cells to sense and utilize the diversity of soluble regulatory factors in specific ways. While it would be well beyond the scope of this review to discuss all the growth factors and cytokines thought to regulate angiogenesis, we will highlight a number of examples to illustrate the complexity by which these molecules contribute to new vessel growth. Given the diversity of soluble factors that impact vessel formation, it is becoming increasingly clear that in order to gain more complete understanding of how growth factor signaling impacts angiogenesis, a greater appreciation of the roles that integrin-mediated sensing of the surrounding microenvironment has on modulating angiogenic growth factor signaling is needed. Therefore, we will focus our discussion on the multitude of ways by which integrins, growth factors, and their receptors function cooperatively to regulate angiogenesis (Fig. 29.1).
Blood Vessel Formation While the formation of functional blood vessels can occur by several processes that share many common mechanisms, it is important to note that distinct differences also exist, especially between normal and pathological vessel formation (St Croix et al. 2000; Alavi et al. 2007; Hida and Klagsbrun 2005; Ghosh et al. 2008). These differences between normal vascular development and pathological angiogenesis may be due in part to the unique tissue microenvironments within which vessels form. For example, the local variability in stromal cell types that express a distinct repertoire of pro- and anti-angiogenic factors, combined with the composition and integrity of ECM molecules associated with blood vessels is likely to be quite different between normal and tumor vessels. Examples of the major ways by which blood vessel formation occurs include embryonic vasculogenesis, which involves the de novo formation of vessels from precursor cells, arteriogenesis, a process by which functional vessels form by a mechanism of activation, dilation and remodeling of small pre-existing nonfunctional vessels and finally angiogenesis, a process by which blood vessels form from pre-existing vessels (Hodivala-Dilke et al. 2003; Fischer et al. 2006). New evidence now suggests that enhanced vascularization of hypoxic microenvironments may also occur by a mechanism involving cell-mediated tissue contraction that physically translocates existing functional vessels into a hypoxic area where
676
J. Roth et al.
Cooperation Between Receptor / Ligand Networks Regulate New Blood Vessel Formation
GF
ECM
CRD CRD
FN FN FN FN III III JM
JM
TK
TK
Fibroblasts/Stromal Cells Inflammatory Cells Bone Marrow Progenitor Cells Tumor Cells Vascular Cells
Pathological Angiogenesis Fig. 29.1 Cooperation between receptor/ligand networks regulate new blood vessel formation. A wide variety of cell types within the tissue microenvironment are thought to contribute to the formation of new blood vessels. Two of the major receptor/ligand signaling networks that function cooperatively to control angiogenesis include the growth factor (GF)/growth factor receptor system and extracellular matrix (ECM)/integrin receptor system
few vessels previously existed (Kilarski et al. 2009). Given the diversity by which blood vessels form, we will limit the remainder of our discussion largely to angiogenesis and in particular pathological angiogenesis.
Pathological Angiogenesis Angiogenesis can be sub-categorized as to whether it occurs by intussuception or sprouting (Burri et al. 2004; Hlushchuk et al. 2008). Intussuception involves the splitting of pre-existing vessels by the formation of translumenal tissue bridges that
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
677
result in the formation of a diversity of branching patterns (Burri et al. 2004; Hlushchuk et al. 2008). This form of angiogenesis is thought to occur more rapidly than sprouting angiogenesis and may involve minimal if any cellular proliferation and extracellular matrix (ECM) remodeling (Burri et al. 2004; Hlushchuk et al. 2008). Sprouting angiogenesis on the other hand can be organized temporally into three general steps, including an initiation phase, an invasive phase and a maturation phase, and both integrins and growth factor signaling play vital roles in each of these steps. The sprouting form of angiogenesis involves growth factor-stimulated proliferation and invasion of tip cells from an existing vessel. Endothelial cells line up and organize into solid cords, which is followed by the process of canulization, which ultimately results in the formation of a functional lumen (Folkman 2007). Understanding the particular type of vascular development occurring may help direct the most appropriate therapeutic approach to control aberrant angiogenesis. For example, targeting a specific angiogenic growth factor to inhibit pathological blood vessel formation that is occurring predominately by intussuception or by biomechanical tissue contraction may result in only limited therapeutic benefit. In fact, new studies are emerging suggesting that blocking one growth factor signaling pathway such as VEGF may initiate an anti-angiogenic rescue program that allows alternative growth factor signaling cascades to substitute or compensate in part for the disrupted pathway (Ebos et al. 2009; Fernado et al. 2008; Puigvert et al. 2009). Thus, more molecular insight into these regulatory feed back loops will likely lead to more efficacious therapeutic intervention.
Integrins and Their ECM Ligands in Angiogenesis It is known that understanding both cell surface receptors and their cognate ligands in growth factor signaling systems is critical to appreciate their biological functions, and this same principal applies to integrins and their ligands. Integrins have been shown to bind a host of different regulatory molecules including: ECM protein, proteolytic enzymes, protease receptors, other cell adhesion molecules, and both growth factors and growth factor receptors (Contois et al. 2009; Avraamides et al. 2008). Given this diversity of binding partners, it is not surprising that a growing number of studies have demonstrated functional roles for integrins in modulating growth factor signaling. Integrins are transmembrane heterodimers composed of a and b chains. Distinct combinations of these separate gene products give rise to at least 24 different integrins with specific and sometimes overlapping ligandbinding capacity (Avraamides et al. 2008). While integrins lack catalytic domains traditionally associated with signaling receptors, they retain the capacity to promote signaling in a bi-directional manner by binding to a wide array of cytosolic adaptor molecules (Wegener and Campbell 2008). These transmembrane heterodimers associate with many cytoplasmic molecules to form large multicomponent signaling complexes that facilitate communication between the extracellular tissue microenvironment and the cell’s interior. Some recent estimates have suggested
678
J. Roth et al.
that integrin cytoplasmic domains may have the capacity to associate with over 100 different proteins, which facilitate their ability to integrate and define specific responses to numerous signaling networks (Wegener and Campbell 2008; Zaidel-Bar et al. 2007). The functional integration of integrins within growth factor signaling cascades may explain in part the rapidly expanding list of different integrins, which have been shown to regulate angiogenesis. For example, integrin such as a1b1, a2b1, a3b1, a4b1, a5b1, a6b1 as well as a6b4 have been implicated in the control of angiogenesis (Avraamides et al. 2008; Akalu et al. 2005). Experimental evidence also indicates members of the av family of integrins such as avb3, avb5, and avb8 in blood vessel formation (Avraamides et al. 2008; Akalu et al. 2005). Interestingly, early studies have linked the functional impact of integrins on angiogenesis to their interactions with insoluble ECM, resulting in enhanced endothelial cell adhesion, migration, and proliferation. However, new studies have now uncovered novel mechanisms by which integrins regulate angiogenesis by processes that are dependent on integrin binding to soluble pro- and anti-angiogenic molecules (Contois et al. 2009). Given the growing list of integrins thought to contribute to angiogenesis, it is clear that the ability of vascular cells to sense their immediate microenvironment plays important roles in new vessel development. Equally important in the contribution of integrins to angiogenesis are their respective ligands. For example, studies have shown that cellular interactions with distinct types of ECM components can alter the expression of integrins by mechanisms such as regulating mRNA stability (Feng et al. 1999; Retta et al. 2001; Delcommenne and Streuli 1995). Moreover, cellular interactions with distinct ECM components can differentially alter a cells response to a particular growth factor (Madri et al. 1988; Walker et al. 2005). In addition, certain ECM molecules have specific binding sites that allow secreted growth factors to become localized within the matrix (Schultz and Wysocki 2009; Hynes 2009). Collectively, these studies imply that different microenvironments composed of distinct ECM molecules may differentially regulate integrin expression and growth factor signaling. The composition and integrity of ECM ligands can vary widely between distinct tissue microenvironments such as those associated with embryonic development and those in the adult. Studies have indicated that embryonic vessels as well as angiogenic tumor vessels express alternatively spliced forms of fibronectin containing EIIIA and EIIIB domains, which are normally lacking in adult vessels (Astrof and Hynes 2009). Moreover, the major type of laminin within basement membranes during murine embryonic development is largely laminin-8 which contains the a4 laminin chain, while in postnatal mice, laminin-10 expression is predominate which contains the a5 laminin chain and a functionally exposed RGD avb3 binding site (Sasaki and Timpl 2001). Thus, developmentally regulated expression of ECM ligands may alter the functional significance of a given signaling pathway triggered by the particular integrin. As an example, if avb3-mediated interactions with laminin-10 contribute to the regulation of vessel development, it would not be surprising to observe only minimal impact from the lack of avb3 on embryonic vascular development since laminin-10 is not highly expressed. Conversely, in adult pathological
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
679
angiogenesis when laminin-10 is strongly expressed along with other RGD containing provisional matrix proteins such as fibrinogen and vitronectin, blocking avb3 disrupts pathological angiogenesis (Brooks et al. 1994). It is interesting to note that during angiogenesis associated with tumors and wound healing, provisional ECM components such as fibrin and denatured collagen are often found and studies suggest that endothelial cell interactions with these ECM proteins can increase the expression of integrin avb3 by a mechanisms associated with enhanced b3 integrin mRNA stability (Feng et al. 1999; Retta et al. 2001).
Modulation of Growth Factor/Growth Factor Receptor Systems within Different Tissue Microenvironments As mentioned above, when examining receptor-mediated signaling systems it is important to have an appreciation of the mechanisms that impact expression, biodistribution, and activity of both the receptors and their ligands. Given the diversity of growth factors and cytokines thought to play roles in regulating tumor angiogenesis, it is likely that tissue-specific processes exist within a particular microenvironment that may alter the mechanisms by which growth factors are presented to cells since many growth factors can be physically bound and stored within the ECM (Clark 2008; Hutchings et al. 2003). Thus, the relative ratio and bioavailability of different growth factors alter the functional impact of these angiogenic molecules during blood vessel development. The ability of secreted growth factors to become deposited and stored within the ECM of a particular tissue microenvironment contributes to their function during angiogenesis and tumor progression (Clark 2008; Hutchings et al. 2003). In fact, studies have suggested that matrix immobilized growth factors as well as proteolytically liberated growth factors can impact angiogenic signaling pathways. Thus, a more detailed understanding of control mechanisms regulating tissue-specific expression and localization of angiogenic growth factors will likely provide important molecular insight into how these receptor/ ligand systems function during new blood vessels formation.
Integrin–ECM Interactions Regulate Growth Factor Expression and Bio-distribution As discussed previously, integrin-mediated reciprocal communication links between cells and ECM can play roles in determining the expression, bio-distribution, and activity of specific components within growth factor signaling systems (Somanath et al. 2008; Streuli and Akhtar 2009; Ivaska and Heino 2010). For example, retinalpigmented epithelial cell interaction with the matricellular protein thrombospondin-1 has been shown to increase expression of VEGF and FGF2 as compared to interactions with collagen and laminin (Mousa et al. 1999). These studies illustrate
680
J. Roth et al.
the differential impact that ECM composition may have on expression of angiogenic growth factors. In other studies, monocyte adhesion to fibronectin resulted in elevated levels of PDGF and IGF-1 as compared to non-adherent cells (Shaw et al. 1990; Jendraschak et al. 1998). These findings are consistent with a potential role for monocyte/macrophage infiltration into sites of hypoxia serving to modulate the relative ratios of growth factors that promote angiogenesis. In this regard, a number of studies have shown a correlation between elevated migration and invasion of macrophages into the tumor microenvironment and enhanced angiogenesis and tumor progression (Toge et al. 2009; Chung et al. 2009). In studies with carcinoma cells, experiments have suggested that the laminin receptor a6b4 may play a role in translation of VEGF by a mechanism involving regulation of e4F1 (Chang et al. 2002). These studies, as well as many others, suggest that the cellular interactions with specific ECM molecules can alter the local growth factor profile within a give tissue microenvironment. Thus, understanding the ECM composition of a particular tissue microenvironment may help direct the most efficacious therapeutic approaches to control pathological angiogenesis occurring at a given site. While it is known that organ-specific differences exist in the composition of ECM components, how these variations in the insoluble microenvironment impacts organ-specific vasculature is not completely understood. Given the striking differences in ECM composition between brain tissue and other organs, the unique brain microenvironment provides multiple examples to help illustrate how angiogenic signaling cascades might be altered by the composition of the local microenvironment. Studies suggest that glioblastoma-derived tumor endothelial cells exhibit enhanced resistance to chemotherapeutic drugs as compared to normal brain endothelial cells and this resistance may be partly due to elevated levels of survivin, which can be regulated by b1-integrin-mediated interactions with ECM (Virrey et al. 2008; Fornaro et al. 2003) In addition, targeting avb3 has been shown to enhance the anti-tumor activity of radiation in brain tumors (Mikkelsen et al. 2009). In this regard, it is interesting to note that while normal vascular development did occur in avb3-integrin null mice defects were noted in brain vasculature (Bader et al. 1998). Collectively, these studies and others provide additional evidence that the specific composition of a given tissue microenvironment may alter the response of vascular cells to angiogenesis regulatory factors. Interestingly, elevated levels of the VEGF121 have been detected in association with angiogenic glioma tumors growing intracranially, while similar findings were not observed within glioma tumors growing within subcutaneous microenvironments (Guo et al. 2001). Recent studies have also identified brain-specific growth factors, some of which have been shown to bind integrins and activated MAP kinase signaling (Staniszewska et al. 2008). In an intriguing new study, evidence was provided that activated avb3 expressed within metastatic breast tumor cells growing intracranially were associated with elevated levels of VEGF (Lorger et al. 2009). This avb3-mediated upregulation of VEGF expression was dependent on phosphorylation-mediated inactivation of 4EBP1, which can suppress translation of VEGF (Lorger et al. 2009). Remarkably, this avb3-dependent upregulation of VEGF was only observed in metastatic tumors growing intracranially and was not
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
681
observed in similar tumors growing in the mammary fat pad, suggesting a specific microenvironmental influence. These important findings along with recent evidence, that the growth of brain tumors may depend in large part on the unique repertoire of ECM, may explain in part the enhanced efficacy for avb3 antagonists for brain tumors in animal models and in human clinical trials (Carbonell et al. 2009; MacDonald et al. 2008).
Modulation of Growth Factor Signaling by Integrins Just as cell–ECM signaling may alter growth factor expression in a tissue-specific way, so can growth factors stimulation in turn, modulate integrin and ECM expression, thereby altering the molecular composition of a particular microenvironment. While it is known that growth factors such as FGF2, IGF-1, VEGF, and TGF-b can differentially regulate expression of integrins and modulate a variety of important cellular parameters critical to angiogenesis and tumor progression, the molecular mechanisms by which they accomplish this may be quite different and depend on the particular tissue microenvironment (Somanath et al. 2008; Streuli and Akhtar 2009; Ivaska and Heino 2010). Therefore, understanding the microenvironmental context within which these signaling pathways operate may provide unique insight to optimize the development of specific drugs and therapeutic strategies. Given the ability of both integrin and ECM components within defined tissue microenvironment to modulate expression and downstream signaling activity of growth factors, it is not surprising that evidence is emerging that distinct integrin-mediated communication links modulate different angiogenic signaling pathways. One of the better-characterized examples of this concept includes the differences between VEGF and FGF in their ability to regulate endothelial cell survival during angiogenesis. Experimental observation suggested that antagonists of integrin avb3 inhibit FGF-induced angiogenesis, while exhibiting minimal impact on VEGFinduced angiogenesis (Friedlander et al. 1995). Conversely, antagonists of integrin avb5 inhibited VEGF-induced angiogenesis while showing limited impact on FGF stimulated angiogenesis. In other studies, investigators have shown distinct differences in the ability of FGF and VEGF to regulate cellular responses in endothelial cells derived from distinct microenvironments. For example, FGF2 but not VEGF stimulated reorganization of the actin cytoskeleton in bovine microvascular endothelial cells (BMEs), but similar results were not observed in bovine aortic endothelial cells (BAEs) (Cavallaro et al. 2001). Moreover, while FGF2 stimulation of BME cells resulted in a reduction in expression of the angiogenesis inhibitor TSP-1, neither VEGF nor FGF2 altered expression of TSP-1 in BAEs (Cavallaro et al. 2001). These studies, along with experiments demonstrating differential chemoresistance in brain tumor endothelial cells as compared to normal endothelial cells, provide further evidence that growth factor-stimulated responses within endothelial cells derived from different tissue microenvironments may be quite different.
682
J. Roth et al.
Since the initial observations concerning the differential integrin dependency for growth factor-stimulated responses, a number of studies have begun elucidating molecular mechanisms to account for these findings. Evidence suggests that c-Src plays a role in VEGF-stimulated angiogenesis in both the chick CAM and mouse models while playing minimal roles in FGF2-induced angiogenesis (Eliceiri et al. 1999). Moreover, VEGF but not FGF stimulation of endothelial cells resulted in Src-mediated phosphorylation of Fak at tyrosine 861 which facilitated FAK association with avb5 (Eliceiri et al. 2002; Ricono et al. 2009). Providing further insight into the molecular mechanisms by which distinct integrins modulate growth factor responses in endothelial cells, comes from studies showing that FGF2 stimulated, p21 activated protein kinase-1 (PAK-1)-dependent phosphorylation of Raf-1 at tyrosine residues 338 and 339, while VEGF stimulated Src-dependent phosphorylation of Raf-1 at tyrosine 340 and 341 (Hood et al. 2003; Alavi et al. 2003). This FGF-mediated phosphorylation of Raf-1 resulted in its translocation to the mitochondria independent from MEK/Erk activation and promoting resistance to apoptosis mediated by activation of the intrinsic pathway. In contrast, VEGF stimulation resulted in Src-mediated suppression of apoptosis mediated by the extrinsic pathway that depended on MEK/Erk activation (Hood et al. 2003; Alavi et al. 2003). In more recent studies, evidence suggested that c-abl may play a functional role in mediating FGF2- and avb3-dependent angiogenesis, while exhibiting little impact on VEGF/avb5-dependent angiogenesis (Yan et al. 2008). Further complicating our understanding of the biological relevance of these two integrin-dependent pathways have on angiogenesis are studies indicating that VEGF stimulation results in association of b3 integrin with the VEGFR2, thereby enhancing VEGF-mediated signaling (Mahabeleshwar et al. 2007; Borges et al. 2000). Surprisingly, in mice lacking b3 integrin, enhanced angiogenesis was observed and this enhanced angiogenesis was associated with elevated VEGF signaling (Reynolds et al. 2002, 2004). In contrast, pathological angiogenesis was significantly inhibited in b3 integrin knockin mice harboring a signaling deficient avb3 receptor (Mahabeleshwar et al. 2006). Thus, given the known complexity and compositional differences that exist between normal and tumor microenvironments, along with signaling feedback mechanisms that exist between distinct growth factor signaling cascades, the cooperative interactions between integrins and growth factors signaling pathways may provide new molecular understanding into these contrasting observations.
Integrin/Growth Factor Cooperation in Angiogenesis As discussed above, pathological angiogenesis can be organized into three general stages including an initiation phase, an invasive phase, and a maturation phase. The cooperative interactions between integrins and growth factor signaling play crucial roles in each of these overlapping steps. A growing body of evidence indicates that integrins may not only influence growth factor signaling indirectly by
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
683
Integrin Associations with Growth Factors VEGF FGF
TGF-β α β
PTN
CTGF
NGF
Ang -1 Cyr61
Fig. 29.2 Integrin associations with growth factors. Integrin receptors are known to associate with a number of distinct types of molecules and secreted growth factors are emerging as an important class of molecules that may form functional associations with integrins. Representative examples of growth factor/integrin associations. VEGF vascular endothelial growth factor, TGF-b transforming growth factor-beta, CTGF connective tissue growth factor, Ang-1 angiopoietin-1, Cyr61 cysteine-rich protein 61, NGF nerve growth factor, PTN pleiotrophin, FGF fibroblast growth factor
modulating growth factor expression and bio-distribution, but integrins may also form functional associations with growth factors (Fig. 29.2), thereby modulating downstream signaling pathways (Mori et al. 2008; Vlahakis et al. 2007; Munger et al. 1998; Mikelis et al. 2009; Chen et al. 2001; Aidoudi et al. 2008). When considering the potential impact of cooperative interactions between integrins and growth factor signaling, it is important to have an appreciation of the particular tissue microenvironment within which these signaling cascades are occurring. During tumor angiogenesis, the local ECM may be enriched with provisional ECM proteins such as denatured collagens, fibronectin, fibrin, and vitronectin. These ECM proteins have been shown to contain functional binding sites for a number of angiogenic growth factors, thus facilitating immobilization of these pro-angiogenic molecules within the ECM (Clark 2008; Rahman et al. 2005). This ECM-mediated storage of secreted growth factors may either enhance or inhibit their activity by altering receptor recognition. Integrins have been shown to have the capacity to bind some immobilized pro-angiogenic growth factors. Integrins such as avb3 and a3b1 have been shown to bind immobilized VEGF-A, which may facilitate endothelial cell adhesion, migration, and survival within a particular microenvironment (Hutchings et al. 2003). Interestingly, antibodies directed to a9b1 were shown to inhibit VEGF but not FGF2-induced angiogenesis suggesting that the repertoire of integrins expressed within a particular tissue may differentially regulate growth factor utilization and signaling (Mori et al. 2008). In addition to VEGF, a number of other angiogenic factors have also been shown to bind integrins including FGFs, angiopoietins (Ang1), connective tissue growth
684
J. Roth et al.
factor (CTGF), cysteine-rich angiogenic protein 61 (Cyr61), semaphorin 7A, and nerve growth factor (NGF) (Mori et al. 2008; Vlahakis et al. 2007; Munger et al. 1998; Mikelis et al. 2009; Aidoudi et al. 2008; Rahman et al. 2005; Hutchings et al. 2003). Studies have shown that avb3 may directly bind immobilized FGF1 and enhanced cell adhesion, migration, and proliferation (Mori et al. 2008). Studies have also demonstrated that a5b1 may associate with angiogenesis regulatory factor Ang-1 and that this interaction promotes signaling leading to vascular cell migration that was not dependent on the Ang-1 receptor Tie-2 (Shim et al. 2007). Experiments continue to provide additional examples by which integrins may associate with angiogenic growth factors and influence there downstream signaling. For example, interesting new work suggests that the heparin-binding angiogenic growth factor pleiotrophin (PTN), which was shown to stimulate endothelial cell migration, may interact with integrin avb3 in a non-RGD-dependent fashion (Mikelis et al. 2009). Moreover, function-blocking antibodies directed to avb3, but not a5b1 blocked PTN-induced endothelial cell migration mediated through the PTN receptor protein tyrosine phosphatase (Mikelis et al. 2009). Interestingly, PTN treatment of glioblastoma cells that lacked avb3 resulted in decreased migration, suggesting the possibility that avb3 may differentially regulate the utilization and activity of growth factors such as PTN in distinct cell types. Finally, a well-studied group of growth factors known to play multiple roles in regulating cellular behavior, angiogenesis, and tumor growth belong to the TGF-b family. Many studies have provided convincing evidence that multiple integrins including members of the av integrin subfamily (avb1, avb3, avb5, avb6, and avb8) interact with the latent form of TGF-b1 by binding to the latency-associated peptide (LAP) within its small latency complex (SLC) (Wipff and Hinz 2008). These TGF-b complexes can associate with TGF-b binding proteins (LTBPs) to form large latent complexes that bind ECM and become integrated within insoluble tissue matrix (Wipff and Hinz 2008). Importantly, some members of this av integrin family are thought to facilitate activation of TGF-b1 by at least two general but incompletely understood mechanisms including protease-dependent and -independent pathways (Wipff and Hinz 2008). Integrin-associated activation of TGF-b is known to promote diverse cellular signaling leading to control of cellular proliferating and survival (Wipff and Hinz 2008). Interestingly, studies have indicated that avb3 can modulate and enhance the TGF-b signaling in fibroblasts and avb6dependent activation of TGF-b was shown to be associated with regulation of pulmonary inflammation and fibrosis (Scaffidi et al. 2004). Finally, a number of anti-angiogenic factors are also thought to bind integrins thereby regulating the balance of pro- and anti-angiogenic signaling activity (Petitclerc et al. 2000; Wickström et al. 2004; Woodal et al. 2008). Recent evidence suggests that the anti-angiogenic chemokine CXCL4 may bind avb3 in endothelial cells and inhibit adhesion and migration (Aidoudi et al. 2008). Furthermore, angiogenesis inhibitors derived from the ECM such as Canstatin and Tumstatin may bind b1 and b3 integrins leading to inhibition of endothelial cell proliferation and survival by altering expression of cell cycle inhibitors and apoptosis regulatory proteins (Magnon et al. 2005; Sudhakar and Boosani 2008). Taken together, evidence supports the notion that integrin
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
685
receptors play important tissue-specific roles in integrating and coordinating the response of many cell types to angiogenic growth factors.
Integrin/Growth Factor Receptor Cooperation in Angiogenesis As can be appreciated from the discussion above, when one considers the molecular and biological impact of a given ligand/receptor signaling system, it is crucial to have an understanding of both the ligand and its cognate receptor. Thus, just as important as growth factor–integrin cooperation is in modulating angiogenesis, it is also crucial that we have a molecular understanding of the consequences of growth factor receptor interactions with integrins in controlling new blood vessel development. A steadily increasing number of studies demonstrate interactions between growth factor receptors and integrins (Fig. 29.3) (Borges et al. 2000; Mikelis et al. 2009; Scaffidi et al. 2004; Elsegood et al. 2006; Cascone et al. 2005; Lazova et al. 2009; Sahni et al. 2005; Kim et al. 2009). In recent years, a considerable amount of attention has been paid to the association of VEGFR2 with avb3. Studies have revealed that following stimulation of endothelial cells with VEGF-A, VEGFR2
Integrin Associations with Growth Factor Receptors VEGFR Tie-2
TGF-βR
α β HGFR
EGFR
FGFR
cFms
IGF-1R
Fig. 29.3 Integrin associations with growth factor receptors. In addition to the ability of integrin receptors to associate with secreted growth factors, integrins also can form functional associations with growth factor receptors. Representative examples of growth factor receptor/integrin associations. VEGFR vascular endothelial growth factor receptor, TGF-bR transforming growth factorbeta receptor, FGFR fibroblast growth factor receptor, IGF-1R insulin-like growth factor receptor-1, cFms c-MCF receptor, EGFR epidermal growth factor receptors, HGFR hepatocyte growth factor receptor, Tie-2 angiopoietin receptor
686
J. Roth et al.
recruits Src, which promotes the phosphorylation of the cytoplasmic tail of b3 integrin (Soldi et al. 1999). This phosphorylation event has been suggested to facilitate the formation of a complex between avb3 and VEGFR2. In turn, this functional complex may enhance downstream signaling promoting endothelial cell proliferation and survival, thereby facilitating angiogenesis (Soldi et al. 1999). While significant attention has been focused on the ability of avb3 to associate with VEGFR2, avb3 is well known to bind and associate with multiple molecules. Thus it is not surprising that avb3 has also been shown to interact with an array of other angiogenic growth factor receptors such as FGFRs, IGF-1Rs, PDGFbRs, and HGFRs (Toledo et al. 2005; Clemmons and Maile 2005; Schneller et al. 1997; Rahman et al. 2005). In a recent study, endothelial cell expression of the angiogenesis regulator nitric oxide (NO) following IL-1b stimulation was significantly enhanced in the presence of the avb3 ligand fibrinogen (Sahni et al. 2005). This elevated expression of NO was dependent on avb3, as anti-avb3 antibodies blocked this effect (Sahni et al. 2005). To this end, a stable association between avb3 and the IL-R was detected and further work suggested that the avb3/IL-1R association may play a role in regulating NO production in endothelial cells (Sahni et al. 2005). While some of these avb3-growth factor receptors interactions have been document in cell types other than endothelial cells, given the well-established contribution of tumor cells, inflammatory cells and stromal cells to blood vessel formation, it is likely that the formation of these interesting complexes in non-endothelial cells may also play tissue-specific roles in pathological angiogenesis. For example, previous studies have suggested that macrophage colony-stimulating factor (M-CSF) plays a role in regulating tumor angiogenesis in part by modulating expression of VEGF as well as recruitment of endothelial progenitor-like cells into the tumor microenvironment (Kubota et al. 2009). Moreover, recent studies have provided evidence that inhibition of M-CSF may selectively inhibit pathological angiogenesis (Kubota et al. 2009). Interestingly, recent studies suggest that M-CSF stimulation in osteoclasts resulted in a stable interaction between avb3 and the M-CSF receptor cFms (Elsegood et al. 2006). Even though the molecular mechanisms governing these associations are not fully understood, these avb3–growth factor receptor complexes likely enhance the activation of numerous signaling cascades such as the MAP/Erk and PI3k/Akt pathways, which are well known to play roles in new blood vessel formation. Finally, additional examples of growth factor receptor association with avb3 include interactions between avb3 and the pleiotrophin (PTN) receptor phosphatase as well as avb3 interactions with TGF-bRII (Mikelis et al. 2009; Scaffidi et al. 2004). Interestingly, a common occurrence in many of these growth factor receptor/integrin associations is the apparent requirement for growth factor stimulation prior to complex formation. For example, in human lung fibroblast, TGF-b1 stimulation was required to induce a functional association between avb3 and TGF-bRII (Scaffidi et al. 2004). Furthermore, VEGF stimulation was necessary to induce VEGFR2/avb3 association in endothelial cells, PTN stimulation was needed to induce association of the PTN receptor phosphatase with avb3 and finally IL-1b stimulation was required to induce an association between IL-1R and avb3 (Mikelis et al. 2009; Sahni et al. 2005; Soldi et al. 1999).
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
687
The biological implication of avb3 association with TGF-b receptors is complex and may be cell- and or tissue-type specific. Studies have shown that TGF-b1induced association of avb3 with the TGF-bIIR within human lung fibroblasts was associated with an amplification of TGF-b1-stimulated proliferation and migration and this cellular response could be inhibited by blocking avb3 (Scaffidi et al. 2004). These findings are surprising given that many reports implicate TGF-b signaling in inhibiting cellular proliferation and promoting differentiation in normal cell types (Wipff and Hinz 2008). In contrast, several studies have suggested that TGF-b may enhance proliferation, migration, invasion, and metastasis in transformed cells (Wipff and Hinz 2008). In this regard, it is well known that the repertoire and activation state of integrins can be significantly modified in tumor cells as compared to that observed in normal cells and thus it would be interesting to speculate that the particular ECM microenvironment and integrin repertoire may contribute to the differential response of cells to TGF-b stimulation. Interestingly, recent studies have suggested that mice deficient in b3 integrin exhibited elevated levels of TGF-b1, TGF-bIR, and TGF-bIIRs (Reynolds et al. 2005). Moreover, b3 integrin knockout fibroblasts derived from these mice exhibited elevated levels of TGF-b1 signaling, enhanced SMAD2 phosphorylation, but interestingly, reduced phosphorylation of SMAD3 and reduced localization of SMAD3 to the nucleus (Reynolds et al. 2005). Importantly, wound re-epithelialization within these b3-deficient mice was significantly enhanced (Reynolds et al. 2005). This enhanced re-epithelialization response was associated with elevated levels of fibronectin and vitronectin within the wound. Finally, increased fibroblast infiltration into the wound as compared to wildtype mice was also observed. These interesting findings suggest functional role for b3 containing integrins in regulating the response of cells to TGF-b in vitro and in vivo. Other studies also suggest that b1-containing integrins including a3b1 may associate with TGF-b1 receptor in a complex with E-cadherin expressed in epithelial cells and this complex may alter the response of epithelial cells to TGF-b1 by a unique b-catenin/SMAD signaling pathway (Kim et al. 2009). Taken together, with previous work demonstrating a role for integrins in activating TGF-b, along with other studies are consistent with the notion that the particular tissue microenvironment and integrin expression profile of cells may contribute to the differential response of cells to TGF-b. In addition to avb3, a number of b1 containing integrins have also been shown to physically associate with growth factor receptors. The fibronectin receptor a5b1 has been shown to interact with the Tie-2 receptor thereby altering Ang-1 signaling (Cascone et al. 2005). These studies indicate the Tie-2/a5b1 association may play a role in reducing the levels of Ang-1 needed to stimulate Tie-2 phosphorylation thereby modulating endothelial cell response to this angiogenic growth factor (Cascone et al. 2005). Moreover, studies have also shown that members of the b1integrin family may associate with the EGFR and promote PI3 kinase signaling, which is thought to play an important role in regulating endothelial cell survival and angiogenesis (Falcioni et al. 1997). Interestingly, the roles of b1 integrins in controlling angiogenesis may depend in part on the particular integrin heterodimer and the ECM ligands present within
688
J. Roth et al.
the specific tissue microenvironment. For example, while the a1b1 and a2b1 integrins have been implicated in regulating angiogenesis and antibodies directed to these collagen binding integrins inhibit VEGF-induced angiogenesis, studies have suggested that collagen binding mediated specifically through a1b1 can limit EGF-stimulated proliferation by activating the protein tyrosine phosphates TCPTP, which in turn lead to inhibition of EGF-induced EGFR signaling by de-phosphorylation (Mattila et al. 2004). In the case of endothelial cells, studies have indicated that collagen-I-mediated endothelial cell adhesion caused a reduction of VEGFR signaling as a result of recruitment of SHP2 and de-phosphorylation of VEGFR (Mittola et al. 2006). Moreover, recent studies have also shown that a1b1 binding to intact collagen, but not denatured collagen can also activate TCPTP resulting in reduction of VEGFR signaling, inhibition of endothelial cell proliferation, migration, and endothelial cell sprouting (Mattila et al. 2008). These studies provide convincing evidence that integrins can have either a positive or negative impact on growth factor receptor signaling and that these changes can further depend on the particular integrin and ECM ligands expressed within the specific tissue microenvironment.
