Molecularly Targeted Therapy for Childhood Cancer
w
Peter J. Houghton ╇ ╇ Robert J. Arceci ●
Editors
Molecularly Targeted Therapy for Childhood Cancer
Editors Peter J. Houghton Elizabeth M. and Richard M. Ross Chair Director, Center for Childhood Cancer The Research Institute Nationwide Children’s Hospital 700 Children’s Drive Columbus, OH 43205
[email protected]
Robert J. Arceci King Fahd Professor of Pediatric Oncology Professor of Pediatrics Oncology and Cellular and Molecular Medicine Kimmel Comprehensive Cancer Center at Johns Hopkins The Bunting and Blaustein Cancer Research Building 1650 Orleans Street, Suite 207 Baltimore, MD 21231 USA
[email protected]
ISBN 978-0-387-69060-5 e-ISBN 978-0-387-69062-9 DOI 10.1007/978-0-387-69062-9 Springer New York Dordrecht Heidelberg London Library of Congress Control Number: 2010930693 © 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 software, 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)
Acknowledgments
We would like to especially thank all the authors and investigators who contributed to the content of this book along with the publisher for believing in the importance of directly addressing important issues concerning children with cancer.
v
w
Contents
Part Iâ•… Hematologic Malignancies The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia..................................................................... William L. Carroll and Rob Pieters
3
Molecular Targeted Therapies in T-CellAcute Lymphoblastic Leukemia................................................................................ Alejandro Gutierrez and A. Thomas Look
19
Molecularly Targeted Therapy for Infant ALL............................................ Patrick A. Brown and Carolyn A. Felix
31
Targeted Therapeutic Approaches for AML................................................. Robert J. Arceci and Donald Small
59
Acute Promyelocytic Leukaemia.................................................................... Andrea Biondi, Anna Maria Testi, and Brenda E.S. Gibson
83
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity to Chemotherapy......................................... 109 Jeffrey W. Taub, Yubin Ge, and Yaddanapudi Ravindranath Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)........................................................................................... 123 Jennifer O’Hara Lauchle and Benjamin S. Braun Chronic Myeloid Leukemia: Pathophysiology and Therapeutics............... 139 Seth J. Corey and Jorge Cortes Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes............................................................................ 155 Lia Gore vii
viii
Contents
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma............................................................................... 177 Megan S. Lim and Mitchell S. Cairo Molecular Targeting of Post-transplant Lymphoproliferative Disorders........................................................................................................... 215 Michael Wang and Thomas G. Gross Part IIâ•… Solid Tumors Molecularly Targeted Therapies for Astrocytomas...................................... 231 Ian F. Pollack Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy for Childhood Cancer...................................................... 267 Yoon-Jae Cho and Scott L. Pomeroy Future Treatments of Ependymoma.............................................................. 291 Richard J. Gilbertson Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions of INI1/hSNF5 Tumor Suppressor........................ 305 Ganjam V. Kalpana and Melissa E. Smith Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors......................................................................... 331 Brigitte Widemann Molecular Therapy for Neuroblastoma......................................................... 351 Yaël P. Mossé and John M. Maris Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows..................................................................................................... 373 Jeffrey A. Toretsky and Aykut Üren Molecular Targeted Therapy for Wilms’ Tumor.......................................... 401 James I. Geller and Jeffrey S. Dome Molecular Therapy for Rhabdomyosarcoma................................................ 425 Raushan T. Kurmasheva, Hajime Hosoi, Ken Kikuchi, and Peter J. Houghton Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?......................................................................... 459 Rosanna Ricafort and Richard Gorlick
Contents
ix
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments Based on a Better Understanding of Disease Biology.................................................................. 499 Stephen X. Skapek Index.................................................................................................................. 521
w
Introduction to Targeted Therapy Text Robert J. Arceci and Peter Houghton
The dream of “magic bullets” to treat patients with cancer, as defined by Paul Ehrlich in the early 1900s during his Harben Lectures, promised great hope for molecules that would specifically react with and eradicate tumors without harming the host. His lecture was entitled, “Experimental Researches on Specific Therapy. On Immunity with Special Reference to the Relationship between Distribution and Action of Antigens.”1 Many decades have come and gone since those concepts were first presented along with many lives prematurely lost to cancer. One might ask then “Where did we go wrong?” As if trying to push Ehrlich’s ideas more rapidly forward, Hollywood and Warner Brothers Studios premiered in 1940 the movie “Dr. Ehrlich’s Magic Bullets,” starring Emanuel Goldenberg, aka Edward G. Robinson. During that same period, under the auspices of the Department of Defense, Drs. Louis Goodman and Alfred Gilman were enlisted to investigate autopsy results of soldiers who had died following mustard gas exposures, leading to one of the earliest insights into the selective ablation of the hematopoietic elements and subsequently to early clinical trials in patients with primarily lymphoid malignancies.2 During the first half of the twentieth century, the Indian investigator, Dr. Yellapragada Subbarao, was successful in synthesizing a spectrum of antimetabolites, leading in turn to the landmark study by Farber et€ al., “Temporary Remissions in Acute Leukemia in Children Produced by Folic Acid Antagonist, 4-aminopteroyl-glutamic acid (Aminopterin), in 1948.3 In 1956, Li, Hertz, and Spencer reported the effect of methotrexate on gestational related choriocarcinoma leading eventually to a curative therapy and demonstrating that indeed a small molecule could cure a widely metastatic cancer.4 These were of course all examples of targeted cancer therapy. Following the death of her husband, Albert Lasker, from cancer in the 1950s, Mrs. Mary Lasker and her Citizens Committee for the Conquest of Cancer, effectively promoted the conviction that the federal government had the ability and resources to develop a national crusade to cure cancer. A second United States National Cancer Act (The first was in 1937 and established the National Cancer institute.) was signed by President Richard Nixon in 1971. There was a tremendous amount of excitement during those decades for the day when Dr. Ehrlich’s magic bullets would cure all cancers. Comprehensive Cancer Centers arose, small molecule drug screening began on a wide scale leading to the development of new therapies, xi
xii
Introduction to Targeted Therapy Text
training in oncologic specialties expanded, the maturation of the clinical cooperative groups for adults and children occurred, the need for multidisciplinary treatment teams and supportive care became apparent, and the initiation of combination chemotherapy, adjuvant and neoadjuvant approaches to treatment along with sanctuary site prophylaxis were developed. Over the course of those decades, many cancer patients of all ages benefited. For children with cancer, cancer evolved from a disease that was nearly always fatal to a disease in which about 75% of patients can be cured. Yet, a large percentage of patients still could not expect to be cured. And for those who were cured, the adverse consequences of treatment were too often severe. Thus, as combinations of more intense doses and delivery schedules were tested, the very real plateau of balancing cure and toxicity became an increasingly difficult challenge. Had Dr. Ehrlich’s dream come to an end? Fortunately, throughout those decades, basic laboratory and translational investigators were pushing forward with novel tools and approaches to extend our knowledge of the molecular basis of different cancers and the hosts they afflict. This basic understanding of genetics and epigenetics has formed the substrate of a resurgence of hope in the form of the next generation of targeted therapies. The new therapeutic agents are based on the distinct genetic and epigenetic signatures that translate into unique protein expression patterns that can be exploited for the benefit of patients. From the dawn of this resurgence, stemming from the demonstration of targeting BCR-ABL with imatinib,5 a plethora of novel tumor specific or selective approaches are being applied. Targeting these tumor specific pathways is based on the development of small molecule signal transduction modulators to monoclonal antibodies to antitumor vaccines to nanoparticles. In addition, a growing recognition of the confluence of stem cell biology to the behavior of tumor initiating stem cells is provided for the first time in history the correct cellular targets. The treatment of children with cancer has been at center stage addressing the basic causes of cancer as well as developing and exploiting novel molecularly targeted therapies. This volume attempts to bring together in one place some of the successes, but mostly the hope of eradicating the morbidity and mortality of both hematopoietic malignancies and solid tumors in pediatric patients. The chapters, written by expert laboratory and clinical investigators, provide updated information on current and future approaches to targeted therapy in pediatric oncology. Further, there is a strong belief that the development of less toxic and more effective combinations of therapeutic approaches will also be important in improving the outcome of the 85% of children with cancer who live in developing countries and too often cannot be successfully treated. While the pace of progress has continued to accelerate, for too many patients the pace remains too slow. As Mario Andretti once said, “If things seem under control, you’re not going fast enough.” It is with that sense of urgency that the current state of targeted therapy should be catapulted to the next level based on an increasingly detailed molecular description of both tumors and hosts as we enter an era of increasing genetically directed therapies. There still remains an immense need for
Introduction to Targeted Therapy Text
xiii
more financial support, more collaboration, more discovery, more innovation, and more translation, but there has never been, nor will there ever be a lack of passion and hard work. Given the opportunities, those dedicated in eradicating this ancient enemy of the young and old will most assuredly succeed and bring into reality the magic bullets of Paul Ehrlich.
References Witkop B. Paul Ehrlich and His magic Bullets, Revisited. Proceedings of the American Philosophical Society. Vol. 143. Philadelphia, PA: American Philosophical Society; 1999; 540–557. Goodman LS, Wintrobe MM, Dameshek W, Goodman MJ, Gilman A, McLennan MT. Nitrogen mustard therapy. Use of methyl-bis(beta-chloroethyl)amine hydrochloride and tris(beta-chloroethyl)amine hydrochloride for Hodgkin’s disease, lymphosarcoma, leukemia and certain allied and miscellaneous disorders. JAMA. 1946; 251:2255–2261. Farber S, Diamond LK, Mercer RD, Sylvester RF Jr, Wolff JA. Temporary Remissions in acute leukemia in children produced by folic acid antagnonist, 4-aminopteroyl-glutamic acid (aminopterin). NEJM. 1948; 238:787–793. Li M, Hertz R, Spencer DB. Effect of methotrexate therapy upon choriocarcinoma and chorioadenoma. Proc Soc Exp Biol Med. 1956; 93:361–366. Druker BJ, Talpaz M, Resta DJ, et€al. Efficacy and safety of a specific inhibitor of the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med. 2001; 344:1031–1037.
w
Contributors
Robert J. Arceci King Fahd Professor of Pediatric Oncology, Professor of Pediatrics, Oncology and Cellular and Molecular Medicine, Kimmel Comprehensive Cancer Center at Johns Hopkins, The Bunting & Blaustein Cancer Research Building, 1650 Orleans Street, Suite 207, Baltimore, MD 21231, USA
[email protected] Andrea Biondi Centro M. Tettamanti, Clinica Pediatrica Università di Milano-Bicocca, Ospedale San Gerardo, Via Pergolesi, 33, 20052 Monza, Italy
[email protected] Patrick A. Brown Sidney Kimmel CCC at Johns Hopkins, Bunting-Blaustein Cancer Research Building, 1650 Orleans Street, CRB I-2M49, Baltimore, MD 21231, USA
[email protected] Mitchell S. Cairo Division of Pediatric Blood and Marrow Transplantation, New York-Presbyterian Morgan Stanley Children’s Hospital, Columbia University, 3959 Broadway, CHN 10-03, New York, NY 10032, USA
[email protected] William L. Carroll New York University Cancer Institute, 522 First Avenue, Smilow 1201, New York, NY 10016, USA
[email protected] Seth J. Corey Departments of Pediatrics and Cell & Molecular Biology, Children’s Memorial Hospital and the Robert H. Lurie Comprehensive Cancer Center, Northwestern University, Chicago, IL
[email protected]
xv
xvi
Contributors
Jeffrey S. Dome Division of Oncology, Center for Cancer and Blood Disorders, Children’s National Medical Center, 111 Michigan Avenue NW, Washington, DC 20010, USA
[email protected] Carolyn A. Felix Division of Oncology, Department of Pediatrics, The Children’s Hospital of Philadelphia, University of Pennsylvania School of Medicine, Colket Translational Research Building, Room 4006, 3501 Civic Center Blvd, Philadelphia, PA
[email protected] Lia Gore The University of Colorado Cancer Center and The Center for Cancer and Blood Disorders, The Children’s Hospital, University of Colorado Denver, Pediatrics Mail Stop 8302, P.O. Box€6511, Aurora, CO 80045, USA
[email protected] Richard Gorlick Department of Pediatrics, Pediatric Hematology/Oncology, The Children’s Hospital at Montefiore and the Albert Einstein College of Medicine, 3415 Bainbridge Avenue, Rosenthal 3rd Floor, Bronx, NY 10467, USA
[email protected] Richard J. Gilbertson Neurobiology and Brain Tumor Program, St. Jude Children’s Research Hospital, Memphis, TN, USA
[email protected] Thomas G. Gross Division of Hematology/Oncology/BMT, The Ohio State University College of Medicine, 700 Children’s Drive, Columbus, OH 43205, USA
[email protected] Peter J. Houghton Department of Director, Center for Childhood Cancer, The Research Institute Nationwide Children’s Hospital, 700 Children’s Drive, Columbus, OH 43205
[email protected] Ganjam V. Kalpana Department of Molecular Genetics and Albert Einstein College Cancer Center, Albert Einstein College of Medicine of Yeshiva University, 1300, Morris Park Ave., Ullman 821, Bronx, NY 10461, USA
[email protected]
Contributors
xvii
Jennifer O’Hara Lauchle University of California, Helen Diller Family Cancer Research Building (Optional), 1450 3rd Street, Room 264 (Lauchle) and 265 (Braun), San Francisco, CA 94158, USA
[email protected] A. Thomas Look Department of Pediatric Oncology, Dana-Farber Cancer Institute and Children’s Hospital Boston, 44 Binney Street, Boston, MA 02115, USA and Harvard Medical School, Boston, MA, USA
[email protected] Yaël P. Mossé Division of Oncology, Department of Pediatrics, Children’s Hospital of Philadelphia, University of Pennsylvania, 3615 Civic Center Blvd., ARC 907C, Philadelphia, PA, USA
[email protected] Rob Pieters Department of Pediatric Oncology and Hematology, Erasmus MC, Sophia Children’s Hospital, Room Sp2456, Dr Molewaterplein 60, 3015 GJ Rotterdam, The Netherlands
[email protected] Ian F. Pollack Department of Neurosurgery, Children’s Hospital of Pittsburgh, 3705 Fifth Avenue, Pittsburgh, PA, USA
[email protected] Scott L. Pomeroy Department of Neurology, Children’s Hospital Boston, 300 Longwood Avenue, Enders 270 , 02115 Boston, MA, USA
[email protected] Yaddanapudi Ravindranath Children’s Hospital of Michigan, 3901 Beaubien Boulevard, Detroit, MI 48201, USA
[email protected] Donald Small Sidney Kimmel CCC at Johns Hopkins, Bunting-Blaustein Cancer Research Building, 1650 Orleans Street, CRB I-252, Baltimore, MD 21231, USA
[email protected]
xviii
Contributors
Stephen X. Skapek Pediatric Hematology/Oncology, The University of Chicago Comer Children’s Hospital, 900 E. 57th Street, Chicago, IL 60637, USA
[email protected] Jeffrey W. Taub Division of Hematology/Oncology, Children’s Hospital of Michigan, 3901 Beaubien Boulevard, Detroit, MI 48201, USA and Department of Pediatrics, Wayne State University School of Medicine, Detroit, MI 48201, USA and Developmental and Therapeutics Program, Barbara Ann Karmanos Cancer Institute, 4100 John R Street, Detroit, MI 48201, USA e-mail:
[email protected] Jeffrey A. Toretsky Departments of Oncology and Pediatrics, Lombardi Comprehensive Cancer Center, Georgetown University, 3970 Reservoir Rd. N.W., Washington, DC 20057-1469, USA
[email protected] Brigitte Widemann Pharmacology and Experimental Therapeutics Section, National Cancer Institute, Pediatric Oncology Branch, 10 Center Drive, 10-CRC, Room 1-5750, MSC 1101, Bethesda, MD 20892, USA
[email protected]
Part I
Hematologic Malignancies
This page intentionally left blank.