Conclusions A growing body of evidence indicates that pathological angiogenesis involves highly complex sets of cellular, biochemical, molecular, and mechanical events that are controlled by an interconnected network of signaling systems that depend in large part on the particular tissue microenvironment within which angiogenesis is occurring. Thus, it is becoming increasingly clear that the development of more effective strategies to control pathological angiogenesis and tumor progression will depend on a more in-depth molecular understanding of how a particular tissue microenvironment modulates the integrated network of angiogenic signaling pathways. As can be appreciated from the discussion above, growth factor signaling networks present within a particular tissue microenvironment need to be sensed, integrated, and responded to in a meaningful way in order for functional blood vessels to form. To this end, integrin receptors are emerging as a central group of molecules that play critical roles in modifying and integrating these diverse signaling inputs from a variety of angiogenic growth factors. Thus, from a network systems perspective, integrins may serve as network hubs, by allowing tissue-specific detection, assimilation, and distribution of these signaling cues within the local microenvironment to allow angiogenesis to proceed in an effective manner. With the appreciation of the ability of integrins to modulate angiogenic growth factor signaling at multiple levels, a crucial challenge in the coming years will be to decipher how to exploit this new molecular understanding of the cooperative interactions of integrins and growth factor signaling networks to improve the efficacy of anti-angiogenic strategies for the treatment of human diseases.
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
689
Acknowledgments This work was supported in part by grant 2ROICA91645 to PCB and grant P20RR15555 to Robert Friesel and subproject to PCB. We would like to apologize to all those investigators whose important work was not discussed due to space limitations.
References Aidoudi S, Bujakowska K, Kieffer N et al (2008) The CXC-chemokine CXCL4 interacts with integrins implicated in angiogenesis. PLoS One. doi:10.137/journal.pone.0002657 Akalu A, Cretu A, Brooks PC (2005) Targeting integrins for the control of tumor angiogenesis. Expert Opin Investig Drugs 14:1475–1486 Alavi AS, Acevedo L, Min W et al (2007) Chemoresistance of endothelial cells induced by basic fibroblast growth factor depends on Raf-1-mediated inhibition of the proapoptotic kinase, ASK1. Cancer Res 67:2766–7272 Alavi A, Hood JD, Frausto R et al (2003) Role of Raf in vascular protection from distinct apoptotic stimuli. Science 301:94–96 Astrof S, Hynes RO (2009) Fibronectins in vascular morphogenesis. Angiogenesis 12:165–175 Avraamides CJ, Garmy-Susini B, Varner JA (2008) Integrins in angiogenesis and lymphangiogenesis. Nat Rev Cancer 8:604–617 Bader BL, Rayburn H, Crowley D et al (1998) Extensive vasculogenesis, angiogenesis, and organogenesis precede lethality in mice lacking all alpha v integrins. Cell 95:507–519 Borges E, Jan Y, Ruoslahti E (2000) Platelet-derived growth factor beta and vascular endothelial growth factor receptor 2 bind to beta 3 integrin through its extracellular domain. J Biol Chem 275:39867–39873 Brooks PC, Clark RA, Cheresh DA (1994) Requirement of vascular integrin avb3 for angiogenesis. Science 264:569–571 Burri PH, Hlushchuk R, Djonov V (2004) Intussuceptive angiogenesis: its emergence, its characteristics, and its significance. Dev Dyn 231:474–488 Carbonell WS, Ansorge O, Sibson N et al (2009) The vascular basement membrane as “soil” in brain metastasis. PLoS One 4:1–14 Cascone I, Napione L, Maniero F et al (2005) Stable interaction between alpha 5 beta 1 integrin and Tie2 tyrosine kinase receptor regulates endothelial cell response to Ang-1. J Cell Biol 170:993–1004 Cavallaro U, Tenan M, Castelli V et al (2001) Response of bovine endothelial cells to FGF-2 and VEGF is dependent on their site of origin: relevance to the regulation of angiogenesis. J Cell Biochem 82:619–633 Chen CC, Chen N, Lau LF (2001) The angiogenic factors Cyr61 and connective tissue growth factor induces adhesive signaling in primary human skin fibroblasts. J Biol Chem 276: 10443–10452 Chung ES, Chauhan SK, Jin Y et al (2009) Contribution of macrophages to angiogenesis induced by vascular endothelial growth factor receptor-3-specific ligands. Am J Pathol 175: 1984–1992 Chang J, Bachelder RE, Lipscomb EA et al (2002) Integrin (alpha 6 beta 4) regulation of eIF-4E activity and VEGF translation: a survival mechanism for carcinoma cell. J Cell Biol 158: 165–174 Clark RA (2008) Synergistic signaling from extracellular matrix-growth factor complexes. J Invest Dermatol 128:1354–1355 Clemmons DR, Maile LA (2005) Interaction between insulin-like growth factor-I receptor and alpha V beta 3 integrin linked signaling pathways: cellular responses to changes in multiple signaling inputs. Mol Endocrinol 19:1–11 Colmone A, Sipkins DA (2008) Beyond angiogenesis: the role of endothelium in the bone marrow vascular niche. Transl Res 151:1–9
690
J. Roth et al.
Contois C, Akalu A, Brooks PC (2009) Integrins as functional hubs in the regulation of pathological angiogenesis. Semin Cancer Biol 19:318–328 Delcommenne M, Streuli CH (1995) Control of integrin expression by extracellular matrix. J Biol Chem 270:26794–26801 Ebos JM, Lee CR, Kerbel RS (2009) Tumor and host-mediated pathways of resistance and disease progression in response to antiangiogenic therapy. Clin Cancer Res 15:5020–5025 Eliceiri BP, Paul R, Schwartzberg PL et al (1999) selective requirement for Src kinases during VEGF-induced angiogenesis and vascular permeability. Mol Cell 4:915–924 Eliceiri BP, Puente XS, Hood JD et al (2002) Src-mediated coupling of focal adhesion kinase to integrin avb5 in vascular endothelial growth factor signaling. J Cell Biol 157:149–159 Elsegood CL, Zhuo Y, Wesolowski GA et al (2006) M-CSF induces the stable interaction of cFms with alpha v beta 3 integrin in osteoclasts. Int J Biochem Cell Biol 38:1518–1529 Engelman JA (2009) Targeting PI3K signaling in cancer: opportunities, challenges and limitations. Nat Rev 9:550–562 Falcioni R, Antonini A, Nisticò P et al (1997) Alpha 6 beta 4 and alpha 6 beta 1 integrins associate with ErbB-2 in human carcinoma cell lines. Exp Cell Res 236:76–85 Feng X, Clark AF, Galanakis D et al (1999) Fibrin and collagen differentially regulate human dermal microvascular endothelial cell integrins: stabilization of av/b3 mRNA by fibrin. J Invest Dermatol 113:913–919 Fernado NT, Koch M, Rothrock C et al (2008) Tumor escape from endogenous, extracellular matrix-associated angiogenesis inhibitors by up-regulation of multiple proangiogenic factors. Clin Cancer Res 14:1529–1539 Fischer C, Schneider M, Carmeliet P (2006) Principles and therapeutic implications of angiogenesis, vasculogenesis and arteriogenesis. Hand Exp Pharmacol 176:157–212 Folkman J (2007) Angiogenesis: an organizing principle for drug discovery? Nat Rev Drug Discov 6:273–286 Fornaro M, Plescia J, Chheang S et al (2003) Fibronectin protects prostate cancer cells from tumor necrosis factor-a-induced apoptosis via the AKT/Survivin pathway. J Biol Chem 278: 50402–50411 Friedlander M, Brooks PC, Shaffer RW et al (1995) Definition of two angiogenic pathways by distinct alpha v integrins. Science 270:1500–1502 Ghosh K, Thodeti CK, Dudley AC et al (2008) Tumor-derived endothelial cells exhibit aberrant Rho-mediated mechanosensing and abnormal angiogenesis in vitro. Proc Natl Acad Sci U S A 105:11305–11310 Greenberg JI, Shields DJ, Barillas SG et al (2008) A role for VEGF as a negative regulator of pericyte function and vessel maturation. Nature 456:809–813 Guo P, Xu L, Pan S et al (2001) Vascular endothelial growth factor isoforms display distinct activities in promoting tumor angiogenesis at different anatomic sites. Cancer Res 61: 8569–8577 Hida K, Klagsbrun M (2005) A new perspective on tumor endothelial cells: unexpected chromosome and centrosome abnormalities. Cancer Res 65:2507–2510 Hlushchuk R, Riesterer O, Baum O et al (2008) Tumor recovery by angiogenic switch from sprouting to intussuceptive angiogenesis after treatment with PTK787/ZK222584 or ionizing radiation. Am J Pathol 173:1173–1185 Hodivala-Dilke KM, Reynolds AR, Reynolds LE (2003) Integrins in angiogenesis: multitalented molecules in a balancing act. Cell Tissue Res 314:131–144 Hood JD, Frausto R, Kiosses WB et al (2003) Differential av integrin-mediated Ras-ERK signaling during two pathways of angiogenesis. J Cell Biol 162:933–943 Huang SC, Fraenkel E (2009) Integrating proteomic, transcriptional, and interactome data reveals hidden components of signaling and regulatory networks. Sci Signal 2:1–10 Hutchings H, Ortega N, Plouët J (2003) Extracellular matrix-bound vascular endothelial growth factor promotes endothelial cell adhesion, migration, and survival through integrin ligation. FASEB J 11:1520–1532 Hynes RO (2009) The extracellular matrix: not just pretty fibrils. Science 326:1216–1219
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
691
Ivaska J, Heino J (2010) Interplay between cell adhesion and growth factor receptors: from plasma membranes to the endosomes. Cell Tissue Res 339:111–120 Jendraschak E, Kaminski WE, Kiefl R et al (1998) IGF-1, PDGF and CD18 are adhesive responsive genes: regulation during monocyte differentiation. Biochim Biophys Acta 1396:320–335 Kilarski WW, Samolov B, Petersson L et al (2009) Biomechanical regulation of blood vessel growth during tissue vascularization. Nat Med 15:657–664 Kim KK, Wei Y, Szekeres C et al (2009) Epithelial cell a3b1 integrin links b-catenin and smad signaling to promote myofibroblast formation and pulmonary fibrosis. J Clin Invest 119: 213–224 Klement GL, Yip TT, Cassiola F et al (2008) Platelets actively sequester angiogenesis regulators. Blood 5:125–135 Kubota Y, Takubo K, Shimizu T et al (2009) M-CSF inhibition selectively targets pathological angiogenesis and lymphangiogenesis. J Exp Med 206:1089–1102 Lazova R, Rothberg G, Rimm D et al (2009) The semaphoring 7A receptor plexin C1 is lost during melanoma metastasis. Am J Dermatopathol 31:177–181 Lorger M, Krueger JS, O’Neal M et al (2009) Activation of tumor cell integrin avb3 controls angiogenesis and metastatic growth in the brain. Proc Natl Acad Sci U S A 106:10666–10671 MacDonald TJ, Stewart CF, Kocak M et al (2008) Phase I clinical trial of cilengitide in children with refractory brain tumors: pediatric brain tumor consortium study PBTC-012. J Clin Oncol 26: 919–924 Madri JA, Pratt BM, Tucker AM (1988) Phenotypic modulation of endothelial cells by transforming growth factor-b depends upon the composition and organization of the extracellular matrix. J Cell Biol 106:1375–1384 Mikkelsen T, Brodie C, Finniss S et al (2009) Radiation sensitization of glioblastoma by cilengitide has unanticipated schedule-dependency. Int J Cancer 124:2719–2727 Magnon C, Galaup A, Mullan B et al (2005) Canstatin acts on endothelial and tumor cells via mitochondrial damage initiated through interaction with alpha v beta 3 and alpha v beta 5 integrins. Cancer Res 65:4353–4361 Mahabeleshwar GH, Feng W, Phillips DR et al (2006) Integrin signaling is critical for pathological angiogenesis. J Exp Med 203:2495–2507 Mahabeleshwar GH, Feng W, Reddy K et al (2007) Mechanisms of integrin-vascular endothelial growth factor receptor cross-activation in angiogenesis. Circ Res 101:570–580 Mattila E, Auvinen K, Salmi M et al (2008) The protein tyrosine phosphatase TCPTP controls VEGFR2 signaling. J Cell Sci 121:3570–3580 Mattila E, Pellinen T, Nevo J et al (2004) Negative regulation of EGFR signaling through integrina1b1-mediated activation of protein tyrosine phosphatase TCPTP. Nat Cell Biol 7:78–85 Mikelis C, Sfaelou E, Koutsioumpa M et al (2009) Integrin avb3 is a pleiotrophin receptor required for pleiotrophin-induced endothelial cell migration through receptor protein tyrosine phosphatase b/z. FASEB J 23:1459–1469 Mittola S, Brenchio B, Piccinini M et al (2006) Type I collagen limits VEGFR-2 signaling by SHP2 protein-tyrosine phosphatase-dependent mechanism. Circ Res 98:45–54 Mori S, Wu CY, Yamaji S et al (2008) Direct binding of integrin avb3 to FGF1 plays a role in FGF1 signaling. J Biol Chem 283:18066–18075 Mousa SA, Lorelli W, Campochiaro (1999) Role of hypoxia and extracellular matrix-integrin binding in the modulation of angiogenic growth factors secretion by retinal pigmented epithelial cells. J Cell Biochem 74:135–143 Munger JS, Harpel JG, Giancotti FG et al (1998) Interactions between growth factors and integrins: latent forms of transforming growth factor-b are ligands for integrin avb1. Mol Biol Cell 9:2627–2638 Murdoch C, Muthana M, Coffelt SB et al (2008) The role of myeloid cells in the promotion of tumor angiogenesis. Nat Rev Cancer 8:618–631 Petitclerc E, Boutaud A, Prestayko A et al (2000) New functions for non-collagenous domains of human collagen type IV: novel integrin ligands inhibiting angiogenesis and tumor growth in vv. J Biol Chem 275:8051–8061
692
J. Roth et al.
Puigvert JC, Huveneers S, Fredriksson L (2009) Cross-talk between integrins and oncogenes modulates chemosensitivity. Mol Pharmacol 75:947–955 Rahman S, Patel Y, Murray J et al (2005) Novel hepatocyte growth factor (HGF) binding domains on fibronectin and vitronectin coordinate a distinct and amplified Met-integrin induced signaling pathway in endothelial cells. BMC Cell Biol 6:8–14 Rapisarda A, Melillo G (2009) Role of hypoxic tumor microenvironment in the resistance to antiangiogenic therapies. Drug Resist Updat 12:74–80 Retta SF, Cassara G, D’Amato M et al (2001) Cross talk between b1 and av integrins: b1 affects b3 mRNA stability. Mol Biol Cell 12:3126–3138 Reynolds AR, Reynolds LE, Nagel TE et al (2004) Elevated Flk1 (vascular endothelial growth factor receptor 2) signaling mediates enhanced angiogenesis in b3-integrin-deficient mice. Cancer Res 64:8643–8650 Reynolds LE, Conti FJ, Lucas M et al (2005) Accelerated re-epithelialization in b3-integrindeficient mice is associated with enhanced TGF-b1 signaling. Nat Med 11:167–174 Reynolds LE, Wyder L, Lively JC et al (2002) Enhanced pathological angiogenesis in mice lacking b3 integrin or b3 and b5 integrins. Nat Med 8:27–34 Ricono JM, Huang M, Barnes LA et al (2009) Specific cross talk between epidermal growth factor receptor and integrin avb5 promotes carcinoma cell invasion and metastasis. Cancer Res 69: 1383–1391 Sahni A, Sahni SK, Francis CW (2005) Endothelial cell activation by IL-1b in the presence of fibrinogen requires avb3. Arterioscler Thromb Vasc Biol 25:2222–2227 Sasaki T, Timpl R (2001) Domain IVa of laminin a5 chain is cell-adhesive and binds b1 and avb3 integrins through Arg-Gly-Asp. FEBS Lett 509:181–185 Scaffidi AK, Petrovic N, Moddley YP et al (2004) avb3 integrin interacts with the transforming growth factor b (TGFb) type II receptor to potentiate the proliferative effects of TGFb1 in living human lung fibroblasts. J Biol Chem 279:37726–37733 Schneller M, Vuori K, Ruoslahti E (1997) Alpha v beta 3 integrin associates with activated insulin and PDGFbeta receptors and potentiates the biological activity of PDGF. EMBO J 16:5600–56007 Schultz GS, Wysocki A (2009) Interactions between extracellular matrix and growth factors in wound healing. Wound Rep Reg 17:153–162 Shaw RJ, Doherty DE, Ritter AG et al (1990) Adherence-dependent increase in human monocyte PDGF(B) mRNA is associated with increase in c-fos, c-jun, and EGR2 mRNA. J Cell Biol 111:2139–2148 Shim SW, Ho IA, Wong PE (2007) Angiopoietin: a TIE(d) balance in tumor angiogenesis. Mol Cancer Res 5:655–665 Soldi R, Mitola S, Strasly M (1999) Role of alpha v beta 3 integrin in the activation of vascular endothelial growth factor receptor-2. EMBO J 8:882–892 Somanath PR, Ciocea A, Byzova TV (2008) Integrin and growth factor receptor alliance in angiogenesis. Cell Biochem Biophys doi:10.1007/s12013-008-9040-5 St Croix B, Rago C, Velculescu V et al (2000) Gene expressed in human endothelium. Science 289:1197–1202 Staniszewska I, Sariyer IK, Lecht S et al (2008) Integrin alpha9 beta1 is a receptor for nerve growth factor and other neurotrophins. J Cell Sci 121:504–513 Streuli CH, Akhtar N (2009) Signal cooperation between integrins and other receptor systems. Biochem J 418:491–506 Sudhakar A, Boosani C S (2008) Inhibition of tumor angiogenesis by tumstatin: insights into signaling mechanisms and implications in cancer regression. Pharm Res 25:2731–2739 Toge H, Inagaki T, Kojimoto Y et al (2009) Angiogenesis in renal cell carcinoma: the role of tumor-associated macrophages. Int J Urol 16: 801–807 Toledo MS, Suzuki E, Handa K et al (2005) Effect of ganglioside and tetraspanins in microdomains on interaction of integrins with fibroblast growth factor receptor. J Biol Chem 280: 16227–16234
29 Cooperative Interactions Between Integrins and Growth Factor Signaling
693
Virrey JJ, Guan S, Li W et al (2008) Increased survivin expression confers chemoresistance to tumor-associated endothelial cells. Am J Pathol 173:575–585 Vlahakis N, Young BA, Atakilit A et al (2007) Integrin a9b1 directly binds vascular endothelial growth factor (VEGF)-A and contributes to VEGF-A-induced angiogenesis. J Biol Chem 282: 15187–15196 Wagner EF, Nebreda AR (2009) Signal integration by JNK and p38 MAPK pathways in cancer development. Nat Rev 9:537–549 Walker JL, Fournier AK, Assoian RK (2005) Regulation of growth factor signaling and cell cycle progression by cell adhesion and adhesion-dependent changes in cell tension. Cyt Growth Factor Rev 16:395–405 Wegener KL, Campbell ID (2008) Transmembrane and cytoplasmic domains in integrin activation and protein–protein interactions. Mol Membr Biol 25:76–87 Wickström SA, Alitalo K, Keski-Oja J et al (2004) An endostatin-derived peptide interacts with integrins and regulates actin cytoskeleton and migration of endothelial cells. J Biol Chem 279:20178–20185 Wipff PJ, Hinz B (2008) Integrins and activation of latent transforming growth factor b1 – an intimate relationship. Eur J Cell Biol 87:601–615 Woodal BP, Nyström A, Iozzo RA et al (2008) Integrin alpha 2 beta 1 is the required receptor for endorepellin angiostatic activity. J Biol Chem 283:2335–2343 Yan W, Bentley B, Shao R (2008) Distinct angiogenic mediators are required for basic fibroblast growth factor- and vascular endothelial growth factor-induced angiogenesis: the role of cytoplasmic tyrosine kinase c-Abl in tumor angiogenesis. Mol Biol Cell 19:2278–2288 Zaidel-Bar R, Itzkovitz S, Ma’ayan A et al (2007) Functional atlas of the integrin adhesome. Nat Cell Biol 9:858–867
Chapter 30
The Extracellular Matrix and the Growth and Survival of Tumors Yves A. DeClerck
Abstract The role of the extracellular matrix (ECM) in the tumor microenvironment extends well beyond the formation of a barrier against tumor invasion. Through a combination of physical forces and chemical signals generated upon contact between cells and ECM proteins, the ECM exerts a control on the proliferation and survival of cancer cells, which were previously considered to be solely affected by genetic alterations. The interaction between tumor cells and the ECM is also a dynamic one that changes upon modification of the ECM by proteases produced by tumor cells and by normal cells in the tumor microenvironment. These proteases modify the ECM generating cryptic epitopes, producing proteolytic fragments that are biologically active and releasing growth factors and cytokines that are trapped in the ECM. Our knowledge of the mechanisms by which the ECM controls cancer cells has significantly improved over the last 10 years and is now leading toward clinical trials testing agents disrupting tumor cell–ECM interaction.
Introduction Although able to proliferate and survive in the absence of exogenous signals, cancer cells remain deeply influenced by extracellular stimuli. The proliferation and survival of tumor cells is influenced by a large variety of external factors of a physical (interstitial pressure, mechanical forces) and chemical (acidosis, hypoxia) nature, as well as by cell–cell contact, soluble factors like growth factors, chemokines and cytokines, and by contact with the extracellular matrix (ECM). The ECM is an integral component of the tumor microenvironment that can affect cancer at multiple stages of progression, from initiation to metastasis. The importance of the interaction Y.A. DeClerck (*) Departments of Pediatrics and Biochemistry and Molecular Biology, University of Southern California and The Saban Research Institute of Childrens Hospital Los Angeles, Los Angeles, CA 20027, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_30, © Springer Science+Business Media, LLC 2010
695
696
Y.A. DeClerck
between the ECM and tumor cells became clear in the late 1990s when the laboratory of M. Bissell made the seminal observation that when grown in three-dimensional (3D) laminin-rich basement membrane matrices, malignant breast epithelial cells reverse to a normal polarized epithelial phenotype in the presence of integrin blocking antibodies (Weaver et al. 1997). The mechanisms by which the ECM affects cancer progression are highly complex and involve physical forces and biochemical signals, based on protein–protein interactions. Contact between the ECM and tumor cells can either inhibit or promote cancer progression. Importantly, this contact is a dynamic one that substantially changes during cancer progression. As cancer cells invade surrounding tissues and enter in contact with different ECM proteins, they modify the ECM by the action of a large variety of proteases that they produce or whose expression they stimulate in non-malignant stromal and inflammatory cells. This proteolytic modification of the ECM releases proteins that are trapped in the ECM and generates proteolytic fragments of ECM proteins that are biologically active and exert diverse functions on cancer progression and angiogenesis. In this chapter, we will focus on reviewing the mechanisms by which contact between tumor cells and ECM proteins and proteolytic modification of the ECM affect the proliferation and survival of tumor cells and influence angiogenesis.
Mechanical Forces and the ECM in Cancer Progression The stiffness of a tumor has long been recognized by clinicians as a sign of malignancy, and the presence of a desmoplastic stroma in a tumor is typically an indicator of a more aggressive behavior (Paszek and Weaver 2004). In breast cancer, a stiffening of the tissue surrounding the tumor is caused by a response of normal fibroblasts to the presence of invasive malignant epithelial cells and is known as the desmoplastic response. Fibroblasts not only increase their production and deposition of ECM proteins like fibronectin, tenascin, and collagen types I, III and IV but also increase collagen cross-linking by overproducing the extracellular enzyme lysyl oxidase (LOX) that catalyses covalent intermolecular cross-linking between collagen molecules and with elastin (Payne et al. 2007). This increases collagen insolubility and ECM stiffness (Payne et al. 2005). A rigid stroma generates three types of physical stress on cells: compressive stress applied by forces perpendicular to the cell surface and resulting in cell compaction, tensile stress applied by perpendicular forces exerting a traction on cells, and shear stress due to forces that are applied parallel to the cell surface. These forces are all sensed by cells via mechanoreceptors (Butcher et al. 2009). On the basis of Newton’s third law of action and reaction, cells respond to these external forces by adjusting intracellular tensions through the network of cytoskeletal proteins, a process called “mechanoreciprocity” (Lopez et al. 2008). Under normal conditions, this mechanoreciprocity maintains a tensional homeostasis in the cells and tissue that prevents an increase in internal forces. However in cancer cells this mechanoreciprocity is often lost, and consequently integrins become activated and cluster. This clustering causes activation of the focal adhesion kinase (FAK)-125, a-actinin tyrosine phosphorylation and interaction
30 The Extracellular Matrix and the Growth and Survival of Tumors
697
with FAK and Src, and stimulation of Rho-GTPases (Craig et al. 2007). Consistent with this concept, it has been demonstrated that when normal epithelial cells are maintained in the presence of a rigid matrix, there is an increase in integrin activation, enhancement of Erk activation, and ROCK-dependent contractility that promote malignant transformation (Paszek and Weaver 2004). A rigid matrix can thus promote malignant transformation. Cell stretching contributes to this phenomenon as it is an important stimulator of cell proliferation. For example, when forced to spread by changing the space between multiple points of focal adhesions, cells proliferate and escape apoptosis regardless of the type of ECM they are in contact with (Chen et al. 1997). For the same level of matrix rigidity, tumor cells stretch much more than normal cells and thus activate integrin-mediated Rho signaling much more effectively. The consequences of Rho-GTPase activation are an increased contractility in tumor cells associated with increased migratory behavior and proliferation via activation of Erk1/2. At the same time, Rho-GTPase activation increases matrix stiffness, providing a positive feedback signal that further activates integrin-mediated signaling (Huang et al. 1998). Forces generated when cells are in contact with the ECM also affect gene expression. For example, the expression of integrin is much higher in fibroblasts and epithelial cells that are in contact with a rigid matrix and signaling pathways like Erk and JNK become hyperactivated. Consequently, transcription factors like AP-1, STATs, MYC, CEBP and NFkB translocate to the nucleus, and their transcriptional activity is increased (Avvisato et al. 2007). A rigid ECM can also affect cancer cell proliferation by having a direct mechanical effect on the cell nucleus. This nuclear mechanotransduction acts like an epigenetic determinant by distorting the shape of the nucleus, altering the structure of the chromatin and affecting gene expression (Dahl et al. 2008). The exact mechanism linking deformation of the nuclear cytoskeleton and gene expression is still poorly understood, but it involves cadherin and integrin and cytoskeletal protein like actin (Gieni and Hendzel 2008). Proteases are also upregulated by stress forces that apply to cells. As a result of their overproduction, the ECM is degraded and its stiffness negatively affected. By decreasing the stiffness of the ECM, ECM degrading proteases may therefore have a negative effect on cancer cell proliferation and provide a negative feedback loop that limits the effect of tissue stiffness on tumor cells. These fundamental concepts on the role of mechanical forces in tumor cell growth have recently reached clinical application. It has long been well recognized that the presence of increased tissue densities in mammographic images of the breast is associated with a significant increase in the risk of developing a breast tumor (Wolfe 1976), an observation that is consistent with the concept that increased tissue stiffness promotes malignant transformation and progression (Tlsty and Coussens 2006). Loss of elastic properties in a tissue can now be accurately measured by elastography, using non-invasive ultrasound and magnetic resonance-based imaging techniques. Preliminary studies aimed at testing the use of such techniques to differentiate a benign tumor from a malignant tumor in prostate and breast cancer have been encouraging but studies on large cohorts will be needed to fully evaluate their potential (Garra 2007; Linden and Halpern 2007; Siegmann et al. 2009; Tse et al. 2009).
698
Y.A. DeClerck
Contact Between ECM and Tumor Cells Regulates Proliferation and Survival Although the ability to proliferate in the absence of exogenous stimulatory signals is a well recognized hallmark of cancer cells, cancer cells remain sensitive to stimuli from ECM proteins. Not every ECM protein however exerts a function on cell proliferation and survival. To distinguish between ECM proteins that have an effect on these cell functions from those that do not, the term matricellular proteins has been proposed (Bornstein and Sage 2002). These matricellular proteins can stimulate or inhibit cell proliferation and can have a pro-survival as well as a proapoptotic effect on cancer cells (Table 30.1) (Reddig and Juliano 2005).
Regulation of Tumor Cell Proliferation by ECM Proteins Most ECM proteins have a growth stimulatory activity on cells, including tumor cells. This growth stimulatory activity is typically mediated by integrins present at the cell surface. The contact between tumor cells and ECM proteins promotes the clustering and activation of integrins at points of focal adhesion. As a result, p130CAS is activated and phosphorylates FAK at Tyr397. Activated FAK recruits the Src family of kinases and activates phosphoinositol-3 kinase (PI3-K). Src also activates p130CAS, which in turn activates Rac, which can also be directly activated by FAK. This ultimately results in activation of paxillin and extracellular Table 30.1 Matricellular proteins Protein Domain/structure Fibronectin
Function Increases survival and proliferation
Mechanism b1 integrin medicated, stimulates Erk1/2, ROCK and PI3-K and Akt
Laminin
Domain 10/11
Increases survival and proliferation
PI3-K and Akt
Collagen type I
Non-fibrillar
Stimulates proliferation
a1 or a3b1 integrin, FAK and Erk1/2
Fibrillar
Inhibits proliferation
DDR-2, increases p27KIP-1, p21CIP-1 and p15INK4B
CCN1
Stimulates apoptosis in EC
a6b1 integrin and syndecan, activates Bax and caspase-3 and -9
TSP-1
Stimulates apoptosis in EC
Binds to CD 36, activates p59fyn and p38MAPK
SPARC
Stimulates apoptosis
Unknown
Decorin
Stimulates apoptosis
Caspase-3 activation
EMILIN2
Stimulates apoptosis in tumor cells
DR4 and DR5 activation and caspase-8 activation
30 The Extracellular Matrix and the Growth and Survival of Tumors
699
signal-related kinase (MEK) and Erk that upregulates cell cycle-dependent kinases (CDK) and downregulates inhibitors of CDKs like p21CIP-1and p27KIP-1(Aplin et al. 1999; Aplin and Juliano 1999). This is the pathway by which fibronectin exerts its well known growth stimulatory effect on cancer cells. Binding of epithelial cells to fibronectin via a5b1 integrin inhibits the expression of p21CIP-1and PTEN, activates Erk1/2, Rho kinase and PI3-K-protein kinase B (PkB/Akt) and increases the expression of c-MYC and cyclin D1 (Han and Roman 2006). Another ECM protein with a well known growth stimulatory activity is laminin, a major component of the basement membrane (Kim et al. 1999). Contact between tumor cells and collagen type I can also stimulate proliferation, however the effect is a function of the native state of the collagen. Collagen type I is typically present in tissues in an insoluble multimeric fibrillar form. In this form it inhibits the proliferation of normal and malignant cells (Henriet et al. 2000; Koyama et al. 1996). This growth inhibitory effect is associated with an upregulation of the expression of p27KIP-1, p21CIP-1, and p15INK-4. p21CIP-1is responsible for blocking of cells in the G1/G0 and p15INK-4for blocking cells at the G2 phase of the cell cycle (Wall et al. 2007). In contrast to growth stimulation, which is mediated by integrin, growth inhibition is mediated by the discoidin domain receptor-2 (DDR-2), a tyrosine kinase cell surface receptor that binds to fibrillar collagen and not to denatured collagen (Wall et al. 2005). When partially denatured (gelatin), monofibrillar collagen, like fibronectin, exerts a growth stimulatory activity by promoting cell spreading, integrin clustering and activation, formation of focal adhesions, actin cytoskeleton organization, and FAK phosphorylation (Hotary et al. 2003). The effect that collagen type I exerts on cell proliferation is therefore complex and highly dependent of its native state. When highly cross-linked with other collagen molecules and elastin though the action of LOX, it exerts a tensile stress on cells that stimulates proliferation. When present in a fibrillar multimeric structure, it inhibits proliferation and when partially denatured it again promotes cell growth. A second mechanism by which integrin activation promotes proliferation involves a crosstalk with growth factor receptors tyrosine kinase (RTKs) (Juliano 1996). Binding of integrin to the ECM enhances the activity of soluble growth factors and their affinity for RTKs (Guo and Giancotti 2004; Larsen et al. 2006), which provides an amplification system where insoluble ECM proteins and soluble growth factors interacting with their corresponding receptors (integrins and RTKs) cooperate to stimulate cell proliferation.
Regulation of Apoptosis by ECM Proteins In the absence of contact with the ECM, most normal cells undergo apoptosis (anoikis). A hallmark of malignant cells is their ability to escape apoptosis, which is commonly achieved by genetic alterations that change the balance between survival and death signals in favor of survival signals such as the overexpression of Bcl-2 associated with 8q24:14q32 translocation in Burkitt’s lymphoma, loss of p53 function
700
Y.A. DeClerck
associated with a majority of human cancers or loss of PTEN function in glioma (Hanahan and Weinberg 2000). However, tumor cells remain sensitive to mechanisms of survival that are triggered by contact with the ECM. Central to this mechanism is PI3-K, which is activated by integrins and integrin-linked kinase (ILK), and generates 3¢ phosphorylated inositides that promote the recruitment of PkB/Akt to the cell membrane and its activation by phosphorylation at Thr 308 by phosphoinositide-dependent kinase 1 (PDK1). As a result, a series of survival pathways become activated that inhibit caspase-9, decrease Bad, and increase the expression of survival proteins like survivin (Reddig and Juliano 2005). A second pathway is via FAK, activated by integrin clustering upon contact with ECM proteins. FAK activates p130CAS, an inhibitor of Bax and interacts with the receptor-interacting protein (RIP), a death domain containing serine/threonine kinase that is a major component of the death receptor complex (Kurenova et al. 2004). Through the adaptor proteins Grb2 and son of sevenless (SOS), FAK also activates the Ras/ Raf-1/MEK1/Erk1/2 pathway that not only stimulates proliferation but also negatively regulates Bim, by promoting its degradation after phosphorylation, thus preventing Bim from antagonizing Bcl-2 (Reginato et al. 2003). Integrin-mediated cell adhesion also protects cells from the extrinsic apoptotic pathway and caspase-8 activation by inhibiting the production of the death-associated protein-3 (DAP-3), that binds to Fas-associated death domain (FADD) and promotes caspase-8 activation (Miyazaki et al. 2004). Integrin-mediated contact between tumor cells and ECM proteins like fibronectin, laminin and vitronectin, protects tumor cells from stress-induced apoptosis (Gu et al. 2002; Trikha et al. 2002), including drug-induced apoptosis, providing therefore a reversible mechanism of acquired chemoresistance (Zutter 2007). There are however a few noticeable exceptions that are discussed in the next section.
Stimulation of Apoptosis by ECM Proteins A few ECM proteins are known to stimulate rather than inhibit apoptosis. Although their function has been primarily studied in normal cells, there is evidence that they could also induce apoptosis in malignant cells (Marastoni et al. 2008). Among these proteins is a secreted ECM-associated heparin-binding protein CCN1. The pro-apoptotic function of this protein is mediated by a6b1 and syndecan-4, an ECM heparan-sulfate proteoglycan. It culminates with activation of Bax which induces cytochrome C release from the mitochondria and activation of caspase-9 and -3. In prostate cancer cells it has a dual function, promoting proliferation and enhancing TRAIL-mediated apoptosis (Franzen et al. 2009). Thrombospondin-1 (TSP-1) is a secreted glycoprotein that interacts with its receptor CD36 present on endothelial cells. This results in association of CD36 with the Src-family tyrosine kinase protein p59fyn and activation of p38MAPK. In endothelial cells, this mechanism is responsible for the antiangiogenic activity of TSP-1 (Armstrong and Bornstein 2003). The effect of
30 The Extracellular Matrix and the Growth and Survival of Tumors
701
TSP-1 on tumor cells remains controversial with stimulatory and inhibitory effects that can be in part explained by conformational changes induced by proteases (Sid et al. 2004). The secreted protein acidic and rich in cysteine (SPARC) stimulates apoptosis in pancreatic tumor cells and inhibits growth (Puolakkainen et al. 2004). Decorin has a similar effect on other cancer cells (Goldoni and Iozzo 2008). A different mechanism of apoptosis is triggered by EMILIN-2, a member of the EMILIN family of ECM proteins. This protein activates the extrinsic pathway of apoptosis through direct binding to the death receptor (DR) 4 and 5 (although to a lower extent) present at the cell surface. This results in receptor clustering, death-induced signal complex (DISC) and caspase-8 activation (Mongiat et al. 2007). This ECM protein thus mimics the activity of a soluble death receptor ligand. Interestingly, in the particular case of EMILIN2, non-malignant cells like fibroblasts do not seem to be sensitive to this pathway of ECM-induced apoptosis, whereas it is active in several cancer cell lines.
Proteolytic Modification of the ECM and Tumor Growth and Survival Proteolytic Modification of the ECM and Cancer Progression During the different stages of cancer progression, the ECM is actively modified by proteases either produced by cancer cells or secreted in the tumor microenvironment by stromal and inflammatory cells. Matrix metalloproteinases (MMPs), serine proteases, cysteine proteases and aspartic proteases, as well as heparanase (an enzyme degrading proteoglycans) all have been shown to contribute to the proteolytic modification and degradation of the ECM (Lopez-Otin and Matrisian 2007). It was initially thought that the primary role of these proteases was to permit the invasion of tumor cells through the basement membrane and underlying connective tissues, a key hallmark of malignancy and a critical step in metastasis. In fact, whether the degradation of the ECM is always a necessary step for invasion and metastasis of cancer cells has been recently questioned as it became clear that cancer cells can adopt amoeboid movements that do not require proteases and ECM degradation to migrate through the ECM (Wolf et al. 2003). It is now clear that ECM degrading proteases do much more than allow tumor cells to invade. The proteolytic cleavage of ECM proteins has multiple effects on the tumor microenvironment. By altering the structure of proteins, these enzymes can make apparent new functional domains that were cryptic. They release cleaved peptides that are biologically active and finally, by dissolving the ECM, they liberate growth factors and cytokines in soluble forms increasing their biological activity. It is therefore more correct to refer to proteolytic modification rather than degradation of the ECM by proteases.
702
Y.A. DeClerck
Proteolytic Modification of the ECM Reveals Cryptic Domains in ECM Proteins (Matricryptins) ECM degrading proteases cleave ECM proteins at specific cleavage sites. In the majority of the cases, this initial proteolytic cleavage denatures the ECM protein and promotes its degradation by other less specific proteases. However in some cases, the cleavage generates a partially denatured protein in which a cryptic domain becomes exposed. This domain can exert a new function. Proteins having such cryptic domains that become active upon proteolytic cleavage are designated matricryptins, a term that underlines their biological activity on cell proliferation, migration, and differentiation. They are recognized by specific integrins expressed by cells. Among those, the integrin avb3, an integrin expressed by angiogenic endothelial cells (EC) and melanoma cells, has been well described for its ability to recognize, in addition to many other proteins, matricryptins that are unfolded upon denaturation or degradation of collagen type IV (Xu et al. 2001). Other matricryptins present in collagen, fibronectin or laminin, and exposed upon partial proteolysis of these ECM proteins interact with growth factor receptors and stimulate cell proliferation (Tran et al. 2005; Watanabe et al. 2000).
Proteolytic Modification of the ECM Releases Soluble Active Peptides (Matrikines) Among peptides that are generated from the proteolytic cleavage of precursor ECM proteins and have a biological activity is a family of anti-angiogenic peptides (Table 30.2). Most of these inhibitors have only an indirect effect on tumor cell growth by their anti-angiogenesis activity. The concept that proteolytic degradation of proteins by proteases like MMPs can generate biologically active peptides came from an original observation made in J. Folkman’s laboratory in the early 1990s. His laboratory had observed that in Lewis Lung carcinoma-bearing mice, the development of lung metastasis was dramatically accelerated upon surgical excision of the primary tumor. In an attempt to understand this process, the laboratory identified in the serum of these mice a factor produced by the primary tumor, that inhibited angiogenesis. This factor was later purified and identified as a 38-kDa protein called angiostatin (O’Reilly et al. 1994). It was then determined that angiostatin was generated by the cleavage of plasminogen by MMPs like MMP-2, and was a potent inhibitor of EC proliferation (Cao et al. 1996; O’Reilly et al. 1999). Since this initial discovery, many other peptides with a similar activity have been identified, all fragments of ECM precursor molecules, mainly collagen (Mundel and Kalluri 2007). Whereas the proteases generating some of these peptides are known, others are still unknown. Endostatin is a 20-kDa C-terminal fragment of collagen type XVIII that enhances apoptosis in EC by suppressing apoptosis inhibitors like Bcl-2. It is generated by cleavage by elastase (Wen et al. 1999). It binds to the integrin a5 and av
30 The Extracellular Matrix and the Growth and Survival of Tumors Table 30.2 Anti-angiogenic matrikines Name Mr (kDa) ECM source Angiostatin 38 Plasminogen Endostatin
20
Neostatin 7
28
Neostatin 14
23
Arresten
26
Canstatin
24
Tumstatin
28
TSP1 and 2 Neostatin
110 and 36 50
Collagen type XVIII NC domain Collagen type XVIII NC domain Collagen type XVIII NC domain Collagen type IV a1 NC domain Collagen type IV a2 NC domain Collagen type IV a3 NC domain Thrombospondin 1 & 2 Fibulin 1
703
Cleavage MMP-2, MMP-9 MMP-12 Elastase MMP-7 MMP-14 Not reported Not reported Not reported ADAMS-TS Cathepsin D
on the surface of human EC (Rehn et al. 2001). Neostatin 7 (28 kDa) and 14 (23 kDa) are two other anti-angiogenic factors derived from the cleavage of collagen type XVIII by MMP-7 and MMP-14, respectively (Chang et al. 2005; Kojima et al. 2008). They inhibit EC proliferation but their mechanism of action has not been fully investigated. Arresten is a 26 kDa proteolytic fragment of the non-collagenous (NC1) N-terminal domain of a1 collagen type IV by MMPs (Colorado et al. 2000). It inhibits the proliferation of basic fibroblast growth factor (bFGF)-stimulated endothelial cells and induces apoptosis (Nyberg et al. 2008). It also inhibits EC migration and tube formation in Matrigel. In vivo it inhibits the growth of xenotransplanted tumor cells and their metastasis. Its activity is mediated by binding to b1 integrin and inhibition of Erk signaling. It has not been shown to have a direct inhibitory effect on tumor cells. Canstatin is a 24-kDa proteolytic fragment of the NC1 domain of a2 chain of collagen type IV that binds avb3 and avb5 integrins and induces apoptosis in EC by inhibiting the phosphorylation of PkB/Akt and the activation of FAK (Magnon et al. 2005). It also induces FasLmediated apoptosis and activation of caspase-8 and -9 cleavage. It does not directly inhibit tumor cell growth (He et al. 2003; Kamphaus et al. 2000). Tumstatin is a 28-kDa proteolytic fragment of the a3 chain of collagen type IV that binds to integrin avb3 and a6b1 and inhibits activation of FAK, PI3-K, PkB/Akt, and a mammalian target of rapamycin (mTOR), in an EC-specific manner (Maeshima et al. 2002). Whereas arresten and canstatin have no direct effect on tumor cells, tumstatin has a direct inhibitory effect on tumor cells and inhibits their growth in a PTEN and PkB/Akt-mediated mechanism (Kawaguchi et al. 2006). The fact that several of these angiogenesis inhibitors are generated by the activity of MMPs on ECM proteins in part explains the failure of MMP inhibitors in clinical trials. In some of these trials, a paradoxical increase in tumor growth in patients
704
Y.A. DeClerck
treated with MMP inhibitors was reported. It later became apparent that these inhibitors have an undesirable and unanticipated effect by decreasing the production of MMP-generated anti-angiogenic peptides from their precursor molecules (Coussens et al. 2002; Houghton et al. 2006). Experiments in mice confirmed that this was in fact the case and investigators demonstrated that inhibition of MMP-9 in tumor-bearing animals increased tumor growth and vascularization and decreased the levels of angiostatin in the serum (Pozzi et al. 2002). The production of antiangiogenic peptides from precursor molecules is however not limited to the activity of MMPs. Cathepsin D digestion of fibulin-1 produces a fragment (also designated Neostatin) with nearly the same molecular weight as fibulin-5 that inhibits EC proliferation (Xie et al. 2008). TSP1 and 2 are also substrates for ADAMTS1. The cleavage of TSP by ADAMTS1 releases polypeptides from the trimeric structure of both TSP1 and 2, generating a pool of anti-angiogenic fragments from matrixbound thrombospondin (Iruela-Arispe et al. 1999; Lee et al. 2006). It is clear that the list of peptides released by the proteolytic remodeling of ECM proteins will continue to increase. Whereas many of those peptides have an anti-angiogenic activity by directly binding to surface receptors on EC (mainly integrins), some have a direct inhibitory activity on tumor cells.
Proteolytic Degradation of the ECM Releases Soluble Growth Factors Many growth factors share in their structure the presence of heparin binding domains and other motifs that mediate their binding to components of the ECM. This property explains that the ECM is an important reservoir of growth factors and cytokines. By doing so, the ECM increases their stability and limits their bioavailability. Upon proteolytic degradation of the ECM, these factors are released in soluble forms and can exert their biological function much more efficiently. This process plays an important role, particularly in bone metastasis and angiogenesis. When arrested in the bone marrow, tumor cells secrete a series of hormones and growth factors that activate pathways that alter the homeostatic balance between the formation of new bone by osteoblasts and the degradation of the bone matrix by osteoclasts (Guise 2000; Roodman 2004). This is mediated by the production of a large variety of stimulatory molecules like parathormone related peptide (PTHrP) that stimulates the production of the receptor activator of NFkB ligand (RANKL) by osteoblasts. By binding to RANK present on the surface of osteoclast precursor cells, RANKL activates their maturation and osteolytic activity. Alternatively, tumor cells secrete osteoclast activating factors like interleukin-1 and 6, granulocytic monocyte colony stimulating factor (GMCSF), and macrophage inflammatory protein (MIP)-1a. Activated osteoclasts demineralize the bone and degrade collagen type I through the combined activities of proteases like MMP-7, 9 and cathepsin K. The degradation of the bone matrix releases growth factors including insulin-like growth factor 1 (IGF-1) and transforming growth factor-b (TGF-b) that
30 The Extracellular Matrix and the Growth and Survival of Tumors
705
directly stimulate the proliferation of tumor cells and further increase the production of RANKL by osteoblasts and the production of PTHrP by tumor cells (Guise and Chirgwin 2003; Yin et al. 1999). This process further fuels bone degradation and is known as the vicious circle of bone metastasis. Degradation of the ECM by proteases like MMP-9 has also been shown to release vascular endothelial growth factor (VEGF) and thus to stimulate angiogenesis. This mechanism explains the critical role of MMP-9 in initiating the angiogenic switch in transgenic mouse models of pancreatic cancers and squamous cell carcinoma (Bergers et al. 2000; Coussens et al. 2000). The degradation of glycosaminoglycans by heparanase is another mechanism releasing many growth factors like VEGF in the microenvironment that is responsible for the pro-tumorigenic and pro-angiogenic activities of heparanase (Vlodavsky et al. 2007).
Clinical Implications As we increasingly appreciate the importance of the interactions between tumor cells and the ECM and how they regulate the proliferation and survival of cancer cells, the possibility to target these interactions in the treatment of cancer has been considered and some agents are currently tested in clinical trials.
Broad Inhibitors of ECM Degradation Blocking the degradation of the bone matrix by tumor-activated osteoclasts has been among the first attempts to affect the interaction between tumor cells and the ECM. It can be effectively achieved by bisphosphonates, a group of phosphoric acid-based molecules that tightly bind to the bone matrix and are potent stimulators of apoptosis in osteoclasts and tumor cells residing in the bone. Several of these bisphosphonates, like ibandronate and zoledronic acid, have been successfully tested in prostate and breast cancer metastatic to the bone. Their use has been a powerful strategy to interrupt the vicious circle of osteolysis in patients with cancer metastasis to the bone (Clezardin 2002; Coleman 2004; Lipton 2003) and they have been successfully tested in myeloma and breast and prostate cancer metastasis (Michaelson and Smith 2005).
Integrin Inhibitors Because integrin plays such a central role in mediating the interactions between tumor cells and the ECM, and in particular by protecting tumor cells from apoptosis, inhibitors of integrin have been first considered as potential anti-cancer agents.
706
Y.A. DeClerck
A small RGD peptide inhibitor of avb3 and avb5 integrin labeled Cilengitide has been developed, and is currently tested in patients, including children, with brain tumors (MacDonald et al. 2008; Reardon et al. 2008). It has shown little toxicity but the results of the clinical trials have not been entirely conclusive. One reason is that, at low concentration, this inhibitor is pro-angiogenic and stimulates tumor cell proliferation (Reynolds et al. 2009). A monoclonal antibody against a5b1 integrin (Volociximab) has been developed as an anti-angiogenic agent and has been shown to be well tolerated in patients with solid tumors (Ricart et al. 2008). It is presently too early to determine the potential value of these inhibitors in the clinic and more testing will be needed.
Protease Inhibitors In the late 1990s, several small molecule inhibitors of MMPs were tested in clinical trials in patients with a variety of cancers. The rationale behind their use was based on their anti-invasive and anti-metastatic activity, as well as their anti-angiogenic and growth inhibitory activities well demonstrated in pre-clinical models. The results of these trials were however disappointing, to the point that by 2001 all trials were abandoned (Coussens et al. 2002). One of the reasons for this has already been discussed and is related to the fact that inhibition of MMP activity also inhibits the release of anti-angiogenic matrikines like endostatin or angiostatin. There are however other reasons for these disappointing results. Most of the inhibitors tested had a broad spectrum of activity and were thus inhibiting most MMPs. Clinical studies were done without a complete understanding of the complex role of MMPs and of their many substrates, and a lack of realization that some MMPs have an anti-tumor activity. Studies did not include the monitoring of their effect on proteolytic activity and unanticipated side effects in the form of an arthritis-like syndrome requiring interruption of the treatment were observed in almost 30% of the cases. As we understand better the complex action of MMPs and the large number of proteins they cleave, the use of MMP inhibitors in cancer clinical trials may be reconsidered (Overall and Kleifeld 2006).
Angiogenesis Inhibitors Several native angiogenic inhibitors like endostatin and angiostatin made as recombinant proteins are currently in clinical phase I trials. Recombinant human angiostatin in combination with paclitaxel and carboplatin has been tested in patients with advanced non-small cell lung cancer and an overall response rate of 39% was reported (Kurup et al. 2006). Recombinant human endostatin (Endosteal) was also tested with or without interferon-a2 in metastatic melanoma but failed to show any activity (Moschos et al. 2007). Although well tolerated, these inhibitors have however not yet shown clinical efficacy (Kulke et al. 2006).
30 The Extracellular Matrix and the Growth and Survival of Tumors
707
Conclusion It is now clear that the ECM is not an innocent bystander in the tumor microenvironment, and that its role is not limited to providing a barrier against the expansion of tumor cells. Through a combination of physical and chemical signals, the ECM controls essential cellular functions like growth, survival and angiogenesis, and therefore controls and can even reverse the malignant behavior of tumor cells. Our understanding of the mechanism by which the ECM controls tumor cell proliferation and apoptosis has significantly improved over the last decade, to a point that such knowledge is now leading to the identification of molecules interfering with tumor cell–ECM interaction or ECM-derived peptides that could have a therapeutic value.
References Aplin AE, Howe AK, Juliano R (1999) Cell adhesion molecules, signal transduction and cell growth [see comments]. Curr Opin Cell Biol 11:737–744 Aplin AE, Juliano RL (1999) Integrin and cytoskeletal regulation of growth factor signaling to the MAP kinase pathway. J Cell Sci 112(Pt 5):695–706 Armstrong LC, Bornstein P (2003) Thrombospondins 1 and 2 function as inhibitors of angiogenesis. Matrix Biol 22:63–71 Avvisato CL et al (2007) Mechanical force modulates global gene expression and beta-catenin signaling in colon cancer cells. J Cell Sci 120:2672–2682 Bergers G et al (2000) Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat Cell Biol 2:737–744 Bornstein P, Sage EH (2002) Matricellular proteins: extracellular modulators of cell function. Curr Opin Cell Biol 14:608–616 Butcher DT, Alliston T, Weaver VM (2009) A tense situation: forcing tumour progression. Nat Rev Cancer 9:108–122 Cao Y et al (1996) Kringle domains of human angiostatin. Characterization of the anti-proliferative activity on endothelial cells. J Biol Chem 271:29461–29467 Chang JH et al (2005) Functional characterization of neostatins, the MMP-derived, enzymatic cleavage products of type XVIII collagen. FEBS Lett 579:3601–3606 Chen CS et al (1997) Geometric control of cell life and death. Science 276:1425–1428 Clezardin P (2002) The antitumor potential of bisphosphonates. Semin Oncol 29:33–42 Coleman RE (2004) Bisphosphonates: clinical experience. Oncologist 4(9 Suppl):14–27 Colorado PC et al (2000) Anti-angiogenic cues from vascular basement membrane collagen. Cancer Res 60:2520–2526 Coussens LM, Fingleton B, Matrisian LM (2002) Matrix metalloproteinase inhibitors and cancer: trials and tribulations. Science 295:2387–2392 Coussens LM et al (2000) MMP-9 supplied by bone marrow-derived cells contributes to skin carcinogenesis. Cell 103:481–490 Craig DH, Haimovich B, Basson MD (2007) Alpha-actinin-1 phosphorylation modulates pressureinduced colon cancer cell adhesion through regulation of focal adhesion kinase-Src interaction. Am J Physiol Cell Physiol 293:C1862–C1874 Dahl KN, Ribeiro AJ, Lammerding J (2008) Nuclear shape, mechanics, and mechanotransduction. Circ Res 102:1307–1318 Franzen CA et al (2009) Matrix protein CCN1 is critical for prostate carcinoma cell proliferation and TRAIL-induced apoptosis. Mol Cancer Res 7:1045–1055
708
Y.A. DeClerck
Garra BS (2007) Imaging and estimation of tissue elasticity by ultrasound. Ultrasound Q 23: 255–268 Gieni RS, Hendzel MJ (2008) Mechanotransduction from the ECM to the genome: are the pieces now in place? J Cell Biochem 104:1964–1987 Goldoni S, Iozzo RV (2008) Tumor microenvironment: modulation by decorin and related molecules harboring leucine-rich tandem motifs. Int J Cancer 123:2473–2479 Gu J et al (2002) Laminin-10/11 and fibronectin differentially prevent apoptosis induced by serum removal via Phosphatidylinositol 3-KInase/AKT and MEK1/ERK-dependent pathways. J Biol Chem 277:19922–19928 Guise TA (2000) Molecular mechanisms of osteolytic bone metastases. Cancer 88:2892–2898 Guise TA, Chirgwin JM (2003) Transforming growth factor-beta in osteolytic breast cancer bone metastases. Clin Orthop Relat Res S32–S38 Guo W, Giancotti FG (2004) Integrin signalling during tumour progression. Nat Rev Mol Cell Biol 5:816–826 Han SW, Roman J (2006) Fibronectin induces cell proliferation and inhibits apoptosis in human bronchial epithelial cells: pro-oncogenic effects mediated by PI3-kinase and NF-kappa B. Oncogene 25:4341–4349 Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100:57–70 He GA et al (2003) Canstatin-N fragment inhibits in vitro endothelial cell proliferation and suppresses in vivo tumor growth. Biochem Biophys Res Commun 312:801–805 Henriet P et al (2000) Contact with fibrillar collagen inhibits melanoma cell proliferation by upregulating p27KIP1. Proc Natl Acad Sci U S A 97:10026–10031 Hotary KB et al (2003) Membrane type I matrix metalloproteinase usurps tumor growth control imposed by the three-dimensional extracellular matrix. Cell 114:33–45 Houghton AM et al (2006) Macrophage elastase (matrix metalloproteinase-12) suppresses growth of lung metastases. Cancer Res 66:6149–6155 Huang S, Chen CS, Ingber DE (1998) Control of cyclin D1, p27(Kip1), and cell cycle progression in human capillary endothelial cells by cell shape and cytoskeletal tension [In Process Citation]. Mol Biol Cell 9:3179–3193 Iruela-Arispe ML et al (1999) Inhibition of angiogenesis by thrombospondin-1 is mediated by 2 independent regions within the type 1 repeats. Circulation 100:1423–1431 Juliano R (1996) Cooperation between soluble factors and integrin-mediated cell anchorage in the control of cell growth and differentiation. Bioessays 18:911–917 Kamphaus GD et al (2000) Canstatin, a novel matrix-derived inhibitor of angiogenesis and tumor growth. J Biol Chem 275:1209–1215 Kawaguchi T et al (2006) The PTEN/Akt pathway dictates the direct alphaVbeta3-dependent growth-inhibitory action of an active fragment of tumstatin in glioma cells in vitro and in vivo. Cancer Res 66:11331–11340 Kim WH et al (1999) Laminin-1-adherent cancer cells show increased proliferation and decreased apoptosis in vivo. Anticancer Res 19:3067–3071 Kojima T, Azar DT, Chang JH (2008) Neostatin-7 regulates bFGF-induced corneal lymphangiogenesis. FEBS Lett 582:2515–2520 Koyama H et al (1996) Fibrillar collagen inhibits arterial smooth muscle proliferation through regulation of Cdk2 inhibitors. Cell 87:1069–1078 Kulke MH et al (2006) Phase II study of recombinant human endostatin in patients with advanced neuroendocrine tumors. J Clin Oncol 24:3555–3561 Kurenova E et al (2004) Focal adhesion kinase suppresses apoptosis by binding to the death domain of receptor-interacting protein. Mol Cell Biol 24:4361–4371 Kurup A et al (2006) Recombinant human angiostatin (rhAngiostatin) in combination with paclitaxel and carboplatin in patients with advanced non-small-cell lung cancer: a phase II study from Indiana University. Ann Oncol 17:97–103 Larsen M et al (2006) The matrix reorganized: extracellular matrix remodeling and integrin signaling. Curr Opin Cell Biol 18:463–471
30 The Extracellular Matrix and the Growth and Survival of Tumors
709
Lee NV et al (2006) ADAMTS1 mediates the release of antiangiogenic polypeptides from TSP1 and 2. EMBO J 25:5270–5283 Linden RA, Halpern EJ (2007) Advances in transrectal ultrasound imaging of the prostate. Semin Ultrasound CT MR 28:249–257 Lipton A (2003) Bisphosphonate therapy in the oncology setting. Expert Opin Emerg Drugs 8: 469–488 Lopez JI, Mouw JK, Weaver VM (2008) Biomechanical regulation of cell orientation and fate. Oncogene 27:6981–6993 Lopez-Otin C, Matrisian LM (2007) Emerging roles of proteases in tumour suppression. Nat Rev Cancer 7:800–808 MacDonald TJ et al (2008) Phase I clinical trial of cilengitide in children with refractory brain tumors: Pediatric Brain Tumor Consortium Study PBTC-012. J Clin Oncol 26:919–924 Maeshima Y et al (2002) Tumstatin, an endothelial cell-specific inhibitor of protein synthesis. Science 295:140–143 Magnon C et al (2005) Canstatin acts on endothelial and tumor cells via mitochondrial damage initiated through interaction with alphavbeta3 and alphavbeta5 integrins. Cancer Res 65:4353–4361 Marastoni S et al (2008) Extracellular matrix: a matter of life and death. Connect Tissue Res 49: 203–206 Michaelson MD, Smith MR (2005) Bisphosphonates for treatment and prevention of bone metastases. J Clin Oncol 23:8219–8224 Miyazaki T et al (2004) Functional role of death-associated protein 3 (DAP3) in anoikis. J Biol Chem 279:44667–44672 Mongiat M et al (2007) Regulation of the extrinsic apoptotic pathway by the extracellular matrix glycoprotein EMILIN2. Mol Cell Biol 27:7176–7187 Moschos SJ et al (2007) Endostatin plus interferon-alpha2b therapy for metastatic melanoma: a novel combination of antiangiogenic and immunomodulatory agents. Melanoma Res 17: 193–200 Mundel TM, Kalluri R (2007) Type IV collagen-derived angiogenesis inhibitors. Microvasc Res 74:85–89 Nyberg P et al (2008) Characterization of the anti-angiogenic properties of arresten, an alpha1beta1 integrin-dependent collagen-derived tumor suppressor. Exp Cell Res 314:3292–3305 O’Reilly MS et al (1994) Angiostatin: a novel angiogenesis inhibitor that mediates the suppression of metastases by a Lewis lung carcinoma. Cell 79:315–328 O’Reilly MS et al (1999) Regulation of angiostatin production by matrix metalloproteinase-2 in a model of concomitant resistance. J Biol Chem 274:29568–29571 Overall CM, Kleifeld O (2006) Tumour microenvironment – opinion: validating matrix metalloproteinases as drug targets and anti-targets for cancer therapy. Nat Rev Cancer 6:227–239 Paszek MJ, Weaver VM (2004) The tension mounts: mechanics meets morphogenesis and malignancy. J Mammary Gland Biol Neoplasia 9:325–342 Payne SL et al (2005) Lysyl oxidase regulates breast cancer cell migration and adhesion through a hydrogen peroxide-mediated mechanism. Cancer Res 65:11429–11436 Payne SL, Hendrix MJ, Kirschmann DA (2007) Paradoxical roles for lysyl oxidases in cancer – a prospect. J Cell Biochem 101:1338–1354 Pozzi A, LeVine WF, Gardner HA (2002) Low plasma levels of matrix metalloproteinase 9 permit increased tumor angiogenesis. Oncogene 21:272–281 Puolakkainen PA et al (2004) Enhanced growth of pancreatic tumors in SPARC-null mice is associated with decreased deposition of extracellular matrix and reduced tumor cell apoptosis. Mol Cancer Res 2:215–224 Reardon DA et al (2008) Randomized phase II study of cilengitide, an integrin-targeting arginineglycine-aspartic acid peptide, in recurrent glioblastoma multiforme. J Clin Oncol 26: 5610–5617 Reddig PJ, Juliano RL (2005) Clinging to life: cell to matrix adhesion and cell survival. Cancer Metastasis Rev 24:425–439
710
Y.A. DeClerck
Reginato MJ et al (2003) Integrins and EGFR coordinately regulate the pro-apoptotic protein Bim to prevent anoikis. Nat Cell Biol 5:733–740 Rehn M et al (2001) Interaction of endostatin with integrins implicated in angiogenesis. Proc Natl Acad Sci U S A 98:1024–1029 Reynolds AR et al (2009) Stimulation of tumor growth and angiogenesis by low concentrations of RGD-mimetic integrin inhibitors. Nat Med 15:392–400 Ricart AD et al (2008) Volociximab, a chimeric monoclonal antibody that specifically binds alpha5beta1 integrin: a phase I, pharmacokinetic, and biological correlative study. Clin Cancer Res 14:7924–7929 Roodman GD (2004) Mechanisms of bone metastasis. N Engl J Med 350:1655–1664 Sid B et al (2004) Thrombospondin 1: a multifunctional protein implicated in the regulation of tumor growth. Crit Rev Oncol Hematol 49:245–258 Siegmann KC et al (2009) Diagnostic value of MR elastography in addition to contrast-enhanced MR imaging of the breast-initial clinical results. Eur Radiol 20:318–325 Tlsty TD, Coussens LM (2006) Tumor stroma and regulation of cancer development. Annu Rev Pathol 1:119–150 Tran KT, Lamb P, Deng JS (2005) Matrikines and matricryptins: implications for cutaneous cancers and skin repair. J Dermatol Sci 40:11–20 Trikha M et al (2002) Role of beta 3 integrins in human melanoma growth and survival. Int J Cancer 101:156–167 Tse ZT et al (2009) Magnetic resonance elastography hardware design: a survey. Proc Inst Mech Eng H 223:497–514 Vlodavsky I et al (2007) Heparanase: structure, biological functions, and inhibition by heparinderived mimetics of heparan sulfate. Curr Pharm Des 13:2057–2073 Wall SJ et al (2005) Discoidin domain receptor 2 mediates tumor cell cycle arrest induced by fibrillar collagen. J Biol Chem 280: 40187–40194 Wall SJ, Zhong ZD, DeClerck YA (2007) The cyclin-dependent kinase inhibitors p15INK4B and p21CIP1 are critical regulators of fibrillar collagen-induced tumor cell cycle arrest. J Biol Chem 282:24471–24476 Watanabe K et al (2000) Interaction with heparin and matrix metalloproteinase 2 cleavage expose a cryptic anti-adhesive site of fibronectin. Biochemistry 39:7138–7144 Weaver VM et al (1997) Reversion of the malignant phenotype of human breast cells in threedimensional culture and in vivo by integrin blocking antibodies. J Cell Biol 137:231–245 Wen W et al (1999) The generation of endostatin is mediated by elastase. Cancer Res 59: 6052–6056 Wolf K et al (2003) Compensation mechanism in tumor cell migration: mesenchymal-amoeboid transition after blocking of pericellular proteolysis. J Cell Biol 160:267–277 Wolfe JN (1976) Risk for breast cancer development determined by mammographic parenchymal pattern. Cancer 37:2486–2492 Xie L et al (2008) Basement membrane derived fibulin-1 and fibulin-5 function as angiogenesis inhibitors and suppress tumor growth. Exp Biol Med (Maywood) 233:155–162 Xu J et al (2001) Proteolytic exposure of a cryptic site within collagen type IV is required for angiogenesis and tumor growth in vivo. J Cell Biol 154:1069–1079 Yin JJ et al (1999) TGF-beta signaling blockade inhibits PTHrP secretion by breast cancer cells and bone metastases development. J Clin Invest 103:197–206 Zutter MM (2007) Integrin-mediated adhesion: tipping the balance between chemosensitivity and chemoresistance. Adv Exp Med Biol 608:87–100
Chapter 31
Secreted Growth Factors as Therapeutic Targets Beverly A. Teicher
Abstract Secreted growth factors directed from malignant cell to malignant cell, malignant cell to stromal cells, vascular cells and immune system cells or from stromal cells, vascular cells and immune system cells to malignant cells are essential for the growth and progression of malignant disease. Proangiogenic factors including members of the vascular endothelial growth factor (VEGF)-A family, placental growth factor, members of the fibroblast growth factor family, semaphorins, ephrins, angiopoietins, stromal-cell derived factor-1, EG-VEGF, Bv8, transforming growth factor-bs (TGFbs), and others act on endothelial cells, endothelial precursor cells and endothelial progenitor cells from the bone marrow to promote tumor growth. Platelet-derived growth factors act on pericytes. The VEGF-C and VEGF-D family members stimulate lymphangiogenesis. TGFb secreted by malignant cells and stromal cells acts on the tumor stromal cells and on immune system cells to promote growth of the malignancy. Urokinase plasminogen activator and tissue plasminogen activator stimulate malignant cell and vascular cell migration as do numerous secreted matrix metalloproteinases. Secreted protein acidic and rich in cysteine (SPARC) stimulates the growth of malignant cells and influences malignant cell invasion and metastasis. In the bone, RANK ligand secretion can be stimulated by malignant disease to stimulate osteoclast-mediated bone resorption and allow metastasis growth. The interleukin family of secreted proteins mediates immune system activity. Immune system damping interleukins are involved in tumor immune evasion as are the chemokines monocyte chemoattractant protein 1 (MCP1) and RANTES. The large epidermal growth factor family stimulates the proliferation of epithelial malignant cells. Similarly, insulin-like growth factors and hepatocyte growth factor increase the growth of malignant tumors. Recently, the Wnt family of secreted growth factors has been identified as deregulated in multiple epithelial tumors. Many secreted growth factors are potential therapeutic targets for neutralizing antibodies or soluble receptor constructs.