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia William L. Carroll and Rob Pieters
One of the most fundamental goals of modern cancer research is to develop more effective therapies that specifically target the cancer cell while sparing normal cells from the collateral damage that is common to conventional therapies. The cornerstone of current cancer treatment depends on drugs associated with a very narrow therapeutic index in that the effective dose and the toxic dose frequently overlap. While progress in pediatric oncology, specifically, improved cure rates for the most common childhood malignancy, acute lymphoblastic leukemia (ALL), has outpaced improvements in other cancer subtypes, treatment for ALL still relies on conventional cytotoxic agents thereby exposing children to considerable short- and long-term side effects. Optimally, targeted therapy would converge on a specific lesion or pathway in the cancer cell, not shared by normal cells, that is essential to the maintenance of the population of leukemia cells. In order to be most effective, the agent must have efficient access to the biological target to maximize tumor kill. Fortunately, in the past few years, a spectrum of agents have been, or are currently being developed, that target cancer cells with increasing specificity. With the initial development of monoclonal antibodies to specific leukocyte antigens, it became clear that leukemia cells display a unique profile of surface proteins shared by a minority of normal cells. These antigens have proved to be useful targets for therapeutic antibodies. Likewise, breakthroughs in methods to analyze gross chromosome structure and the subsequent introduction of recombinant DNA technology led to the recognition that all human cancers were the result of disrupted biological pathways due to somatic, and sometimes germline, alterations in DNA copy number (e.g., amplifications, deletions) and/or structure (translocations, mutations, etc.). Thus, there are many examples of where altered biological pathways that drive cancer initiation and maintenance can be traced back to these sentinal genetic lesions.
W.L. Carrollâ•›(*) New York University Cancer Institute, 522 First Avenue, Smilow 1201, New York, NY 10016, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_1, © Springer Science+Business Media, LLC 2010
3
4
W.L. Carroll and R. Pieters
The identification of the specific genes and pathways involved in transformation and their subsequent validation in various experimental models now provide the first bona fide molecular targets. These studies have also determined that the discovery of a somatically mutated gene in a cancer cell does not necessarily indicate that it will prove to be a useful target, and subsequent analysis to distinguish “driver” from “passenger” mutations will prove key to the prioritization potential targets (Frohling et€ al. 2007). This complex process is now even more daunting with the massive amount of data coming from more recent advances in high throughput sequencing, proteomic approaches, epigenetic profiling, and analysis of non-coding RNAs. While many approaches to targeted therapy are being considered, this overview will focus on those targets and their corresponding targeted agents that are currently in clinical trials that best illustrate the challenges and opportunities for this, the next generation of clinical trials in ALL.
The BCR/ABL Tyrosine Kinase and Ph+ ALL The t(9;22)(q34;q11) or Philadelphia chromosome was the first nonrandom chromosomal abnormality detected in cancer and is identified in 2% of childhood ALL samples and in approximately 20% of adult cases (Moorman et€al. 2007; Schultz et€al. 2007). The translocation results in the well known BCR-ABL protein fusion that is a constitutively active tyrosine kinase (Chan et€al. 1987; Clark et€al. 1988). Historically, the presence of the Ph+ genotype correlates with an extremely poor outcome and has been used in risk stratification (Secker-Walker et€al. 1997; Schultz et€al. 2007). While age, WBC, and early response may select for patients with a better outcome, stem cell transplantation (SCT) strategies have been usually considered for this subgroup of patients (Ribeiro et€ al. 1997; Schrappe et€ al. 1998). However, even if stem cell transplantation is performed after successful induction, many patients will still succumb to the disease. In the largest analysis to date in children, the 5€ year disease-free survival is 25±4% with chemotherapy alone compared to 65±8% when SCT is performed (Arico et€al. 2000). The discovery of the biological function of the fusion protein as well as recognition that BCR-ABL was essential for tumor maintenance led to efforts to screen for inhibitors. The ATP binding pocket proved to be an attractive target for selective inhibition. Imatinib mesylate was the first such compound to enter clinical trials. Considerable preclinical data showed that it bound to the inactive confirmation of the kinase domain, blocked kinase activity, and inhibited proliferation of leukemia cells harboring the t(9;22). The results from early trials in chronic myelogenous leukemia led to considerable enthusiasm for integration into treatment strategies for Ph+ ALL (Druker et€al. 1996, 2001, 2006). In phase I studies performed in children and adults, the drug was well tolerated and the majority of patients showed a response including a complete response rate of approximately 20% (Ottmann et€al. 2002). However, these responses were short-lived with most patients showing
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
5
disease progression within weeks of their initial response. Studies incorporating imatinib into a backbone of chemotherapy soon followed these initial observations. A number of protocols using a combination of imatinib and chemotherapy in adults showed that this approach improved the induction rate, led to a lower disease burden, increased the number of patients proceeding to SCT, and most importantly, showed an impact on EFS and overall survival. For example, a study by Thomas and colleagues from MD Anderson Cancer Center investigated the efficacy of imatinib given concurrently with the hyper-CVAD regimen (Thomas et€al. 2004). The CR rate on this study was 93% and the 3€year relapse free and overall survival rate was 62% and 55%, respectively, far better than historical controls. The Children’s Oncology Group COG initiated a study for Ph+ ALL (AALL0031) in 2002. Children and adolescents 1 to 21€years of age with Ph+ ALL were eligible after initial induction using standard chemotherapy. Imatinib was introduced in a stepwise fashion over five cohorts of patients to assess potential toxicity when used with aggressive chemotherapy. Patients with an HLA identical sibling proceeded to stem cell transplantation after two consolidation courses. Other patients received three additional intensification and re-induction blocks followed by 12 8-week cycles of maintenance chemotherapy for a total protocol duration of just over 2€ years. Patients in cohort one received 42€ days of imatinib before maintenance whereas patients in cohort 5 received 280€days. All groups received 336€days of imatinib in maintenance and patients who received SCT had 6€months of imatinib post transplant. Imatinib was well tolerated in the context of aggressive chemotherapy on AALL0031. There was modest ALT elevation noted in maintenance that led to a shortening of the duration of imatinib from 21 to 14€days for each month of maintenance. In addition, there was an increase in grade III and IV neutropenia during re-induction block 2 and a lower total WBC in consolidation. The 3€year EFS of patients on cohort 5 who received continuous imatinib was 80.5±11.2%, well above the 35.0±4.4% EFS reported on previous cooperative group studies. Twenty one patients received matched sibling transplants, and there was no difference in outcome among those receiving chemotherapy versus SCT. Furthermore, end induction minimal residual disease was not prognostic in those patients exposed to continuous imatinib indicating that targeted therapy can overcome the adverse prognostic significance of slow early response to chemotherapy. These striking results, while early, are extremely encouraging and represent an example of the promise of targeted therapy. Has been well documented that imatinib resistance is most commonly due to the development of BCR-ABL mutants (usually in the P loop or ATP-binding site) that alter affinity for imatinib (Branford et€al. 2003; O’Hare et€al. 2007). Newer generation tyrosine kinase inhibitors such as dasatinib and nilotinib, among others, inhibit such mutants, show more potent suppression of BCR-ABL kinase activity, bind to both the inactive and active conformation of BCR-ABL (e.g., dasatinib), and inhibit important collateral pathways such as those involving Src that are known to play a role in Ph+ ALL (Shah et€al. 2004; Weisberg et€al. 2005; Talpaz et€al. 2006). Preliminary data from an ongoing European phase I/II study with dasatinib show promising activity in Â�children with
6
W.L. Carroll and R. Pieters
Ph+ ALL who are refractory or intolerant to imatinib (Zwaan et€ al. 2008). Therefore, dasatinib is being used in the current COG study (AALL0622) and is integrated earlier in treatment during induction to maximize early tumor kill.
The FLT-3 Pathway in MLL Rearranged Infant ALL Infants under 1€year of age constitute an especially high risk subgroup of patients with ALL. The primary biological basis for this association is that leukemic blasts from such children harbor an 11q23 rearrangement involving the MLL gene, usually t(4;11), t(11;19) or t(9:11) (Silverman 2007). The 6€ year event free survival (EFS) for these infants ranges from 22% to 43% and aggressive attempts to improve outcome using augmented chemotherapy including stem cell rescue have largely failed (Pui et€ al. 2002; Hilden et€ al. 2006; Pieters et€ al. 2007). Thus, novel approaches are clearly warranted. The FMS-like tyrosine kinase-3 (FLT-3) is a class III receptor tyrosine kinase, a family that includes the c-KIT and platelet derived growth factor (PDGF) receptors (Agnes et€al. 1994) (Fig.€1). FLT-3 is expressed in almost all cases of AML and B-lineage ALL, and up to a third of T ALL (Carow et€al. 1996; Rosnet et€al. 1996).
FLT3L-Binding
Extracellular Domain
Transmembrane Domain Juxtamembrane Domain Kinase Domain
Internal Tandem Duplications Activation Loop Mutations P
P
Fig.€1╅ The FLT-3 Receptor. The FLT-3 receptor contains an extracellular domain consisting of five immunoglobulin like domains. The juxtamembrane portion is a target for tandem duplications. Two kinase domains are linked through a tyrosine kinase insert. The kinase domain is mutated in hematological malignancies and most mutations are single point substitutions with insertions and deletions occuring less frequently. Binding of FLT-3 ligand leads to receptor dimerization and phophorylation of the kinase domain that triggers the activation of downstream pathways. (Revised based on reference (Stirewalt and Radich 2003))
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
7
Constitutive activation of FLT-3 is observed through somatic mutation and/or a through co-expression of the FLT-3 ligand as part of an autocrine loop (Brasel et€al. 1995; Drexler 1996; Nakao et€al. 1996). Activating mutations of FLT-3, either in the juxtamembrane region or kinase domain, are seen in 22% of childhood AML samples but are rare in childhood ALL cases (Schnittger et€al. 2002; Thiede et€al. 2002). Juxtamembrane mutations are characterized by internal tandem duplications (FLT/ITD) and insertion/deletion mutations. Importantly, FLT/ITD constitutes an adverse prognostic variable in AML (Meshinchi et€al. 2001). High level mRNA and protein expression of FLT-3 is observed in infant and childhood ALL with rearrangements of MLL and cases associated with hyperdiploidy – two subtypes with distinctly different outcomes (Armstrong et€ al. 2002; Yeoh et€al. 2002). FLT-3 mutations are absent in most subtypes of childhood ALL except for MLL-rearranged and hyperdiploid cases where up to 18 and 25% of samples, respectively, show either kinase mutations (MLL and hyperdiploid) and/ or juxtamembrane mutations (hyperdiploid) (Armstrong et€al. 2004; Taketani et€al. 2004). Other studies showed a much lower incidence of FLT-3 mutations in these subtypes and that the level of expression may be related to outcome in MLL rearranged ALL (Stam et€ al. 2007a, b). Finally, several studies have demonstrated constitutive phosphorylation of FLT-3 in MLL rearranged infant ALL and hyperdiploid cases even in the absence of mutation suggesting an autocrine loop (Brown et€al. 2005; Stam et€al. 2005). Given the importance of FLT-3 in childhood leukemia, there has been great interest in exploring the therapeutic potential of FLT-3 inhibitors. A number of such agents have been developed including CEP-701 (lestaurtinib), SU112248 (sunitinib malate), and PKC412 (midostaurin) among others (Stirewalt and Radich 2003). Lestuartinib is of particular interest because it is currently being evaluated in two Children’s Oncology Group (COG) trials for MLL rearranged infant ALL, as well as, relapsed FLT-3 mutant AML. Preclinical work has demonstrated the therapeutic potential of this agent (Brown et€ al. 2005, 2006). Phase II trials in adults with refractory/relapsed AML demonstrate tolerability and clinical responses, especially in patients with FLT-3 mutant AML (Smith et€al. 2004). A phase I trial in children with heavily pretreated neuroblastoma (lestuartinib also inhibits the Trk neurotropin receptors) has also demonstrated the feasibility of administration. It is noteworthy that preclincal studies with combination chemotherapy show that synergism is achieved when chemotherapy is delivered first followed by FLT-3 inhibition with lestuartinib while the reverse sequence is antagonistic (Brown et€ al. 2006). This important observation underscores the need for careful preclinical evaluation when combining conventional chemotherapy with biologically targeted agents. COG AALL0631 will test whether FLT-3 inhibition with lestuartinib administered on a backbone of chemotherapy will improve outcome for children with MLLrearranged ALL. The trial has a limited institution safety phase where two cohorts of children (<90€days and ³90€days) will receive lestuartinib with chemotherapy, and after this portion of the study is completed, the full randomized study will commence group wide.
8
W.L. Carroll and R. Pieters
A second interesting FLT3 inhibitor is midostaurin. This drug was identified in 2002 and was been shown to be active against MLL rearranged ALL cells in a mouse models (Weisberg et€al. 2002; Armstrong et€al. 2003). Also, midostaurin is differentially cytotoxic to primary ALL cells with MLL rearrangements derived from infants compared to cells from other ALL patients (Stam et€al. 2005). In these studies, target validation was shown by dephosphorylation of constitutively active FLT3. Clinical studies in refractory/relapsed adult AML have demonstrated moderate clinical activity of midostaurin as a single drug (Stone et€al. 2005). In combination, these data have led to the design of a phase I/II study with midostaurin in children with a relapse of MLL rearranged ALL in Europe.