B.A. Teicher (*) Genzyme Corporation, 49 New York Avenue, Framingham, MA 01701-9322, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_31, © Springer Science+Business Media, LLC 2010
711
712
B.A. Teicher
Pro-angiogenic and Lymphangiogenic Factors Secreted growth factors are ideal targets for antibody therapeutics which can neutralize the target. One of the most successful anticancer drugs available is a humanized antibody, bevacizumab, which neutralizes the pro-angiogenic secreted growth factor vascular endothelial growth factor (VEGF)-A. VEGF signaling is a critical potentially rate-limiting step in physiological angiogenesis. The VEGF family includes seven secreted glycoproteins which have been designated VEGF-A, VEGF-B, VEGF-C, VEGF-D, VEGF-E, placental growth factor (PlGF) and VEGF-F. These proteins stimulate angiogenesis through a signaling process initiated by binding to their receptors which have been designated VEGF receptor (VEGFR)-1, VEGFR-2, VEGFR-3 and neuropilins (NP-1 and NP-2) (Otrock et al. 2007). VEGFs are potent, multifunctional cytokines that with important, possibly independent actions on blood and lymph vascular endothelium. VEGF-A was discovered as a tumor-secreted protein which caused large and small veins to be hyperpermeable to circulating macromolecules. VEGF-B promotes angiogenesis; however, its role has not been elucidated. VEGF-C drives in lymphangiogenesis during embryogenesis and functions in the maintenance of differentiated lymphatic endothelium in adults. VEGF-D stimulates the growth of vascular and lymphatic endothelial cells. Although initially characterized in the placenta, PlGF is expressed by a variety of cells, tissues, and organs. PlGF is involved in angiogenesis, wound healing, and the inflammatory response (Roskoski 2008). In cell culture VEGF induces tube formation, proliferation, transient accumulation of calcium, shape change, cell division, migration, and invasion of mature and precursor endothelial cells. In addition, exposure to VEGF alters endothelial cell gene expression and in vivo induces angiogenesis (Dvorak et al. 1995; Folkman 1995). VEGF neutralizing therapies were hypothesized to inhibit new blood vessel growth and inducing hypoxia and “starving” the tumor of oxygen and nutrients. It is becoming clear that the efficacy of VEGF neutralizing agents arises from multiple effects on varied cell types (Ellis and Hicklin 2008). The majority of anti-angiogenic agents approved and currently in clinical trial are directed toward neutralizing VEGFA or blocking VEGFR intracellular signaling. However, although many patients benefit from, for example, bevacizumab, many patients derive little or no benefit from treatment. Those patients who initially benefit through disease stabilization and prolonged progression-free and overall survival after treatment with VEGF-directed agents, their tumors eventually become non-responsive to continued anti-VEGF treatment so that the increased survival duration is months rather than years. Various mechanisms may contribute to initial lack of response or decreased in responsiveness after initial response to VEGF-directed therapy including dependence upon or up-regulation of other pro-angiogenic factors (Fischer et al. 2008). One pro-angiogenic factor which is emerging as a potentially important therapeutic target is PlGF. PlGF is a pleiotropic cytokine that stimulates endothelial cell growth, migration, and survival. PlGF is a chemoattractant for angiocompetent
31 Secreted Growth Factors as Therapeutic Targets
713
macrophages and bone marrow (BM)-derived endothelial progenitor cells. PlGF signals through binding to VEGFR-1 and the co-receptors neuropilin-1 and -2. VEGFR-1 is expressed not only by endothelail cells but also by macrophages, bone marrow-derived endothelial progenitor cells, and some tumor cells (Fischer et al. 2007). While VEGF signal primarily through VEGFR-2, PlGF binds selectively to VEGFR-1 and thus produces a specific signal distinct from VEGF. Upon binding to VEGFR-1, PlGF amplifies VEGF-induced signaling through VEGFR-2 by intrareceptor crosstalk, resulting in potentiation of the pro-angiogenic effect of VEGF. PlGF is a key player in resistance mechanisms, used by tumors to escape anti-VEGF therapies. PlGF was upregulated when pancreatic tumors, initially responsive to blockade of VEGFR-2, became resistant and broke through the treatment. PlGF homozygous knockout mice grow up normally and give birth to healthy litters. However, when challenged with cancer, ischemia, inflammation, or wounds, PlGF homozygous knockout mice demonstrate defective angiogenesis and plasma extravasation. A monoclonal anti-PlGF has been tested as an anticancer agent in several syngeneic mouse tumor models. The anti-PlGF was an effective treatment at a dose of 50 mg/kg. The anti-PlGF inhibited neo-angiogenesis and induced regression of existing tumor vessels, inhibited lymphangiogenesis, inhibited recruitment of angiocompetent macrophages, and had a direct cytostatic effect on certain tumor cells (Loges et al. 2008). PlGF seems to be highly inducible in the clinic especially during treatment of patients with VEGF neutralizing agents or inhibitors of VEGFR kinase activity (Kopetz et al. 2009; Motzer et al. 2006; Rosen et al. 2007). Semaphorins are a large family of secreted, transmembrane and GPI-linked proteins initially characterized in the development of the nervous system and axonal guidance. Semaphorins are expressed in many tissues where they regulate normal development, organ morphogenesis, immunity, and angiogenesis. They affect the cytoskeleton, actin filament organization, microtubules, and cell adhesion. Several semaphorins and their receptors (neuropilins and plexins) participate in vascular development, angiogenesis, and cancer. Neuropilins, which are high-affinity receptors for class-3 semaphorins, are also co-receptors for VEGF and other growth factors, and their expression is often abnormal in cancer. Class-3 semaphorin signaling is transduced by neuropilin receptors. In cancer, semaphorins have both tumor suppressor and tumor promoting functions. In tumor-bearing mice, treatment with anti-neuropilin antibodies inhibited tumor angiogenesis and was additive when combined with an anti-VEGF antibody (Potiron et al. 2009). The large family of fibroblast growth factors (FGF) regulates a wide range of developmental processes, including brain patterning, branching morphogenesis and limb development. The functions of individual members of the FGF family differ in nearly every aspect including cell types affected and cellular functions affected. The angiogenic properties of FGF2, also known as basic FGF (bFGF) are well known. Exogenous FGF2 stimulates migration and proliferation of endothelial cells in vivo, and enhances survival and proliferation of smooth muscle cells and fibroblasts, which induces the development of large. Mice engineered to over-express of FGF2 have normal vasculature, possibly due to balancing by negative regulators of
714
B.A. Teicher
vascular growth. Interferon-a (IFNa) and IFNb downregulate FGF2 in human kidney, bladder and prostate tumor cell lines. Exposure of bladder carcinoma to IFNa or b inhibits FGF2 expression and cellular proliferation. There is evidence that inhibition of FGF2 signaling can slow tumor growth by inhibiting angiogenesis. However, FGF2 levels do not correlate with tumor microvessel density indicating that the anti-tumor effects of interferons mediated through FGF2 blockade may not be due only to anti-angiogenesis (Beenken and Mohammadi 2009). In a fastgrowing malignant tissue, tumor blood vessels are exposed to multiple growth factors and cytokines. Although the roles of individual factors and their signaling pathways in regulation of tumor angiogenesis and neovascularization are well-established, understanding the interactions amongst these factors in stimulating tumor angiogenesis and metastasis remains an area of intense investigation. Quiescent vascular endothelial cells remain non-responsive to platelet-derived growth factor (PDGF)-BB stimulation alone; however, exposure to the combination of PDGF-BB and FGF2 switches on the PDGF receptor and activates endothelial cells. PDGF-BB transduces positive feedback signals to the FGF2 signaling system by amplifying FGF receptor expression in vascular pericytes. These reciprocal interactions in the tumor microenvironment lead to the formation of the tumor vasculature; thus, providing a rationale for development of anti-angiogenic agent combinations for the treatment of cancer (Beenken and Mohammadi 2009; Cao et al. 2008). Growth factors that stimulate lymphangiogenesis have been identified. The growth of lymphatic vessels is controlled by proteins related to VEGF, namely VEGF-C and VEGF-D. Increased expression of VEGF-C and VEGF-D is associated with several human tumor types. The experimental inhibition of VEGF-C and VEGF-D signaling in animal models suggests that lymphangiogenic growth factors facilitate the metastatic spread of tumor cells via lymphatics. Anti-lymphangiogenesis has not yet been tested as a therapeutic approach. Lymphangiogenesis inhibitors may block the metastatic spread of cancers and may be very useful when cancer is detected earlier in the course of the disease (Stacker and Achen 2008). It is well-established that the metastatic spread of tumor cells frequently occurs via lymphatic vasculature. Experimental tumor data and human clinicopathologic data indicate that growth of lymphatic vessels near a tumor mass is often associated with lymph node metastasis. Changes in the adhesive properties of lymphatic endothelium near tumors facilitate metastatic spread. Lymphangiogenic growth factors that promote formation of tumor lymphatics and metastatic spread of tumor cells to lymph nodes are known. These include VEGF-C and VEGF-D which act through binding to receptor tyrosine kinase VEGFR-3 located on lymphatic endothelial cells. Several other signaling molecules promote lymphangiogenesis and/or lymphatic metastases in cancer include VEGFA, PDGF-BB, hepatocyte growth factor (HGF), and certain chemokines. Recently, it was shown that lymphangiogenic growth factors secreted by a primary tumor can induce lymphangiogenesis in nearby lymph nodes, prior to migration by tumor cells, thus further facilitating metastasis (Achen and Stacker 2008). The invasion of lymphatic vessels by colorectal cancer and subsequent spread of the tumor to draining lymph nodes is a key determinant of prognosis in colon cancer. Although lymphangiogenesis most likely contributes to this process a simple
31 Secreted Growth Factors as Therapeutic Targets
715
correlation between lymphatic vessel density and colorectal cancer metastasis has been difficult to prove. Attempts to correlate expression of VEGF-C and VEGF-D with the lymphatic metastasis in colon cancer have provided contradictory results. Evidence from models of tumor metastasis suggests that interactions between the tumor microvasculature and varied cell types within the tumor microenvironment are involved in the spread of metastasizing tumors. Strategies similar to those which have been instrumental to increasing understanding of vascular angiogenesis are being applied to tumor lymphangiogenesis and suggest tumor lymphatics have gene expression profiles distinct from those of normal lymphatic vessels and that promote metastasis (Royston and Jackson 2009). As with all anticancer therapies, most tumors development resistance to treatment with anti-angiogenic agents, while a fraction of tumors are non-responsive to these therapies from the outset. These two forms of resistance have been termed acquired resistance and intrinsic resistance. Both forms of resistance have been associated with the genetic instability of malignant disease. Similarly, two modes of resistance have been described for tumors treated with anti-angiogenic therapies: evasive resistance, an adaptation to circumvent the specific angiogenic blockade; and intrinsic or pre-existing indifference. Multiple mechanisms can be invoked that manifest in evasive or intrinsic resistance. The two modes of resistance in response to anti-angiogenic therapy imply adaptive evasion and intrinsic non-responsiveness of tumors. Adaptive or evasive resistance refers to the ability of a tumor, after an initial response phase, to evade the therapeutic blockade by altering the secreted proangiogenic factors or by reducing dependence on such growth of new blood vessels such growth along host vessels. Intrinsic non-responsiveness is defined by the absence of benefit from anti-angiogenic therapy. Acquired/adaptive resistance may occur by activation and/or upregulation of non-VEGF pro-angiogenic signaling pathways such as PlGF, FGF, ephrin, and angiopoietins. Recruitment of bone marrow-derived endothelial precursor cells (EPC) through secretion of certain chemokines such as stromal cell-derived factor 1 can restore neovascularization (Bergers and Hanahan 2008). Tumor stromal and vascular cells can be derived from bone marrow-derived progenitor cells including mesenchymal stem cells (MSC), EPC, pericyte progenitor cells, VEGFR-1+ progenitor cells, and Tie2-expressing monocytes, which are mobilized into circulation and incorporate into tumor microenvironment. In the tumor microenvironment, MSC may promote tumor growth by secretion of growth factors, participation in vessel formation and possibly development of tumor stem cell niches. It is clear that bone marrow-derived cells can functionally contribute to tumor growth. Precursor and progenitor cells mobilized from the bone marrow into circulation migrate to tumor sites and incorporate into the tumor microenvironment. The contribution of these cells to tumor growth seems to be highly variable and whether blockade of recruitment of cells from the bone marrow to the tumor can be of therapeutic value (Roorda et al. 2009). Bone marrow-derived cells participate in tumor angiogenesis. The secreted growth factors, EG-VEGF and Bv8 proteins (also called prokineticin1 (Prok1) and prokineticin 2 (Prok2)), promote both tissue-specific angiogenesis and hematopoietic cell mobilization. Bv8 and EG-VEGF bind two highly related G-protein coupled
716
B.A. Teicher
receptors (GPCRs), EG-VEGFR/PKR1 and EG-VEGFR/PKR2. Both EG-VEGF and Bv8 are mitogens selective for endothelial cell types. Bv8 and EG-VEGF can induce hematopoietic cell mobilization in vivo and stimulate production of granulocytic and monocytic colonies in vitro. Bv8 is expressed in the bone marrow and granulocyte colony-stimulating factor up-regulates Bv8 expression. Anti-Bv8 antibodies reduced CD11b+Gr1+ myeloid cell mobilization elicited by granulocyte colony-stimulating factor. Anti-Bv8 inhibited growth of several tumors in mice and suppressed tumor angiogenesis. Anti-Bv8 treatment also reduced the number of CD11b+Gr1+ myeloid cells in circulation and in tumors. Bv8 may be interesting candidate for a mediator of inflammatory-cell-dependent angiogenesis (Shojaei et al. 2007). Pro-angiogenic bone marrow cells include subsets of hematopoietic cells that provide vascular support and EPC, which under certain conditions can differentiate into functional vascular cells. The chemokine stromal-cell derived factor-1 (SDF-1; CXCL12) recruits and retains CXCR4+ bone marrow cells to angiogenic niches supporting ischemic tissue revascularization and tumor growth. The mechanism by which activation of CXCR4 modulates angiogenesis is not clear. SDF-1 promotes revascularization by binding CXCR4 expressed on vascular cells and supports mobilization of pro-angiogenic CXCR4+VEGFR1+ hematopoietic cells, thus stimulating ischemic site revascularization. SDF-1–CXCR4 signaling pathway has multiple functions in the regulation of vascularization during acute ischemia and tumor growth. By modulating plasma SDF-1 levels, the short-lived CXCR4 antagonist AMD3100 acutely promotes, while chronic AMD3100 treatment inhibits, mobilization of pro-angiogenic cells (Petit et al. 2007). In malignant disease, SDF-1+ or CXCR4+ various lineages cells are found in tumor tissues. Blockade of the SDF-1/CXCR4 axis decreases growth of preclinical gastrointestinal tumors through an anti-angiogenic effect. CXCR4 Neutralization can suppress the growth of mouse Colon38 and PancO2 tumors. The suppression of tumor growth was independent of CXCR4 expression by the malignant cells. CD31+ intratumoral vessels were decreased 55% and intratumor blood flow was decreased 35% by CXCR4 blockade. Intratumoral VEGF was not altered by CXCR4 neutralization. CXCR4+ endothelial cells were detected in the tumors suggesting that the anti-angiogenic effects of CXCR4 blockade lead to decreased tumor vessels in a VEGF independent process (Guleng et al. 2005). In patients with multiple myeloma, elevated SDF-1 levels in peripheral blood plasma are associated with osteolysis and with tumor angiogenesis. High SDF-1 levels produced by multiple myeloma plasma cells promote osteolysis and angiogenesis (Martin et al. 2006). Cellular niches are key in regulating normal stem cell differentiation and regeneration and appear to support malignant metastasis. Using dynamic in vivo confocal imaging, it has been shown that murine bone marrow contains unique anatomic regions defined by specialized endothelium which express the adhesion molecule E-selectin and the chemoattractant SDF-1 in discrete, discontinuous areas. Circulating leukemic cells can engraft around these regions suggesting that these may be a microenvironment for metastatic tumor spread in bone marrow. Purified hematopoietic stem/progenitor cells and lymphocytes localize to the same microdomains. Disruption of the
31 Secreted Growth Factors as Therapeutic Targets
717
interaction between SDF-1 and CXCR4 blocked migration of Nalm-6 acute lymphoblastic leukemia cells to these regions (Sipkins et al. 2005). Multiple myeloma plasma cells produce significant levels of SDF-1. Multiple myeloma patients have elevated SDF-1 plasma which positively correlates with the presence of multiple radiological bone lesions, suggesting a potential role for SDF-1 in osteoclast precursor recruitment and activation. There is increased osteoclast motility and activation in the presence of SDF-1 as well as an increase in osteoclast activation-related gene expression including receptor activator of nuclear factor-kB ligand (RANKL), RANK, TRAP, matrix metalloproteinases (MMP)-9, CA-II, and Cathepsin K. A small molecule CXCR4 inhibitor effectively blocked osteoclast formation stimulated by RPMI-8226 myeloma cells. In multiple myeloma patients, high circulating SDF-1 may recruit osteoclast precursors to local sites within the bone marrow and enhance bone resorbing activity (Zannettino et al. 2005). A large array of small molecule VEGF receptor tyrosine kinase inhibitors and a few biologic therapeutic primarily directed toward VEGF are in clinical trial. Drugs which target VEGF and PDGF pathways have revolutionized the treatment of patients with metastatic renal cell cancer. Approximately 75% of patients with clear cell renal cell carcinoma have mutations or silencing of the von Hippel Lindau gene producing an accumulation of HIF 1 alpha. These tumors produce high levels of proangiogenic factors such as VEGF and PDGF leading to angiogenesis and endothelial stabilization. Two small molecules and one biological therapeutic targeting these pathways have reached FDA approval. Both sunitinib and sorafenib target VEGF and PDGF receptor tyrosine kinases and bevacizumab, a VEGFtargeting monoclonal antibody. These three agents have superior progression free survival in patients with metastatic renal cell carcinoma compared with interferon or placebo. Several additional receptor tyrosine kinase inhibitors such as axitinib, pazopanib, and cediranib are under investigation and may provide further treatment options (Heng and Bukowski 2008). Several clinical studies have investigated treatment with bevacizumab for patients with recurrent malignant glioma. Treatment with bevacizumab is commonly combined with cytotoxic chemotherapy. These combination regimens often produce marked responses radiographically, prolongation of progression-free survival, and decreased corticosteroids use. Small molecule VEGFR receptor tyrosine kinase inhibitors, such as cediranib can achieve similar results. Anti-angiogenic treatment is generally well tolerated but common adverse effects include hypertension and proteinuria, whereas the potentially more serious adverse effects, such as thromboembolic disease and haemorrhage, occur infrequently. Approximately 50% of patients fail to respond to anti-angiogenic treatment and the response duration is variable. Resistance to anti-VEGF therapy implicates alternative pro-angiogenic factors such as bFGF, SDF-1, the angiopoietin receptor Tie2, and PlGF in the angiogenic process (Norden et al. 2008). Several agents in development for the treatment of non-small cell lung cancer (NSCLC) target the VEGF pathway. Most clinical trial data generated to date are with either bevacizumab or small-molecule inhibitors of VEGF receptor tyrosine kinase activity (sunitinib, sorafenib, and ZD6474). VEGF Trap, an engineered soluble receptor made from extracellular domains of VEGFR1 and VEGFR2, binds to
718
B.A. Teicher
VEGF and to PlGF. VEGF Trap binds to VEGFA with higher affinity than does bevacizumab. The toxicities in phase I trials with subcutaneous and intravenous administration of VEGF Trap, hypertension and proteinuria, are similar to those seen with other molecules that target the VEGF pathway. Ongoing phase I trials are evaluating combinations of VEGF Trap with platinum-based doublets and single agent docetaxel. The activity of single agent VEGF Trap in NSCLC is being assessed in a multicenter phase II trial (Riely and Miller 2007). VEGF Trap is in clinical trial in a wide variety of malignant diseases.
Pro-stromal Factors Transforming growth factor-b (TGFb) is a key player in malignant disease through its actions on host tissues and cells. Malignant cells often secrete large amounts of TGFb that act on non-transformed cells present in the tumor mass as well as distal cells in the host to suppress antitumor immune responses creating an environment of immune tolerance, augmenting angiogenesis, invasion and metastasis, and increasing tumor extracellular matrix (ECM) deposition (Pinkas and Teicher 2006; Teicher 2007). The tumor stroma actively contributes to tumorigenesis; it is more than a bystander. The involvement of stroma may occur early during epithelial cell transformation. Stroma consists of many cell types including fibroblasts, immune cells, endothelial cells, and pericytes lining blood and lymphatic vessels, which are embedded in ECM and produce soluble factors. The percentage and composition of tumor stroma varies amongst malignant diseases but is different from normal stroma. Malignant cells produce growth factors which induce a “reactive” stroma supporting tumor cell proliferation, migration and invasion, and angiogenesis (Mueller and Fusenig 2004). “Reactive stroma” is also called desmoplasia. Tumor stroma resembles the stroma usually found in injured tissues (Dvorak et al. 1995). Communication between malignant cells and tumor stroma components is facilitated by a network of growth factors including TGFb (Naber et al. 2008). Fibroblasts are the predominant stromal cells in carcinomas suggesting that fibroblasts may support or induce tumor progression. Progression to invasive carcinoma involves differentiation of fibroblasts to myofibroblasts or so-called cancer-associated fibroblasts (CAF) (Zeisberg et al. 2007). CAFs secrete growth factors including insulinlike growth factor (IGF), VEGF, epithelial growth factor (EGF), HGF, and TGFb, as well as MMP and ECM components. TGFb is critical to communication between epithelial malignant cells and stromal cells. TGFb stimulates production of ECM components such as collagens, fibronectin, tenascin and basement membrane components such as laminin. TGF regulates expression of MMPs and induces expression of protease inhibitors such as tissue inhibitors of metalloproteases (TIMPs) and plasminogen activator inhibitor-1 (PAI-1). TGFb is important in regulation of immune system response to malignancy. Tumor associated macrophages (TAM) are a major source of pro-angiogenic factors and are involved in tumor invasion. TGFb stimulates chemotaxis of monocytes, macrophage precursors (Naber et al. 2008). In malignant disease, the host immune
31 Secreted Growth Factors as Therapeutic Targets
719
system can promote tumor progression, invasion, and dissemination to distant sites. TGFb secreted by the tumor cells and stromal cells recruits leukocytes to secrete chemokines, growth factors, cytokines, and proteases which support the tumor and neutralizes the immune response to the tumor. As a potent immunosuppressant, TGFb, both directly and through the generation of regulatory T cells supports tumor growth (Moutsopoulos et al. 2008). Cells of the innate immune system contribute to the high concentrations of TGFb found in tumor masses. In addition, dendritic cell subpopulations secreting TGFb contribute to the generation of regulatory T cells that actively inhibit the activity of other T cells. Elevated levels of plasma TGFb are associated with advanced stage disease and may separate patients into prognostically high-risk populations. Anti-TGFb therapy could reverse the immunosuppressive effects of this cytokine on the host as well as decrease ECM formation, decrease angiogenesis, decrease osteolytic activity, and increase the sensitivity of the malignant cells to cytotoxic therapies and immunotherapies. Clinical trials of an inhibitor of TGFb receptor type I kinase activity and a TGFb neutralizing antibody are underway (Teicher 2007). In normal tissues, fibrosis is a failure of tissue remodeling that is the result of an excessive inflammatory response, representing an imbalance between enhanced production and deposition and impaired degradation of ECM components. Urokinase-type plasminogen activator (uPA) and TGFb1 are critical in ECM deposition and degradation (Philippou et al. 2008). The plasminogen activator system controls intravascular fibrin deposition and a variety of other physiologic and pathologic processes. In cancer, components of the plasminogen activator system are involved in tumor growth, invasion and metastasis through effects on angiogenesis and cell migration. Both plasminogen activators, tPA and uPA, are expressed by tumor cells. uPA with its receptor (uPAR) are involved in cellular functions, while tPA with its receptor annexin II on the endothelial cell surface regulates intravascular fibrin deposition. Among the inhibitors of fibrinolysis, PAI-1 has a major role in the pathogenesis of many vascular diseases as well as in cancer. Therapeutic intervention, either using plasminogen activators or agents directed against PAI-1, have shown encouraging results in experimental tumors (McMahon and Kwaan 2007). The human MMP family includes 23 enzymes which degrade the ECM and are associated with cancer invasion and metastasis. In recent studies, some MMPs such as collagenase 2 (MMP8), macrophage metalloelastase (MMP12) and matrilysin 2 (MMP26), showed a protective effect in different stages of cancer progression. Stromelysin1 (MMP3), gelatinase B (MMP9), stromelysin 3 (MMP11), and MMP19 may also be protective enzymes in some specific situations (Lopez-Otin and Matrisian 2007). MMPs are produced in latent forms are prototypically modified to be active. MMPs are classified into groups designated matrilysins, collagenases, stromelysins, gelatinases, membrane-type metalloproteinases, and others. Matrilysins (MMP-7, MMP-26) have the simplest structure, consisting of a signal peptide, a prodomain, and a catalytic domain with a zinc-binding site and are expressed in physiological and pathological cells in the endometrium, small intestine, breast, pancreas, liver, and prostate. Collagenases (MMP-1, MMP-8, MMP-13) and stromelysins have a hemopexin-like domain and regulate fetal bone
720
B.A. Teicher
development and fetal wound repair by rapidly remodeling ECM. MMP-13 is expressed in malignant melanoma and breast, head and neck, and bladder carcinoma. The gelatinases (MMP-2 and MMP-9) have three-fibronectin type II repeats within their catalytic domains and catalyze the degradation of fibral collagens after initial cleavage by collagenases. Gelatinases are key proteases in malignant tumor metastasis. The membrane-type metalloproteinases (MMP-14, MMP-15, MMP-16, MMP-17, MMP-24, MMP-25) have a glycosylphosphtidylinositol anchor that provides cell surface binding. MMP-2 is important in gastric, pancreatic, prostate, and breast cancer cell migration. MMP-2 expression is dependent on extracellular matrix metalloproteinase inducer (EMMPRIN). The detection of active MMP-2 alone or the rate of pro-MMP-2 and active MMP-2 has been used as an indicator of cancer metastasis. Modulation of MMP-2 expression and activation through specific inhibitors and activators may thus provide a new mechanism for breast cancer treatment. Degradation of the cellular network established by adhesion molecules such as E-cadherin or ALCAM/CD166 causes tumor tissue relaxation, increases metastasis, and correlates with shortened survival in patients with primary breast carcinoma (Jezierska and Motyl 2009). Secreted protein acidic and rich in cysteine (SPARC) is a prototypic matricellular protein. Matricellular proteins modulate cell–cell and cell–matrix interactions and are expressed during morphogenesis, development, tissue injury, and tissue remodeling. The SPARC-related protein sub-family share three domains: (1) an N-terminal acidic and low-affinity calcium-binding domain; (2) a disulfide-bonded, copper-binding follistatin domain (homologous to TGFb inhibitors activin and inhibin); and (3) a C-terminal extracellular calcium-binding domain. SPARC is counter-adhesive that is, SPARC impairs cell attachment to the ECM in a concentration-dependent manner. Upon exposure to SPARC, primary cultured cells lose focal adhesions and exhibit decreased cell spreading (Ledda et al. 1997). SPARC may affect the tumor–host interaction in the metastatic niche, where metastatic tumor cells migrate through the ECM. Understanding SPARC influence on tumor invasion and metastasis requires continued investigation (Clark and Sage 2008). In a clinical study, increased SPARC expression in pancreatic tumors resulted in improved response to albumin-bound paclitaxel (Smith et al. 2008). RANKL is the primary mediator of osteoclast formation, function, and survival. RANK and RANKL are expressed by cells involved in bone remodeling, by cells of the immune system, and by cells in other tissues. Bone loss is mediated by osteoclasts, cells whose formation, function and survival depend on the RANKL. RANKL binds RANK on pre- and mature osteoclasts and activates and maintains osteoclast-mediated bone resorption. In preclinical inflammatory disease models, inhibition of RANKL prevents bone loss and has no detectable effect on immune mediators or inflammation. A phase 2 clinical study of the fully human RANKLneutralizing antibody, denosumab, in postmenopausal women with low bone mineral density using showed increased bone mineral density similar adverse event frequency as placebo and open-label alendronate. A subset of patients in the trial tested for immunological markers had no differences in T, B, or NK cell numbers or in immunoglobulin levels across dose or treatment groups (Stolina et al. 2007).
31 Secreted Growth Factors as Therapeutic Targets
721
In another clinical trail, women with non-metastatic breast cancer and low bone mineral density who were receiving adjuvant aromatase inhibitor therapy, twiceyearly administration of denosumab led to significant increases in bone mineral density over 24 months at trabecular and cortical bone, with overall adverse event rate similar to those of placebo (Ellis et al. 2008). Denosumab has shown benefit in other breast cancer clinical trials (Fizazi et al. 2009; Lipton et al. 2008). Blockade of the RANKL–RANK pathway may be of benefit in additional cancers (Schwartz 2008; Sung et al. 2009; Roodman 2009).
Immune System Modulators Solid tumors are infiltrated by leukocytes. Interaction between malignant and blood borne cells have profound effects on tumor progression. Leukocytes account for up to 50% of the tumor cellular component; these are mainly lymphocytes and macrophages. The presence of immunocompetent cells in tumors is believed to indicate a host immunological response to the malignancy. In some tumor types (colon, ovarian, melanoma), high T lymphocyte density can be an indication of a better clinical outcome. In most tumors, a high macrophage density is an indication of increased angiogenesis, tumor invasion, and poor prognosis (Sica et al. 2008). Activated immune T cells can kill cancer cells. Cancer vaccines function by increasing the activated immune T cells in the tumor. Immunotherapeutic agents that can increase the number of activated tumor-directed T cells include: (1) dendritic cell activators and growth factors, (2) vaccine adjuvants, (3) T-cell stimulators and growth factors, (4) immune checkpoint inhibitors, and (5) agents to neutralize or inhibit suppressive cells, cytokines, and enzymes (Cheever 2008). A few immune system stimulators are secreted factors. Interleukin-15 (IL-15) is a T-cell growth factor, similar to interleukin-2 (IL-2). Preclinical data suggest IL-15 could improve the therapeutic effect of cancer vaccine and adoptive T-cell regimens. IL-15 inhibits antigen-induced T cells death, whereas IL-2 promotes antigen-induced cell death. IL-15 is secreted by dendritic cells, macrophages, and stromal cells. IL-15 acts on CD8 T cells, CD4 T cells, natural killer (NK) cells, and mast cells. IL-15 is required to maintain CD8 memory cells and NK cell development. It stimulates development of long-lived and highavidity CD8 T cells that kill tumor cells effectively (Cheever 2008). Interleukin-7 (IL-7) is a T-cell growth factor responsible for homeostatic expansion of naive cells. In mice and primates, IL-7 treatment induces dramatic expansion in peripheral T-cell numbers without obvious toxicity. In preclinical bone marrow transplant studies, IL-7 increases the rate and degree of immune reconstitution. In humans, IL-15 can reverse T-cell anergy. A phase I clinical trial was completed in patients with cancer (Rosenberg et al. 2006). Treatment with human recombinant IL-7 increased in total CD4 and CD8 T cells with modest increases in NK cells and with no change in mature B cells or increase in T regulatory cells (Tregs). IL-12 is a
722
B.A. Teicher
potent immune adjuvant. IL-12 promotes interferon-g release by IL-12R-expressing T and NK cells and induces T-helper 1 cell (Th1) polarization and proliferation of interferon-g-expressing T cells. IL-12 plays a central role in resistance to mycobacterial and intracellular pathogens. Phase I and II clinical trials IL-12 showed very modest efficacy alone with a handful of melanoma and renal cell cancer responses and proved challenging to administer safely (Younces et al. 2004). Flt3 ligand (Flt3L) is a hematopoietic growth factor that stimulates proliferation and differentiation of dendritic cell progenitors, especially interferon-producing killer and plasmacytoid dendritic cells. Administration of Flt3L increases dendritic cells in circulating blood, secondary lymphoid tissues and tumors. A Phase I study examined Flt3L alone, with peptide vaccines, as dendritic cell stimulators, and after bone marrow transplant and increases in circulating dendritic cells occur; however, studies were too small and variable to determine clinical efficacy (Fong et al. 2001). When activated CD4+ T cells develop into varied T helper cell subsets with specific cytokine profiles and effector functions. IL-17-producing effector T helper cells, Th17 cells, produce IL-17, IL-17F, IL-21, and IL-22, resulting in massive tissue reaction because IL-17 and IL-22 receptors are widely expressed. Because TGF-b is involved in the differentiation of Th17 cells, similar to CD4+CD25+Foxp3+ regulatory T cells (Tregs) (Korn et al. 2009). Chemokines are key molecules involved in the migration and homeostasis of immune cells (Krieg and Boyman 2009). The predominant tumor infiltrating cells are macrophages and CD8+ T lymphocytes (Sica et al. 2008; Balkwill 2004). Specifically, monocyte chemoattractant protein 1 (MCP1; CCL2) localizes to tumor epithelial areas. MCP1 levels correlate with the numbers of lymphocytes and macrophages that localize in the same area. RANTES (CCL5) localizes with tumor-infiltrating leukocytes and the CCL5 concentrations reflect the extent of CD8+ T-lymphocyte infiltration. High numbers of tumor-infiltrating leukocytes and production of chemokines that attract leukocytes is a poor prognostic sign in human breast cancer, where CCL5 and CCL2 correlate with tumor progression, macrophage infiltrates, lymph-node metastasis, and clinical aggressiveness. In esophageal squamous-cell carcinoma, CCL2 expression was associated with macrophage infiltration, tumor cell invasion and tumor vascularity.