The Notch Pathway in T-ALL The Notch family of proteins are transmembrane receptors for cell surface ligands of the Delta-Serrate-Lag2 (DSL) family (Demarest et€ al. 2008). There are four members of the Notch family and structurally they exist as heterodimers (Fig.€2). The extracellular domain has an extended domain of epidermal growth factors repeats. Two regions at the C terminus of the extracellular domain, a cysteine rich
ICN MAML
EGF-like repeats
Lin12/Notch repeats
CSL
Myc
AKT PEST domain
mTOR
Heterodimerization domain Transmembrane domain
Notch intracellular domain
PTEN
HES1
SCF FBXW7
Proteosome Degradation
Fig.€2╅ The Notch Pathway. The Notch receptor contains an extracellular domain with a series of epidermal growth factor receptor (EGF) repeats followed by a negative regulatory region composed of three Lin12/Notch repeats and a heterodimerization domain. Upon ligand binding two cleavage events (first via ADAMS like protease and the second by g secretase) occur leading to release of the intracellular notch domain (ICN) that functions as a transcription factor. ICN degradation is mediated by phophorylation of the PEST domain followed by degradation in the proteosome. ICN translocates to the nucleus where it binds a transcriptional repressor CSL (CBF-1/ Suppressor of Hairless/Lag1). This complex then recruits MAML (Mastermind like) transcriptional activators leading to transcriptional activation of key target genes. (Figure based on (Nefedova and Gabrilovich 2008; Palomero and Ferrando 2008))
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
9
region and the heterodimerization domain (HD), function in a negative fashion to prevent receptor activation (Gordon et€al. 2007). The intracellular domain functions as a transcription factor and includes a region that binds to the DNA-binding CSL protein, as well as a PEST domain that mediates the protein half life of intracellular Notch (Fortini and Artavanis-Tsakonas 1994; Oberg et€al. 2001). Upon ligand binding the negative regulatory region of the extracellular domain undergoes a conformational change and exposes a proteolytic site that is cleaved by an ADAM metalloproteinase. This is followed by a second cleavage mediated by a g-secretase complex, thus releasing intracellular Notch (ICN) from the plasma membrane (Schroeter et€al. 1998). ICN translocates to the nucleus where it associates with the CSL protein and converts a transcriptional repressor complex into an activator, thus activating downstream biological pathways. Notch activation is terminated by phosphorylation in the PEST region and binding of the F-box protein Fbw7 that is part of an SCF-E3 ubiquitin ligase complex (SCFFBX7) (Oberg et€al. 2001; Fryer et€al. 2004). Notch was first implicated in T-ALL through the description of a t(7;9) (q34;q34.3) that is observed in <1% of all patients with the disease (Ellisen et€al. 1991). By virtue of the translocation, the 3¢ region of Notch is juxtaposed to the TCR-b locus resulting in constitutive expression of intracellular Notch. However, more recently, it was discovered that over 50% of T-ALL cell lines and samples show point mutations in the HD domain (26%), the PEST domain (12.5%), or both (17.7%) (Weng et€al. 2004; Malecki et€al. 2006). Mutations in the HD domain result in ligand independent proteolytic cleavage of Notch, whereas the PEST mutations prevent Fbx7 interaction, thus prolonging half life of the protein. Furthermore, 16% of T-ALL samples show mutations in Fbw7 leading to loss of recognition of Notch, thus leading to prolonged half life of the protein (Thompson et€ al. 2007). Many samples revealed mutations at relapse that were not present at initial diagnosis(Weng et€ al. 2004; Thompson et€ al. 2007). The importance of Notch activation in the T-ALL transformation pathway is underscored by the observation that forced expression of ICN in murine hematopoietic cells leads to T-ALL in 100% of recipient mice (Pear et€ al. 1996). In summary, these data collectively indicate that the activation of Notch plays a dominant role in T-ALL transformation and maintenance of the malignant phenotype. The fact that the Notch pathway is involved in both hematological and solid neoplasms prompted many investigators to examine approaches to inhibit the pathway. g-secretase complex mediated activation of Notch has drawn the most attention. The complex is composed of many proteins including presenelin and it processes many important proteins including b-amyloid precursor protein (Lleo 2008). There was great interest in g-secretase inhibitors (GSIs) for the treatment of Alzheimer’s disease and GSIs were already in clinical development at the time of the initial discoveries of widespread Notch mutations in T-ALL. Early work showed that the GSI, compound E, was able to elicit cell cycle arrest in 5 of 30 T-ALL cell lines prompting an early clinical trial (Weng et€al. 2004). While the full results of this clinical trial have yet to be published, it was marked by the lack of clinical responses and severe gastrointestinal toxicity. A recent article showed that GSI
10
W.L. Carroll and R. Pieters
restored dexamethasone sensitivity of T-ALL blasts and importantly glucocorticoid treatment protected mice from GSI-induced gut toxicity. Thus, these results support future trials incorporating glucocorticoids and GSIs in the treatment of refractory T-cell disease (Real et€al. 2009). The downstream pathways modulated by Notch are just beginning to be discovered and provide additional nodal points for inhibition. c-Myc is a direct target of Notch, and in turn, c-Myc activates the m-TOR pathway in at least a subset of T-ALLs (Palomero et€al. 2006; Chan et€al. 2007). In vitro data showed synergistic inhibition of T-ALL growth with simultaneous application of GSI (Compound E) and the m-TOR inhibitor rapamycin. Notch also positively regulates the NF-kB signaling pathway and inhibition of NF-kB with bortezomib results in apoptotic cell death in T-ALL cell lines and combined therapy with a GSI (LY411575) and bortezomib causes synergistic cytotoxicity. Finally, work comparing GSI sensitive and resistant cell lines showed equivalent Notch inhibition indicating corruption of downstream pathways in GSI resistant lines. Notch is known to downregulate the tumor suppressor PTEN (mediated by the direct Notch target genes MYC and HES1) that in turn inhibits AKT (Palomero et€al. 2007). The PI3K-AKT pathway drives many cellular processes including cell growth, proliferation, and survival. GSI resistant cell lines showed mutational loss of PTEN and this finding was also observed in 17% of T-ALL samples. In this context, inhibitors of the PI3K-AKT signaling pathway already in development may be excellent candidates for therapy.
Monoclonal Antibodies in the Treatment of ALL Leukemia-associated antigens are promising targets for therapy for ALL and the impact of monoclonal antibodies (mAbs) is just being realized in spite of the fact that they have been in development for decades. While not truly tumor-specific, expression is restricted to narrow subsets of normal hematopoietic cells, and therefore, they are unlikely to be associated with the broad collateral damage to normal tissues so characteristic of conventional chemotherapy. Therapeutic Mabs, while of murine origin, are frequently engineered so that the bulk of the murine backbone is replaced with human immunoglobulin regions. mAbs can be delivered in an unconjugated format or linked to conventional agents, immunotoxins, or radioactive elements. In principle, unconjugated Mabs may be more safely integrated into chemotherapy regimens. Rituximab, directed against the CD20 antigen, was the first mAb approved by the Food and Drug Administration. It is a chimeric, unconjugated antibody whose mechanism of action depends on antibody and complement dependent cell-mediated cytotoxicity (Harris 2004). Rituximab also induces direct tumor cell apoptosis. The greatest experience with rituximab has been in adult indolent and aggressive B-cell non-Hodgkins lymphoma where it provides significant advantage over chemotherapy alone (Feugier et€al. 2005; Pfreundschuh et€al. 2006). In addition, it has
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
11
been used in children with recurrent/refractory B-cell NHL and mature B-cell ALL with encouraging results (Griffin et€al. 2009). Expression of CD-20 is associated with a poor prognosis in adult ALL but no such relationship is seen in pediatric ALL (Jeha et€al. 2006; Thomas et€al. 2008). CD-20 is expressed in only 50% of pediatric ALL but a recent study indicates that the antigen is upregulated during induction such that CD-20 positivity was 45% in initial samples but increased to 81% in residual blasts at end-induction (Dworzak et€al. 2008). Moreover the relative level of expression is also increased. Thus, a clinical trial could be designed to incorporate rituximab for those with high levels of MRD at end induction. Given the prognostic significance of CD-20 expression in adult ALL, many trials have incorporated rituximab into chemotherapy protocols. Thomas et€ al. reported their experience with chemoimmunotherapy with Hyper-CVAD and rituximab in adult Burkitt and Burkitt Type lymphoma and ALL(Thomas et€ al. 2006). Overall 3€year survival (89%) and event free survival (80%) was superior to their experience with hyper-CVAD alone. Experience in children is limited to case reports and small series. Morris et€al. described successful remission induction using rituximab alone in a patient with multiply relapsed ALL (Morris and Vora 2007). An additional three children with refractory ALL have been treated with rituximab-based chemoimmunotherapy with encouraging results (Claviez et€al. 2006). CD-22 is another B-cell surface antigen that is expressed in over 90% of children with ALL (Raetz et€al. 2008). Epratuzumab is a humanized mAb that has also shown encouraging results in adult indolent and aggressive NHL and has been used in combination with rituximab (Carnahan et€ al. 2003; Leonard et€ al. 2005). The Children’s Oncology Group has completed a pilot protocol using epratuzumab in combination with chemotherapy for children with ALL in first or later marrow relapse (Raetz et€al. 2008). Therapy started with a 2€week window of epratuzumab alone administered twice weekly followed by four weekly doses combined with standard re-induction chemotherapy. Effective targeting of CD-22 was validated in 10 of 11 patients and epratuzumab was well-tolerated. Grade 1 and 2 infusion reactions were observed in 10 of 15 patients during the reduction phase but quickly resolved with temporary stopping of the infusions and administration of steroids and/or meperidine. All patients tolerated the rest of the initial as well as additional infusions. Nine patients achieved a complete remission in this highly refractory population and seven were MRD negative at end induction. A larger COG study using epratuzumab twice a week during induction for first early marrow relapsed ALL is underway. Alemtuzumab is another humanized mAb that recognizes CD52, which is expressed on most B and almost all T malignancies (Xia et€al. 1993; Frampton and Wagstaff 2003). It has been approved for use in chronic lymphocytic leukemia and is also part of myelablative conditioning regimens. Limited experience with single agent alemtuzumab in adult ALL has been disappointing (Tibes et€al. 2006). A COG phase II trial in children with ALL in second or greater relapse was suspended early due to low accrual, but only one of the 13 patients showed a significant response (Personal communication, Anne Angiollilo). Alemtuzumab is
12
W.L. Carroll and R. Pieters
associated with a higher number of infectious complications that may limit its use in chemoimmunotherapy protocols. Gemtuzumab ozogamicin (GO) is a humanized anti-CD33 mAb that has significant activity in relapsed AML and is currently being evaluated as part of frontline therapy for that disease (Sievers et€al. 2001). The CD33 antigen is expressed on 90% of AML samples but is restricted to 15 to 20% of ALL (Putti et€al. 1998; Sievers et€al. 2001). In vitro and in€vivo mouse model activity has been observed for GO in CD-33+ ALL and case reports demonstrate that the agent has activity either alone or in combination with additional agents in relapsed ALL (Balduzzi et€al. 2003; Cotter et€al. 2003; Zwaan et€al. 2003; Golay et€al. 2005). GO is associated with hepatotoxicity and its use following BMT may be problematic. Moreover, its restriction to a small minority of ALL cases will limit its application to ALL in the future. CD-19 is a B lineage antigen that is expressed earlier in differentiation when compared to CD-22 and CD-20 and is expressed in almost all B-precursor ALLs. Many immunotherapeutic strategies have been developed including the use of unmodified antibodies, antibody-drug, and immunotoxin combinations as well as bi-specific antibodies (Grossbard et€al. 1993; Rowland et€al. 1993; Molhoj et€al. 2007; Horton et€al. 2008). Phase I studies have demonstrated safety but shortlived responses (Hekman et€al. 1991). A particularly novel approach has focused on the development of bi-specific antibodies to promote the engagement of T-cells. A recent report documented the clinical utility of blinatumomab, a CD-19/CD-3 single chain bi-specific T cell engager, in patients with relapsed B NHL (Bargou et€al. 2008). Eleven major responses were seen among 38 patients including four complete responses. The doses required to achieve these responses were far below those used for conventional mAb therapy highlighting the extraordinary lytic potential of cytotoxic T cells. This agent is currently being studied for use in ALL.
Cancer Stem Cells in ALL The existence of a small population of self-renewing cancer cells that are responsible for the proliferation and dissemination of the bulk population has now been documented for both hematological and solid tumors. Failure to eradicate such cancer stem cells may explain the emergence of resistant disease. Although recent studies indicate that the assay used to functionally characterize such cancer stem cells may be imperfect, they have been documented in CML, AML, central nervous system tumors, breast cancer, and colon cancer (Bonnet and Dick 1997; Wang et€al. 1998; Al-Hajj et€al. 2003; O’Brien et€al. 2007). Work in many laboratories to identify leukemic stem cells LSCs, also known as leukemia initiating cells (LICs) in CML and AML has defined a phenotype associated with self renewal. However, similar work in ALL has led to conflicting results about the surface phenotype of
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
13
LSCs (Cobaleda et€al. 2000; Cox et€al. 2004; Castor et€al. 2005). However, some of these discordant results may be due to the precision of the technique used to isolate these cells (e.g., different monoclonal antibodies), the functional assay (e.g., mouse strain used to demonstrate self renewal), and the biologic subtype of the samples being analyzed (e.g., Ph+ vs. ETV6-RUNX1). Available data favor that a CD34+/CD38− ALL subpopulation is capable of reconstituting hematopoiesis in a xenograft mouse model and that this population can be further enriched with CD133, an antigen present on many cancer stem cells (Cox et€ al. 2009). There is controversy, however, about whether LSCs express CD19 and this is crucial since this cell surface receptor is the target for monoclonal antibody therapy as described above (Castor et€al. 2005; Hong et€al. 2008). A recent study by Cox et€al. demonstrates the critical issue of targeting the LSC (Cox et€al. 2009). In a series of experiments it was noted that CD133+/CD19− cells were enriched during long-term in€ vitro culture and that this subpopulation led to engraftment of NOD/SCID animals. Furthermore, the engrafted cells showed high expression of CD10 and CD19 as well as a karyotype similar to the original bulk leukemia. Further refinement of leukemia stem cells could be facilitated with three color flow cytometry since most investigators have shown that these stem cells lack CD38 expression (e.g., CD133+/CD38−/CD19−). Most importantly, CD133+/ CD19− leukemia initiating cells in the study by Cox et€al. were resistant to glucorticoids and vincristine in contrast to the unsorted bulk leukemia. Thus new strategies to identify unique pathways in leukemia stem cells and assays to determine their sensitivity to pharmacological treatment will be essential to prevent re-emergence of the leukemia in selected cases.
Summary The development of truly tumor-specific therapy in childhood ALL is now a reality with the success of imatinib and other agents directed against the BCR-ABL fusion product. It is expected that other examples will soon follow and the challenge is how best to evaluate the scores of new agents entering the drug pipeline based on an understanding of the biology of childhood ALL. Such agents will undoubtedly make their initial impact when used with conventional chemotherapy. The relatively uncommon incidence of childhood ALL will limit the number of new agents that can be evaluated in a timely fashion. It is, therefore, incumbent on investigators to develop a rigorous prioritization scheme based on preclinical in€vitro data and in€vivo mouse models to promote the most promising compounds. In addition, the development of novel clinical trial designs based using a high bar for success with early stopping rules for those agents who fail to meet a meaningful standard is clearly needed. Nonetheless, it is clear that the gap between laboratory and clinical biologists is as narrow as it has ever been and the future holds great promise for major advances in the field.
14
W.L. Carroll and R. Pieters
Acknowledgementsâ•… This work was supported by the Penelope London Foundation, the Friedman Fund for Childhood Leukemia, and the Walter Family Pediatric Leukemia Fund. The authors thank Drs. Teena Bhatla and Elizabeth Raetz for review of the manuscript.
References Agnes, F., B. Shamoon, et€al. (1994). “Genomic structure of the downstream part of the human FLT3 gene: exon/intron structure conservation among genes encoding receptor tyrosine kinases (RTK) of subclass III.” Gene 145(2): 283–8. Al-Hajj, M., M. S. Wicha, et€al. (2003). “Prospective identification of tumorigenic breast cancer cells.” Proc Natl Acad Sci USA 100(7): 3983–8. Arico, M., M. G. Valsecchi, et€al. (2000). “Outcome of treatment in children with Philadelphia chromosome-positive acute lymphoblastic leukemia.” N Engl J Med 342(14): 998–1006. Armstrong, S. A., A. L. Kung, et€al. (2003). “Inhibition of FLT3 in MLL. Validation of a therapeutic target identified by gene expression based classification.” Cancer Cell 3(2): 173–83. Armstrong, S. A., M. E. Mabon, et€al. (2004). “FLT3 mutations in childhood acute lymphoblastic leukemia.” Blood 103(9): 3544–6. Armstrong, S. A., J. E. Staunton, et€al. (2002). “MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia.” Nat Genet 30(1): 41–7. Balduzzi, A., V. Rossi, et€al. (2003). “Molecular remission induced by gemtuzumab ozogamicin associated with donor lymphocyte infusions in t(4;11) acute lymphoblastic leukemia relapsed after transplantation.” Leukemia 17(11): 2247–8. Bargou, R., E. Leo, et€al. (2008). “Tumor regression in cancer patients by very low doses of a T cell-engaging antibody.” Science 321(5891): 974–7. Bonnet, D. and J. E. Dick (1997). “Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell.” Nat Med 3(7): 730–7. Branford, S., Z. Rudzki, et€al. (2003). “Detection of BCR-ABL mutations in patients with CML treated with imatinib is virtually always accompanied by clinical resistance, and mutations in the ATP phosphate-binding loop (P-loop) are associated with a poor prognosis.” Blood 102(1): 276–83. Brasel, K., S. Escobar, et€al. (1995). “Expression of the flt3 receptor and its ligand on hematopoietic cells.” Leukemia 9(7): 1212Â�–8. Brown, P., M. Levis, et€ al. (2006). “Combinations of the FLT3 inhibitor CEP-701 and chemotherapy synergistically kill infant and childhood MLL-rearranged ALL cells in a sequencedependent manner.” Leukemia 20(8): 1368–76. Brown, P., M. Levis, et€al. (2005). “FLT3 inhibition selectively kills childhood acute lymphoblastic leukemia cells with high levels of FLT3 expression.” Blood 105(2): 812–20. Carnahan, J., P. Wang, et€al. (2003). “Epratuzumab, a humanized monoclonal antibody targeting CD22: characterization of in€vitro properties.” Clin Cancer Res 9(10 Pt 2): 3982S–90S. Carow, C. E., M. Levenstein, et€al. (1996). “Expression of the hematopoietic growth factor receptor FLT3 (STK-1/Flk2) in human leukemias.” Blood 87(3): 1089–96. Castor, A., L. Nilsson, et€al. (2005). “Distinct patterns of hematopoietic stem cell involvement in acute lymphoblastic leukemia.” Nat Med 11(6): 630–7. Chan, L. C., K. K. Karhi, et€al. (1987). “A novel abl protein expressed in Philadelphia chromosome positive acute lymphoblastic leukaemia.” Nature 325(6105): 635–7. Chan, S. M., A. P. Weng, et€al. (2007). “Notch signals positively regulate activity of the mTOR pathway in T-cell acute lymphoblastic leukemia.” Blood 110(1): 278–86. Clark, S. S., J. McLaughlin, et€al. (1988). “Expression of a distinctive BCR-ABL oncogene in Ph1positive acute lymphocytic leukemia (ALL).” Science 239(4841 Pt 1): 775–7. Claviez, A., C. Eckert, et€al. (2006). “Rituximab plus chemotherapy in children with relapsed or refractory CD20-positive B-cell precursor acute lymphoblastic leukemia.” Haematologica 91(2): 272–3.