Malignant Cell Growth Factors Epidermal growth factor receptor (EGFR, HER1), HER2/neu (ErbB2), HER3 (ErbB3) and HER4 (ErbB4), the EGFR family of receptor tyrosine kinases and their secreted protein ligands comprise an important growth factor pathway in cancer. A subset of breast cancers are driven by overactive EGFR or HER2 tyrosine kinases and continued activity of these pathways stimulates cancer progression (Sergina et al. 2007). In humans, there are more than 30 ligands and the EGFR family of four receptors initiates a complex, multi-layered signal-transduction network. The varied activated receptor–ligand complexes produce cellular responses differing strength and type. Multiple processes can modulate EGFR
31 Secreted Growth Factors as Therapeutic Targets
723
signal transduction including receptor heterodimerization and endocytosis. The multiple ligands and receptors allow the EGFR signaling system with greater specificity and a large repertoire of responses. The four receptors can potentially form ten distinct homo- and heterodimers that are activated by different ligands (Fig. 31.1) (Yarden 2001). Epidermal growth factor (EGF) and transforming growth factor-a (TGFa) are two predominant EGFR ligands. The binding of these ligands to the EGFR extracellular domain results in receptor monomer dimerization forming either EGFR homodimers or heterodimers with other HER proteins. Under normal conditions, ligand binding results in EGFR tyrosine kinase activation. In tumor cells a variety of other events can initiate EGFR tyrosine kinase activity, for example, certain mutations in the EGFRgene such as the EGFRvIII mutation, result in constitutive activation of EGFR kinase (Riely et al. 2006). The activated EGFR pathway supports solid tumor malignant progression. TGFa and EGF induce angiogenesis by up-regulating the expression of VEGF in tumor cells. Increased microvessel density is found in tumors with an activated EGFR pathway. An activated EGFR pathway also alters cell–cell adhesion, in a manner which favors the up-regulation or activation of MMP and tumor cell motility and invasion of adjacent and distal tissues (Herbst and Bunn 2003; Ciardiello et al. 2001). Although the functioning of signaling pathways requires the presence both ligand and receptor, many preclinical and clinical studies test only for the presence of the receptor. Measurements of EGFR expression do not reliably predict therapeutic responses to EGFR inhibitors (Gusterson and Hunter 2009). In one study, the ligand TGFa was detected in only 2.7% of tumors and the low rate of expression of TGFa was associated with the low rate of response to EGFR tyrosine kinase inhibitors. It is likely that the patients most likely to benefit from treatment with EGFR tyrosine kinase inhibitors are those whose tumors express both the receptor(s) and the ligands (Onn et al. 2004). Interestingly, a connection between response to EGFR inhibitors and KRAS mutational status has been made. Now, all patients with metastatic colorectal carcinoma who are candidates for anti-EGFR antibody therapy must have their tumor tested for KRAS mutations. If KRAS mutation in codon 12 or 13 is detected, then patients with metastatic colorectal carcinoma should not receive anti-EGFR antibody therapy as part of their treatment (Allegra et al. 2009). The neuregulins are the largest subclass of EGFR ligands. Neuregulins were identified by searching for HER2 receptor activators, thus, neuregulins are also called heregulins (Montero et al. 2008). The neuregulins can up-regulate the number of acetylcholine receptors in neuromuscular junctions and therefore these factors are also called acetylcholine receptor-inducing activity (ARIA). Other names for neuregulins include Neu differentiation factor or glial growth factors due to their breast and glial cells activities. The neuregulins are synthesized as membrane-bound, biologically active growth factors that bind to the HER/ErbB receptors. Preclinically, increased expression and function of neuregulins can induce cancer. They are potent mitogens for cell expressing HER receptors. In mice overexpression of neuregulins in the mammary tissue can generate adenocarcinomas and stimulate breast cancer metastatic spread. Neuregulin expression is detected in several neoplasias and tends to correlate with response to HER receptor-targeting treatments such as trastuzumab. The growth and progression of many ovarian cancers are regulated by the HER/EGFR
X
3
Src
1
EGF
Cbl
2
SP1
PKC
PLC
1
Migration
1
Epiregulin
2
BAD
Akt
PI3K
S6K
2
2
Growth
FOS
ELK
RAF MEK MAPK JUN
SHC
X
3
2
STAT
Adhesion
EGR1
NCK
X
3
NGR1
PAK JNKK JNK
SOS
RAS-GDP GRB2
4
Amphiregulin
RAS-GTP
Gap
1
HB-EGF
SHP2
4
MYC
Beta-cellulin
4
RAC
VAV
NGR2
GRB7
4
Crk
3
XX
3
NGR4
Differentiation
ABL
4
NGR3
Fig. 31.1 The epidermal growth factor pathway includes of many ligands and a system of receptors which can homo-dimerize and hetero-dimerize. These ligand– receptor complexes can trigger multiple intracellular signal transduction routes to reach activation of nuclear transcription factors that usually result in cell growth (adapted from Yarden (2001))
Apoptosis
1
TGF-α α
724 B.A. Teicher
31 Secreted Growth Factors as Therapeutic Targets
725
family of receptor tyrosine kinase pathway. The receptors are activated by the ligands especially TGFa (activating HER1) and heregulin (activating HER3 and HER4) (Fig. 31.1). HER2 (HER2/neu) has no direct ligand and is activated via either homodimerization or heterodimerization with HER1, HER3 or HER4. Through heterodimerization with other members of the HER family, TGFa and heregulin (HRGh1) stimulate the growth of ovarian cancer cells (Mullen et al. 2007). HER2/neu overexpression is associated with increased tumor progression and metastasis. HER2/neu receptors are important in heregulin-induced angiogenesis. Overexpression of the HER2/neu receptor alone results in increased secretion of VEGF and exposure to heregulin enhances VEGF secretion in breast cancer cells (Konecny et al. 2004). The neuregulins are implicated in resistance to anti-HER therapies, so targeting neuregulins therapeutically useful in neoplastic diseases in which these ligands contribute to tumor progression (Montero et al. 2008). The hypothalamic decapeptide gonadotropin-releasing hormone (GnRH) controls pituitary gonadotropin secretion. GnRH hormone and receptor are also expressed in extra-pituitary tissues and tumor cells including epithelial ovarian cancers. Outside of the CNS, the GnRH pathway may function as an autocrine regulatory system. GnRH and several synthetic analogs have a direct antiproliferative effect on ovarian cancer cell lines. This effect is attributed to multiple steps in the GnRH signaling cascade such as cell cycle arrest at G0/G1. In ovarian cancer cells, GnRH receptors appear to couple to the pertussis toxin-sensitive protein Gai activating a protein phosphatase, which blocks growth factor-induced mitogenic signals. Recently, a regulatory role of GnRH analogs in ovarian cancer cell migration and invasion emerged. GnRH actions in ovarian cancer indicate that clinical application of GnRH analogs in ovarian cancer patients may be useful (So et al. 2008). Activation of the HGF/MET signaling pathway often has a role in oncogenesis, cancer metastasis, and drug resistance. The MET gene encodes a high-affinity receptor for HGF (also called scatter factor). The transforming properties of MET were identified first in a human osteosarcoma cell line following chemically induced mutagenesis. HGF is secreted by mesenchymal cells and MET is widely expressed by epithelial cancer cells. HGF binding to MET induces receptor homodimerization, phosphorylation of the tyrosine kinase domain and activation of MET-mediated signaling. HGF/MET signaling is essential during embryogenesis and is important in normal adult cells including hepatocytes, renal tubule cells, and myoblasts. Deregulation of the MET pathway occurs in many human malignancies. Sustained MET activation has been well characterized in preclinical models. Activation of HGF/MET signaling promotes cell invasiveness and triggers metastases through direct involvement of angiogenic pathways. HGF can stimulate endothelial cell proliferation and migration through induction of VEGF expression and down-regulation of thrombospondin-1 (You and McDonald 2008). Altered HGF secretion occurs in both solid and hematologic malignancies. Both malignant and mesenchymal cells can secrete increased HGF leading to paracrine and/or autocrine mechanisms of receptor activation. Inhibition of HGF/MET signaling represents a promising cancer treatment either alone or in combination therapy regimens. Clinical trials with agents targeting HGF/MET signaling are underway. A key issue
726
B.A. Teicher
in the clinical studies is appropriate patient selection strategies. Methods for assessment of HGF/MET overexpression and MET gene amplification are under development. Understanding the other key activated signaling pathways that occur concurrently with HGF/MET activation may allow rational combination therapeutic strategies (Toschi and Janne 2008). Insulin and IGFs differ from many other regulatory peptides that are relevant to cancer because they regulate whole organism cellular functions. Insulin and IGFs are tissue growth factors and hormones regulating whole organism growth and energy metabolism. Insulin is produced by pancreatic b-cells and is distributed to tumors through the circulation. Circulating insulin-like growth factor 1 (IGF1) and insulin-like growth factor 2 (IGF2) are produced in the liver as well as by tumor cells and stromal fibroblasts. IGF1 and 2 function through autocrine, paracrine, or endocrine mechanisms (Pollack 2008). The IGF1 receptor (IGF1R) is activated by IGF1 and IGF2. In cells that express both IGF1R and insulin receptor heterodimers form by association of insulin half-receptor with IGF half-receptor. These heterodimers are similarly activated by IGF1 and IGF2 (Chitnis et al. 2008). Insulin and IGF are relevant to cancer. Several therapeutic candidates that target IGF1 signaling have antineoplastic activity in preclinical tumor models as single agents and in combination regimens. Many tumors have altered levels of IGF1R, IGF2 and/or IGF binding proteins. Tumors can also express insulin receptor and heterodimers. Changes in IGF may be initiating events in tumorigenesis. Individuals with highnormal circulating IGF1 are at increased risk of later development of common solid tumors, possibly because IGFs stimulate neoplastic progression of occult lesions. Tumor growth can be stimulated by increasing circulating IGFl in experimental models. Rigorous prospective studies provided evidence for a relationship between circulating IGF1 and risk of developing prostate, breast, colorectal or other cancers, such that individuals at the high end of the normal range of serum IGF1 had more than double the risk of a cancer diagnosis than those at the low end of the normal range (Shukla et al. 2008; Kleinberg et al. 2009; Kawada et al. 2006; Rowlands et al. 2009). Several investigational antibody therapeutics directed toward IGFR are in early clinical trial including AmG479 (Amgen), AvE1642 (Sanofi-Aventis), A12 (Lilly), mK0646 (Merck) and R1507 (Roche). A compensatory increase in the circulating concentrations of growth hormone and IGF1 occurs on administration of IGF1R-specific antibodies (Pollack 2008; Rodon et al. 2008). Netrin-1 is a member of a large family of conserved 60–80 kDa proteins with sequence homology to laminins. Secreted netrins, netrin-1, -3, and -4 are expressed in mammals. The functions of netrin-1 are mediated by two families of receptors: the UNC5 family (human UNC5H1-H4 or UNC5A-D) and DCC, including the paralog neogenin, which are members of the immunoglobulin superfamily. Netrin-1 and the receptors DCC and UNC5 are involved in normal development in the brain– gut axis by regulating axonal guidance and survival, pancreatic, and mammary gland morphogenesis and angiogenesis. Netrin receptors are expressed by vascular endothelial cells and are implicated in the control of vascular endothelial and smooth muscle cell morphogenesis. Tumor development is associated with the loss of netrin-1 receptors and increased production of netrin-1, thus escaping the control
31 Secreted Growth Factors as Therapeutic Targets
727
of homeostasis mediated by the balance between receptor and ligand providing a selective advantage for tumor development. Overexpression of netrin-1 in the gastrointestinal tract of transgenic mice is associated with tumor initiation and progression. In human colorectal tumors netrin-1 is only rarely overexpressed (7%); however, loss of DCC/UNC5 is more frequent. Bernet et al. showed that 60% of breast cancers with distant metastases at diagnosis and 37% of tumors with axillary node involvement show increased netrin-1 mRNA expression and immunohistochemistry (Bernet and Fitamant 2008). Netrin-1 may be a useful therapeutic target in cancer. S100A4 (mts1, pEL-98, 18A2, p9Ka, CAPL, calvasculin, and FSPl) belongs to the S100 family of calcium binding proteins. S100A4 expression in tumor cells is correlated with an aggressive metastatic phenotype. Up-regulation of S100A4 in tumor and stroma cells is associated with poor prognosis and survival of patients with cancer. Immune cells including macrophages, neutrophils, certain lymphocytes, dendritic and mast cells, and human endothelial cells express and release S100A4 into the extracellular space (Grigorian et al. 2008). S100A4 does not have a signal sequence and is released from the cell through an atypical pathway. Extracellular S100A4 can affect tumor progression by stimulation of angiogenesis, cell motility, upregulation of MMPs, modulation of tumor-related transcription factors, and stromal factors. It is not known whether S100A4 has a cell surface receptor. However, the short exposure time (1–2 min) needed to generate the response of neurons to S100A4 indicate receptor-mediated signaling is likely (Helfman et al. 2005). The Wnt family of secreted growth factors has key roles in directing cell patterning both during development and in adult tissues (Musgrove 2004). There are 19 Wnt proteins with a highly regulated pattern of expression and distinct roles in development and tissue homoeostasis. Wnts activate b-catenin signaling and consequently modulate the expression of specific target genes that regulate cell proliferation, apoptosis, and cell fate. Wnts are involved in oncogenesis. Early, Wnt-1 was found to be overexpressed in mammary epithelial adenocarcinomas. Since then, deregulated Wnt expression and Wnt signaling has been found in multiple epithelial cancers (Giles et al. 2003). The Wnt signaling pathways are activated through binding of a Wnt to one of nine seven-pass transmembrane receptors called the Frizzled receptor proteins (Fz) and LRP5/6 coreceptors. Following receptor binding, Wnt signals are transmitted by Wnt receptors and Dishevelled (Dvl) which triggers the disruption of the complex containing APC, Axin, GSK-3, and b-catenin, preventing phosphorylation-dependent b-catenin degradation (Jin et al. 2008). The Wnts are divided into two classes by signaling through the “canonical” or the “non-canonical” signaling pathway. Canonical Wnts are thought to activate a signal-transduction pathway that induces nuclear accumulation and transcriptional activation of b-catenin (Giles et al. 2003). The non-canonical signaling pathway describes all Wnt-activated cellular signaling pathways that do not promote b-catenin-mediated transcription. Emerging data suggest that these pathways are not as autonomous (Hopfner et al. 2008; Sirica 2008; McDonald and Silver 2009). For example, Wnts activate erbB signaling in addition to stimulating Wnt and b-catenin signaling, elucidating another level of regulation by Wnt family members and providing
728
B.A. Teicher
increased interest in these and other developmental pathways in breast cancer (Musgrove 2004).
References Achen MG, Stacker SA (2008) Molecular control of lymphatic metastasis. Ann N Y Acad Sci 1131:225–234 Allegra CJ, Jessup M, Somerfield MR, Hamilton SR, Hammond EH, Hayes DF, McAllister PK, Morton RF, Schilsky RL (2009) American society of clinical oncology provisional clinical opinion: testing for KRAS gene mutations in patients with metastatic colorectal carcinoma to predict response to anti-epidermal growth factor receptor monoclonal antibody therapy. J Clin Oncol 27:2091–2096 Balkwill F (2004) Cancer and the chemokine network. Nat Rev Cancer 4:540–550 Beenken A, Mohammadi M (2009) The FGF family: biology, pathophysiology and therapy. Nat Rev Drug Discov 8:235–253 Bergers G, Hanahan D (2008) Modes of resistance to anti-angiogenic therapy. Nat Rev Cancer 8:592–603 Bernet A, Fitamant J (2008) Netrin-1 and its receptors in tumor growth promotion. Expert Opin Ther Targets 12:995–1007 Cao Y, Cao R, Hedlund E-M (2008) Regulation of tumor angiogenesis and metastasis by FGF and PDGF signaling pathways. J Mol Med 86:785–789 Cheever MA (2008) Twelve immunotherapy drugs that could cure cancers. Immunol Rev 222:357–368 Chitnis MM, Yuen JSP, Protheroe AS, Pollack M, Macaulay VM (2008) The type 1 insulin-like growth factor receptor pathway. Clin Cancer Res 14:6364–6370 Ciardiello F, Caputo R, Bianco R, Damiano V, Fontanini G, Cuccato S, de Placido S, Bianco AR, Tortora G (2001) Inhibition of growth factor production and angiogenesis in human cancer cells by ZD1839 (Iressa), a selective epidermal growth factor receptor tyrosine kinase inhibitor. Clin Cancer Res 7:1459–1465 Clark CJ, Sage EH (2008) A prototypic matricellular protein in the tumor microenvironment – where there’s SPARC, there’s fire. J Cell Biochem 104:721–732 Dvorak HF, Brown LF, Detmar M, Dvorak AM (1995) Vascular permeability factor/vascular endothelial growth factor, microvascular hyperpermeability, and angiogenesis. Am J Pathol 146:1029–1039 Ellis LM, Hicklin DJ (2008) VEGF-targeted therapy: mechanisms of anti-tumor activity. Nat Rev Cancer 8:579–596 Ellis GK, Bone HG, Chlebowski R, Paul D, Spadafora S, Smith J, Fan M, Jun S (2008) Randomized trial of denosumab in patients receiving adjuvant aromatase inhibitors for nonmetastatic breast cancer. J Clin Oncol 26:4875–4882 Fischer C, Jonckx B, Mazzone M, Zacchigna S, Loges S, Pattarini L, Chorianopoulos E, Liesenborghs L, Koch M, De Mol M, Autiero M, Wyns S, Plaisance S, Moons L, van Rooijen N, Giacca M, Stassen JM, Dewerchin M, Collen D, Carmeliet P (2007) Anti-PlGF inhibits growth of VEGF(R)inhibitor-resistant tumors without affecting healthy vessels. Cell 131:463–475 Fischer C, Mazzone M, Jonckx B, Carmeliet P (2008) FLT1 and its ligands VEGFB and PlGF: drug targets for anti-angiogenic therapy? Nat Rev Cancer 8:942–956 Fizazi K, Lipton A, Mariette X, Body JJ, Rahim Y, Gralow JR, Gao G, Wu L, Sohn W, Jun S (2009) Randomized phase II trial of denosumab in patients with bone metastases from prostate cancer, breast cancer or other neoplasms after intravenous bisphosphonates. J Clin Oncol 27:1564–1571 Folkman J (1995) Angiogenesis in cancer, vascular, rheumatoid and other diseases. Nat Med 1:27–31
31 Secreted Growth Factors as Therapeutic Targets
729
Fong L, Hou Y, Rivas A, Benike C, Yuen A, Fisher GA, Davis MM, Engleman EG (2001) Altered peptide ligand vaccination with Flt3 ligand expanded dendritic cells for tumor immunotherapy. Proc Natl Acad Sci U S A 98:8809–8814 Giles RH, van Es JH, Clevers H (2003) Caught up in a Wnt storm: Wnt signaling in cancer. Biochim Biophys Acta 1653:1–24 Grigorian M, Ambartssumian N, Lukanidin E (2008) Metastasis-inducing S100A4 protein: implication in non-malignant human pathologies. Curr Mol Med 8:492–496 Guleng B, Tateishi K, Ohta M, Kanai F, Jazag A, Ijichi H, Tanaka Y, Washida M, Morikane K, Fukushima Y, Yamori T, Tsuruo T, Kawabe T, Miyagishi M, Taira K, Sata M, Omata M (2005) Blockade of the stromal cell-derived factor-1/CXCR4 axis attenuates in vivo tumor growth by inhibiting angiogenesis in a vascular endothelial growth factor-independent manner. Cancer Res 65:5864–5871 Gusterson BA, Hunter KD (2009) Should we be surprised at the paucity of response to EGFR inhibitors? Lancet Oncol 10:522–527 Helfman DM, Kim EJ, Lukanidin E, Grigorian M (2005) The metastasis associated protein S100A4: role in tumor progression and metastasis. Br J Cancer 92:1955–1958 Heng DYC, Bukowski RM (2008) Anti-angiogenic targets in the treatment of advanced renal cell carcinoma. Curr Cancer Drug Targets 8:676–682 Herbst RS, Bunn PA (2003) Targeting the epidermal growth factor receptor in non-small cell lung cancer. Clin Cancer Res 9:5813–5824 Hopfner M, Schuppan D, Scherubl H (2008) Targeted medical therapy of biliary tract cancer: recent advances and future perspectives. World J Gastroenterol 14:7021–7032 Jezierska A, Motyl T (2009) Matrix metalloproteinase-2 involvement in breast cancer progression: a mini-review. Med Sci Monit 15:RA32–RA40 Jin T, Fantus IG, Sun J (2008) Wnt and beyond Wnt: multiple mechanisms control the transcriptional property of b-catenin. Cell Signal 20:1697–1704 Kawada M, Inoue H, Masuda T, Ikeda D (2006) Insulin-like growth factor I secreted from prostate stromal cells mediate tumor-stromal cell interactions of prostate cancer. Cancer Res 66:4419–4425 Kleinberg DL, Wood TL, Furth PA, Lee AV (2009) Growth hormone and insulin-like growth factor-I in the transition from normal mammary development to preneoplastic mammary lesions. Endocrine Rev 30:51–74 Konecny GE, Meng YG, Untch M, Wang HJ, Bauerfeind I, Epstein M, Stieber P, Vernes JM, Gutierrez J, Hong K, Beryt M, Hepp H, Slamon DJ, Pegram MD (2004) Association between HER-2/neu and vascular endothelial growth factor predicts clinical outcome in primary breast cancer patients. Clin Cancer Res 10:1706–1716 Kopetz S, Hoff PM, Eng C, Overman MJ, Glover KY, Chang DZ, Wolff RA, Abbruzzese JL, Ellis LM, Heymach JV (2009) Levels of angiogenic cytokines prior to disease progression in metastatic colorectal cancer patients treated with bevacizumab. Proceedings of the ASCO 2009; Gastrointestinal cancer symposium, Abstr 292 Korn T, Bettelli E, Oukka M, Kuchroo VK (2009) IL-17 and Th17 cells. Annu Rev Immunol 27:485–517 Krieg C, Boyman O (2009) The role of chemokines in cancer immune surveillance by the adaptive immune system. Semin Cancer Biol 19:76–83 Ledda MF, Adris S, Bravo AI, Kairiyama C, Bover L, Chernajovsky Y, Mordoh J, Podhajcer OL (1997) Suppression of SPARC expression by antisense RNA abrogates the tumorigenicity of human melanoma cells. Nat Med 3:171–176 Lipton A, Steger GG, Figueroa J, Alvarado C, Solal-Celigny P, Body JJ, de Boer R, Berardi R, Gascon P, Tonkin KS, Coleman RE, Paterson AHG, Gao GM, Kinsey AC, Peterson MC, Jun S (2008) Extended efficacy and safety of denosumab in breast cancer patients with bone metastases not receiving prior bisphosphonate therapy. Clin Cancer Res 14:6690–6696 Loges S, Roncal C, Carmeliet P (2008) Development of targeted angiogenic medicine. J Thromb Haemost 7:21–33 Lopez-Otin C, Matrisian LM (2007) Emerging roles of proteases in tumor suppression. Nat Rev Cancer 7:800–808
730
B.A. Teicher
Martin SK, Dewar AL, Farrugia AN, Horvath N, Gronthos S, To LB, Zannettino ACW (2006) Tumor angiogenesis is associated with plasma levels of stromal-derived factor-1a in patients with multiple myeloma. Clin Cancer Res12:6973–6977 McDonald SL, Silver A (2009) The opposing roles of Wnt5a in cancer. Br J Cancer 101:209–214 McMahon B, Kwaan HC (2007) The plasminogen activator system and cancer. Pathophysiol Haemost Thromb 36:184–194 Montero JC, Rodriguez-Barrueco R, Ocana A, Diaz-Rodriguez E, Esparis-Ogando A, Pandiella A (2008) Neuregulins and cancer. Clin Cancer Res 14:3237–3241 Motzer RJ, Michaelson MD, Redman BG, Hudes GR, Wilding G, Figlin RA, Ginsberg MS, Kim ST, Baum CM, DePrimo SE, Li JZ, Bello CL, Theuer CP, George DJ, Rini BI (2006) Activity of SU11248, a multitargeted inhibitor of vascular endothelial growth factor receptor and platelet-derived growth factor receptor, in patients with metastatic renal cell carcinoma. J Clin Oncol 24:16–24 Moutsopoulos NM, Wen J, Wahl SM (2008) TGF-b and tumors – an ill-fated alliance. Curr Opin Immunol 20:234–240 Mueller MM, Fusenig, NE (2004) Friends or foes – bipolar effects of the tumour stroma in cancer. Nat Rev Cancer 4:839–849 Mullen P, Cameron DA, Hasmann M, Smyth JF, Langdon SP (2007) Sensitivity to pertuzumab (2C4) in ovarian cancer models: cross-talk with estrogen receptor signaling. Mol Cancer Ther 6:93–100 Musgrove EA (2004) Wnt signaling via the epidermal growth factor receptor: a role in breast cancer? Breast Cancer Res 6:65–68 Naber HPH, ten Dijke P, Pardali E (2008) Role of TGF-b in the tumor stroma. Curr Cancer Drug Targets 8:466–472 Norden AD, Drappatz J, Wen PY (2008) Novel anti-angiogenic therapies for malignant gliomas. Lancet Neurol 7:1152–1160 Onn A, Correa AM, Gilcrease M, Isobe T, Massarelli E, Bucana CD, O’Reilly MS, Hong WK, Fidler IJ, Putnam JB, Herbst RS (2004) Synchronous overexpression of epidermal growth factor receptor and HER2-neu protein is a predictor of poor outcome in patients with stage I non-small cell lung cancer. Clin Cancer Res 10: 136–143 Otrock ZK, Makarem JA, Shamseddine AI (2007) Vascular endothelial growth factor family of ligands and receptors: review. Blood Cells Mol Dis 38:258–268 Petit I, Jin D, Rafii S (2007) The SDF-1-CXCR4 signaling pathway: a molecular hub modulating neo-angiogenesis. Trends Immunol 28:300–307 Philippou A, Maridaki M, Koutsilieris M (2008) The role of urokinase-type plasminogen activator (uPA) and transforming growth factor beta 1 (TGFb1) in muscle regeneration. In Vivo 22:735–750 Pinkas J, Teicher BA (2006) TGF-b in cancer and as a therapeutic target. Biochem Pharmacol 72: 523–529 Pollack M (2008) Insulin and insulin-like growth factor signaling in neoplasia. Nat Rev Cancer 8:915–928 Potiron VA, Roche J, Drabkin HA (2009) Semaphorins and their receptors in lung cancer. Cancer Lett 273:1–14 Riely GJ, Miller VA (2007) Vascular endothelial growth factor trap in non-small cell lung cancer. Clin Cancer Res 13(15 suppl):4623s–4627s Riely GJ, Politi KA, Miller VA, Pao W (2006) Update on epidermal growth factor receptor mutations in non-small cell lung cancer. Clin Cancer Res 12:7232–7241 Rodon J, DeSantos V, Ferry RJ Jr, Kurzrock R (2008) Early drug development of inhibitors of the insulin-like growth factor-I receptor pathway: lessons from the first clinical trials. Mol Cancer Ther 7:2575–2588 Roodman GD (2009) Pathogenesis of myeloma bone disease. Leukemia 23:435–441
31 Secreted Growth Factors as Therapeutic Targets
731
Roorda BD, ter Elst A, Kamps WA, de Bont ESJM (2009) Bone marrow-derived cells and tumor growth: contribution of bone marrow-derived cells to tumor micro-environments with focus on mesenchymal stem cells. Crit Rev Oncol Hematol 69:187–198 Rosenberg SA, Sportes C, Ahmadzadeh M, et al (2006) IL-7 administration to humans leads to expansion of CD8 and CD4 cells but a relative decrease of CD4T-regulatory cells. J Immunother 29:313–319 Rosen LS, Kurzrock R, Mulay M, Van Vugt A, Purdom M, Ng C, Silverman J, Koutsoukos A, Sun YN, Bass MB, Xu RY, Polverino A, Wiezorek JS, Chang DD, Benjamin R, Herbst RS (2007) Safety, pharmacokinetics and efficacy of AMG 706, an oral multikinase inhibitor, in patients with advanced solid tumors. J Clin Oncol 25:2369–2376 Roskoski R Jr (2008) VEGF receptor protein-tyrosine kinases: structure and regulation. Biochem Biophys Res Commun 375:287–291 Rowlands MA, Gunnell D, Harris R, Vatten LJ, Holly JMP, Martin RM (2009) Circulating insulinlike growth factor peptides and prostate cancer risk: a systematic review and meta-analysis. Int J Cancer 124:2416–2429 Royston D, Jackson DG (2009) Mechanisms of lymphatic metastasis in human colorectal adenocarcinoma. J Pathol 217:608–619 Schwartz GG (2008) prostate cancer, serum parathyroid hormone and the progression of skeletal metastases. Cancer Epidemiol Biomarkers Prev 17:478–483 Sergina NV, Rausch M, Wang D, Blair J, Hann B, Shokat KM, Moasser MM (2007) Escape from HER-family tyrosine kinase inhibitor therapy by the kinase-inactive HER3. Nature 445:437–441 Shojaei F, Wu X, Zhong C, Yu L, Liang XH, Yao J, Blanchard D, Bais C, Peale FV, van Bruggen N, Ho C, Ross J, Tan M, Carano RAD, Meng YG, Ferrara N (2007) Bv8 regulates myeloidcell-dependent tumor angiogenesis. Nature 450:825–834 Shukla V, Coumoul X, Vassilopoulos A, Deng CX (2008) IGF signaling pathway as a selective target of familial breast cancer therapy. Curr Mol Med 8:727–740 Sica A, Allavena P, Mantovani A (2008) Cancer related inflammation: the macrophage connection. Cancer Lett 264:204–215 Sipkins DA, Wei X, Wu JW, Runnels JM, Cote D, Means TK, Luster AD, Scadden DT, Lin CP (2005) In vivo imaging of specialized bone marrow endothelial microdomains for tumor engraftment. Nature 435:969–973 Sirica AE (2008) Role of ErbB family receptor tyrosine kinases in intrahepatic cholangiocarcinoma. World J Gastroenterol 14:7033–7058 Smith LS, Drengler RL, Wood TE, Laheru DA, Hidalgo M, Borad MJ, Trieu V, Knauer D, Desai N, Von Hoff DD (2008) SPARC and CA19-9 as biomarkers in patients with advanced pancreatic cancer treated with nab paclitaxel plus gemcitabine. J Clin Oncol 26 (Abstr 15592) So WK, Cheng JC, Poon SL, Leung PCK (2008) Gonadotropin-releasing hormone and ovarian cancer: a functional and mechanistic overview. FEBS J 275:5496–5511 Stacker SA, Achen MG (2008) From anti-angiogenesis to anti-lymphangiogenesis: emerging trends in cancer therapy. Lymphatic Res Biol 6: 165–172 Stolina M, Kostenuik PJ, Dougall WC, Fitzpatrick LA, Zack DJ (2007) RANKL inhibition: from mice to men (and women). In: Choi Y (ed.) Osteoimmunology. Springer, NY, pp 143–150 Sung B, Murakami A, Oyajobi BO, Aggarwal BB (2009) Zerumbone abolishes RANKL-induced NF-kB activation, inhibits osteoclastogenesis, and suppresses human breast cancer-induced bone loss in athymic nude mice. Cancer Res 69:1477–1484 Teicher BA (2007) Transforming growth factor-b and the immune response to malignant disease. Clin Cancer Res 13:6247–6251 Toschi L, Janne PA (2008) Single-agent and combination therapeutic strategies to inhibit hepatocyte growth factor/MET signaling in cancer. Clin Cancer Res 14:5941–5946 Yarden Y (2001) The EGFR family and its ligands in human cancer: signaling mechanisms and therapeutic opportunities. Eur J Cancer 37:3–8
732
B.A. Teicher
You WK, McDonald DM (2008) The hepatocyte growth factor/c-Met signaling pathway as a therapeutic target to inhibit angiogenesis. BMB Rep 41:833–839 Younces A, Pro B, Robertson MJ, Flinn IW, Romaguera JE, Hagemeister F, Dang NH, Fiumara P, Loyer EM, Cabanillas FF, McLaughlin PW, Rodriguez MA, Samaniego F (2004) Phase II clinical trial of interleukin-12 in patients with relapsed and refractory non-Hodgkin’s lymphoma and Hodgkin’s disease. Clin Cancer Res 10:5432–5438 Zannettino ACW, Farrugia AN, Kortesidis A, Mannavis J, To LB, Martin SK, Diamond P, Tamamura H, Lapidot T, Fujii N, Gronthos S (2005) Elevated serum levels of stromal-derived factor-1a are associated with increased osteoclast activity and osteolytic bone disease in multiple myeloma patients. Cancer Res 65:1700–1709 Zeisberg EM, Potenta S, Xie L, Zeisberg M, Kalluri R (2007) Discovery of endothelial to mesenchymal transition as a source for carcinoma-associated fibroblasts. Cancer Res 67:10123–10128
Chapter 32
Adrenomedullin Rebecca G. Bagley
Abstract Adrenomedullin is a multifunctional regulatory and vasoactive peptide that was identified in the early 1990s. Since then, research into this secreted hormone related to malignant disease has been emerging. Adrenomedullin is present in some normal tissues and is upregulated in a multitude of cancers. Adrenomedullin overexpression has been detected in human adrenal, breast, endometrial, lung, ovarian, pancreatic, prostate, and renal cancers. In certain indications, adreno-medullin expression correlated to disease progression and therefore adrenomedullin could possibly serve as a prognostic biomarker. Adrenomedullin promoted tumor development in preclinical models that can be attributed in part to an increase in angiogenesis. Both cancer cell lines and endothelial cells exposed to exogenous adrenomedullin displayed enhanced proliferative, migratory, and invasive properties. Efforts to target adrenomedullin under certain conditions have generated results which suggest that neutralizing adrenomedullin activity or interfering with receptor activity may have potential as a novel strategy to treat cancer.
Structure and Function Adrenomedullin (AM) is a secreted peptide hormone that was originally isolated from human pheochromocytoma, a rare tumor of the adrenal gland, and AM is also abundant in the normal adrenal medulla (Kitamura et al. 1993). AM has vasodilatory properties that were discovered by monitoring elevated platelet cAMP. Adrenomedullin consists of 52 amino acids with one intramolecular disulfide bond resulting from the cleavage of a 185 amino acid protein (pre-proadrenomedullin), which also produces proadrenomedullin N-terminal 20 peptide (PAMP).
R.G. Bagley (*) Genzyme Corporation, 49 New York Ave, Framingham, MA 01701, USA e-mail:
[email protected] R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5_32, © Springer Science+Business Media, LLC 2010
733
734
R.G. Bagley
Adrenomedullin belongs to the calcitonin gene peptide superfamily due to the shared homology with the calcitonin gene-related peptide (CGRP) and amylin (Poyner et al. 2002). Several receptors for adrenomedullin have been reported. The first putative receptor described as L1 was isolated from rat lung. In COS-7 cells transfected with L1, there was an increase in cAMP upon exposure to AM (Kapas et al. 1995). However, these findings were subsequently thrown into question when the results could not be duplicated by other researchers (Kennedy et al. 1998). A second putative AM receptor, RDC-1, was recognized from a dog thyroid library and was shown to be able to bind both AM and CGRP (Kapas and Clark 1995). Additional research indicated that RDC-1 and AM were upregulated under hypoxic conditions in rat astrocytes in the blood–brain barrier (Ladoux and Frelin 2000). Subsequently, the calcitonin receptor-like receptor (CRLR) was shown to bind to AM (Njuki et al. 1993). CRLR is a seven-transmembrane, G-protein coupled receptor that forms heterodimers with receptor activity modifying protein (RAMP)-2 and -3 (Chakravarty et al. 2000; McLatchie et al. 1998). The vasodilatory properties of adrenomedullin were elucidated upon the discovery of AM in human pheochromocytoma tissue (Kitamura et al. 1993). Soon thereafter, the multifunctional characteristics of adrenomedullin emerged. The physiological and pathological roles of adrenomedullin were studied in genetically engineered mice that either lacked or overexpressed AM (Caron and Smithies 2001; Imai et al. 2001; Shindo et al. 2001). In homozygous knockout mice, the lack of AM proved to be embryonic lethal. However, the heterozygotes were viable and had a hypertensive phenotype compared with the wild-type mice. In transgenic mice that overexpressed AM, blood pressure was significantly lower than in the wild-type mice. The AM-transgenic mice were resistant to vascular injuries and lipopolysaccharide-induced septic shock. These results implied a significant role for AM in circulatory homeostasis and also indicated the importance of AM in vascular development during gestation. The role of AM in angiogenesis is further described below. In addition to maintaining physiological vascular homeostatsis and to supporting embryogenesis during pregnancy, adrenomedullin can regulate insulin secretion and blood glucose metabolism. A survey of pancreatic tissues from multiple species identified adrenomedullin protein in the pancreas of the rat, hamster, dog, and guinea pig (Martinez et al. 1996). mRNA transcripts for AM and its receptor were also present in multiple human adrenal and pancreatic cell lines that produce insulin. In the rat pancreas, in situ hybridization confirmed the expression of the AM receptor in the islets of Langherans. Using isolated rat islets, Martinez et al. demonstrated that AM can inhibit insulin secretion in a concentration-dependent manner and that exposure to an AM-neutralizing monoclonal antibody can result in an increase in insulin secretion. The co-expression of three receptors (L1, RDC-1, and CRLR) for AM in pancreatic b-cells is further evidence linking AM to the regulation of insulin release (Martinez et al. 2000). Adrenomedullin is a ubiquitous peptide with numerous biological functions that can also influence the CNS and renal function, and exert effects on the
32 Adrenomedullin
735
endocrine system. Several reports describe in detail the synthesis and secretion of AM, AM receptors and intracellular signal transduction pathways triggered by AM binding to its receptor, and AM-mediated activity in various organ systems (Hinson et al. 2000; Martinez and Cuttitta 1998). The remainder of this chapter will focus on the properties of AM related to angiogenesis, expression in tumors and in the tumor microenvironment, and the potential value of targeting AM with novel therapeutics.
Angiogenesis The ability of adrenomedullin to induce and regulate angiogenesis is well-documented in multiple preclinical models and in clinical specimens (Nikitenko et al. 2006). Transgenic mice were generated that lacked the AM gene or were heterozygous for AM to further elucidate significant functions of AM. In knockout mice, the absence of adrenomedullin gene expression was embryonic lethal (Caron and Smithies 2001). The mice died at mid-gestation with extreme hydrops fetalis and cardiovascular abnormalities. These abnormalities included overdeveloped ventricular trabeculae and underdeveloped arterial walls. In a separate study, investigators obtained similar results, noting a high mortality rate of embryos and abnormal vasculature causing hemorrhage and high blood pressure in the surviving embryos (Shindo et al. 2001). Detailed examination of the vessels revealed detached endothelial cells and incomplete basement membranes. The CRLR mediates AM signaling when the receptor is associated with RAMP-2 (Fernandez-Sauze et al. 2004). Additional evidence supporting the observation that adrenomedullin signaling is necessary for embryogenesis was revealed in knockout mice lacking CRLR or RAMP-2 genes (Fritz-Six et al. 2008). These mice also died mid-gestation after interstitial lymphedema occurred. In the knockout mice that were deficient in CRLR expression, the lack of AM signaling resulted in reduced lymphatic endothelial cell proliferation that caused abnormal jugular lymphatic vessels. Thus, these results establish a role for AM in the development of lymphatic vasculature. The placenta is a highly angiogenic tissue. Placental expression of adrenomedullin during pregnancy was investigated at the mRNA and protein levels (Moriyama et al. 2001). AM mRNA was detected in human placental trophoblastic tissues obtained during the first, second, and third trimesters. AM peptide was identified by immunohistochemical methods in the cytotrophoblasts but not in the syncytiotrophoblasts or endothelial cells of fetal stroma. AM peptide expression in the cytotrophoblasts was the highest during the first trimester and then decreased throughout the course of the pregnancy. In rare instances, choriocarcinomas, a trophoblastic cancer of the placenta, can develop during pregnancy. A choriocarcinoma cell line, JAr, was included in this study for analysis. The JAr cells expressed AM mRNA, secreted AM in culture, and also expressed a receptor for AM as demonstrated by the specific binding of radiolabeled rat AM to the cells.
736
R.G. Bagley
The effects of endothelial cell exposure in culture to adrenomedullin were investigated in several studies. The addition of AM to human umbilical vein endothelial cells (HUVEC) increased migration and tube formation in culture (FernandezSauze et al. 2004). Experiments with HUVEC and AM revealed that AM signaled through the Akt, ERK, and focal adhesion kinase pathways (Kim et al. 2003). Inhibitors to those pathways were able to suppress partially AM-induced HUVEC tube formation. In rat endothelial cells, AM functioned via autocrine and paracrine mechanisms as an apoptosis survival factor (Kato et al. 1997). In addition to adrenomedullin, the 20 amino acid peptide (PAMP) that results from the cleavage of pre-proadrenomedullin also possesses angiogenic properties. In a series of angiogenesis assays, PAMP was able to enhance the sprouting of new vessels in the chick embryo aortic ring assay and increase migration and cord formation in human microvascular endothelial cells (HMVEC) that expressed PAMP receptors (Martinez et al. 2004). Exposure of HMVEC to PAMP also increased molecular levels of other potent angiogenic factors such as VEGF, bFGF, and PDGF-C. In vivo, the angiogenic potential of adrenomedullin was demonstrated in a rabbit hindlimb ischemia model via gene delivery with a plasmid DNA (Nagaya et al. 2005). Transgenic expression of adrenomedullin resulted in an increase in capillary density, calf muscle blood pressure ratio, and blood flow measured by laser Doppler methodology. Similarly, recombinant human AM protein enhanced blood flow to the ischemic limb in the mouse hindlimb ischemia model and upregulated VEGF production in the affected tissue (Iimuro et al. 2004). These results suggest that AM could offer therapeutic value in ischemic disease. In mice, adrenomedullin stimulated angiogenesis in “angioreactors”, silicone tubes filled with matrix proteins, that were implanted subcutaneously in nude mice (Martinez et al. 2004). AM stimulated neo-vascularization in a traditional Matrigel plug assay and increased blood flow that was characterized by immunohistochemistry (IHC), hemoglobin assays, and laser Doppler perfusion image analysis (Kim et al. 2003; Miyashita et al. 2003). Analysis of the cellular components of the newly formed vessels in Matrigel plugs whereby angiogenesis was induced by exogenous AM revealed the recruitment of not only endothelial cells and pericytes but also myeloid precursor cells and macrophages (Kaafarani et al. 2009). This influx of cells that contribute to the tumor microenvironment was inhibited by an antibody directed toward the CLRL receptor. Adrenomedullin is involved in angiogenesis not only during embryogenesis but also during tumor development in a wide variety of human cancers. In renal cell carcinoma (RCC) cell lines and human biopsies, a positive correlation existed between AM and VEGF mRNA levels. Both AM and VEGF mRNA correlated with microvessel density (MVD) (Fujita et al. 2002; Deville et al. 2009). Under hypoxic conditions, the increase in AM mRNA was significantly greater than the increase in VEGF mRNA levels (10.6- to 26.7-fold increase vs. 1.5- to 1.9-fold) in several RCC cell lines. Similarly, hypoxia induced AM expression in hepatocellular carcinoma (HCC) (Park et al. 2008). AM expression in the HCC study also correlated with vascular invasion and N-cadherin expression by IHC, factors that may be associated with a poor prognosis.
32 Adrenomedullin
737
Adrenomedullin in Cancer Expression of adrenomedullin in normal and malignant tissues is not limited to pheochromocytomas, the adrenal glands, or the pancreas (Kitamura et al. 1993; Martinez et al. 1996). Adrenomedullin mRNA transcripts were detected by RT-PCR in 18 of 20 major organs including the brain, heart, and lung (Miller et al. 1996). AM transcripts were not found in normal thyroid or thymus tissues. In the same survey, 55 of 58 (95%) of human cancer cell lines were positive for AM mRNA. The few tumor cell lines that were negative included the H187 small cell lung carcinoma (SCLC), H23 lung adenocarcinoma, and H460 lung large cell carcinoma. The human cancer cell lines where AM was detected were derived from malignancies of the bone marrow, breast, cartilage, colon, lung, nervous system, ovary, and prostate. Based upon these findings, the study of AM expression in malignant tissues was expanded to the examination of multiple samples from cancer patients and preclinical studies in tumor models.
Breast Cancer There have not been many reports focused on the role of adrenomedullin in breast cancer. One study investigated the effects of AM overexpression on a subline of the human breast cancer cell line T47D that was stably transfected to overexpress AM (Martinez et al. 2002). The morphology of T47D cells that overexpressed AM changed by becoming more pleiotropic and the cells also demonstrated an increased angiogenic potential in vitro and in vivo. In vitro, the transfected T47D breast cancer cells served as a feeder layer in the chick aortic ring assay and promoted more sprouting of new vessels compared to the control cells. In vivo silicone tubes were filled with Matrigel-containing transfected T47D cells, or conditioned medium collected in culture was used to further evaluate the angiogenic potential of AM. After intravenous injection of FITC-dextran, there was an increase in fluorescence in the implants containing the cells that overexpressed AM as well as in those exposed to conditioned medium compared to that in the controls. Under serum deprivation, the T47D cells that overexpressed AM were more resistant to apoptosis and had lower levels of pro-apoptotic proteins such as Bax, Bid, and caspase 8 compared to T47D cells transfected with an empty vector. The T47D cells that overexpressed AM also had higher levels of proteins associated with oncogenic signal transduction pathways such as Ras, Raf, PKC, and MAPKp49, and were more tumorigenic when injected subcutaneously in nude mice. These data indicate that AM can confer a survival advantage in breast cancer cells, a finding that was also observed with human prostate cancer cells (Abasolo et al. 2006). The preclinical data on the role of adrenomedullin in breast cancer are supported by clinical data that examined tissue and plasma expression of AM in patients with breast cancer, some of whom had metastatic disease (Oehler et al. 2003). Immunohistochemically detected AM peptide expression was present with
738
R.G. Bagley
moderate to strong staining intensity that was homogeneous and cytoplasmic in 82% of the breast tumor specimens (27 of 33). Importantly, AM expression significantly correlated with axillary lymph node metastasis. When plasma concentrations of AM were compared between breast cancer patients and healthy controls, no meaningful differences were found, although there was a significant positive correlation between tumor size and AM plasma levels. In addition, elevated circulating plasma levels of AM in patients with breast cancer correlated with the presence of lymph node metastasis. These findings suggested a possible role for AM in metastatic disease, although AM offered little value as a diagnostic biomarker for breast cancer since there was no significant difference between plasma levels between healthy volunteers and breast cancer patients in this study.
Central Nervous System Adrenomedullin expression has been studied in glioblastoma, one of the more aggressive malignancies (Ouafik et al. 2002). AM mRNA transcripts were detected by RT-PCR and Northern blot in multiple human glioblastoma cell lines. Western blot confirmed protein expression in the conditioned media of U87, SW1088, and U373 glioblastoma cultures. RT-PCR methods also detected mRNA transcripts for CLRL/RAMP-2 and -3 in glioblastoma cell lines and human glioma tissues. A survey of AM mRNA in human brain tumor biopsies revealed that AM expression correlated with tumor type and grade. While high expression was present in glioblastomas, lower AM mRNA levels were found in anaplastic astrocytomas. In low-grade astrocytomas and oligodendrogliomas, levels of AM mRNA were negligible. The signal transduction pathway of adrenomedullin was further investigated in human glioblastoma cells (Ouafik et al. 2009). In the study, AM was found to promote cell transit with a concomitant increase in cyclin D1 protein level and proliferation through the activation of the cJun/JNK pathway. Adrenomedullin peptide was also detected by radioimmunoassay in human ganglioneuroblastomas and neuroblastomas (Satoh et al. 1995). The same study revealed high levels of adrenomedullin in every region of the human brain, with the highest concentrations present in the thalamus and hypothalamus. These data suggest that adrenomedullin could function as a neurotransmitter, neuromodulator, or a neurohormone.
Endometrial Cancer Endometrial cancer is an indication where AM has been shown to correlate with more aggressive disease both in preclinical tumor models and from the study of clinical samples. In mouse models, adrenomedullin promoted the growth of endometrial xenograft tumors. Two human endometrial carcinoma cell lines were
32 Adrenomedullin
739
transfected with a cDNA encoding AM and then implanted subcutaneously into mice (Oehler et al. 2002). AM expression enhanced in vitro proliferation of RL95.2 cells and in vivo tumor growth compared to controls. IHC of RL95.2 tumors with an antibody against CD31 revealed greater MVD when AM was overexpressed. Transfected Ishikawa endometrial tumor cells did not exhibit a faster growth rate in vitro but tumor growth in vivo was enhanced. These cells were also resistant to apoptosis under hypoxic conditions (Oehler et al. 2001). The oncoprotein Bcl-2 was upregulated in Ishikawa endometrial tumor cells overexpressing AM, and therefore AM may confer resistance to hypoxia-induced cell death. The potential contribution of AM to endometrial cancer growth was investigated in a collection of human tumor biopsies (Nunobiki et al. 2009). A total of 180 clinical endometrial specimens were surveyed for AM expression and MVD by an immunoperoxidase method. The panel of endometrial cancers included 30 proliferative phase specimens, 30 simple hyperplasias, 30 atypical simple hyperplasias, 30 complex hyperplasias, 30 atypical complex hyperplasias, and 30 grade 1 adenocarcinoma specimens. Significant differences existed between AM expression between the normal proliferation samples and those representing hyperplasia without atypia. The levels of expression of AM, the area of the venules, and the MVD increased in a stepwise manner from normal, simple, or complex hyperpasia with or without atypia to specimens of grade 1 adenocarcinomas. These results not only suggest a link between AM expression and stage of the disease but also illustrate the role of AM as an angiogenic protein.
Lung Cancer Little research has been carried out in lung cancer in associated with adrenomedullin but a retrospective study was conducted using archived lung tumor specimens. In a survey of 22 squamous cell carcinomas, 15 adenocarcinomas, and 13 small cell carcinomas, AM expression was more prevalent in the non-small cell carcinomas (Buyukberber et al. 2007). Of the biopsies examined, 91% of squamous cell carcinomas and 87% of adenocarcinomas had moderate to strong levels of AM compared to non-neoplastic lung tissue. By comparison, there was immunoreactivity in only 23% of small cell carcinomas where AM was weakly expressed. These results indicate that adrenomedullin is overexpressed in lung cancer; however, in this study, there was no correlation between AM expression and tumor differentiation, cancer stage, or overall survival.
Mast Cells and the Tumor Microenvironment In human tumors, mast cells can secreted and responded to adrenomedullin (Zudaire et al. 2006). Under hypoxic conditions in culture, differentiated human mast cells increased production of AM mRNA and AM protein expression. AM was also
740
R.G. Bagley
chemotactic for human mast cells and stimulated the production of VEGF, monocyte chemoattractant protein-1, and basic fibroblast growth factor at the molecular level. Mast cells also released histamine or b-hexosaminidase upon exposure to AM. Mast cells expressing AM were identified by immunohistochemical methods in human breast and lung carcinoma tumors.
Ovarian Cancer Adrenomedullin expression has been studied in ovarian cancer both in preclinical experiments and from clinical specimens. One report described incorporating the human ECV ovarian tumor cell line into in vitro experiments, the results of which demonstrated that the addition of exogenous AM to ECV cells promoted in an increase in motility (Martinez et al. 2002). Data generated in preclinical studies with CAOV3 ovarian epithelial cancer cells demonstrated that CAOV3 cells express both AM mRNA and AM protein in vivo (Zhang et al. 2009). The study also indicated that bFGF can induce increased AM expression through the JNK-AP-1 pathway. Ovarian cancer is an indication where adrenomedullin expression was significantly associated with poor prognosis. Survival data and RT-PCR data on AM levels were generated from 60 fresh surgical specimens of epithelial ovarian cancer (29 serous, 14 mucinous, 13 endometrioid, 3 clear cell, and 1 undifferentiated carcinoma) (Hata et al. 2000). In addition to survival, a distinct association existed between the histological grade of the tumors and AM gene expression. IHC methods applied to the ovarian cancer biopsies detected AM protein in the cytoplasm or outer cell membrane of the malignant cells, and in the endothelial cells of the tumor vasculature. In a separate investigation, adrenomedullin expression was evaluated in ovarian granulosa cell tumors and in fibrothecomas, stromal tumors of the ovary that tend to be rich in lipids (Liu et al. 2009). The levels of AM mRNA transcripts as analyzed by Northern blots were higher in the granulosa cell tumors than that in the fibrothecomas and normal ovaries. Within normal ovaries, AM protein expression was localized in both granulosa and theca cells, and in the oocyte of a preovulatory follicle. However, any immunoreactivity of the anti-adrenomedullin antibody with AM protein in the granulosa cell tumors or fibrothecomas was typically considered to be weak. The conclusions from this report are that any correlation between AM levels and the prognosis of ovarian cancer patients may depend upon the histology of the tumor. The findings from these studies that incorporated human ovarian cancer specimens are intriguing, and further research is warranted into the role of AM in ovarian cancer.
Pancreatic Cancer AM and its receptors were identified in normal pancreatic tissue, where it can regulate insulin secretion and blood glucose metabolism (Martinez et al. 1996, 2000).
32 Adrenomedullin
741
To determine if AM may be involved in pancreatic cancer, a panel of pancreatic cancer cell lines were surveyed for AM mRNA expression by RT-PCR, and numerous human pancreatic tumor specimens were analyzed by IHC for AM protein expression (Ramachandran et al. 2007). All eight pancreatic cancer cell lines expressed AM mRNA, and the IHC results revealed AM protein in the epithelium of 90% (43 of 48) of human pancreatic ductal adenocarcinomas. In culture, all five pancreatic cancer cell lines secreted AM that was detected by ELISA. When several of those cell lines (Panc-1, BxPC3, and MPanc96) were exposed to exogenous AM, cellular proliferation, invasion, and nuclear factor kB activity increased. In vivo, silencing AM expression with shRNA in MPanc96 pancreatic cancer cells that expressed relatively high levels of AM resulted in slower growth of orthotopic tumors and metastasis to the lung and liver. Human Panc-1 pancreatic cancer cells that expressed relatively low levels of AM and which were genetically engineered to overexpress AM resulted in larger orthotopic tumors. These observations imply that AM overexpression can promote more aggressive disease in patients with pancreatic cancer, which overall is associated with a poor outcome. The expression levels of the AM receptor L1 and the calcitonin receptor-like receptor (CRLR) were further investigated in pancreatic cancer cells and the cells that contribute to the tumor microenvironment (Ramachandran et al. 2007, 2009). Human pancreatic cancer cells were found to express the AM receptor but not CRLR by RT-PCR and Western blotting methods. Human pancreatic stellate cells (HPSC) derived from pancreatic adenocarcinoma samples, HUVEC, and mouse lung endothelial cells (MLEC) expressed both the receptors as determined by the same techniques. Similar to the human pancreatic cancer cells, HPSC, HUVEC, and MLEC also secreted AM in culture that could be quantified by ELISA. These cells responded to exogenous AM with an increase in proliferation and in vitro tube formation on extracellular matrix proteins. Using shRNA to silence the AM receptor by lentiviral infection in several pancreatic cancer cell lines resulted in a reduction in growth of BxPC3, MPanc96, and Panc-1 cells and invasion of MPanc96 compared to controls when exposed to exogenous AM. Knockdown of the AM receptor or CRLR by siRNA in HPSC, HUVEC, and MLEC indicated that the silencing of the AM receptor but not CRLR reduced the growth of the cells in vitro and polygon formation by endothelial cells. The results generated from experiments with human pancreatic cancers cells and cells of the tumor microenvironment indicated that the autocrine effects of AM are mediated by the AM receptor but not by CRLR. Three receptors associated with AM were co-expressed in pancreatic b-cells: L1, RDC1, and CRLR (Martinez et al. 2000). Overexpression of adrenomedullin was also associated with pancreatic cancer, and the expression of adrenomedullin in pancreatic cancer was observed in clinical samples. In a survey that quantified mRNA levels by QRT-PCR in pancreatic adenocarcinoma vs. normal pancreatic tissue, AM transcripts were 1.5- to 2.4-fold higher in the malignant samples, particularly when lymph node metastasis was present (Keleg et al. 2007). Serum levels of AM were significantly increased in pancreatic cancer patients compared to that in healthy controls or those with chronic pancreatitis. Hypoxia induced AM expression
742
R.G. Bagley
in multiple human pancreatic cell lines and although exogenous recombinant human AM inhibited proliferation slightly, the invasiveness of the cells increased in vitro.
Prostate Cancer Adrenomedullin is expressed in both the normal and malignant prostate tissues. In the healthy human prostate tissue, AM protein was detected by immunocytochemistry primarily in the basal cells of the glandular epithelium and the utriculus (Jimenez et al. 1999). The same study found that PAMP, the other peptide resulting from the cleavage of pre-proadrenomedullin, was expressed in neuroendocrine (NE) cells throughout the epithelium of the prostate glands. QRT-PCR methods applied to human prostate cancer specimens generated results indicating that AM mRNA levels were 3-fold higher in prostate adenocarcinomas with a high Gleason’s score compared to those in benign prostate hyperplasia (Rocchi et al. 2001). The AM protein detected in the prostate tumor tissues was localized in the carcinomatous epithelial compartment. Adrenomedullin expression was specifically associated with androgen-independent prostate cancer. The production of adrenomedullin by prostate cancer cells was investigated in the hormone-dependent LNCaP and hormone-independent PC3 and DU145 cell lines (Rocchi et al. 2001). AM protein was detected in both the cell extracts and conditioned medium of androgen-nonresponsive PC3 and DU145 cells derived from therapeutic hormone-derived prostate cancer but not the androgenresponsive LNCaP cells. AM mRNA was present in xenograft tumors generated from the subcutaneous implantation of PC3 and DU145 cells but not the LNCaP cells. Exogenous AM added to the three prostate cancer cells in culture resulted in enhanced proliferation of the DU145 cells only. In contrast, overexpression of AM by PC3 cells inhibited proliferation through a G0/G1 cell cycle arrest (Abasolo et al. 2004). The transfected PC3 cells also resulted in slower tumor growth in vivo compared to that in the mock transfected cells, suggesting that in certain prostate cancers, AM may inhibit the growth of prostate cancer cells. PC3, DU145, and LNCaP cells expressed the CLRL/RAMP-2 receptor complex, indicating that AM functions in an autocrine manner in prostate cancer cells. Additional experiments conducted with human prostate cancer cells in culture indicated that adrenomedullin can prevent apoptosis (Abasolo et al. 2006). PC3, DU145, and LNCaP human prostate cancer cell lines were transfected with an expression vector to overexpress AM. The overexpression of AM conferred a survival advantage under conditions when the parental cells, cultured with exogenous AM, or the three genetically engineered prostate cancer cell lines were serum deprived or exposed to the topoisomerase II inhibitor etoposide. AM exposure prevented apoptosis in the DU145 and PC3 prostate cancer cells after serum removal and prevented apoptosis in the PC3 and LNCaP cells exposed to etoposide. Western blotting methods applied to the parental PC3 cell line after treatment with etoposide
32 Adrenomedullin
743
indicated that protein levels of phosphorylated ERK1/2, kinases that regulate cell cycle growth and progression, increased. However, the PC3 cells that overexpressed AM had lower phosphorylated ERK1/2 basal levels that did not increase upon etoposide exposure. Similarly, control PC3 cells treated with etoposide resulted in a significant increase in PARP, a protein involved in programmed cell death, but the PC3 clones that overexpressed AM only showed a modest increase in fragmented PARP under the same conditions. The findings from these experiments imply that AM may have a protective role against drug-induced apoptosis in prostate cancer cells. NE differentiation in prostate cancer is of interest as it may be an early marker associated with androgen independence and more aggressive disease (Abrahamsson 1999). Adrenomedullin has recently been implicated in NE differentiation under conditions of androgen withdrawal in models of prostate cancer. The earlier investigations with the LNCaP prostate cancer cell line did not find significant levels of AM produced in vitro or in vivo (Rocchi et al. 2001). However, a more recent study revealed that levels of AM increased 4- to 7-fold in androgen-sensitive LNCaP cells after androgen withdrawal in vitro and in xenograft tumors following castration (Berenguer et al. 2008). The expression of the AM receptors did not change under these conditions. In culture, LNCaP cells that were exposed to AM adopted a NE phenotype including extension of neuritic processes and expression of the neuronspecific enolase (NSE). The administration of AM also increased NSE levels in the serum of noncastrated mice bearing LNCaP xenograft tumors, although there were no changes in tumor growth. However, when exogenous AM was delivered to castrated animals, a significant increase in LNCaP tumor volume was observed 36 days after treatment. These data demonstrate that AM mediates the NE phenotype and can promote tumor regrowth in a hormone-independent manner. These findings have potentially significant implications for the clinic and warrant further investigation.
Renal Cancer The role of adrenomedullin was investigated in human renal cancer cell lines and in clinical specimens (Deville et al. 2009). In culture, the BIZ and 786-O clear-cell renal carcinoma (cRCC) cell lines secreted AM into the medium. Addition of exogenous AM to the cultures stimulated cellular proliferation, migration, and invasion. In human tumor specimens, levels of AM mRNA transcripts were higher in advanced cRCC and in chromophobe renal cell carcinomas (chRCC) compared to that in normal renal tissue. The molecular expression of AM in these samples correlated with levels of VEGF-A mRNA transcripts. Immunohistochemical analyses showed that in RCC tumors, the AM receptors CRLR/RAMP-2 were expressed by the malignant carcinoma cells, whereas RAMP-3 was expressed by infiltrating inflammatory cells, suggesting cross-talk within the tumor microenvironment. Clinically, AM demonstrated a prognostic value as high AM mRNA levels were
744
R.G. Bagley
associated with an increased risk of relapse after curative nephrectomy for cRCC. The results generated from this study in renal cancer implicate AM in the metastatic process and support the potential value of AM as a prognostic biomarker in RCC.
Adrenomedullin as a Therapeutic Target The overexpression of AM in human cancer cells increased angiogenic activity in vitro and in vivo. In preclinical tumor models, studies have demonstrated that AM can promote tumor growth. In clinical specimens, high AM levels correlated with disease grade in some malignancies. These findings suggest that AM should be evaluated as a novel therapeutic target. Preclinically, several strategies have been employed to interfere with AM function. In vitro, an antibody that neutralizes adrenomedullin was effective at inhibiting MCF-7 breast cancer cell growth (Miller et al. 1996). In vivo, intratumoral injections of an anti-AM antibody inhibited the growth of well-established subcutaneous tumors in the U87 glioblastoma xenograft model compared to a vehicle or an irrelevant antibody (Ouafik et al. 2002). Additional studies in several preclinical models using an antibody against the AM receptors (CLRL, RAMP-2, and RAMP-3) either alone or in combination further demonstrated the therapeutic potential of targeting the AM pathway. The antibodies were effective when delivered by intraperitoneal injection to mice bearing subcutaneous U87 glioblastoma, HT29 colon carcinoma, or A549 lung carcinoma xenograft tumors (Kaafarani et al. 2009). In the U87 tumors, immunofluorescent methods employing anti-CD31 and anti-SMA antibodies revealed depletions in endothelial cell and pericyte coverage. Immunohistochemical analysis utilizing anti-vWF antibody confirmed a reduction in mean vessel area; furthermore, an increase in apoptosis was quantified in the U87 model following treatment with the antibodies by the detection of single-stranded DNA. Pancreatic cancer is an indication where targeting adrenomedullin may offer clinical benefits. An AM antagonist (AM 22–52) demonstrated that AM promotes pancreatic cancer by both enhancing angiogenesis and stimulating the malignant cells (Ishikawa et al. 2003). When delivered in vivo to mice bearing PCI-43 xenograft tumors, the AM antagonist delayed tumor formation compared to control. Angiogenesis was reduced with the resulting blood vessels being smaller in diameter. In addition, the tumor proliferation index was lower. In an orthotopic model of pancreatic cancer, sublines of BxPC3 and MPanc 96 pancreatic carcinoma cells stably bearing shControl or shADMR vectors were developed. Sublines of both tumor lines transfected with shADMR generated tumors that were reduced by ~90% in tumor volume compared to tumors grown following the implantation of shControl-transfected cells (Ramachandran et al. 2009). The formation of lung and liver metastasis was also inhibited in the shADMR-transfected models. Nanoliposomes were used to deliver siADMR to mice with orthotopic MPanc96 pancreatic tumors. Both human and mouse siRNAs to ADMR were delivered to target both the host stroma and human cancer cells.
32 Adrenomedullin
745
Bioluminescence imaging indicated that the siRNAs resulted in decreased tumor burden. Subsequent staining of the tumors with an anti-CD31 antibody and IHC revealed collapsed blood vessels in the siADMR-treated tumors compared to the vasculature in the control tumors that had well-developed lumen formation. Targeting pro-adrenomedullin (PAMP) or the adrenomedullin pathway has resulted in anti-angiogenic effects in other preclinical models as well. A peptide fragment of PAMP that inhibited PAMP signaling slowed tumor growth in the A549 xenograft tumor model and decreased angiogenesis in “angioreactors” implanted subcutaneously into nude mice (Martinez et al. 2004). In HUVEC cells, AM antagonists and inhibitors for PKA or PI3K were effective at inhibiting the neo-vascularization induced by AM (Miyashita et al. 2003). In the Matrigel plug assay, angiogenesis was inhibited by an antibody against the AM receptor (Kaafarani et al. 2009). In a xenograft model, AM promoted the growth of subcutaneous mouse sarcoma S180 tumors that could then be inhibited with an antiangiogenic competitive inhibitor of AM (Iimuro et al. 2004). The increases in capillary density and weight of the tumors in mice that were treated with AM were comparable in mice injected with VEGF. S180 tumors in the heterozygous AM+/mice were smaller and had a lower capillary density compared to tumors grown in wild-type mice.