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
15
Cobaleda, C., N. Gutierrez-Cianca, et€al. (2000). “A primitive hematopoietic cell is the target for the leukemic transformation in human philadelphia-positive acute lymphoblastic leukemia.” Blood 95(3): 1007–13. Cotter, M., S. Rooney, et€al. (2003). “Successful use of gemtuzumab ozogamicin in a child with relapsed CD33-positive acute lymphoblastic leukaemia.” Br J Haematol 122(4): 687–8. Cox, C. V., P. Diamanti, et€al. (2009). “Expression of CD133 on leukemia initiating cells in childhood ALL.” Blood 113(14): 3287–96. Cox, C. V., R. S. Evely, et€al. (2004). “Characterization of acute lymphoblastic leukemia progenitor cells.” Blood 104(9): 2919–25. Demarest, R. M., F. Ratti, et€al. (2008). “It’s T-ALL about Notch.” Oncogene 27(38): 5082–91. Drexler, H. G. (1996). “Expression of FLT3 receptor and response to FLT3 ligand by leukemic cells.” Leukemia 10(4): 588–99. Druker, B. J., F. Guilhot, et€ al. (2006). “Five-year follow-up of patients receiving imatinib for chronic myeloid leukemia.” N Engl J Med 355(23): 2408–17. Druker, B. J., C. L. Sawyers, et€ al. (2001). “Activity of a specific inhibitor of the BCR-ABL tyrosine kinase in the blast crisis of chronic myeloid leukemia and acute lymphoblastic leukemia with the Philadelphia chromosome.” N Engl J Med 344(14): 1038–42. Druker, B. J., S. Tamura, et€al. (1996). “Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells.” Nat Med 2(5): 561–6. Dworzak, M. N., A. Schumich, et€al. (2008). “CD20 up-regulation in pediatric B-cell precursor acute lymphoblastic leukemia during induction treatment: setting the stage for anti-CD20 directed immunotherapy.” Blood 112(10): 3982–8. Ellisen, L. W., J. Bird, et€al. (1991). “TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms.” Cell 66(4): 649–61. Feugier, P., A. Van Hoof, et€al. (2005). “Long-term results of the R-CHOP study in the treatment of elderly patients with diffuse large B-cell lymphoma: a study by the Groupe d’Etude des Lymphomes de l’Adulte.” J Clin Oncol 23(18): 4117–26. Fortini, M. E. and S. Artavanis-Tsakonas (1994). “The suppressor of hairless protein participates in notch receptor signaling.” Cell 79(2): 273–82. Frampton, J. E. and A. J. Wagstaff (2003). “Alemtuzumab.” Drugs 63(12): 1229–43; discussion 1245–6. Frohling, S., C. Scholl, et€al. (2007). “Identification of driver and passenger mutations of FLT3 by high-throughput DNA sequence analysis and functional assessment of candidate alleles.” Cancer Cell 12(6): 501–13. Fryer, C. J., J. B. White, et€al. (2004). “Mastermind recruits CycC:CDK8 to phosphorylate the Notch ICD and coordinate activation with turnover.” Mol Cell 16(4): 509–20. Golay, J., N. Di Gaetano, et€ al. (2005). “Gemtuzumab ozogamicin (Mylotarg) has therapeutic activity against CD33 acute lymphoblastic leukaemias in€vitro and in€vivo.” Br J Haematol 128(3): 310–7. Gordon, W. R., D. Vardar-Ulu, et€al. (2007). “Structural basis for autoinhibition of Notch.” Nat Struct Mol Biol 14(4): 295–300. Griffin, T. C., S. Weitzman, et€al. (2009). “A study of rituximab and ifosfamide, carboplatin, and etoposide chemotherapy in children with recurrent/refractory B-cell (CD20+) non-Hodgkin lymphoma and mature B-cell acute lymphoblastic leukemia: a report from the Children’s Oncology Group.” Pediatr Blood Cancer 52(2): 177–81. Grossbard, M. L., J. M. Lambert, et€al. (1993). “Anti-B4-blocked ricin: a phase I trial of 7-day continuous infusion in patients with B-cell neoplasms.” J Clin Oncol 11(4): 726–37. Harris, M. (2004). “Monoclonal antibodies as therapeutic agents for cancer.” Lancet Oncol 5(5): 292–302. Hekman, A., A. Honselaar, et€al. (1991). “Initial experience with treatment of human B cell lymphoma with anti-CD19 monoclonal antibody.” Cancer Immunol Immunother 32(6): 364–72. Hilden, J. M., P. A. Dinndorf, et€al. (2006). “Analysis of prognostic factors of acute lymphoblastic leukemia in infants: report on CCG 1953 from the Children’s Oncology Group.” Blood 108(2): 441–51.
16
W.L. Carroll and R. Pieters
Hong, D., R. Gupta, et€ al. (2008). “Initiating and cancer-propagating cells in TEL-AML1associated childhood leukemia.” Science 319(5861): 336–9. Horton, H. M., M. J. Bernett, et€al. (2008). “Potent in€vitro and in€vivo activity of an Fc-engineered anti-CD19 monoclonal antibody against lymphoma and leukemia.” Cancer Res 68(19): 8049–57. Jeha, S., F. Behm, et€al. (2006). “Prognostic significance of CD20 expression in childhood B-cell precursor acute lymphoblastic leukemia.” Blood 108(10): 3302–4. Leonard, J. P., M. Coleman, et€al. (2005). “Combination antibody therapy with epratuzumab and rituximab in relapsed or refractory non-Hodgkins lymphoma.” J Clin Oncol 23(22): 5044–51. Lleo, A. (2008). “Activity of gamma-secretase on substrates other than APP.” Curr Top Med Chem 8(1): 9–16. Malecki, M. J., C. Sanchez-Irizarry, et€ al. (2006). “Leukemia-associated mutations within the NOTCH1 heterodimerization domain fall into at least two distinct mechanistic classes.” Mol Cell Biol 26(12): 4642–51. Meshinchi, S., W. G. Woods, et€al. (2001). “Prevalence and prognostic significance of Flt3 internal tandem duplication in pediatric acute myeloid leukemia.” Blood 97(1): 89–94. Molhoj, M., S. Crommer, et€al. (2007). “CD19-/CD3-bispecific antibody of the BiTE class is far superior to tandem diabody with respect to redirected tumor cell lysis.” Mol Immunol 44(8): 1935–43. Moorman, A. V., C. J. Harrison, et€al. (2007). “Karyotype is an independent prognostic factor in adult acute lymphoblastic leukemia (ALL): analysis of cytogenetic data from patients treated on the Medical Research Council (MRC) UKALLXII/Eastern Cooperative Oncology Group (ECOG) 2993 trial.” Blood 109(8): 3189–97. Morris, E. S. and A. Vora (2007). “Remission induction with single agent Rituximab in a child with multiply relapsed precursor-B ALL.” Br J Haematol 139(2): 344–5. Nakao, M., S. Yokota, et€al. (1996). “Internal tandem duplication of the flt3 gene found in acute myeloid leukemia.” Leukemia 10(12): 1911–8. Nefedova, Y. and D. Gabrilovich (2008). “Mechanisms and clinical prospects of Notch inhibitors in the therapy of hematological malignancies.” Drug Resist Updat 11(6): 210–8. O’Brien, C. A., A. Pollett, et€al. (2007). “A human colon cancer cell capable of initiating tumour growth in immunodeficient mice.” Nature 445(7123): 106–10. O’Hare, T., C. A. Eide, et€al. (2007). “Bcr-Abl kinase domain mutations, drug resistance, and the road to a cure for chronic myeloid leukemia.” Blood 110(7): 2242–9. Oberg, C., J. Li, et€ al. (2001). “The Notch intracellular domain is ubiquitinated and negatively regulated by the mammalian Sel-10 homolog.” J Biol Chem 276(38): 35847–53. Ottmann, O. G., B. J. Druker, et€al. (2002). “A phase 2 study of imatinib in patients with relapsed or refractory Philadelphia chromosome-positive acute lymphoid leukemias.” Blood 100(6): 1965–71. Palomero, T. and A. Ferrando (2008). “Oncogenic NOTCH1 control of MYC and PI3K: challenges and opportunities for anti-NOTCH1 therapy in T-cell acute lymphoblastic leukemias and lymphomas.” Clin Cancer Res 14(17): 5314–7. Palomero, T., W. K. Lim, et€al. (2006). “NOTCH1 directly regulates c-MYC and activates a feedforward-loop transcriptional network promoting leukemic cell growth.” Proc Natl Acad Sci USA 103(48): 18261–6. Palomero, T., M. L. Sulis, et€al. (2007). “Mutational loss of PTEN induces resistance to NOTCH1 inhibition in T-cell leukemia.” Nat Med 13(10): 1203–10. Pear, W. S., J. C. Aster, et€al. (1996). “Exclusive development of T cell neoplasms in mice transplanted with bone marrow expressing activated Notch alleles.” J Exp Med 183(5): 2283–91. Pfreundschuh, M., L. Trumper, et€ al. (2006). “CHOP-like chemotherapy plus rituximab versus CHOP-like chemotherapy alone in young patients with good-prognosis diffuse large-B-cell lymphoma: a randomised controlled trial by the MabThera International Trial (MInT) Group.” Lancet Oncol 7(5): 379–91.
The Emerging Era of Targeted Therapy in Childhood Acute Lymphoblastic Leukemia
17
Pieters, R., M. Schrappe, et€al. (2007). “A treatment protocol for infants younger than 1€year with acute lymphoblastic leukaemia (Interfant-99): an observational study and a multicentre randomised trial.” Lancet 370(9583): 240–50. Pui, C. H., P. S. Gaynon, et€al. (2002). “Outcome of treatment in childhood acute lymphoblastic leukaemia with rearrangements of the 11q23 chromosomal region.” Lancet 359(9321): 1909–15. Putti, M. C., R. Rondelli, et€al. (1998). “Expression of myeloid markers lacks prognostic impact in children treated for acute lymphoblastic leukemia: Italian experience in AIEOP-ALL 88-91 studies.” Blood 92(3): 795–801. Raetz, E. A., M. S. Cairo, et€al. (2008). “Chemoimmunotherapy reinduction with epratuzumab in children with acute lymphoblastic leukemia in marrow relapse: a Children’s Oncology Group Pilot Study.” J Clin Oncol 26(22): 3756–62. Real, P. J., V. Tosello, et€al. (2009). “Gamma-secretase inhibitors reverse glucocorticoid resistance in T cell acute lymphoblastic leukemia.” Nat Med 15(1): 50–8. Ribeiro, R. C., A. Broniscer, et€al. (1997). “Philadelphia chromosome-positive acute lymphoblastic leukemia in children: durable responses to chemotherapy associated with low initial white blood cell counts.” Leukemia 11(9): 1493–6. Rosnet, O., H. J. Buhring, et€al. (1996). “Human FLT3/FLK2 receptor tyrosine kinase is expressed at the surface of normal and malignant hematopoietic cells.” Leukemia 10(2): 238–48. Rowland, A. J., G. A. Pietersz, et€al. (1993). “Preclinical investigation of the antitumour effects of anti-CD19-idarubicin immunoconjugates.” Cancer Immunol Immunother 37(3): 195–202. Schnittger, S., C. Schoch, et€al. (2002). “Analysis of FLT3 length mutations in 1003 patients with acute myeloid leukemia: correlation to cytogenetics, FAB subtype, and prognosis in the AMLCG study and usefulness as a marker for the detection of minimal residual disease.” Blood 100(1): 59–66. Schrappe, M., M. Arico, et€al. (1998). “Philadelphia chromosome-positive (Ph+) childhood acute lymphoblastic leukemia: good initial steroid response allows early prediction of a favorable treatment outcome.” Blood 92(8): 2730–41. Schroeter, E. H., J. A. Kisslinger, et€al. (1998). “Notch-1 signalling requires ligand-induced proteolytic release of intracellular domain.” Nature 393(6683): 382–6. Schultz, K. R., D. J. Pullen, et€al. (2007). “Risk- and response-based classification of childhood B-precursor acute lymphoblastic leukemia: a combined analysis of prognostic markers from the Pediatric Oncology Group (POG) and Children’s Cancer Group (CCG).” Blood 109(3): 926–35. Secker-Walker, L. M., H. G. Prentice, et€al. (1997). “Cytogenetics adds independent prognostic information in adults with acute lymphoblastic leukaemia on MRC trial UKALL XA. MRC Adult Leukaemia Working Party.” Br J Haematol 96(3): 601–10. Shah, N. P., C. Tran, et€al. (2004). “Overriding imatinib resistance with a novel ABL kinase inhibitor.” Science 305(5682): 399–401. Sievers, E. L., R. A. Larson, et€ al. (2001). “Efficacy and safety of gemtuzumab ozogamicin in patients with CD33-positive acute myeloid leukemia in first relapse.” J Clin Oncol 19(13): 3244–54. Silverman, L. B. (2007). “Acute lymphoblastic leukemia in infancy.” Pediatr Blood Cancer 49(7 Suppl): 1070–3. Smith, B. D., M. Levis, et€al. (2004). “Single-agent CEP-701, a novel FLT3 inhibitor, shows biologic and clinical activity in patients with relapsed or refractory acute myeloid leukemia.” Blood 103(10): 3669–76. Stam, R. W., M. L. den Boer, et€al. (2007a). “D-HPLC analysis of the entire FLT3 gene in MLL rearranged and hyperdiploid acute lymphoblastic leukemia.” Haematologica 92(11): 1565–8. Stam, R. W., M. L. den Boer, et€ al. (2005). “Targeting FLT3 in primary MLL-gene-rearranged infant acute lymphoblastic leukemia.” Blood 106(7): 2484–90. Stam, R. W., P. Schneider, et€al. (2007b). “Prognostic significance of high-level FLT3 expression in MLL-rearranged infant acute lymphoblastic leukemia.” Blood 110(7): 2774–5.
18
W.L. Carroll and R. Pieters
Stirewalt, D. L. and J. P. Radich (2003). “The role of FLT3 in haematopoietic malignancies.” Nat Rev Cancer 3(9): 650–65. Stone, R. M., D. J. DeAngelo, et€al. (2005). “Patients with acute myeloid leukemia and an activating mutation in FLT3 respond to a small-molecule FLT3 tyrosine kinase inhibitor, PKC412.” Blood 105(1): 54–60. Taketani, T., T. Taki, et€ al. (2004). “FLT3 mutations in the activation loop of tyrosine kinase domain are frequently found in infant ALL with MLL rearrangements and pediatric ALL with hyperdiploidy.” Blood 103(3): 1085–8. Talpaz, M., N. P. Shah, et€al. (2006). “Dasatinib in imatinib-resistant Philadelphia chromosomepositive leukemias.” N Engl J Med 354(24): 2531–41. Thiede, C., C. Steudel, et€al. (2002). “Analysis of FLT3-activating mutations in 979 patients with acute myelogenous leukemia: association with FAB subtypes and identification of subgroups with poor prognosis.” Blood 99(12): 4326–35. Thomas, D. A., S. Faderl, et€al. (2004). “Treatment of Philadelphia chromosome-positive acute lymphocytic leukemia with hyper-CVAD and imatinib mesylate.” Blood 103(12): 4396–407. Thomas, D. A., S. Faderl, et€al. (2006). “Chemoimmunotherapy with hyper-CVAD plus rituximab for the treatment of adult Burkitt and Burkitt-type lymphoma or acute lymphoblastic leukemia.” Cancer 106(7): 1569–80. Thomas, D. A., S. O’Brien, et€al. (2008). “Prognostic significance of CD20 expression in adults with de€novo precursor B-lineage acute lymphoblastic leukemia.” Blood 113: 6330–37. Thompson, B. J., S. Buonamici, et€ al. (2007). “The SCFFBW7 ubiquitin ligase complex as a tumor suppressor in T cell leukemia.” J Exp Med 204(8): 1825–35. Tibes, R., M. J. Keating, et€al. (2006). “Activity of alemtuzumab in patients with CD52-positive acute leukemia.” Cancer 106(12): 2645–51. Wang, J. C., T. Lapidot, et€al. (1998). “High level engraftment of NOD/SCID mice by primitive normal and leukemic hematopoietic cells from patients with chronic myeloid leukemia in chronic phase.” Blood 91(7): 2406–14. Weisberg, E., C. Boulton, et€al. (2002). “Inhibition of mutant FLT3 receptors in leukemia cells by the small molecule tyrosine kinase inhibitor PKC412.” Cancer Cell 1(5): 433–43. Weisberg, E., P. W. Manley, et€al. (2005). “Characterization of AMN107, a selective inhibitor of native and mutant Bcr-Abl.” Cancer Cell 7(2): 129–41. Weng, A. P., A. A. Ferrando, et€al. (2004). “Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia.” Science 306(5694): 269–71. Xia, M. Q., G. Hale, et€al. (1993). “Structure of the CAMPATH-1 antigen, a glycosylphosphatidylinositol-anchored glycoprotein which is an exceptionally good target for complement lysis.” Biochem J 293(Pt 3): 633–40. Yeoh, E. J., M. E. Ross, et€al. (2002). “Classification, subtype discovery, and prediction of outcome in pediatric acute lymphoblastic leukemia by gene expression profiling.” Cancer Cell 1(2): 133–43. Zwaan, C. M., V. H. J. van der Velden et€al. (2008). “Dasatinib in children and adolescents with relapsed or refractory leukemia: interim results of the CA180-018 phase I study from the ITCC consortium.” Blood 112: 3241. Zwaan, C. M., D. Reinhardt, et€al. (2003). “Gemtuzumab ozogamicin in pediatric CD33-positive acute lymphoblastic leukemia: first clinical experiences and relation with cellular sensitivity to single agent calicheamicin.” Leukemia 17(2): 468–70.
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia Alejandro Gutierrez and A. Thomas Look
Introduction T cell precursors arise from hematopoietic progenitors in the bone marrow and migrate to the thymus, where they undergo a series of proliferation and differentiation steps that include the somatic recombination of T cell receptor (TCR) gene loci (Market and Papavasiliou 2003). TCR gene rearrangements are followed by positive and negative selection steps that allow the survival of T cells only if their TCR functions appropriately within the context of an individual’s immune microenvironment. This highly regulated developmental process results in the generation of a population of T cells with a wide range of somatically acquired T cell receptor variation, which forms the foundation of a fully competent adaptive immune system that can respond to a countless variety of foreign antigens. Genetic alterations involving oncogenes or tumor suppressors can result in the aberrant proliferation, differentiation arrest, and clonal expansion of T cell precursors that is characteristic of T cell acute lymphoblastic leukemia (T-ALL). T cell lymphoblastic malignancies can be characterized based on the pattern of surface antigen expression in T cell lymphoblasts in relation to the pattern of surface antigen expression that occurs in normal thymocyte development (Reinherz and Schlossman 1980; Rothenberg and Taghon 2005; Staal et al. 2001). The earliest T cell precursors migrate from the bone marrow to the thymus, where they proceed through a series of so-called double-negative (CD4-, CD8-) differentiation stages, during which rearrangements of the T-cell receptor d, g and b genes occur. This is
A.T. Lookâ•›(*) Department of Pediatric Oncology, Dana-Farber Cancer Institute and Children’s Hospital Boston, 44 Binney Street, Boston, MA 02115, USA and Harvard Medical School, Boston, MA, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_2, © Springer Science+Business Media, LLC 2010
19
20
A. Gutierrez and A.T. Look
followed by the intermediate single-positive (CD4+, CD8-, surface CD3-), and the subsequent double-positive (CD4+, CD8+) stage, during which TCR a rearrangements occur. Intrathymic T cell development ends with the generation of singlepositive CD4+ or CD8+ T cells that migrate outside the thymus, although additional differentiation steps occur in T cells upon antigen recognition. T-ALL accounts for 10 to 15% of acute lymphoblastic leukemias in children, and it is most common in older children and adolescents. This disease typically presents with distinctive clinical features, including very high numbers of lymphoblasts in the peripheral blood, involvement of the central nervous system, and an anterior mediastinal mass. T cell lymphoblastic lymphoma is a closely related malignancy that typically presents with a thymic anterior mediastinal mass and is differentiated from T-ALL based on the degree of bone marrow involvement, with greater than 25% lymphoblasts in the bone marrow defining T-ALL. These diseases exhibit a high degree of similarity in morphology, immunophenotype, and response to therapy, and are generally thought to represent different clinical presentations of the same disease (Cairo et€al. 2005). Most of the insights into the molecular pathogenesis of T-cell lymphoblastic malignancies have come through the study of T-ALL rather than T cell lymphoblastic lymphoma, probably due to the greater ease of obtaining primary tumor specimens. This chapter is focused on the key aspects of the molecular pathogenesis of T-ALL that are most relevant to the development of molecular targeted therapies for this disease.
Oncogenic Transcription Factors Many of the early insight into the pathobiology underlying T-ALL came through the study of genes affected by recurrent chromosomal translocations. In contrast to B-precursor ALL, where recurrent translocations often lead to the expression of a fusion gene product, T-ALL translocations typically lead to the aberrant overexpression of structurally intact transcription factors, due to the placement of these genes under the control of regulatory elements that are highly active in T cell precursors, such as the gene regulatory elements that normally drive the expression of€T-cell receptor (TCR) genes. Most often, the transcription factors involved play prominent roles in normal hematopoietic development. Oncogenic transcription factors dysregulated by this mechanism in T-ALL include basic region helix-loophelix (bHLH) genes such as TAL1, TAL2, LYL1, and MYC, cysteine-rich genes such as LMO1 and LMO2, and homeodomain genes such as HOX11 and HOX11L2. Furthermore, although chromosomal translocations involving these genes are present in only 25% of cases of T-ALL, gene expression profiling has identified overexpression of these genes in 80% of cases (Ferrando et€al. 2004; Ferrando and Look 2003), strongly suggesting that the dysregulated expression of these proto-oncogenic transcription factors plays a fundamental role in the molecular pathogenesis of most cases of T-ALL.
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia
21
Although of great biologic interest due to their central role in the pathobiology of T-ALL, dysregulated transcription factor activity, which relies on intracellular protein–protein and protein–DNA interactions, has been difficult to target with currently available pharmacologic strategies. Small molecules with favorable pharmacologic properties can readily be designed to target intracellular proteins whose activity relies on small hydrophobic pockets, such as enzymes. However, it is estimated that only about 12% of the proteins encoded by the human genome are “druggable” by small molecules with current technology, due to structural and mechanistic considerations (Hopkins and Groom 2002; Hopkins and Groom 2003; Russ and Lampel 2005); most transcription factors are currently considered “undruggable.” On the other hand, antibodies can be used as therapeutic agents to inhibit protein–protein interactions and thus have a much broader range of potential targets; however, antibodies generally cannot cross the cell membrane, and are thus poorly suited as targeted modulators of intracellular proteins (Verdine and Walensky 2007). However, recent advances in peptide medicinal chemistry include the development of chemically stabilized alpha-helical peptides that can successfully target intracellular protein–protein interactions and that are taken up in vivo by intact cells (Gavathiotis et€al. 2008; Walensky et€ al. 2004). This strategy may allow the development of specific therapies targeting dysregulated oncogenic transcription factor activity in the human leukemias.
NOTCH1 The NOTCH1 Pathway Notch, which was initially identified by the characterization of a Drosophila melanogaster mutant strain with notched wings (Morgan 1917), plays central roles in the development of multicellular organisms, where it acts by regulating cell fate, proliferation, survival, migration, and differentiation (Lai 2004). The first human ortholog of Notch, NOTCH1, was initially identified as a partner gene in a t(7;9) translocation found in very rare cases of T-ALL, in which the expression of a truncated NOTCH1 protein is driven by T cell receptor b gene regulatory elements (Ellisen et€ al. 1991). NOTCH1 plays central roles in T cell development, where it directs lymphoid progenitors toward a T cell fate at multiple stages of lymphoid development (Grabher et€ al. 2006; Maillard et€ al. 2006; Radtke et€ al. 2004; Rothenberg and Taghon 2005). Additionally, constitutive NOTCH1 signaling transforms T cell precursors in experimental systems (ZweidlerMcKay and Pear 2004). Despite the very low frequency of translocations involving NOTCH1, activating mutations in the heterodimerization or PEST domains of the full-length NOTCH receptor have been found in greater than 50% of cases of T-ALL (Weng et€al. 2004).
22
A. Gutierrez and A.T. Look
NOTCH1 is a cell surface receptor with an unusual mechanism of action (reviewed in (Bray 2006; Grabher et€al. 2006)). After its synthesis, the NOTCH1 protein undergoes cleavage by a furin-like protease into extracellular and transmembrane subunits that remain associated via noncovalent interactions at the cell membrane. The binding of ligand to the extracellular subunit of NOTCH1 triggers two additional cleavage events in the transmembrane subunit, the first catalyzed by ADAM-family metalloproteases and the second by the g-secretase enzyme complex, which consists of presenilin, nicastrin, PEN2, and APH1. This final g-secretase-mediated cleavage allows the release of the intracellular domain of NOTCH1, known as ICN, into the cytoplasm. The ICN then translocates to the nucleus, where it binds to the CSL/RBPJ transcription factor, leading to the displacement of transcriptional corepressors and the recruitment of coactivators, thus resulting in the transcriptional upregulation of target genes (Fig.€1). NOTCH1 activating mutations in T-ALL generally occur in two distinct regions of the protein, and result in the accumulation of nuclear ICN protein via distinct mechanisms. Missense mutations in the heterodimerization domain lead to constitutive proteolytic activation of ICN, whereas truncating mutations in the C-terminal PEST domain impair the degradation of nuclear ICN. Interestingly, mutations in
Fig.€1╅ Activation of NOTCH signaling via proteolytic cleavage and nuclear translocation of the intracellular NOTCH domain (ICN). Interaction with delta serrate ligand (DSL) stimulates proteolytic cleavage of NOTCH by metalloproteases and g-segretase. This leads to the release of the intracellular ICN domain, which translocates to the nucleus, displaces corepressors and recruits coactivators (MAM1), thereby converting CSL from a repressor to an activator of gene expression. Reprinted with permission from Armstrong and Look (2005)
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia
23
both regions of the protein frequently occur in cis in the same NOTCH1 allele, resulting in the synergistic activation of NOTCH1 signaling (Weng et€ al. 2004). Prominent transcriptional targets of the NOTCH1 ICN include MYC, which has been shown to be critical for the growth-promoting effects of NOTCH1 in T-ALL cell lines (Palomero et€ al. 2006; Sharma et€ al. 2006; Weng et€ al. 2006), while another NOTCH1 target, HES1, is a transcriptional repressor that is a critical mediator of the ICN-induced transcriptional repression of the PTEN tumor suppressor (Palomero et€al. 2007), as discussed in more detail below.
Strategies for NOTCH1 Inhibition A number of strategies have been devised for the inhibition of NOTCH. Wild-type NOTCH1 signaling can be targeted at the receptor-ligand interaction, but such strategies are unlikely to be effective against T-ALL cells expressing mutant NOTCH1 because the mutant receptor undergoes ligand-independent receptor activation. Other strategies, including the development of monoclonal antibodies that lock NOTCH1 in an inactive conformation, are in preclinical development (reviewed in (Rizzo et€al. 2008)). The identification of a requirement for enzymatic activity of the g-secretase complex for the proteolytic activation of ICN, both upon ligand binding and in the presence of activating NOTCH1 mutations, generated considerable excitement for the application of g-secretase inhibitors (GSI) in T-ALL. Small molecule inhibitors of g-secretase were developed prior to the discovery of NOTCH1 activation in T-ALL because the g-secretase complex also plays a central role in the proteolytic generation of the insoluble amyloid-b peptide that accumulates in the brain in Alzheimer’s disease (reviewed in (Gotz and Ittner 2008). Interestingly, toxicity to normal T cells was found to be a toxicity of GSI during their preclinical development for the treatment of Alzheimer’s disease (Wong et€al. 2004). Clinical trials of GSIs have been undertaken in T-ALL; however, the clinical utility of these inhibitors has been hindered by the development of serious gastrointestinal toxicity (DeAngelo et€ al. 2006; Rizzo et€ al. 2008). GSI therapy induces apoptosis of intestinal epithelial cells and goblet cell metaplasia (Milano et€al. 2004), an on-target toxicity resulting from altered intestinal cell differentiation toward a goblet cell fate in the absence of NOTCH signaling (van Es et€al. 2005). Efforts to overcome this toxicity through alterations in drug schedules and doses are the focus of current investigations. It has been reported that patients with breast cancer treated with a combination of tamoxifen and GSI had greatly alleviated intestinal toxicity (Rizzo et€al. 2008). Furthermore, another report has demonstrated that the treatment with a combination of dexamethasone and GSI dramatically ameliorates the GSI-mediated intestinal toxicity in an animal model, while GSI therapy could simultaneously reverse NOTCH-mediated glucocorticoid resistance in human T-ALL cell lines (Real et€ al. 2009). This important finding suggests that combination therapy with dexamethasone and GSIs could
24
A. Gutierrez and A.T. Look
lead to synergistic antileukemic activity while simultaneously abrogating the dose-limiting gastrointestinal toxicity of GSIs, a possibility that should be tested in T-ALL clinical trials in the near future.