Conclusion Adrenomedullin is a multifunctional peptide that is important for embryogenesis and vascular homeostasis, and for regulating endocrine functions as well as exerting effects on the renal and central nervous systems. AM is upregulated in many cancers from patients in the clinic and in preclinical tumor models. The data from many investigations have demonstrated that AM is an angiogenic peptide and that it conferred a survival advantage in malignant cells. Several strategies that target AM or the AM receptors have been tested in human tumor xenograft models with the resulting data suggesting that neutralizing AM signaling may be a valuable therapeutic approach in cancer patients.
References Abasolo I, Wang Z, Montuenga LM, Calvo A (2004) Adrenomedullin inhibits prostate cancer cell proliferation through a cAMP-independent mechanism. Biochem Biophys Res Commun 322:878–886 Abasolo I, Montuenga LM, Calvo A (2006) Adrenomedullin prevents apoptosis in prostate cancer cells. Regul Pept 133(1–3):115–122 Abrahamsson PA (1999) Neuroendocrine differentiation in prostatic carcinoma. Prostate 39:135–148 Berenguer C, Foudouresque F, Dussert C, Daniel L, Muracciole X, Grino M, Rossi D, Mabrouk K, Figarella-Branger D, Martin P-M Ouafik L’H (2008) Adrenomedullin, an auocrine/paracrine
746
R.G. Bagley
factor induced by androgen withdrawal, stimulates ‘neuroendocrine phenotype’ in LNCaP prostate tumor cells. Oncogene 27:506–518 Buyukberber S, Sari I, Camci C, Buyukberber NM, Sevinc A, Turk HM (2007) Adrenomedullin expression does not correlate with survival in lung cancer. Med Oncol 24(2):245–249 Caron KM, Smithies O (2001) Extreme hydrops fetalis and cardiovascular abnormalities in mice lacking a functional adrenomedullin gene. Proc Natl Acad Sci U S A 98(2):615–619 Chakravarty P, Suthar TP, Coppock HA, Nicholl CG, Bloom SR, Legon S, Smith DM (2000) CGRP and adrenomedullin binding correlates with transcript levels for calcitonin receptor-like receptor (CRLR) and receptor activity modifying proteins (RAMPs) in rat tissues. Br J Pharmacol 130(1):189–195 Deville J-L, Bartoli C, Berenguer C, Fernandez-Sauze S, Kaafarani I, Delfino C, Fina F, Salas S, Muracciole X, Mancini J, Lechevallier E, Martin P-M, Figarella-Branger D, Ouafik L’H, Daniel L (2009) Expression and role of adrenomedullin in renal tumors and value of its mRNA levels as prognostic factor in clear-cell renal carcinoma. Int J Cancer 125(10):2307–2315 Fernandez-Sauze S, Delfino C, Mabrouk K, Dussert C, Chinot O, Martin P-M, Grisoli F, Ouafik L’H, Boudouresque F (2004) Effects of adrenomedullin on endothelial cells in the multistep process of angiogenesis: involvement of CRLR/RAMP2 and CRLR/RAMP3 receptors. Int J Cancer 108:797–804 Fritz-Six KL, Dunworth WP, Li M, Caron KM (2008) Adrenomedullin signaling is necessary for murine lymphatic vascular development. J Clin Inv 118(1):40–50 Fujita Y, Mimata H, Nasu N, Nomura T, Nomura Y, Nakagawa M (2002) Involvement of adrenomedullin induced by hypoxia in angiogenesis in human renal cell carcinoma. Int J Urol 9:285–295 Hata K, Takebayashi Y, Akiba S, Fujiwaki R, Iida K, Nakayama K, Nakayama S, Fukumoto M, Miyazaki K (2000) Expression of adrenomedullin gene in epithelial ovarian cancer. Mol Hum Reprod 6(10):867–872 Hinson JP, Kapas S, Smith DM (2000) Adrenomedullin, a multifunctional regulatory peptide. Endocr Rev 21(2):138–167 Iimuro S, Shindo T, Moriyama N, Amaki T, Niu P, Takeda N, Iwata H, Zhang Y, Ebihara A, Nagai R (2004) Angiogenic effects of adrenomedullin in ischemia and tumor growth. Circ Res 95:415–423 Imai Y, Shindo T, Maemura K, Kurihara Y, Nagai R, Kurihara H (2001) Evidence for the physiological and pathological roles of adrenomedullin from genetic engineering in mice. Ann N Y Acad Sci 947:26–34 Ishikawa T, Chen J, Wang J, Okada F, Sugiyama T, Kobayashi T, Shindo M, Higashino F, Katoh H, Asaka M, Kondo T, Hosokawa M, Kobayashi M (2003) Adrenomedullin antagonist suppresses in vivo growth of human pancreatic cancer cells in SCID mice by suppressing angiogenesis. Oncogene 22:1238–1242 Jimenez N, Calvo A, Martinez A, Rosell D, Cuttitta F, Montuenga LM (1999) Expression of adrenomedullin and proadrenomedullin N-terminal 20 peptide in human and rat prostate. J Histochem Cytochem 47(9):1167–1177 Kaafarani I, Fernandez-Sauze S, Berenguer C, Chinot O, Delfino C, Dussert C, Metellus P, Boudouresque F, Mabrouk K, Grisoli F, Figarella-Branger D, Martin P-M, Ouafik L’H (2009) Targeting adrenomedullin receptors with systemic delivery of neutralizing antibodies inhibits tumor angiogenesis and suppresses growth of human tumor xenografts in mice. FASEB J 23:3424–3435 Kapas S, Clark AJL (1995) Identification of an orphan receptor gene as a type 1 calcitonin generelated peptide receptor. Biochem Biophys Res Commun 217(3):832–838 Kapas S, Catt KJ, Clark AJL (1995) Cloning and expression of cDNA encoding a rat adrenomedullin receptor. J Biol Chem 270:25344–25347 Kato H, Shichiri M, Marumo F, Hirata Y (1997) Adrenomedullin as an autocrine/paracrine apoptosis survival factor for rat endothelial cells. Endocrinology 138(6):2615–2520
32 Adrenomedullin
747
Keleg S, Kayed H, Jiang X, Penzel R, Giese T, Buchler MW, Friess H, Kleeff J (2007). Adrenomedullin is induced by hypoxia and enhances pancreatic cancer cell invasion. Int J Cancer 121:21–32 Kennedy SP, Sun D, Oleynek JJ, Hoth CF, Kong J, Hill RJ (1998) Expression of the rat adrenomedullin receptor or a putative human adrenomedullin receptor does not correlate with adrenomedullin binding or functional response. Biochem Biophys Res Commun 244(3):832–837 Kim W, Moon S-O, Sung MJ, Kim SH, Lee S, So J-N, Park SK (2003) Angiogenic role of adrenomedullin through activation of Akt, mitogen-activated protein kinase, and focal adhesion kinase in endothelial cells. FASEB J 17:1937–1939 Kitamura K, Kangawa K, Kawamoto M, Ichiki Y, Nakamura S, Matsuo H, Eto T (1993) Adrenomedullin: a novel hypotensive peptide isolated from human pheochromocytoma. Biochem Biophys Res Commun 192(2):553–560 Ladoux A, Frelin C (2000) Coordinated up-regulation by hypoxia of adrenomedullin and one of its putative receptors (RDC-1) in cells of the rat blood-brain barrier. J Biol Chem 275(51):39914–39919 Liu J, Butzow R, Hyden-Granskog C, Voutilainen R (2009) Expression of adrenomedullin in human ovaries, ovarian sex cord-stromal tumors and cultured granulosa-luteal cells. Gynecol Endocrinol 25(2):96–103 Martinez A, Cuttitta F (1998) Adrenomedullin. IOS Press, Netherlands Martinez A, Weaver C, Lopez J, Bhathena SJ, Elsasser TH, Miller M-J, Moody TW, Unsworth EJ, Cuttitta F (1996) Regulation of insulin secretion and blood glucose metabolism by adrenomedullin. Endocrinology 137(6):2626–2632 Martinez A, Kapas S, Miller M-J, Ward Y, Cuttitta F (2000) Coexpression of receptors for adrenomedullin, calcitonin gene-related peptide, and amylin in pancreatic b-cells. Endocrinology 141(1):406–411 Martinez A, Vos M, Guedez L, Kaur G, Chen Z, Garayoa M, Pio R, Moody T, Stetler-Stevenson WG, Kleinman HK, Cuttitta F (2002) The effects of adrenomedullin overexpression in breast tumor cells. J Nat Cancer Inst 94(16):1226–1237 Martinez A, Zudaire E, Portal-Nunez S, Guedez L, Libutti SK, Stetler-Stevenson WG, Cuttitta F (2004) Proadrenomedullin NH2-terminal 20 peptide is a potent angiogenic factor, and its inhibition results in reduction in tumor growth. Cancer Res 64:6489–6494 McLatchie LM, Fraser NJ, Main MJ, Wise A, Brown J, Thompson N, Solari R, Lee MG, Foord SM (1998) RAMPs regulate the transport and ligand specificity of the calcitonin-receptor-like receptor. Nature 393:333–339 Miller MJ, Martinez A, Unsworth EJ, Thiele CJ, Moody TW, Elsasser T, Cuttitta F (1996) Adrenomedullin expression in human tumor cell lines. J. Biol. Chem. 271(38):23345–23351 Miyashita K, Itoh H, Sawada N, Fukunaga Y, Sone M, Yamahara K, Yurugi-Kobayashi T, Park K, Nakao K (2003) Adrenomedullin provokes endothelial Akt activation and promotes vascular regeneration both in vitro and in vivo. FEBS Lett 544:86–92 Moriyama T, Otani T, Maruo T (2001) Expression of adrenomedullin by human placental cytotrophoblasts and choriocarcinoma JAr cells. J Clin Endocrinol Metab 86(8):3958–3961 Nagaya N, Mori H, Murakami S, Kangawa K, Kitamura S (2005) Adrenomedullin: angiogenesis and gene therapy. Am J Physiol Regul Integr Comp Physiol 288:R1432–R1437 Nikitenko LL, Fox SB, Kehoe S, Rees MCP, Bicknell R (2006) Adrenomedullin and tumour angiogenesis. Br J Cancer 94(1):1–7 Njuki F, Nicholl CG, Howard A, Mak JC, Barnes PJ, Girgis SI, Legon S (1993) A new calcitonin receptor-like sequence in rat pulmonary blood vessels. Clin Sci 85:385–388 Nunobiki O, Nakamura M, Taniguchi E, Utsunomiya H, Mori I, Tsubota Y, Mabuchi Y, Kakudo K (2009) Adrenomedullin, Bcl-2 and microvessel density in normal, hyperplastic and neoplastic endometrium. Pathol Int 59:530–536 Oehler MK, Norbury C, Hague S, Rees MCP, Bicknell R (2001) Adrenomedullin inhibits hypoxic cell death by upregulation of Bcl-2 in endometrial cancer cells: a possible promotion mechanism for tumour growth. Oncogene 20:2937–2945
748
R.G. Bagley
Oehler MK, Hague S, Rees MCP, Bicknell R (2002) Adrenomedullin promotes formation of xenografted endometrial tumors by stimulation of autocrine growth and angiogenesis. Oncogene 21:2815–2821 Oehler MK, Fischer DC, Orlowska-Volk M, Herrle F, Kieback DG, Rees MCP, Bicknell R (2003) Tissue and plasma expression of the angiogenic peptide adrenomedullin in breast cancer. Br J Cancer 89:1927–1933 Ouafik L’H, Sauze S, Boudouresque F, Chinot O, Delfino C, Fina F, Vuaroqueaux V, Dussert C, Palmari J, Dufour H, Grisoli F, Casellas P, Brunner N, Martin P-M (2002) Neutralization of adrenomedullin inhibits the growth of human glioblastoma cell lines in vitro and suppresses tumor xenograft growth in vivo. Am J Pathol 106(4):1279–1292 Ouafik L’H, Berenguer-Daize C, Berthois Y (2009) Adrenomedullin promotes cell cycle transit and up-regulates cyclin D1 protein level in human glioblastoma cells through the activation of c-Jun/JNK/AP-1 signal transduction pathway. Cell Signal 21(4):597–608 Park SC, Yoon JH, Lee JH, Yu SJ, Myung SJ, Kim W, Gwak GY, Lee SH, Lee SM, Jang JJ, Suh KS, Lee HS (2008) Hypoxia-inducible adrenomedullin accelerates hepatocellular carcinoma cell growth. Cancer Lett 271(2):314–322 Poyner DR, Sexton PM, Marshall I, Smith DM, Quirion R, Born W, Muff R, Fischer JA, Foord SM (2002) The mammalian calcitonin gene-related peptides, adrenomedullin, amylin, and calcitonin receptors. Pharmacol Rev 54:233–246 Ramachandran V, Arumugam T, Hwang RF, Greenson JK, Simeone DM, Logsdon CD (2007) Adrenomedullin is expressed in pancreatic cancer and stimulates cell proliferation and invasion in an autocrine manner via the adrenomedullin receptor, ADMR. Cancer Res 67(6):2666–2675 Ramachandran V, Arumugam T, Langley R, Hwang RF, Vivas-Mejia P, Sood AK, Lopez-Berestein G, Logsdon CD (2009) The ADMR Receptor mediates the effects of adrenomedullin on pancreatic cancer cells and on cells of the tumor microenvironment. PLoS One 4(10):e7502 Rocchi P, Bodouresque F, Zamora AJ, Muracciole X, Lechevallier E, Martin P-M, Ouafik L’H (2001) Expression of adrenomedullin and peptide amidation activity in human prostate cancer and in human prostate cancer cell lines. Cancer Res 61:1196–1206 Satoh F, Takahashi K, Murakami O, Totsune K, Sone M, Ohneda M, Abe K, Miura Y, Hayashi Y, Sasano H (1995) Adrenomedullin in human brain, adrenal glands, and tumor tissues of pheochromocytoma, ganglioneuroblastoma and neuroblastoma. J Clin Endocrinol Metab 80(5):1750–1752 Shindo T, Kurihara Y, Nishimatsu H, Moriyama N, Kakoki M, Wang Y, Imai Y, Ebihara A et al (2001) Vascular abnormalities and elevated blood pressure in mice lacking adrenomedullin gene. Circ Res 104:1964–1971 Zhang Y, Zhang S, Shang H, Pang X, Zhao Y (2009) Basic fibroblast growth factor upregulates adrenomedullin expression in ovarian epithelial carcinoma cells via JNK-AP-1 pathway. Regul Pept 157(1–3):44–50 Zudaire E, Martinez A, Garayoa M, Pio R, Kaur G, Woolhiser MR, Metcalfe DD, Hook WA, Siraganian RP, Guise TA, Chirgwin JM, Cuttitta F (2006) Adrenomedullin is a cross-talk molecule that regulates tumor and mast cell function during human carcinogenesis. Am J Pathol 168(1):280–291
Index
A Aberrant DNA methylation altered methylation involvement cancer stromal cells, 129–130 EMT, 128–129 field cancerization, 128 somatic mutations, 129 unique natures, 127–128 characteristics DNA methylation status maintenance, 122–123 gene transcription regulation, 124–125 maintenance and de novo DNA methylases, 125 epigenetic changes, 122 epilogue, 130 methylation alterations CpG islands, 126–127 driver methylation and passenger methylation, 127 genome-overall hypomethylation, 125–126 Acetylcholine receptor-inducing activity (ARIA), 723 Acid-extruding mechanisms, 29 Acid production, 27 Acid transport H+ ions buffering, 30 extrusion, 30–31 nonrespiratory sources, 31 intracellular and extracellular pH, 28 metabolic acid efflux, 29–30 pH regulation, 28–29 Activator protein-1 (AP-1), 152–153 Adhesion receptors, family, 295
Adipokine angiopoietin-like 4 (ANGPTL4), 647–648 Adrenomedullin (AM) angiogenesis, 735–736 breast cancer, 737–738 central nervous system, 738 endometrial cancer, 738–739 lung cancer, 739 mast cells, 357, 361, 739–740 ovarian cancer, 740 pancreatic cancer, 740–742 prostate cancer, 742–743 renal cancer, 743–744 structure and function blood glucose metabolism, 734 definition of AM, 733 insulin secretion, 734 intracellular signal transduction, 735 PAMP, 733 vasodilatory properties, 733, 734 therapeutic target, 744–745 Adult stem cells (ASCs), 170 Aerobic ATP generation, 99 Aerobic glycolysis. See Warburg effect Alkaline pHe restoration, 35 AM. See Adrenomedullin Anatomic distribution, NK cell, 433 Angiogenic CXC chemokines, 283 Angioreactor angiogenesis, 736 therapeutic target, 745 Anti-a5b1 integrin function-blocking antibody, 469 Antiangiogenic drug therapy, 264–265
R.G. Bagley (ed.), The Tumor Microenvironment, Cancer Drug Discovery and Development, DOI 10.1007/978-1-4419-6615-5, © Springer Science+Business Media, LLC 2010
749
750 Apurinic/apyrmidinic endonuclease 1 (APE1) AP-1, 152–153 APE1-regulated transcription factors, 148–149 BER, 139–140 GRX/GSH system, 146 HIF-1a, 153–154 p53 APE1, role, 150, 151 BER pathway, 151 DNA-binding domain, 149 DR pathway, 152 glycosylases, 151 HR and NHEJ pathways, 152 MMR pathway, 152 polymerase b (Pol b), 151 proteins interaction, 149 redox functions, 148 Arg-Gly-Asp (RGD), 297 Autophagy, 216 B Baculovirus system, 62–63 Base excision repair (BER), 139–140 Bioluminescence imaging, 318 Blood-brain-barrier (BBB), 358 BMDMSCs. See Bone marrow derived mesenchymal stem/stromal cells Bmi-1, 178–179 Bone marrow derived cell (BMDCs), 257 Bone marrow derived mesenchymal stem/stromal cells (BMDMSCs) activation and metastasis, 283–284 activation and tumor growth breast cancer, 281–282 DARC expression, 282–283 as source, 277–278 tumor stroma interactions, 280–281 Bovine aortic endothelial cells (BAEs), 681 Bovine microvascular endothelial cells (BMEs), 681 Bowman-Birk trypsin inhibitor (BBI), 656–657 C CA. See Carbonic anhydrase CAFs. See Carcinoma associated fibroblasts CA9 gene expression, 66–68 CA IX. See Carbonic anhydrase IX
Index Calcitonin gene-related peptide (CGRP), 734 Calcitonin receptor-like receptor (CRLR), 734, 735, 741 Cancer progression CA IX, 72–74 disseminated tumor cells cellular dormancy, 236–237 chromosomal aberrations, 230–231 dissemination theory, 230 gene expression analysis, 230 immunocytochemical techniques, 231 metastatic cascade, 230 micrometastasis (See Micrometastasis) mechanical forces, 696–697 proteolytic modification, 701 Cancer stem cells (CSC) ASCs, 170 Bmi-1, 178–179 CAF, 180 carcinogenesis, 178 clonal selection, 169 durotaxis, 178 dynamic equilibrium, 180 ESCs cell cycle regulatory mechanisms, 174 characteristics, 172 ChIP-on-chip analysis, 173 epigenetic control, 173 mitogen signaling, MAPK pathway, 176 molecular cues, 177 pluripotent cell characteristics, 174, 175 POU5FI, 172–173 Rb/p105, 174 R point, 174–176 telomerase, 172 transcriptional regulators, 172–173 hypothesis, 170, 171 mechanobiology, 177–178 self-renewal process deregulation, 170 stem cell niche, 170, 177 tumor hypoxia, 179 tumorigenesis, 170 tumor progenitor, 179–180 Cancer stromal cells, 129–130 Carbonic anhydrase (CA) activity, 32 cytosolic isoforms, 61 dominant role, 35 facilitated CO2 diffusion, 32–34
Index facilitated H+ diffusion, 34 human CA isoforms, 60 intracellular and extracellular anhydrase isoforms, 32 metabolons, 61 N-terminal proteoglycan-like region, 61 transport metabolon, 34 zinc-binding enzymes, 59 Carbonic anhydrase IX (CA IX) clinical value anticancer treatment, 72 biopsies, 75 cancer progression, prognosis, and treatment outcome, 72–74 hypoxia marker, 72 pre-surgical diagnosis, 75 RCC, 72, 75 in vivo imaging, 75 molecular features baculovirus system, 62–63 biochemical and crystallographic analysis, 63–64 catalytic domain, 62–63 MN antigen structure, 62 molecular weight, CA IX, 63 regulation CA9 gene expression, 66–68 CA IX ECD shedding, 68 extracellular acidosis, 67 HIF, 65, 66 HRE, genomic position, 66–67 normoxic conditions, 67 posttranslational level, 68 transcription, 65 role in cancer acidosis, 69 catalytic domain, 69 cell adhesion, 71 cell migration, 71 ectopic expression, 70–71 extracellular acidification, 71 oncogenic metabolism, 69 pH regulation, 68–70 targeting strategies cancer-related CA, 76–80 FITC-CAI binding, 79 pH control, 79 pre-clinical and clinical experiment, 76 SK-RC-52 (RCC) xenograft model, 80 sulfonamide inhibitors, 77–79 therapeutic target stems, 75 tissue distribution, 64–65
751 Carcinoma associated fibroblasts (CAFs) BMDMSCs activation and metastasis, 283–284 activation and tumor growth, 281–283 as source, 277–278 tumor stroma interactions, 280–281 cell-autonomous fashion, 331 characteristics, 276–277 counterpart and normal fibroblasts, 329, 330 CSC, 180 genetic and karyotype analysis, 329 heterogeneous cellular origins, 333–334 high-grade malignancy and poor prognosis, 328 immunohistochemical analysis, 330, 331 a-SMA expression, 329 TGF-b, 647 tumor associated stromal cells, 279–280 tumor stroma, 276 tumour-promoting stromal myofibroblasts, 327 CD4+CD25+ regulatory T cells, 396–397 CD34 cell, 259 CD9 interferes EGFR, 567–568 homoclustering, 566 integrin-mediated motility, 567 MMP2 expression, 569 positive correlation, 566 transendothelial migration, 569 tumorigenicity, 566, 569 WAVE2, 568 Wnt proteins, 567, 568 CD16 receptor, 436–437 CD44s cell surface-docking molecule, 537 CSPG, 485, 486 epidermal melanocytes, 484 galactosylation, 486 melanoma motility, 485 MMPs, 486–487 proteolytic removal, 485 shedding, 487 CD8+ T cells CTLs cell memory response, 391–392 MHC class I and MHC class II, 387–388 murine model, 388 TGF-b, 653 Cell culture method, 280 Cellular fibronectin (cFN), 459
752 Chemokines biological and molecular mechanisms, 19 intrinsic properties, tumor cells, 602 leukocyte recruitment CCL2, 609 immunosuppressive type 2 macrophage, 609 inflammatory and pathological stimuli, 608 lymphoreticular infiltration, 610 macrophages, 608, 609 primary tumor eradication, 601 receptors CC chemokines, 605–607 C chemokines, 606 classification, 603 CXC chemokines, 603, 604, 606 CX3C chemokines, 604, 606–607 leukocyte trafficking, 603 promiscuity, 607 targeting and chemotherapy, 617 tumor angiogenesis angiostasis, 613 CXCL5, 611 CXCL8, 610–611 CXCL10, 612 CXCL12, 611–612 CXCR3, 612 definition, 610 fractalkine, 612–613 renal cell carcinoma, 611 tumor growth and metastasis CCR4, 613–614 CCR7, 616 CXCL8, 615 CXCL9 and CXCL10, 616, 617 CXCR3, 616–617 CXCR4, 614–615 CXCR1 and CXCR2, 615 multifaceted role, 602, 603 SCH-479833 and SCH-527123, 615–616 tumor-host interaction, 613 Chemokine stromal-derived factor 1, 277 Chondroitin sulfate proteoglycan (CSPG), 485, 486 Chromatin, 173 Chronic inflammation, 124 vs. acute inflammation leukocyte infiltration, 411 pro-inflammatory response phase, 412–413
Index resolution phase, 414 vasculitis and pre-neoplastic syndromes, 412 dynamic renaissance, 410 Helicobacter pylori colonization, 410, 411 neoplastic transformation, 410 soluble mediator, immune response, 417 tumor immune evasion and progression, 415–416 Circulating endothelial cells (CECs), 264–265 Circulating endothelial progenitors (CEPs), 264–265 c-kit positive mast cells, 360 Collagen adhesion receptor binding CD44s, 484–487 DDRs, 487–488 a2b1 integrin, 483–484 RTKs, 487 THP, 483, 484, 488 angiogenesis, 482 cell homeostasis, 481 ECM cell proliferation, 479 b1 integrin and dystroglycan-dependent process, 481 intermolecular self-assembly, 481 pericytes, 480 fibronectin fragments, 496–497 fragments biological activities, 493, 494 cryptic sites, 493, 496 endostatin, 495 homotypic multimerization, 496 integrins, 493, 495 pentastatin-1, 495 synthetic RGD-containing peptides, 495 matrix metalloproteinases collagenolytic MMPs, 492 heterotrimeric type I and II collagen, 492 LOX, 482 metastatic melanoma, 491 tumor progression, 492 VEGF, 492, 493 mechanotransduction and oncogenic signaling pathways, 482 protein–collagen interactions decorin, 490–491 FN, 488–490 LN, 490
Index proteins C-terminal propeptides, 478–479 glycosylation, 478 homotrimeric and heterotrimeric collagens, 477 left-handed polyPro II helix, 477 NC1 domain, 479 nonfibrillar type IV collagen, 479 Connective tissue mast cells (CTMC), 356 Cortocotropin-releasing hormone (CRH), 356–358, 363 CO2 transmembrane efflux, 28 CpG site, 122–123 Cre-lox techniques, 315 CRH. See Cortocotropin-releasing hormone CRLR. See Calcitonin receptor-like receptor Crystallographic analysis, 63–64 CSC. See Cancer stem cells CSPG. See Chondroitin sulfate proteoglycan C-terminal Cys, 145 CTLs. See Cytotoxic T lymphocytes CXCL8, 282–283 CXCL12, 416. See also Stromal cell-derived factor 1 CXCR4, 416 Cytokine IL-17, 386 Cytotoxic agent, 318 Cytotoxic T lymphocytes (CTLs) CD8+ T cell cell memory response, 391–392 MHC class I and MHC class II, 387–388 murine model, 388 CD4+ T cells induction, 389–390 maintenance, 391 T helper 1 vs. T helper 2 responses, 392–393 D DDRs. See Discoidin domain receptors Decathlon, 214–215 2-Deoxyglucose (2-DG), 14 Desmoplasia, 461, 718 Diethylnitrosamine (DEN), 376 Discoidin domain receptors (DDRs), 242, 487–488 Disseminated tumor cells cancer progression cellular dormancy, 236–237 chromosomal aberrations, 230–231 dissemination theory, 230
753 gene expression analysis, 230 immunocytochemical techniques, 231 metastatic cascade, 230 micrometastasis (See Micrometastasis) dormancy models bone marrow, 246 cancer xenograft tumor cell model, 247 mouse models, 245–246 MYC oncogene, 247 phenotypic change, 246 solitary dormant tumor cells, 246 stroma-associated factors collagen matrix signaling, 241–242 hormone depletion dormancy, 242–243 TGFb signaling, 243–245 tumor microenvironment cellular dormancy, 240 deregulated ligand-dependent signaling, 240–241 ERK signaling, 241 genetic ablation, b1 integrin, 238 mitogenic signaling, 240 p38a activation, 241 paracrine tumor-stroma interactions, 238 DNA repair APE1, redox activity (See Apurinic/ apyrmidinic endonuclease 1) BER, 139–140 cancer therapeutic approach, 159–161 DR, 137–139 global influences, APE1 angiogenesis, 156–157 cell survival, 155–156 inflammation, 157–158 HR, 143–144 MMR, 140–141 NER, 141–142 NHEJ, 143 redox signaling GRX/GSH system, 146–147 Trx system, 145 schematics, 137, 138 tumor microenvironment, 158–159 Driver methylation, 127 Drosophila melanogaster, 190 Duffy antigen receptor for chemokines (DARC) expression cancer, 283 circulating chemokines and pathological conditions, 282–283 lymph node metastasis, 282 Durotaxis, 178
754 E ECM. See Extracellular matrix EDA-FN. See Extra Domains A-fibronectin EDB-FN. See Extra Domains B-fibronectin EGFR. See Epidermal growth factor receptor Embryonic stem cells (ESCs) cell cycle regulatory mechanisms, 174 characteristics, 172 ChIP-on-chip analysis, 173 epigenetic control, 173 mitogen signaling, MAPK pathway, 176 pluripotent cell characteristics, 174, 175 POU5FI, 172–173 Rb/p105, 174 R point, 174–176 telomerase, 172 transcriptional regulators, 172–173 EMT. See Epithelial–mesenchymal transition Endothelial cell colony forming units (CFU-ECs), 263 Endothelial colony-forming cells (ECFCs), 263 Endothelial progenitor cells (EPCs) BMDCs, 257 cellular players, 258 definition ECFCs, 263 hematopoietic marker, 263 multiparametric flow cytometry techniques, 262 VE-cadherin (CD144), 263 functions angiogenesis, 259 confocal laser scan microscopy, 261 FISH staining techniques, 260–261 Id proteins, 259 SDF-1 expression, 262 sprouting angiogenesis, 261 tumor vasculature, 260 VEGFR2 cells, 261 identification, 259 neoangiogenesis, 340 surrogate biomarker, 264–265 therapy-induced cell mobilization and tumor vessel incorporation antiangiogenic drug, 266 CEPs impact, 265–266 colonization, 266, 267 cytotoxic therapy, 267 macroscopic metastases, 268 MTD chemotherapy, 265–266 surface markers, 268–269 VDAs, 266
Index Epidermal growth factor receptor (EGFR) CD9 interferes, 567–568 CD82 interferes, 564, 565 malignant cell growth factors, 722–723 M-CSF receptor, 378 Epilogue, 130 Epithelial–mesenchymal transition (EMT) biomarkers, 188, 189 characteristics and classifications, 188, 517 definition, 649 differentiated epithelial cells, 517 DNA methylation, 128–129 E-cadherin cell–cell adhesion molecule, 188–189 gene expression and protein, 518–520 Rb down-regulation, 651 FBLN5, 649 FOXC2, 650 HMW-tropomyosins, 650 human hepatocellular carcinoma, 652 MECs, 649 metastatic process, 49–50 MMPs, 519 N-cadherin expression, 518–519 perspective, 204–205 phenotypic conversion, 187–188 signaling molecules, 518 Smad gene expression, 651 tetraspanins, 562, 567 tumour-promoting stromal myofibroblasts, 341 type 1 blastula stage, 189–190 mesoderm formation, 190–191 MET, 191 neural crest formation, 191 type 2 fibrosis, 191–192 re-epithelialization, wounded skin, 192 type 3 CSC generation, 202 cytokines, 199–201 EMT stimuli, 193–195 genetic and epigenetic control, 202–203 hypoxia, 201 micro RNA, 203–204 molecular regulation, EMT, 196 signaling pathways, 197–199 vimentin, 189 vimentin expression, 650–651 Epithelial ovarian cancer (EOC) metastasis model, 515, 516
Index ESCs. See Embryonic stem cells Extracellular acidosis, 67 Extracellular carbonic anhydrase, 32, 33 Extracellular matrix (ECM) adipocytes, 542 angiogenesis inhibitors, 706 apoptosis regulation, 699–700 stimulation, 700–701 broad inhibitors, 705 cell proliferation, 479 cellular regulation, 481 collagen fragments, 493, 496 collagen-mediated tumor cell behavior, 488 components, 538 3D cell migration, 545 decorin, 491 definition, 695 degradation, 542, 548 fibrillar collagen, 538 fibrillar proteins, 543 FN polymerization and incorporation, 489 b1 integrin and dystroglycan-dependent process, 481 integrin inhibitors, 705–706 intermolecular self-assembly, 481 invasive tumor and transformed cell growth, 544 LOX, 547 matrix component hydrolysis, 482 mechanical forces, 696–697 MMPs, 490, 491 pericytes, 480 pro-stromal factor, 718–720 protease inhibitors, 706 protein-protein interactions, 696 protein vitronectin, 497 proteolytic modification cancer progression, 701 cryptic domains, 702 soluble active peptides, 702–704 soluble growth factors, 704–705 remodeling, 536 sequestered TGFb, 539 tumor cell proliferation regulation, 698–699 type IV collagen, 538 type VII collagen, 490 VEGF, 543 Extra Domains A-fibronectin (EDA-FN) gene structure and FN-knock out, 458 tumor growth and angiogenesis, 463–464