Resistance to NOTCH1 Inhibitors and Therapeutic Strategies to Overcome Resistance Despite their initial promise, g-secretase inhibitors are not effective against all T-ALL cell lines harboring activating NOTCH1 mutations. The mechanism mediating the resistance of T-ALL cells to g-secretase inhibition has been the target of intense investigation, and at least two distinct mechanisms of resistance have been identified. The first of these involves the inactivation of the FBW7 tumor suppressor, which appears to mediate resistance to g-secretase inhibition by impairing the downregulation of NOTCH1 signaling. FBW7 is an E3 ubiquitin ligase that targets both the ICN and MYC for proteasomal degradation (O’Neil et€al. 2007; Thompson et€al. 2007). Despite the effective inhibition of NOTCH1 activation at the cell surface by g-secretase inhibitor therapy, cells harboring inactivation of FBW7 apparently fail to effectively downregulate NOTCH1 signaling because the proteasomal degradation of ICN, and of its prominent target MYC, is impaired by the inactivation of the FBW7 E3 ubiquitin ligase. The inactivation of PTEN, an important tumor suppressor that negatively regulates PI3K-AKT signaling, has been identified as a second prominent mechanism mediating resistance to g-secretase inhibitor therapy (Palomero et€al. 2007). In contrast to FBW7 mutations, the inactivation of PTEN does not appear to exert its effect by promoting ICN transcriptional activity. Under physiologic conditions, active NOTCH1 signaling leads to the transcriptional upregulation of HES1, which then acts as a transcriptional repressor of PTEN. Thus, in the setting of active NOTCH1 signaling, active PI3K-AKT signaling is promoted due to the transcriptional repression of PTEN (Fig. 2a). Upon inhibition of g-secretase activity, NOTCH1 signaling is inhibited, the transcriptional repression of PTEN is released, and the resultant upregulation in PTEN expression leads to the inhibition of oncogenic PI3K-AKT signaling. Thus, g-secretase inhibitor therapy in a cell with intact PTEN results in the inhibition of both NOTCH1 and PI3K-AKT signaling (Fig.€2b). However, deletions and inactivating mutations of PTEN have been identified in 36% of primary T-ALL samples, as discussed in more detail below. PTEN-negative T-ALL cells with NOTCH1 activating mutations maintain active signaling through both NOTCH1 and PI3K-AKT pathways (Fig.€ 2c). Upon effective g-secretase inhibitor therapy, NOTCH1 signaling is inhibited in PTEN-negative cells. However, PTEN expression cannot be upregulated in these cells despite the downregulation of HES1, and PI3K-AKT signaling remains active in PTEN-negative cells. The failure to downregulate PI3K-AKT signaling thus leads to resistance to GSI treatment in PTEN-negative cells (Fig.€ 2d). Interestingly, Palomero et€ al. have shown that, although PTEN loss induces resistance to NOTCH1 inhibition, PTEN-negative
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia
a
PTEN-positive T-ALL, no NOTCH1 inhibition
b
25
PTEN-positive T-ALL, NOTCH1 inhibited by GSI
N N
TC O
TC O
H
H
1
GSI
PI3K
1
PI3K
PTEN AKT
AKT ICN
PTEN
PTEN
SURVIVAL PROLIFERATION
c
PTEN-negative T-ALL, no NOTCH1 inhibition
PROLIFERATION ARREST APOPTOSIS
d
PTEN-negative T-ALL, NOTCH1 inhibited by GSI
1
H
1
H
TC
O
TC
N
O
N
PI3K
GSI
AKT
AKT ICN
PTEN
SURVIVAL PROLIFERATION
PI3K
PTEN
SURVIVAL PROLIFERATION
Fig.€ 2╅ PTEN inactivation leads to resistance to NOTCH1 inhibition but dependence on AKT signaling. (a) T-ALL lymphoblasts with activating mutations of NOTCH1 constitutively generate intracellular NOTCH1 (ICN), whose transcriptional targets include MYC and HES1. MYC appears to be a transcriptional activator of PTEN, but HES1-mediated repression predominates under ICN signaling conditions. Low expression of PTEN leads to incomplete inhibition of the PI3K-AKT pathway. (b) Inhibition of proteolytic release of ICN from the NOTCH1 receptor by g-secretase inhibitors (GSI) blocks ICN-mediated proliferative and survival signals, leading to cell cycle arrest and apoptosis. Additionally, the HES1-mediated repression of PTEN expression is relieved, and PTEN can therefore inhibit prosurvival signaling mediated by the PI3K-AKT pathway. Thus, PTEN-positive T-ALL cells with activating NOTCH1 mutations depend on NOTCH1 activity for survival and proliferation. (c and d). In the absence of PTEN, uninhibited AKT activation leads to aberrant prosurvival and proliferative signaling independent of NOTCH1 pathway activity, thus leading to resistance to NOTCH1 inhibition. PTEN-null T-ALL cells that are resistant to NOTCH1 inhibition appear to be dependent on PI3K-AKT pathway signaling. Reprinted with permission from Gutierrez and Look (2007)
26
A. Gutierrez and A.T. Look
NOTCH1-mutant T-ALL cells are instead dependent on AKT signaling (Palomero et€al. 2007). A number of inhibitors of the PI3K-AKT pathway are currently undergoing clinical development, and it seems likely that these will be particularly effective in T-ALL when used in combination with NOTCH1 inhibitors.
PTEN, PI3K-AKT, and mTOR The role of the PTEN tumor suppressor in T-ALL has been suggested by murine studies demonstrating that T-ALL develops in mice after the conditional deletion of PTEN in hematopoietic stem cells (Guo et€al. 2008; Yilmaz et€al. 2006). The tumor suppressor activity of PTEN is mediated in large part due to its role as a negative regulator of PI3K-AKT signaling (Chow and Baker 2006), an oncogenic signal transduction pathway activated by growth factor signaling and by a variety of oncogenic mutations in human cancer (Brugge et€ al. 2007). Deletions of PTEN have recently been described in approximately 8% of primary T-ALL patient samples (Maser et€ al. 2007; Mullighan et€ al. 2007). Additionally, we have recently completed an analysis of PTEN as well as PI3K and AKT genes in primary T-ALL samples, and have identified the presence of genetic lesions in the PTEN-PI3KAKT pathway in 48% of diagnostic specimens from a series of 44 children with T-ALL (Gutierrez et al. 2009). Activation of signaling through the PI3K-AKT pathway, as a result of PTEN loss or the introduction of a constitutively active AKT transgene, has been shown to mediate resistance to NOTCH1 inhibition, but this appears to induce sensitivity to PI3K-AKT pathway inhibitors, as discussed in the section “Resistance to NOTCH1 Inhibitors and Therapeutic Strategies to Overcome Resistance” (Palomero et€al. 2007). Additionally, evidence suggests that the mammalian target of rapamycin (mTOR), which is activated by the PI3K-AKT pathway, mediates glucocorticoid resistance in at least some ALL cells, and the inhibition of mTOR effectively reversed glucocorticoid resistance in T-ALL and B-precursor ALL cell lines (Wei et€ al. 2006). Furthermore, combination therapy with a glucocorticoid and a g-secretase inhibitor has been shown to both reverse glucocorticoid resistance and to mitigate the serious gastrointestinal toxicity of g-secretase inhibitors (Real et€al. 2009), as discussed in the section “Strategies for NOTCH1 Inhibition.” Taken together, these findings suggest that combination therapy with an inhibitor of PI3KAKT-mTOR signaling, a g-secretase inhibitor, and a glucocorticoid may represent a particularly effective therapeutic combination in T-ALL, a possibility that will likely be tested by future clinical trials.
Tyrosine Kinase Genes The ABL tyrosine kinase normally plays physiologic roles in the regulation of proliferation, cell adhesion and migration, and apoptosis (Sirvent et€al. 2008). Translocations resulting in the expression of a BCR-ABL fusion oncoprotein represent the characteristic
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia
27
genetic lesion in chronic myeloid leukemia (CML) and also occur in B-precursor ALL. Compared to the wild-type ABL tyrosine kinase, BCR-ABL exhibits aberrant subcellular localization and constitutive kinase activity. The ability of imatinib, a small molecule inhibitor of BCR-ABL kinase activity, to provide long-term disease control in CML has revolutionized therapy for this disease, and imatinib also has activity against BCR-ABL-positive acute lymphoblastic leukemia. Despite the rarity of BCR-ABL translocations in T-ALL, amplified episomes containing NUP214-ABL fusion genes have been described in approximately 6% of adults and children with T-ALL (Graux et€al. 2004). The ABL breakpoint in these cases occurs in intron 1, the same location as in BCR-ABL translocations, and these fusions impart constitutive ABL tyrosine kinase activity to the fusion oncoprotein. Therapy with tyrosine kinase inhibitors designed to target BCR-ABL, including imatinib, dasatinib, and nilotinib, inhibits proliferation and induces apoptosis in T-ALL cell lines harboring NUP214-ABL fusions (Quintas-Cardama et€al. 2008). Furthermore, ABL amplifications and EML1-ABL fusions have also been described in rare cases of T-ALL (Bernasconi et€al. 2005; De Keersmaecker et€al. 2005). Taken together, these findings suggest that ABL kinase inhibitors may have therapeutic utility in a subset of cases of T-ALL, and clinical trials should be initiated to test this possibility. Mutational activation of JAK1 has recently been reported in 18% of T-ALL cases, and these mutations were associated with poor response to therapy (Flex et€al. 2008). The JAK1 gene encodes a nonreceptor tyrosine kinase that associates with cytokine receptors and mediates the activation of STAT proteins upon cytokine receptor activation. JAK1 has been shown to play critical roles in lymphocyte development, and JAK1 knockout mice show a severely hypocellular thymus. JAK1 mutations identified in T-ALL lead to the ligand-independent activation of STAT signaling. A number of inhibitors of JAK/STAT signaling are currently under clinical development, and the determination of their clinical activity against JAK1mutated T-ALL will be of great interest.
RAS The RAS proto-oncogenes encode proteins that play central roles in mediating growth factor signaling, and missense mutations at codons 12, 13, or 61 are frequent events in human cancer. Mutant RAS proteins are potently oncogenic and lead to the induction of signaling though the MAP kinase, PI3K-AKT, and RAL-GDS pathways (Schubbert et€ al. 2007). Mutations activating NRAS or KRAS have been identified in approximately 15% of T-ALL and B-precursor ALL cases (Liang et€al. 2006; Perentesis et€ al. 2004). Due to their role in the pathogenesis of numerous cancers, a number of targeted RAS inhibitors are under development, many of which target the post-translational modifications required for the generation of functional RAS proteins, including farnesylation, palmitoylation and geranylgeranylation. Furthermore, recent studies in KRAS-dependent cell lines derived from diverse tumor types revealed inhibition of the STK33 kinase as effective therapy against KRAS-dependent tumors, indicating the need for preclinical studies of STK33
28
A. Gutierrez and A.T. Look
inhibitors in RAS-dependent T-ALL (Scholl et al. 2009). Targeted inhibitors of RAS, or of signal transduction pathways activated by RAS, including the MAP kinase and the PI3K-AKT pathways, may prove to have a role in the therapy of RAS-mutated T-ALL.
Conclusions Recent advances in our understanding of the molecular pathogenesis of T-ALL have uncovered a number of therapeutic targets in this disease, including proteins in the NOTCH and PI3K-AKT pathways. Despite initial challenges, and the early difficulties with gastrointestinal toxicity with the use of NOTCH inhibitors, remarkable recent advances have suggested rational combination therapies that may lead to both synergistic antileukemic activity and reduced toxicity. Our understanding of T-ALL oncogenic pathways almost certainly remains incomplete, and new advances should lead to the successful clinical application of molecularly targeted agents in T-ALL.
References Armstrong SA, Look AT (2005) Molecular genetics of acute lymphoblastic leukemia. Journal of Clinical Oncology 23:6306–6315. Bernasconi P et€ al. (2005) ABL1 amplification in T-cell acute lymphoblastic leukemia. Cancer Genet Cytogenet 162:146–50. Bray SJ (2006) Notch signalling: a simple pathway becomes complex. Nat Rev Mol Cell Biol 7:678–89. Brugge J et€al. (2007) A new mutational AKTivation in the PI3K pathway. Cancer Cell 12:104–7. Cairo MS et€al. (2005) Childhood and adolescent non-Hodgkin lymphoma: new insights in biology and critical challenges for the future. Pediatr Blood Cancer 45:753–69. Chow LM, Baker SJ (2006) PTEN function in normal and neoplastic growth. Cancer Lett 241:184–96. De Keersmaecker K et€al. (2005) Fusion of EML1 to ABL1 in T-cell acute lymphoblastic leukemia with cryptic t(9;14)(q34;q32). Blood 105:4849–52. DeAngelo DJ et€al. (2006). A phase I clinical trial of the notch inhibitor MK-0752 in patients with T-cell acute lymphoblastic leukemia/lymphoma (T-ALL) and other leukemias. Abstract, American Society of Clinical Oncology 2006 Annual Meeting. Ellisen LW et€al. (1991) TAN-1, the human homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell 66:649–61. Ferrando AA et€al. (2004) Biallelic transcriptional activation of oncogenic transcription factors in T-cell acute lymphoblastic leukemia. Blood 103:1909–11. Ferrando AA, Look AT (2003) Gene expression profiling in T-cell acute lymphoblastic leukemia. Semin Hematol 40:274–80. Flex E et€al. (2008) Somatically acquired JAK1 mutations in adult acute lymphoblastic leukemia. J Exp Med 205:751–8. Gavathiotis E et€ al. (2008) BAX activation is initiated at a novel interaction site. Nature 455:1076–81.
Molecular Targeted Therapies in T-Cell Acute Lymphoblastic Leukemia
29
Gotz J, Ittner LM (2008) Animal models of Alzheimer’s disease and frontotemporal dementia. Nat Rev Neurosci 9:532–44. Grabher C et€al. (2006) Notch 1 activation in the molecular pathogenesis of T-cell acute lymphoblastic leukaemia. Nat Rev Cancer 6:347–59. Graux C et€al. (2004) Fusion of NUP214 to ABL1 on amplified episomes in T-cell acute lymphoblastic leukemia. Nat Genet 36:1084–9. Guo W et€al. (2008) Multi-genetic events collaboratively contribute to Pten-null leukaemia stem-cell formation. Nature 453:529–33. Gutierrez A, Look AT (2007) NOTCH and PI3K-AKT pathways intertwined. Cancer Cell 12:411–413. Gutierrez A et al. (2009) High frequency of PTEN, PI3K and AKT abnormalities in T cell acute lymphoblastic leukemia. Blood 114:647–50. Hopkins AL, Groom CR (2002) The druggable genome. Nat Rev Drug Discov 1:727–30. Hopkins AL, Groom CR (2003) Target analysis: a priori assessment of druggability. Ernst Schering Res Found Workshop:11–7. Lai EC (2004) Notch signaling: control of cell communication and cell fate. Development 131:965–73. Liang DC et€al. (2006) K-Ras mutations and N-Ras mutations in childhood acute leukemias with or without mixed-lineage leukemia gene rearrangements. Cancer 106:950–6. Maillard I et€al. (2006) Notch-dependent T-lineage commitment occurs at extrathymic sites following bone marrow transplantation. Blood 107:3511–9. Market E, Papavasiliou FN (2003) V(D)J recombination and the evolution of the adaptive immune system. PLoS Biol 1:E16. Maser RS et€al. (2007) Chromosomally unstable mouse tumours have genomic alterations similar to diverse human cancers. Nature 447:966–71. Milano J et€ al. (2004) Modulation of notch processing by gamma-secretase inhibitors causes intestinal goblet cell metaplasia and induction of genes known to specify gut secretory lineage differentiation. Toxicol Sci 82:341–58. Morgan TH (1917) The Theory of the Gene. The American Naturalist 51:513. Mullighan CG et€al. (2007) Genome-wide analysis of genetic alterations in acute lymphoblastic leukaemia. Nature 446:758–64. O’Neil J et€al. (2007) FBW7 mutations in leukemic cells mediate NOTCH pathway activation and resistance to gamma-secretase inhibitors. J Exp Med 204:1813–24. Palomero T et€al. (2006) NOTCH1 directly regulates c-MYC and activates a feed-forward-loop transcriptional network promoting leukemic cell growth. Proc Natl Acad Sci U S A 103:18261–6. Palomero T et€al. (2007) Mutational loss of PTEN induces resistance to NOTCH1 inhibition in T-cell leukemia. Nat Med 13:1203–10. Perentesis JP et€al. (2004) RAS oncogene mutations and outcome of therapy for childhood acute lymphoblastic leukemia. Leukemia 18:685–92. Quintas-Cardama A et€al. (2008) Activity of tyrosine kinase inhibitors against human NUP214ABL1-positive T cell malignancies. Leukemia 22:1117–24. Radtke F et€al. (2004) Notch regulation of lymphocyte development and function. Nat Immunol 5:247–53. Real PJ et€al. (2009) Gamma-secretase inhibitors reverse glucocorticoid resistance in T cell acute lymphoblastic leukemia. Nat Med 15:50–8. Reinherz EL, Schlossman SF (1980) Current concepts in immunology: Regulation of the immune response--inducer and suppressor T-lymphocyte subsets in human beings. N Engl J Med 303:370–3. Rizzo P et€al. (2008) Rational targeting of Notch signaling in cancer. Oncogene 27:5124–31. Rothenberg EV, Taghon T (2005) Molecular genetics of T cell development. Annu Rev Immunol 23:601–49. Russ AP, Lampel S (2005) The druggable genome: an update. Drug Discov Today 10:1607–10. Scholl C et al. (2009) Synthetic lethal interaction between oncogenic KRAS dependency and STK33 suppression in human cancer cells. Cell 137:821–34. Schubbert S et€al. (2007) Hyperactive Ras in developmental disorders and cancer. Nat Rev Cancer 7:295–308.
30
A. Gutierrez and A.T. Look
Sharma VM et€al. (2006) Notch1 contributes to mouse T-cell leukemia by directly inducing the expression of c-myc. Mol Cell Biol 26:8022–31. Sirvent A et€al. (2008) Cytoplasmic signalling by the c-Abl tyrosine kinase in normal and cancer cells. Biol Cell 100:617–31. Staal FJ et al. (2001) Transcriptional control of T-lymphocyte differentiation. Stem Cells 19:165–79. Thompson BJ et€al. (2007) The SCFFBW7 ubiquitin ligase complex as a tumor suppressor in T cell leukemia. J Exp Med 204:1825–35. van Es JH et€ al. (2005) Notch/gamma-secretase inhibition turns proliferative cells in intestinal crypts and adenomas into goblet cells. Nature 435:959–63. Verdine GL, Walensky LD (2007) The challenge of drugging undruggable targets in cancer: lessons learned from targeting BCL-2 family members. Clin Cancer Res 13:7264–70. Walensky LD et€al. (2004) Activation of apoptosis in€vivo by a hydrocarbon-stapled BH3 helix. Science 305:1466–70. Wei G et€al. (2006) Gene expression-based chemical genomics identifies rapamycin as a modulator of MCL1 and glucocorticoid resistance. Cancer Cell 10:331–42. Weng AP et€al. (2004) Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 306:269–71. Weng AP et€al. (2006) c-Myc is an important direct target of Notch1 in T-cell acute lymphoblastic leukemia/lymphoma. Genes Dev 20:2096–109. Wong GT et€al. (2004) Chronic treatment with the gamma-secretase inhibitor LY-411,575 inhibits beta-amyloid peptide production and alters lymphopoiesis and intestinal cell differentiation. J Biol Chem 279:12876–82. Yilmaz OH et€al. (2006) Pten dependence distinguishes haematopoietic stem cells from leukaemia-initiating cells. Nature 441:475–82. Zweidler-McKay PA, Pear WS (2004) Notch and T cell malignancy. Semin Cancer Biol 14:329–40.
Molecularly Targeted Therapy for Infant ALL Patrick A. Brown and Carolyn A. Felix
Rational combinations of molecularly targeted agents with synergistic conventional cytotoxic chemotherapeutic agents or, ultimately, with one another are urgently needed for infants with acute leukemia. Leukemia is the commonest malignancy during infancy, comprises 2.5 to 5% of ALL and 6 to 14% of AML in pediatrics overall, (Gurney et€ al. 1999; Smith et€ al. 1999a; Pui et€ al. 1995; SEER Cancer Statistics Review 1975–2006) and represents a special leukemia subtype characterized by MLL (Mixed Lineage Leukemia) gene translocations. MLL translocations with heterogeneous partner genes, of which there are >60, (Meyer et€al. 2009) are the primary molecular aberrations in infant ALL and infant AML alike; approximately 75 to 80% of infant ALL cases and myelomonocytic/monoblastic AML feature MLL translocations (Pui et€al. 1995; Rubnitz et€al. 1994; Robinson et al. 2009). Within infant ALL cases with MLL translocations, 70 and 13% involve the AF4 (ALL-1 fused gene from chromosome 4) or ENL (Eleven-nineteen leukemia) partner genes, respectively, (Pui et€al. 2003) whereas the partner genes in AML are more diverse. In one recent infant ALL treatment study, these more common partner genes were associated with only approximately 30% 5-year event free survival, with a poorer outcome associated with the CD10− immunophenotype and younger age at diagnosis (Hilden et€al. 2006). The treatment options for acute leukemia in infants have been limited to intensive chemotherapy and hematopoietic stem cell transplantation. A deeper biological understanding of MLL disease in infants that becomes translatable to better treatment options would be of utmost impact because the prognosis is dismal due to profound vulnerabilities to treatment-related toxicities unique to this population, frequent relapses, and inherent chemotherapy resistance as a consequence of the translocations. While infant ALL and AML are orphan diseases with annual incidences of 19 and 10 per million, respectively, MLL translocations occur in 15 to 20%
C.A. Felixâ•›(*) Division of Oncology, Department of Pediatrics, The Children’s Hospital of Philadelphia, University of Pennsylvania School of Medicine, Colket Translational Research Building, Room 4006, 3501 Civic Center Blvd, Philadelphia, PA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_3, © Springer Science+Business Media, LLC 2010
31
32
P.A. Brown and C.A. Felix
of pediatric AML (Balgobind et€al. 2009) and comprise 5 to 10% of acquired chromosomal rearrangements in pediatric ALL and adult ALL and AML (Pui et€al. 2003). Nearly all current-era anticancer regimens specify chemotherapy with one or more topoisomerase II poisons (i.e., agents that convert native topoisomerase II to a cellular toxin by altering the topoisomerase II cleavage–religation equilibrium, causing DNA breakage); examples are etoposide, anthracyclines, dactinomycin, and mitoxantrone (Felix et€al. 2006). Various secondary leukemias with balanced translocations are a potential untoward complication of these agents, the most common of which involve MLL (Smith et€al. 1999b). The complex circuitry of relevant target aberrations comprising resultant heterogeneous MLL fusion oncoproteins and proteins with which they are associated (Liedtke and Cleary 2009) has been gradually unraveling since the initial discovery of MLL as the gene disrupted by the translocations approximately 18€years ago (Ziemin-van der Poel et€al. 1991; Djabali et€al. 1992; Gu et€al. 1992; Tkachuk et€al. 1992). Such observations form the basis for the current development of targeted therapies for these highly aggressive and usually incurable diseases. Infant ALL and AML are embryonal diseases that originate in utero from DNA damage to MLL that results in the formation of the translocations (Ford et€al. 1993; Megonigal et€al. 1998). The native MLL protein is a complex transcriptional regulator that undergoes proteolytic cleavage into amino and carboxyl terminal fragments with transcriptional repression and activation properties, which reassemble in a multiprotein complex (Yokoyama et€al. 2002) that controls epigenetic modifications of nucleosomes and histones and maintains HOX expression (Hess 2004). Amino terminal structural domains include AT hook motifs, SNL motifs that direct subnuclear localization, and a cysteine-rich CXXC region that recognizes CpG dinucleotides similar to the CXXC region in DNA methyltransferase 1 (Hess 2004). The adjacent more central PHD mediates homodimerization and protein interactions (Hess 2004). The carboxyl terminal SET domain interacts with the SWI/SNF chromatin remodeling complex and has histone H3K4 methyltransferase activity (Hess 2004). MLL is important in cell cycle regulation; MLL proteolytic cleavage by Taspase1 (Takeda et€al. 2006) and degradation by specific E3 ligases are critical for proper cell cycle progression (Liu et€al. 2007). MLL translocations disrupt a breakpoint cluster region and fuse 5¢ MLL with the 3¢ portion of the partner gene, producing diverse fusion oncoproteins comprising the amino-terminus of MLL and the carboxyl-terminus of the partner protein, which cause leukemia in mice (Hess 2004; Meyer et€al. 2006). N-MLL-Partner-C fusion proteins are believed to mediate leukemogenesis by altering transcription, (Hess 2004) possibly through the gain of function mechanisms (e.g., binding to transcriptional coactivators, fusion protein dimerization, etc.) since HOX expression is increased even though the MLL histone H3K4 transactivation activity of the SET domain is missing (Hess 2004; Armstrong et€ al. 2002a; Ayton and Cleary 2003; Martin et€ al. 2003). However, in murine models, sustained HoxA gene expression is necessary but insufficient for altered differentiation and clonal expansion, (Ayton and Cleary 2003) and levels of overexpression of rate-limiting Meis1/ Pbx cofactors vary with the partner genes (Wong et€al. 2007).
Molecularly Targeted Therapy for Infant ALL
33
Until 2008, no molecularly targeted agents had been used in infant ALL. The pathways and target genes that are critically affected by MLL fusion oncoprotein transcriptional deregulation have been challenging to resolve and remain incompletely understood, and there has been considerable controversy surrounding the nature of the transformed MLL leukemia stem cell (Somervaille and Cleary 2006; Krivtsov et€al. 2006; Wang et€al. 2005). Therefore, in the current landscape of infant leukemia therapy, the molecularly targeted agents that have begun advancing through the drug development pipeline (FDA 2004) are predominantly at the preclinical in€vitro and in€vivo stages of development. The initial focus in bringing molecularly targeted agents to the clinic for this disease has been on compounds with demonstrable activity in other leukemias and cancers that might also silence cooperating events in MLL leukemogenesis, rather than on agents directly targeting the fusion oncoproteins and interacting factors. A major milestone was achieved when comprehensive preclinical analyses formed the basis for the recently opened COG AALL0631 clinical trial, which is testing the combination of chemotherapy and the FLT3 tyrosine kinase inhibitor lestaurtinib for MLL-rearranged infant ALL; this is the first trial to incorporate a signal transduction pathway targeted agent into frontline therapy for infant ALL. In this chapter, the limited targeted agents that have advanced to the clinic, of which there are now two, will be reviewed in detail and promising targets and molecularly targeted agents progressing through the drug development pipeline will also be addressed.
FLT3 Tyrosine Kinase Inhibition The receptor tyrosine kinase FLT3 (FMS-Like Tyrosine Kinase 3) has been strongly implicated in the pathogenesis of infant leukemia (Brown and Small 2004a; Levis and Small 2003). With the development of FLT3-targeted agents, the hypothesis that FLT3-targeted therapy can improve the outcome for infant leukemia is now being tested. FLT3 is aberrantly expressed in greater than 90% of AML and nearly 100% of B-lineage ALL (Brown and Small 2004a; Carow et€al. 1996; Rosnet et€al. 1996; Birg et€ al. 1992). Microarray studies have shown that in primary samples from children with leukemia, the highest levels of FLT3 expression occur in cases of infant and childhood ALL with MLL gene rearrangements (Armstrong et€al. 2002b; Yeoh et€ al. 2002; Ross et€ al. 2003). Nearly all leukemia cells also express FLT3 ligand (FL), leading to constitutive FLT3 phosphorylation and activation through autocrine, paracrine, or intracrine mechanisms (Zheng et€al. 2004; Drexler 1996; Meierhoff et€al. 1995; Brasel et€al. 1995). FLT3 can also be activated through two types of activating mutation: internal tandem duplications (ITD) in the juxtamembrane region and point mutations (PM) within the activation loop of the kinase domain (Yamamoto et€al. 2001; Abu-Duhier et€al. 2001; Nakao et€al. 1996). Both mutations result in ligand-independent FLT3 autophosphorylation and subsequent activation of downstream targets, including
34
P.A. Brown and C.A. Felix
proteins in the STAT, MAP kinase, and AKT pathways, which are involved in proliferation, differentiation, and survival (Lavagna-Sevenier et€ al. 1998; Zhang et€al. 1999, 2000 Zhang and Broxmeyer 2000; Kiyoi et€al. 1998, 2002). FLT3 mutations have been found in many adult and pediatric leukemias, most notably in AML, where they are the most common somatic genetic alteration (Brown and Small 2004a; Levis and Small 2003; Meshinchi et€al. 2001). FLT3/ITD is present in approximately 25% of adult AML and 15% of pediatric AML. FLT3/PM occurs in 7 to 9% of both adult and pediatric AML. In childhood ALL, FLT3/ITD is rare (1%), but FLT3/PM occurs commonly in ALL with rearrangements of the MLL gene in infants and children, where FLT3/PM are present in approximately 15% of cases (Armstrong et€al. 2004a; Taketani et€al. 2004a). Constitutively activated FLT3 signaling has been shown in many model systems to contribute to the process of leukemogenesis, but constitutive FLT3 activation alone is insufficient to fully transform primary hematopoietic cells (Kelly et€ al. 2002a, b; Li et€al. 2008a). Furthermore, a number of clinical studies in adult and pediatric AML have shown that activating mutations, particularly FLT3/ITD, are associated with inferior prognosis (Brown and Small 2004a; Levis and Small 2003; Meshinchi et€al. 2001, 2006). The demonstrated importance of FLT3 signaling in the pathogenesis of human leukemia has led to the development of FLT3-targeted agents. CEP-701 (lestaurtinib) is a highly selective small molecule FLT3 tyrosine kinase inhibitor (TKI) with excellent oral bioavailability and an IC50 of 3€nM (Levis et€al. 2002). Lestaurtinib is selectively cytotoxic to primary pediatric AML samples with FLT3/ITD mutations (Brown et€ al. 2004). Further, in AML cells, lestaurtinib exhibits scheduledependent synergy with multiple chemotherapeutic cytotoxic agents (Levis et€al. 2004). Prior adult phase II trials of lestaurtinib monotherapy for relapsed/refractory FLT3 mutant AML demonstrated clinical activity with clearance of peripheral blasts and, less frequently, bone marrow responses; now lestaurtinib is being tested in combination with standard chemotherapy (Smith et€al. 2004; Stone et€al. 2005). Based on recent evidence that lestaurtinib is selectively cytotoxic against MLL rearranged infant and childhood ALL with high levels of constitutively activated FLT3 expression, (Brown et€ al. 2005a) the Children’s Oncology Group (COG) is currently studying lestaurtinib in combination with chemotherapy in FLT3-mutant AML and infant ALL. Brown et€al. reported the effects of a FLT3 inhibitor (lestaurtinib, or CEP-701) on primary pediatric ALL (Brown et€ al. 2005a) and AML (Brown et€ al. 2004) blasts. The ALL study examined 36 childhood ALL samples, including seven infant samples, with known cytogenetics, quantitative FLT3 expression determined by microarray analysis and/or Western blotting, and FLT3 genotyping. MTT cytotoxicity and annexin V binding (AVB) apoptosis assays were done at 48€hours using increasing doses of CEP-701. CEP-701 was shown to be selectively cytotoxic to samples with high FLT3 expression; within this group, MLL-rearranged and high hyperdiploid samples were especially sensitive to FLT3 inhibition. A significant linear correlation between the MTT and AVB assays in the CEP-701-sensitive samples (Nâ•›=â•›18) indicated that the predominant mechanism of cytotoxicity involved
Molecularly Targeted Therapy for Infant ALL
35
Fig.€1╅ Sensitivity of infant and childhood ALL samples to FLT3 inhibition is based upon marked overexpression of constitutively activated FLT3
apoptosis. To investigate the molecular basis for the variability in CEP-701 cytotoxicity, FLT3 immunoprecipitation and immunoblotting (IP/IB) were performed on CEP-701 untreated and treated samples in cases where there were sufficient cells (Fig.€1); five of the six insensitive samples did not express appreciable FLT3 levels. The one sample that did express a high level of FLT3 had no constitutive FLT3 phosphorylation. By contrast, all seven sensitive samples expressed high levels of FLT3, and FLT3 was constitutively activated. Moreover, all but one of the sensitive samples expressed wild-type FLT3; suggesting the possibility of autocrine activation of FLT3 by FL coexpression, as has been demonstrated in some cases of AML expressing wild-type FLT3 (Zheng et€al. 2004). In all seven cases, FLT3 phosphorylation was potently inhibited by 50€ nM CEP-701. Thus, there appears to be a strong correlation between the cytotoxic response to CEP-701 and high-level constitutively activated FLT3 expression, suggesting that analyses of FLT3 expression and activation status may prospectively identify patients likely to benefit from FLT3 TKI. In the AML study, 44 cryopreserved diagnostic infant (N╛=╛4) and childhood (N╛=╛40) AML samples from the COG leukemia bank were studied. The FLT3 mutation status had been determined by PCR on genomic DNA isolated from the blasts at diagnosis (Meshinchi et€al. 2001, 2003). Cases were selected to evaluate approximately equal numbers of FLT3/ITD, FLT3/PM, and FLT3/WT samples; investigators were blinded to FLT3 mutation status until CEP-701 activity was assayed. MTT assays indicated selective cytotoxicity of CEP-701 for FLT3/ITD
36
P.A. Brown and C.A. Felix