755 Extra Domains B-fibronectin (EDB-FN) gene structure and FN-knock out, 458 potential function, 463 tumor growth and angiogenesis angiogenic switch, 463 cryptic site exposure, 464–465 EDA/EDB-double null mutants, 464 extensive immunohistochemical data, 462 F FAK. See Focal adhesion kinase Fibroblast growth factors (FGF), 713–714 Fibronectin (FN) fibrils, 467 knock out gene structure, 458–459 mice and phenotype, 460–461 molecular structure, 458 MSF, 465–466 potential function, EDB-domain, 463 synthesis and matrix assembly, 459, 460 therapeutic interventions anti-a5b1 integrin function-blocking antibody, 469 endogenous inhibitors, angiogenesis, 469–470 isoforms, 468–469 therapeutic mode, 467, 468 tumor angiogenesis definition, 462 EDA-FN, 463–464 EDB-FN, 462–463 tumor dormancy, 466–467 tumor growth autocrine/paracrine mitogenic factor, 461 desmoplasia, 461 EDA-FN, 463–464 EDB-FN, 462–465 TAMs, 461 tumor invasion and metastasis, 465 tumor stroma, 457 Fibulin-5 (FBLN5), 649 FKN. See Fractalkine Flt3 ligand (Flt3L), 722 Fluorescein-conjugated carbonic anhydrase inhibitor thioureidohomosulfanilamide (FITC-CAI), 79 Fluorescent in situ hybridization (FISH), 260–261
756 Fluorodeoxyglucose positron emission tomography, 27 FN. See Fibronectin Focal adhesion kinase (FAK) cell cycle progression, 513 cytoplasmic tails, 510, 512 downstream signaling pathways, 512 expression, 522 integrin signaling, 297–299 intracellular signals, 510 nuclear localization, 513 Src activation, 512 Forkhead box C2 (FOXC2), 650 Foxp3+ Tregs, 373 Fractalkine (FKN), 612–613 18
G Gastrulation, 190 Gene promoter region, 124 Glucose-6-phosphate (G6P), 92 Glucose transporters (GLUT1), 95–96 Glutaredoxin/glutathione (GRX/GSH) system, 146–147 Glycolysis ATP generation, 93 cancer AMP-activated protein kinase (AMPK), 104 cell growth and proliferation, 94 c-Myc, 104–105 GLUT1, 95–96 HIF-1, 100–103 HKII, 96 LDH-A, 97–98 mitochondrial dysfunction, 99–100 PFK1, 96–97 PI3K/Akt Pathway, 105–106 PPP, 98 p53 regulation, mitochondrial respiration, 103–104 putative mechanisms, 95 pyruvate kinase (PKM2), 97 Ras, 105 chemotherapy, 93 FDG-PET scan, 94 G6P, 92 metabolic alterations, 94 NAD+, 93 overview, 92 tumor inhibition 3-bromopyruvate, 107–108 2-deoxyglucose, 107
Index dichloroacetate, 109 lonidamine, 108 metabolic modulators, 110 metabolic regulatory mechanisms, 102, 107 oxythiamine and 6-aminonicotinamide, 109 Glycosylases, 151 Gonadotropin-releasing hormone (GnRH), 725 Graft-versus-host disease (GVHD), 438 Growth factor cytokines, 675, 679 expression and bio-distribution avb3-dependent upregulation, 681–682 ECM composition, 680 integrin-mediated reciprocal communication, 679 monocyte/macrophage infiltration, 680 VEGF, 680, 681 integrin cooperation ECM proteins, 683 functional associations, 683 initiation, invasive and maturation phase, 682 integrin receptors, 684–685 PTN receptor, 684 pulmonary inflammation and fibrosis regulation, 684 relative ratio and bioavailability, 679 signaling, 681–682 H Hedgehog (Hh) signaling, 336–337 Helicobacter pylori, 410, 411, 418 Helix-loop-helix transcription factor, 50 Hepatocarcinogenesis, 421 Hepatocellular carcinoma (HCC), 736 Hepatocyte growth factor (HGF), 714, 718, 725–726 HER, 722–723, 725 Hexokinases (HKII), 96 HGF. See Hepatocyte growth factor HIF-1. See Hypoxia inducible factor-1 Histidine decarboxylase (HDC), 362 HMC-1. See Human leukemic mast cells Homologous recombination (HR), 143–144 Human leukemic mast cells (HMC-1), 357, 358, 363
Index Human Leukocyte Antigen (HLA), 435–436 Human microvascular endothelial cells (HMVEC), 736 Human umbilical vein endothelial cells (HUVEC), 736, 741, 745 HUVEC. See Human umbilical vein endothelial cells Hybrid resistance model, 440 Hyperproliferating tumor cells, 213 Hypoxia, gene expression, and metastasis causes and consequences, 45–46 CSC CD133 expression, 52–53 iPS cell state, 51 leukocyte trafficking control, 53 PTEN tumor, 52 resistance, 51 striking correlation, CSC and cancer cell, 51 HIF regulation by genetic alterations of upstream regulators, 47–49 regulation by oxygen, 46–47 target genes, 49–50 metastatic process, 44, 45 Hypoxia inducible factor-1 (HIF-1) cancer cell metabolism and tumor microenvironment aerobic glycolysis, 14 HIF-1 inhibitors, 14–15 chemotherapy, 10–11 glycolysis, 100–101 heterodimeric protein, 4 human cancers, 4, 6 intratumor hypoxia anti-angiogenic therapies and HIF-1 inhibitors, 13 tumor vasculature normalization, 12–13 MDR-1, 10 mitochondria, 101–103 molecularly targeted agents, 9 oncogenic signaling pathways, 8 radiation therapy, 11–12 regulation, 4, 5 small molecule inhibitors cell-based HTS, 5 mechanism of action, 7, 8 protein–protein interaction, 7 signaling pathways, 6, 7 xenograft models, 8 Hypoxia inducible factor (HIF) CA IX, 65, 66 degradation, 25
757 HIF-a molecules, 46 regulation genetic alterations of upstream regulators, 47–49 oxygen, 46–47 signaling, complexity, 12 target genes, 49–50 Hypoxia-response element (HRE), 66 I IkB protein, 375 IKKb, 375–377 IL-15 cytokines, 434 Imatinib, 110 Induced pluripotent stem (iPS) cell state, 51 Inducible nitric oxide synthase (iNOS), 372 Infinite replication paradigm, 216 Insulin-like growth factor family of binding proteins (IGFBPs), 659 b3 Integrin, 682 Integrins EMT characteristics and classifications, 517 differentiated epithelial cells, 517 E-cadherin protein, 518–520 MMPs, 519 N-cadherin expression, 518–519 signaling molecules, 518 endothelial cells migration, proliferation, and survival, 296 expression and function, 294–296 growth factor receptor cooperation a1b1 and a2b1 integrins, 688 ligand/receptor signaling system, 685 nitric oxide production, 686 SMAD2 phosphorylation, 687 TCPTP, 688 TGF-bIIR, 686, 687 Tie-2/a5b1 association, 687 avb3, 685–687 VEGFR2, 685, 686 ligand specificity, 296–297 lymphangiogenesis (see Lymphangiogenesis) mechanotransduction, 520–521 motility, and invasion actin-rich protrusion, 514 ACTN4 and CTGF, 517 EOC metastasis model, 515, 516 filopodia and lamellipodia, 514 gene expression, 516–517 MT1-MMP, 515–516
758 Integrins (cont.) provisional tumor stroma, 514 receptor occupancy vs. aggregation, 515 pathological angiogenesis ECM ligands, 678 functional hubs, 675 functional integration, 678 growth factor (see Growth factor) growth factor receptor cooperation, 685–688 laminin-10 expression, 678–679 transmembrane heterodimers, 677 avb3, 678, 679 signaling antiadhesion strategy, 522 antitumor efficacy, 521 biomimetic scaffolds, 523 FAK, 297–299, 522 ILK expression, 522–523 LECs and ECM, interactions, 290 paxillin, 301 Rho family, GTPases, 300 SHC, 299–300 talin, 300 vinculin, 301 structure and function cell–matrix contact, 510 domain structure and conformational alteration, 510, 511 FAK, 510, 512–513 integrin-linked kinase, 513–514 monomeric ligand, 510 Interferon-a (IFN-a), 362 Interferon g, 434 Interferon-g (IFN-g), 362 Interleukin-7 (IL-7), 721 Interleukin-12 (IL-12), 721–722 Interleukin-15 (IL-15), 721 Interleukin-17 (IL-17), 722 Interleukin-like EMT inducer (ILEI), 652 Interwoven signaling network, 674 Intracellular oxidative stress, 146 Isocitrate dehydrogenase 1 (IDH1), 48 K Killer cell immunotherapy adoptive therapy adjunctive strategies, 444–445 adoptively transferred NK cells, 444 alternative NK cell sources, 443 haploidentical NK cells, 441–442 host factors, 445
Index NK cell lines, 443–444 in vitro NK cell expansion, 442–443 LAK cells, 441 Killer immunoglobulin-like receptors (KIR) gene, 435–436 KM12SM cells, 284 Krogh radius, 25 KTI. See Kunitz trypsin inhibitor Kunitz trypsin inhibitor (KTI), 656–657 L Lactate dehydrogenase (LDH-A), 97–98 Lewis lung carcinoma (LLC) cell line, 379 Loss of heterozygosity (LOH), 48 LOX. See Lysyl oxidase Lymphangiogenesis a4b1, 302–303 a5b1, 302 a9b1, 301–302 a1b1 and a2b1, 302 angiogenesis, 220 induction, 291–293 lymphatic makers, 291 lymphatic vasculature, 290 pathology, 293–294 Lymphatic vessels, 290 Lymphedema, 293 Lymphocytes, 433 Lymphoid progenitors, 432–433 Lymphokine activated killer (LAK) cells, 441 Lysyl oxidase (LOX) ECM, 547 HIF target genes, 50 hypoxia, 201 MMPs, 482 Lytic granule exocytosis, 435 M Macromolecules biosynthesis, 98 Macrophage-colony stimulating factor (M-CSF), 372, 378, 686 Macrophages angiogenesis, 376–377 anti-tumor potential/therapeutic implications, 379–380 chemoattractants, 372 chemokines, 372, 374, 378 immunosuppressive phenotype, 373–374 inflammation, 375–376 iNOS, 372
Index metastasis definition, 377 LLC cell line, 379 M-CSF and EGF receptor, 378 MMP-9, 379 MMTV-PyMT mice, 378 TAMs, 378–379 monocytes, 371 Maladaptive response, tissue or organ, 214 Mammary carcinogenesis, 359 Mammary epithelial cells (MECs), 649 Mast cells benefits chemoattractants, 357 histamine, 357, 358 immunosuppression, 358, 359 metalloproteinases, 357, 358 tumor angiogenesis, 357 biology, 356–357 breast cancer, 359–360 cancer cell proliferation, 355 degranulation nitric oxide (NO), 355 pro-tumor molecules, 361 SCF, 357 tumor-derived blockers, 362 heparan sulfate proteoglycans, 362 IFN-a and IFN-g, 362 lung cancer, 361–362 melanoma and basal cell carcinoma, 360 metastasis, 354 pancreatic cancer, 360–361 TRAIL, 362–363 tumor enhancing effect, 356 W/Wv mast cell deficient mice, 356 Matrikines angiostatin, 702, 704 anti-angiogenic peptides, 702, 703 arresten and canstatin, 703 endostatin, 702–703 fibulin-1 and fibulin-5, 704 MMP inhibitors, 703–704 TSP1 and 2, 704 tumstatin, 703 Matrix metalloproteinases (MMPs) anticancer effects, 540–541 biology, 532–533 cathepsin K, 704 chemistry, 533–534 chemokines, 538 collagenolytic MMPs, 492 fibrillar collagen, 538 function regulation
759 CD44, 537 epigenetic mechanism, 536 furin-like recognition domain, 536 gene expression, 535–536 hypothetical model, proMMP-2 activation, 536, 537 MT1-MMP, 536–537 in vivo activity, 535 gene expression signatures, 539–540 heterotrimeric type I and II collagen, 492 HMGA1, 539 inflammation and cancer, 547 inhibitors, 703–704 integrins, 519 invasive/metastatic cancer phenotype cancer cell invasion, three-dimensional matrix, 543–545 epithelial-to-mesenchymal transition, 545–546 premetastatic niche, 546–547 protease-independent cell invasion, 545 LOX, 482 malignant transformation, 531 metastatic melanoma, 491 migration, invasion, and metastasis, 217–219 natural inhibitors, 535 PAR-1, 539 physiologic vs. neoplastic invasion, 539 pro-stromal factors, 719–720 protease classification, 532 stromal cell production fibroblasts, 541, 542 immunohistochemical examinations, 541 squamous carcinoma cells, 542 tumor angiogenesis, 542–543 therapeutic targets catalytic Zn2+, 548 exocyte binding and alosteric inhibitors, 549 impressive tumor regression, 548 MT1-MMP, 548–549 RNAi technology, 549–550 type IV basement membrane collagen, 547 tumor progression, 492 VEGF anti-VEGF antibody, 549 ECM and glycosaminoglycan degradation, 705 primary tumors, 547 tumor angiogenesis, 542–543 type IV collagen, 492, 493
760 Maximum tolerated dose (MTD), 264 M-CSF. See Macrophage-colony stimulating factor Mechanism regulating tumor, 291, 292 Mechanoreciprocity, 696 Mechanosensitivity, 520 MEK/Erk activation, 682 Membrane transporters, role, 29 Membrane type 1 matrix metalloproteinase (MT1-MMP), 515–516 Mesenchymal–epithelial transition (MET), 188 Metastasis cascades decathlon, 214–215 infinite replication paradigm, tumor cells, 216 potential mechanisms, 215 CD81 and CD63, 569–570 CD9 interferes EGFR, 567–568 homoclustering, 566 integrin-mediated motility, 567 MMP2 expression, 569 positive correlation, 566 transendothelial migration, 569 tumorigenicity, 566, 569 WAVE2, 568 Wnt proteins, 567, 568 chemokines CCR4, 613–614 CCR7, 616 CXCL8, 615 CXCL9 and CXCL10, 616, 617 CXCR3, 616–617 CXCR4, 614–615 CXCR1 and CXCR2, 615 multifaceted role, 602, 603 SCH-479833 and SCH-527123, 615–616 tumor-host interaction, 613 disease, 214 EMT, 562 hypoxia, 44–45 macrophages definition, 377 LLC cell line, 379 M-CSF and EGF receptor, 378 MMP-9, 379 MMTV-PyMT mice, 378 TAMs, 378–379 migration and invasion angiogenesis process, 219–220 chemotaxis, 216
Index classic metastatic cascade, 217–218 clinical trials, MMP, 218 emboli formation, 217–218 heparin-binding factors, 217 lymphangiogenesis, 220 mechanistic similarities, 219, 220 MMP, blockage, 217 tumor-cell-centric view, 218 promoting activities, 579–580 rethinking metastasis angioprevention, 223–224 born metastatic hypothesis, 221–222 metastatic decathlon, 221 metronomic therapy, 224 pre-metastatic niche, 223 stem/cancer-initiating cells, 222 transcriptome profiles, 221 tumor cell colonization, 223 suppressor gene CD82/KAI1 c-Met signaling, 564, 565 cysteines, 563 EGFR, 564, 565 gangliosides contribution, 565 integrins and cadherins, 563 KITENIN, 565 rescue, 578–579 TM interactions, 566 uPAR and a5b1, 564 xenogeneic system, 564 Tspan8, 572–573 tumors as tissues, 213–214 Metastatic promoter, 243 Metformin, 110 Micrometastasis angiogenic dormancy, 231–234 immunity-driven dormancy apoptosis resistance mechanism, 236 cytotoxic role, 235 immunesurveillance, 234 non-Hodgkin’s lymphoma, 234 tumor dormancy vs. immune-mediated control, 234–235 Migration-stimulating factor (MSF), 465–466 Mismatch repair (MMR), 140–141 Mitochondrial DNA (mtDNA), 99 Mitochondrial fumarate hydratase, 100 Mitochondrial genomic integrity, 99 MMPs. See Matrix metalloproteinases MN antigen, 62. See also Carbonic anhydrase IX MSF. See Migration-stimulating factor MT1-MMP. See Membrane type 1 matrix metalloproteinase
Index Mucosal mast cells (MMC), 356 Multidrug resistance gene 1 (MDR-1), 10 Multivesicular bodies (MVB), 558, 573–574 MVB. See Multivesicular bodies Myeloid-derived suppressor cells (MDSCs) EMT stimuli, 195 mast cells, 359 myeloid cell recruitment, 423 N NADPH, 98 Na+/H+ exchangers (NHE), 30–31 Natural killer cells (NKC) activation, 434 anti-tumor response, 437 cytokine secretion, 434 cytotoxicity, 435 haploidentical transplantation allogeneic HCT, 438 GVHD risk, 438–439 non-myeloablative transplantation, 440 potential benefits, NK cell, 439 umbilical cord transplantation, 440 immunophenotype adhesion glycoprotein CD56, 432 ontogeny, 432–433 killer cell immunotherapy adoptive therapy, 441–445 LAK cells, 441 licensing, 437 localization and trafficking, 433 receptors activation, 436–437 signals activation and inhibitory, 435–436 tumor infiltrating lymphocytes, 438 Neoangiogenesis angiogenesis APE1, 156 metastasis, 376, 377 EPCs controversy surrounding functions, 261 Matrigel plugs, 262 integrins, 296 MMPs, 542 tumour-associated stroma anti-angiogenic therapy, 337 human breast MCF-7-ras tumours, 338–339 PDGF, 338, 340
761 proangiogenic cytokines and chemokines, 337, 338 therapy-resistant tumour, 338 VEGF, 337–338 tumourigenesis, 327 Neovascularization, 610 Neuropilin, 318 NF-kB. See Nuclear factor-kappa B NKC. See Natural killer cells NKG2D receptor, 437 Non-homologous end joining (NHEJ), 143 Non-small cell lung cancer (NSCLC), 362, 717, 718 Non-steroidal anti-inflammatory drugs (NSAIDs), 375 Notch signal, 191 Nuclear factor-kappa B (NF-kB) pro-inflammatory milieu, 419 proteins, 375–376 Nucleosome-free region (NFR), 124 Nucleotide excision repair (NER), 141–142 O Oncogenes activation c-Myc, 104–105 PI3K/Akt pathway, 105–106 Ras, 105 Oncogenesis, 9 Osseous metastasis, 659 Oxygen-dependent HIF regulation, 46, 47 P PAMP. See Proadrenomedullin N-terminal 20 peptide Pancreatic ductal adenocarcinoma (PDAC), 360 Parathormone related peptide (PTHrP), 704, 705 Passenger methylation, 127 Pasteur effect, 26 Patched 1 (PTCH1) receptor, 336 Pathological angiogenesis blood vessel formation arteriogenesis, 675 ECM molecules, 673, 674 embryonic vasculogenesis, 675 interconnected networks, 673 intussuception/sprouting, 676, 677 normal and pathological vessel formation, 675
762 Pathological angiogenesis (cont.) integrins ECM ligands, 678 functional hubs, 675 functional integration, 678 growth factor (see Growth factor) growth factor receptor cooperation, 685–688 laminin-10 expression, 678–679 transmembrane heterodimers, 677 avb3, 678, 679 neovascularization control, 674 receptor/ligand networks, 675, 676 Paxillin, 301 PDGF pathways, 312, 313 Pentose phosphate pathway (PPP), 98 Perforin, 435 Pericytes biology, physiology, and pathology, 312–313 resistance-antiangiogenic therapy adaptive resistance, 316 antivascular strategies, 317–318 bioavailability, tumor, 316 HIF1-a, 317 metastasis risk, 316 PDGF signaling, 316–317 VEGF pathway, 315 tumor angiogenesis, 313–315 Pharmacokinetics, CECs, 265 Phosphatase and tensin homologue (PTEN) signalling, 335 Phosphatidylinositol 3-kinase (PI3K), 105–106 Phosphofructokinase (PFK1), 96–97 Phosphoinositide 3 kinase (PI-3K) pathway, 48–49 Plasmacytoid and myeloid dendritic cells, 434 Platelet-derived growth factor (PDGF)-BB stimulation, 714 Pleiotrophin (PTN) receptor, 684, 686 Pluripotency, 172 31 P nuclear magnetic resonance (NMR), 28 POU5FI, 172–173 Proadrenomedullin N-terminal 20 peptide (PAMP) angiogenesis, 736 prostate cancer, 742 structure and function, 733 therapeutic target, 745 Progenitors, 312
Index Pro-inflammatory milieu bone marrow-derived cells, 424 chronic inflammation vs. acute inflammation, 411–414 dynamic renaissance, 410 Helicobacter pylori colonization, 410, 411 neoplastic transformation, 410 soluble mediator, immune response, 417 tumor immune evasion and progression, 415–416 malignant epithelial initiation, 414–415 matrix remodeling proteases, 418 myeloid cell recruitment chemokines, 422, 423 immune cell subpopulation, 422 MDSCs, 423 M2 polarization, 423, 424 TAM (See Tumor associated macrophages) oxidative stress species, 417–418 transcription factors and primary inflammatory cytokines IL-1, 421–422 IL-6, 420–421 NF-kB, 419 TGF-b, 420 TNF, 422 tumor progression, metastatic potential, and inflammation, 416–417 Proly hydroxylase, 101 Protease-activated receptor-1 (PAR-1), 539 Prox1, 291 PTHrP. See Parathormone related peptide Pyruvate kinase (PKM2), 97 R RANKL. See Receptor activator of nuclear factor-kB ligand RCC. See Renal cell carcinomas Receptor activator of nuclear factor-kB ligand (RANKL), 704, 705, 720–721 Receptor activity modifying protein (RAMP)2, 734, 735, 744 Receptor tyrosine kinases (RTKs), 487 Redox signaling DNA repair APE1, redox activity (See Apurinic/ apyrmidinic endonuclease 1) BER, 139–140 cancer therapeutic approach, 159–161 DR, 137–139
Index global influences, APE1, 154–158 HR, 143–144 MMR, 140–141 NER, 141–142 NHEJ, 143 schematics, 137, 138 tumor microenvironment, 158–159 GRX/GSH system, 146–147 Trx system, 145 Re-epithelialization, 192 Regional hypermethylation, 126–127 Regulatory CD4+ cells CD4+CD25+ regulatory T cells, 396–397 markers, 394–395 suppression, 397 tumor-induced CD4+ regulatory T cells anti-inflammatory mechanisms, 395 immune response, 396 origin, 398 types, 394 Renal cell carcinomas (RCC) adrenomedullin cell lines and human biopsies, 736 immunohistochemical analysis, 743 prognostic biomarker, 744 CA IX, 72, 75 CXCL1, CXCL3, CXCL5, and CXCL8, 611 CXCR4, 614 RENCAREX®, 76 Rho family, GTPases, 300 Rho-GTPase activation, 697 RNA interference (RNAi) technology, 549–550 S SDF-1. See Stromal cell-derived factor 1 Secreted growth factors immune system modulators, 721–722 malignant cell growth factors EGFR, 722–723 GnRH, 725 heregulin, 725 HGF/MET signaling, 725–726 homo- and heterodimers, 723, 724 insulin and IGFs, 726 KRAS mutation, 723 netrin-1, 726–727 neuregulins, 723 S100A4 protein, 727 Wnt family, 727–728
763 pro-angiogenic and lymphangiogenic factors adaptive/evasive resistance, 715 bevacizumab, 712, 717, 718 CXCR4, 716–717 EG-VEGF and Bv8 proteins, 715–716 FGF, 713–714 IFNa and IFNb, 714 lymphangiogenesis, 714, 745 metastatic renal cell cancer, 717 multiple myeloma plasma cells, 716 NSCLC, 717, 718 PDGF-BB stimulation, 714 PlGF, 712–713 receptor tyrosine kinase inhibitors, 9–10 signaling process, 712 tumor stromal and vascular cells, 715 pro-stromal factors MMP, 719–720 RANKL, 720–721 reactive stroma, 718 SPARC, 720 TGFb, 718–719 uPA, 719 Secreted protein acidic and rich in cysteine (SPARC), 720 SHC, 299–300 Small cell lung carcinomas (SCLC), 362 Squamous cell carcinoma (SCC), 279, 616 STAT3, 190–191 Stem cell factor (SCF), 357 Stroma-derived prognostic predictor (SDPP), 280 Stroma fibroblasts, 213 Stromal cell-derived factor 1 (SDF-1), 335 Stromal matrix metalloproteinases, 359 S180 tumor, 745 Succinate dehydrogenase (SDH), 100 Suppressor gene CD82/KAI1 c-Met signaling, 564, 565 cysteines, 563 EGFR, 564, 565 gangliosides contribution, 565 integrins and cadherins, 563 KITENIN, 565 rescue, 578–579 TM interactions, 566 uPAR and a5b1, 564 xenogeneic system, 564 Surrogate biomarker, 264–265 Synthesis of cytochrome c oxidase 2 (SCO2), 103–104
764 T Talin, 300 TAMs. See Tumor associated macrophages ab T cells, 386–387 Telomerase, 172 Tetraspanins angiogenesis CD151, 576, 577 mRNA and miRNA, 577 plasma protein extravasation, 562 platelet-derived exosomes, 577 Tspan8, 576–577 Tspan32, 577 CD151 and tumor cell motility adhesion process, 572 FAK competent and deficient fibroblasts, 571 human epidermoid carcinoma line, 570 knockdown cells, 571 TEM location, 572 egg–sperm fusion, 561, 579 exosomes coagulation and homeostasis, 575 mRNA and microRNA, 574 MVB, 573–574 proteins, 574 exosome–target cell interaction, 561 major functional activities cellular penetration, invasion, and fusion, 560 exosomal tetraspanins, 561 functional divergence, 562 MMP transcription and secretion, 561 molecule trafficking and biosynthesis, 560 metastasis CD81 and CD63, 569–570 CD9 interferes, 566–569 EMT, 562 promoting activities, 579–580 suppressor gene CD82/KAI1, 563–566, 578–579 Tspan8, 572–573 premetastatic niche, 575–576 structure, 556, 557 web CD82, 560 CD63 and CD151, 559 cytosolic signal transduction molecules, 557 exosome composition, 558, 559 MVB, 558 PKC, PI4KII and PLCg, 557–558 protein-protein interactions, 558
Index TEM, 558 transmembrane and cytosolic proteins, 556 TGF-b. See Transforming growth factor-b TGFb1, 192 TGF-b-response signature (TBRS), 647 Thioredoxin (Trx) system, 145 THP. See Triple-helical peptide Thrombospondin-1 (TSP-1), 700–701, 704 TILs. See Tumor-infiltrating lymphocytes Tissue fibroblasts generation, 188 Toll-like-receptors (TLRs), 412, 419 Transcription-coupled repair (TCR), 12 Transforming growth factor-b (TGF-b) adenocarcinoma, 634 ANGPTL4, 647–648 animal models CRT-PCR amplification, 661 doxycycline-mediated induction, 662 HT and WT mice, 661–662 immunostaining, 660, 661 MCS-1 cells, 664 MDA-MB-435 cells, 663 MMTV/PyVmT transgenic model, 662 Neu-induced mammary tumorigenesis and metastasis, 662 4NQO-induced rat malignant oral keratinocytes, 663 bronchioloalveolar carcinoma (BAC), 634 CAFs, 647 CD8+ T cells, 653 drugs, treatments, and therapies BBI and KTI, 656–657 bryostatin 1 and phorbol-12-myristate13-acetate, 655 calcium homeostasis, 657 13-cis-retinoic acid, 656 cyclooxygenase-2 metabolism, 654 decorin, 657–658 doxorubicin, 658 Fujimycin, 655 LM8-DCN cell, 657–658 NHBE and NSCLC cells, 655–656 rapamycin, 656 retinoid metabolism, 654 tranilast (N-[3,4dimethoxycinnamonyl]-anthranilic acid), 655 EMT, 649–652 ER-negative breast tumor, 647 genomics bone metastasis, 658, 659 gene expression, 659 HARA cells, 658–659
Index host-derived and cell autonomous effects, 660 p53 protein-associated genes, 659 PTHrP and ezrin identification, 658 immune system, 653–654 intratumoral histologic heterogeneity, 634 isoforms A549 cells, 639 immunohistochemical staining, 638 MTLn3 cells, 640 PC3 cells, 640 R3327-MATLyLu cells, 639 TGF-b1, 639–640 TGF-b3, 638–639 metastasis, 633–634 molecular cloning and functional analysis, 635 polypeptide growth factor, 646–647 primary carcinomas, 646 pro-stromal factors, 718–719 receptors Cre/LoxP technology, 644 Fc:TGF-b RII fusion protein, 644 integrins, 642 MDA-MB-435-F-L cells, 641–642 microarray analysis, 642 overexpressing cells, 643 potent immunosuppressive cytokine, 643 Smad2/3 signaling, 640–641 TGF-b RI gene mutation, 641 TGF-b RIIDN, 643–644 TGF-b RII expression, 642–643 TGF-b sRIII, 645 seed and soil theory, 636 signaling antiproliferative effects, 243 BHLHB3, 244–245 cancer cell invasion and metastasis, 341 cysteine knot, 636 growth response, 244 mechanisms, cellular dormancy, 239, 244 normal stroma-derived tumoursuppressive signalling, 335–336 role, DTC dormancy, 243–244 Smad family protein, 637–638 transcriptional signatures, 245 type I and type II receptors, 636–637 Smads, 645–646 TBRS, 647 tumor-associated host cells, 636 uPAR, 648 Transketolase and transaldolase, 98 Tregs. See T regulatory cells
765 T regulatory cells (Tregs), 373–374, 378 Triple combination therapy, 318 Triple-helical peptide (THP), 483, 484, 488 TSP-1. See Thrombospondin-1 Tumor associated fibroblast (TAFs), 277 Tumor associated macrophages (TAMs) EMT stimuli, 193–194 fibronectin, 461, 462, 465 macrophages anti-tumor potential/therapeutic implication, 379–380 immunosuppressive phenotype, 373–374 inflammation, 375, 376 iNOS and arginase, 372 metastasis, 378–379 myeloid cell recruitment functional and phenotypic properties, 423 IL-23/IL-17 axis, 424 leukocytes, 422 MDSCs, 423 M2 polarization, 423–424 pro-angiogenic factor, 718 squamous cell carcinoma invasion, 355 tumor angiogenesis and growth, 354 Tumor-conditioned medium (TCM), 278 Tumor hypoxia, 25, 26 Tumor-infiltrating lymphocytes (TILs) antitumor immunity, 387, 392–393 CD4+ T cells adaptive immunity, 388 CTLs (see Cytotoxic T lymphocytes) hallmark phenotype, 388 predominant effector mechanism, 389 T helper 1 vs. T helper 2 response, 392–393 CD8+ T cells cell memory response, 391–392 MHC class I and MHC class II, 387–388 murine models, 388 cytokine IL-17, 386 malignant tumors, 385 myeloid cell recruitment, 386 prognosis, 399–401 regulatory CD4+ cells CD4+CD25+ regulatory T cells, 396–397 markers, 394–395 suppression, 397 tumor-induced CD4+ regulatory T cells, 395–396, 398 types, 394 targeted tumor tissues, 401–402 ab T cells, 386–387 Th17 cells, 398–399
766 Tumor milieu, 281 Tumor pH acid transport H+ ion buffering, 30 H+ ions extrusion, 30–31 H+ ions, nonrespiratory sources, 31 intracellular and extracellular pH, 28 metabolic acid efflux, 29–30 pH regulation, 28–29 CA dominant role, 35 facilitated CO2 diffusion, 32–34 facilitated H+ diffusion, 34 intracellular and extracellular anhydrase isoforms, 32 transport metabolon, 34 cellular acid sources cellular respiration, 25–26 Warburg effect, 26–27 pH biological importance hydrogen nucleus, 23 pHi displacements, 24 pHi regulation, 24–25 weak acids/bases, 24 Tumor radioresistance, 11 Tumor suppressor p53, 279–280 Tumour-promoting stromal myofibroblasts CAFs, 327 cell-autonomous fashion, 331 counterpart and normal fibroblasts, 329, 330 genetic and karyotype analysis, 329 heterogeneous cellular origins, 333–334 high-grade malignancy and poor prognosis, 328 immunohistochemical analysis, 330, 331 a-SMA expression, 329 cancer cell invasion and metastasis EMT, 341 metastatic carcinoma cells, 340 pre-metastatic niche, 342 TGF-b-Smad signalling, 341 tumour cell-derived paracrine soluble factors, 342 desmoplastic/reactive stroma, 326 drug resistance, 343 epithelium, 325 neoangiogenesis anti-angiogenic therapy, 337 EPCs, 340 human breast MCF-7-ras tumours, 338–339
Index PDGF, 338, 340 proangiogenic cytokines and chemokines, 337, 338 therapy-resistant tumour, 338 VEGF, 337–338 normal stroma-derived tumour-suppressive signalling carcinoma cells, 337 Hh and Smo signaling pathways, 336–337 Notch and PTEN signalling, 335 SDF-1, 335 TGF-b signalling, 335–336 somatic genetic and epigenetic alterations, 331–333 tissue fibrosis share characteristics, 327–328 Twist, 50 Tyrosine kinase inhibitors, 314 U Umbilical cord blood (UCB) transplants, 440 uPA receptor (uPAR), 648 Urokinase-type plasminogen activator (uPA), 719 V Vascular disrupting agents (VDAs), 266 Vascular endothelial growth factor (VEGF) MMPs anti-VEGF antibody, 549 ECM and glycosaminoglycan degradation, 705 primary tumors, 547 tumor angiogenesis, 542–543 type IV collagen, 492, 493 neoangiogenesis, 337–338 pro-angiogenic and lymphangiogenic factors anti-VEGF antibody, 713 cell culture, 712 EG-VEGF, 715–716 glycoproteins, 712 lymphatic vessel growth, 714–715 PDGF, 717 PlGF, 712–713 Trap, 717–718 type IV collagen, 492, 493 VEGFR-3, 291 VEGF receptor (VEGFR) bone marrow-derived progenitor cells, 715 CXCR4+VEGFR1+ hematopoietic cells, 716
Index kinase activity, 713 PlGF signal, 713 receptor tyrosine kinase, 714, 717 signaling process, 712 Vimentin, 189 Vinculin, 301
767 Volociximab. See Anti-a5b1 integrin functionblocking antibody Von Hippel-Lindau syndrome, 48 W Warburg effect, 14, 26–27, 100, 280