samples compared to FLT3/PM or FLT3/WT samples. These studies provide preliminary data supporting the potential for FLT3 inhibitors for infant leukemia, although more samples will need to be studied. Since any molecularly targeted agent is unlikely to be curative as monotherapy in acute leukemia, how FLT3 inhibitors might be incorporated most effectively into standard chemotherapy regimens was also investigated. These studies have shown that lestaurtinib and chemotherapy agents consistently result in synergistic killing of FLT3/ITD+ AML and MLL-rearranged ALL cell lines and primary patient samples when used in the specific sequence of exposure to chemotherapy followed by lestaurtinib (Levis et€al. 2004; Brown et€al. 2006a). Other sequences of exposure (simultaneous and lestaurtinib followed by chemotherapy) are less synergistic and sometimes antagonistic due to the cell-cycle inhibition by lestaurtinib. The results of these experiments have informed the design of the ongoing adult clinical trials of the combination of chemotherapy and lestaurtinib, as well as the ongoing COG trials for pediatric patients (discussed below), illustrating the potential for well-designed preclinical laboratory studies to guide the clinical development of novel agents. Anti-FLT3 monoclonal antibodies represent another potential approach for targeting FLT3, including a fully humanized anti-FLT3 monoclonal antibody (IMC-EB10) that avidly binds FLT3 and blocks FL-mediated receptor activation (Piloto et€al. 2005, 2006). Because the primary mechanism of FLT3 activation in infant leukemia is FL activation of the overexpressed wild type receptor, EB10 may be particularly effective in infant leukemia. Other potential advantages include inhibiting FLT3 more specifically than is possible with small molecule inhibitors, and eradicating FLT3-expressing leukemia cells through antibody-dependent cellular cytotoxicity (ADCC) and complement fixation independently of effects on FLT3 signaling. Thus, antibodies may prove efficacious in leukemias where FLT3 is expressed, regardless of the degree to which leukemia cell survival is FLT3 dependent. A NOD/SCID xenograft model has been used to explore the effects of FLT3targeted agents on the leukemic stem cell (LSC) in infant leukemia in€vivo (Piloto et€al. 2006). Leukemic blasts from a 3-month-old infant with MLL-rearranged ALL (MLL-AF4 fusion) and high level constitutively phosphorylated FLT3 expression, which also exhibited pronounced in€ vitro cytotoxicity with lestaurtinib, were injected by tail vein (1â•›×â•›106 cells/mouse) into sublethally-irradiated 6-week-old NOD/SCID mice. The mice then were treated with subcutaneous injections of vehicle control vs. CEP-701 twice daily, or intraperitoneal injections of C225 (antiEGFR control antibody) vs. EB10 (anti-FLT3 antibody) three times weekly. After 14€ weeks, the mice were sacrificed and assessed by flow cytometry for human leukemia cell engraftment in the bone marrow. CEP-701 and EB10 significantly reduced the marrow engraftment of primary infant ALL cells in NOD/SCID mice. To further examine EB10 activity, a luciferase-GFP construct was packaged in a lentiviral vector and transfected into the SEM-K2 cell line, which was derived from a case of childhood MLL-AF4 ALL at relapse and has high level expression of activated wild-type FLT3. After sorting for GFP+ cells, 1â•›×â•›106 cells were injected into the tail vein of sublethally irradiated 6-week-old NOD/SCID mice. On day 11,
Molecularly Targeted Therapy for Infant ALL
37
in€vivo imaging showed early bone marrow engraftment and a single IP EB10 injection was administered, resulting in marked reversal of established engraftment and prevention of further engraftment as demonstrated 3€days later by reimaging. Thus, FLT3 targeted therapy demonstrates activity not only against bulk leukemia cells, but also against the self-renewing leukemia stem cells responsible for long-term engraftment in xenograft models. An important consideration for the ensuing clinical development of FLT3 inhibitors was whether sustained FLT3 inhibition could be achieved at doses that are tolerable in patients. Small molecules such as lestaurtinib are often highly (>99%) plasma protein bound, such that total drug levels measured by standard pharmacokinetic assays give little information regarding drug activity. Furthermore, direct assessment of FLT3 inhibition in leukemic blasts in patients is particularly unfeasible in clinical trials in which the chemotherapy administered beforehand results in clearance of blasts from the blood. Therefore, a plasma inhibitory activity (PIA) was developed as a surrogate for direct assessment of in€ vivo FLT3 inhibition (Levis et€al. 2006). Plasma is collected from patients prior to and at various trough time points after CEP-701 is administered. Aliquots of TF-1/ITD cells (TF-1 human leukemia cell line engineered to express a FLT3/ITD mutation) are exposed to the plasma, then lysed and subjected to FLT3 IP and P-FLT3 Western blotting. The percent inhibition of P-FLT3 compared to cells exposed to the patient’s pretreatment plasma (untreated control) is calculated to give an estimate of FLT3 inhibitory activity present in the patient’s plasma (PIA) at each time point. Validation of this approach has been done by comparing results in cases where FLT3 inhibition could simultaneously be measured directly in leukemic blasts and in plasma for the same patient at the same time point. The PIA is considered adequate if it is >80% at trough time points, corresponding to the degree of prolonged FLT3 inhibition required for cytotoxicity. The adult AML clinical trials with lestaurtinib have successfully correlated clinical responses with FLT3 PIA levels (Smith et€al. 2004; Levis et€al. 2005). PIA analysis is being performed in the ongoing COG clinical trials of lestaurtinib in childhood AML and infant ALL, and excellent levels of inhibitory activity at tolerable dose levels have been observed in some patients, suggesting that an adequate PIA can be achieved in infants and children also. Lestaurtinib is now being tested in the COG for children with relapsed/refractory FLT3-mutant AML. Lestaurtinib is given in sequential combination with reinduction chemotherapy (HiDAC and idarubicin), where the sequence of exposure (chemotherapy followed by FLT3 inhibitor) is based upon preclinical studies demonstrating maximal synergy between FLT3 inhibition with this sequence, compared to only additive effects with simultaneous exposure and antagonism with FLT3 inhibitor followed by chemotherapy (Levis et€al. 2004; Brown et€al. 2006a). The antagonism seen with the sequence of FLT3 inhibition followed by chemotherapy is due to the cell cycle arrest from FLT3 inhibitors, which protects the cells from the cytotoxic effects of chemotherapeutic agents that are cell-cycle dependent. Plans to incorporate FLT3 inhibitors into COG phase III clinical trials for patients with de€novo FLT3/ITD positive AML are in progress.
38
P.A. Brown and C.A. Felix Key
Enroll
Induction
A
Post-induction Arm A (SR chemo)
B
Post-induction Arm B (IR/HR chemo)
C
Post-induction Arm C (IR/HR chemo + lestaurtinib)
MLL-G
MLL-R
A
Randomize
B
C
Fig.€2╅ Design of COG study AALL0631. MLL-G, germline (wt) MLL; MLL-R, rearranged MLL
Lestaurtinib is also being tested in the COG for infants with newly diagnosed MLL-rearranged ALL. Lestaurtinib is being added in a randomized fashion to the multicourse chemotherapy regimen used in the previous clinical trial for infant ALL (COG P9407). The design of this study (Fig.€2) takes into account the preclinical data regarding combinations of lestaurtinib and chemotherapy, as lestaurtinib will be given immediately following the exposure to standard cytotoxic chemotherapy in an effort to maximize potential synergy, and will not be given for at least 24€hours prior to chemotherapy to avoid potential for antagonism.
Targeting Anti-apoptotic BCL-2 Family Members Deregulated apoptosis due to imbalanced expression of BCL-2 family proteins is a general avenue to chemotherapy resistance because the cytotoxicity of most anticancer chemotherapy occurs through this pathway. The focus on silencing antiapoptotic BCL2 family members in MLL leukemia in infants grew primarily from protein and gene expression level data on abundant expression of BCL-2 and other pro-survival BCL-2 family members in MLL leukemia in infants (Robinson et€al. 2008; Zhang et€ al. 2008a) coupled with preclinical and clinical adult forerunner data in different cancer cell types suggesting the utility of apoptosis pathway modulation, with selection of compounds to silence anti-apoptotic BCL-2 family members that were already in the clinical phases of development and showed minimal toxicities in adults.
Molecularly Targeted Therapy for Infant ALL
39
BCL-2 family members fall into three classes: multidomain anti-apoptotic proteins homologous in all four BCL-2 homology (BH) domains; multidomain pro-apoptotic proteins homologous in BH1, BH2, and BH3 domains; and BH3only proteins, which all are pro-apoptotic. BCL-2 family protein interactions occur via docking of the BH3 domain with a deep binding groove on the surface of the binding partner (Green 2005; Reed 2002, 2003). Pro-apoptotic multidomain BAX and BAK exist as inactive monomers, but they homo-oligomerize upon BH3-only BIM or BID activation, leading to mitochondrial outer membrane permeabilization and release of apoptotic cofactors cytochrome c and Smac/DIABLO to the cytosol (Danial and Korsmeyer 2004; Nguyen et€al. 2007). Cytochrome c is then bound by Apaf-1, which forms Apaf-1 oligomers and recruits pro-caspase 9. Caspase-9 activation activates effector caspases that execute apoptosis (Danial and Korsmeyer 2004). Many targeted strategies, including antisense therapies, siRNAs, peptides, and small molecule inhibitors have emerged for silencing anti-apoptotic BCL-2 family proteins at the transcript and the protein levels (Reed 2003; Gewirtz et€al. 1998; Letai et€al. 2002; Garber 2005). Since quantitative real-time PCR and Western blot analysis demonstrated that BCL-2 mRNA and protein are abundant and suggested that BCL-2 provides a cell survival mechanism in cases of pediatric/infant ALL and AML with the t(4;11) or other MLL translocations, (Robinson et€al. 2008) the first foray to modulate apoptosis in infant leukemias involved preclinical in€ vitro studies on Genasense™ (Oblimersen, G3139; Genta, Inc.), an 18-mer phosphorothioate antisense oligodeoxynucleotide (ODN), which forms a DNA–RNA hybrid with the first six codons of the BCL-2 mRNA (Nicholson 2000). Among the forerunner adult studies on this agent was an encouraging Phase I trial of Genasense™ with chemotherapy for refractory/relapsed adult acute leukemia, in which 45% of patients had disease response and BCL-2 mRNA decreased in 75% of cases; (Marcucci et€al. 2003) in addition, the toxicities (low-grade fever, transient liver function abnormalities, fatigue, thrombocytopenia) were limited (Marcucci et€al. 2003). Its clinical activity in refractory and relapsed adult B-cell malignancies (O’Brien et€al. 2005; Waters et€al. 2000) also was encouraging. Abundant BCL-2 expression was detected in the cell lines RS4:11, MV4-11, and SEM-K2 with t(4;11), similar to primary MLL leukemia cases, enabling drug activity and mechanistic assays on the cell line models (Robinson et€ al. 2008). Genasense™ abrogated BCL-2 mRNA expression in RS4:11 cells and significantly reduced the protein, though to recognize the changes these molecular analyses were performed at high concentrations (Robinson et€ al. 2008). However, studies of cytotoxicity and effects on apoptosis and cell cycle suggested that lower concentrations were sufficient to change the apoptosis threshold (Robinson et€al. 2008). In MTT assays of cytotoxicity, the most sensitive of the three cell lines to single-agent Genasense™ was RS4:11, and SEM-K2 was the least sensitive (Robinson et€al. 2008). Applying pharmacostatistical response surface modeling of drug interactions, Genasense™ at low concentrations was also found to sensitize RS4:11 and MV4-11 cells to select anti-leukemia cytotoxic drugs (ADR, VP16, ARAC, 6-TG in RS4:11; VP16, 6-TG in MV4-11), indicating synergistic
40
P.A. Brown and C.A. Felix
interactions (Robinson et€al. 2008). In further flow cytometric mechanistic studies, combining Genasense™ with doxorubicin in RS4:11 cells increased active caspase-3 and TUNEL staining compared to doxorubicin alone, consistent with greater apoptosis (Robinson et€ al. 2008). In addition, Genasense™ increased S-phase progression, (Robinson et€al. 2008) which is consistent with an effect on the BCL-2 target transcript since the anti-apoptotic effects of BCL-2 are linked to retardation of the G1/S transition. (Deng et€al. 2003) This evaluation of the effects of the pro-apoptotic BCL-2 antisense compound Genasense™ designed to decrease BCL-2 protein levels via selective BCL-2 mRNA degradation as a prototypic strategy to modulate cell death, indicated that abundant BCL-2 affords a molecular target in leukemias with t(4;11). Although the cardinal anti-apoptotic regulator in the intrinsic cell death pathway, BCL-2, has a general role in chemotherapy resistance (Danial and Korsmeyer 2004), the formation of homo- and heterotypic dimers by many different BCL-2 family proteins with opposing anti- or pro- apoptotic actions together determines the apoptosis threshold and regulates the intrinsic (mitochondrial) cell death pathway. Therefore, an increasing number of small molecule inhibitors also have emerged to restore balance in this pathway. Obatoclax (GX15-070; GeminX, Inc.) is one that antagonizes a broad spectrum of anti-apoptotic BCL-2 family proteins (BCL-2, BCL-XL, A1, BCL-w, MCL-1) (Nguyen et€al. 2007). Removal of the mitochondrial block by obatoclax also may enhance TRAIL-mediated extrinsic apoptosis (Huang et€ al. 2009). Thus, there has been significant impetus for preclinical and clinical obatoclax development to potentiate chemotherapy by activating apoptosis mechanisms in malignancies with inherent or acquired chemotherapy resistance. Obatoclax represents the first molecularly targeted agent for silencing anti-apoptotic BCL-2 family members that will be tested in children. The recently approved COG trial ADVL0816, a Phase I study of obatoclax in combination with vincristine/ doxorubicin/dexrazoxane in children with relapsed/refractory solid tumors or leukemia, incorporates a separate stratum for patients with MLL disease on the basis of preclinical activity data elaborated on below. The pan-anti-apoptotic BCL-2 family small molecule inhibitor obatoclax has the ability to bind to the BH3 binding pocket of the anti-apoptotic BCL-2 family members and block their anti-apoptotic function (Garber 2005). A synthetic derivative of the bacterial prodiginines family of red-pigmented tripyrrolic compounds developed on the basis of structure activity relationships; (Williamson et€al. 2007) obatoclax has been reported to cause apoptosis in€vitro in a wide range of human cancer cells derived from adult leukemias, lymphomas, and solid tumors, (Campas et€al. 2006; Galan et€al. 2005; Perez-Galan et€al. 2007; Li et€al. 2007; Li et€al. 2008b; Martinez-Paniagua et€al. 2007; Trudel et€al. 2007a; Witters et€al. 2007) and single agent responses of several human solid tumor xenografts occurred. In solid tumor models, obatoclax showed synergy with cisplatin in nonsmall cell lung cancer cell lines (Li et€al. 2008b). Synergy with cytosine arabinosine was reported in AML cell lines (Konopleva et€ al. 2008). In mantle cell lymphoma lines, obatoclax induced apoptosis and enhanced the cytotoxicity of vincristine, (Bebb et€al. 2006) proteasome inhibitors and doxorubicin (Yazbeck et€ al. 2006). In addition to inducing
Molecularly Targeted Therapy for Infant ALL
41
synergistic cytotoxicity when combined with cisplatin or doxorubicin (Martinez-Paniagua et€ al. 2007; Hernandez-Ilizaliturri et€ al. 2006a), obatoclax increased rituximabmediated antibody dependent cellular cytotoxicity and complement mediated cytotoxicity in non-Hodgkins lymphoma cell lines (Hernandez-Ilizaliturri et€al. 2006b). Obatoclax also exhibited synergy in combination with dexamethasone in a dexamethasone sensitive multiple myeloma cell line and additive activity in several other multiple myeloma lines (Trudel et€al. 2007a). Therefore, obatoclax rapidly advanced into adult Phase I and II solid tumor and leukemia clinical trials (Borthakur et€ al. 2006; Goy et€ al. 2007a; O’Brien et€ al. 2009; Schimmer et€al. 2007a; Verstovsek et€al. 2007). The salient observations of adult Phase I trials evaluating various infusion durations and schedules have been that obatoclax is well tolerated with minimal toxicities, and that its primary toxicities, which affect the central nervous system (i.e., somnolence, euphoria, confusional state, ataxia), are reversible, infusion-related, and self-limited once the infusion stops. Importantly, modest obatoclax activity was suggested against various relapsed/refractory adult leukemias and solid tumors with minimal toxicity (O’Brien et€ al. 2009; Goy et€ al. 2007b). Remarkably, single agent obatoclax resulted in a complete cytogenetic remission, resolution of cytopenias, and transfusion independence in a patient with secondary AML with the t(9;11) translocation (Schimmer et€al. 2007b). Thus, the attractive toxicity profile and the activity in MLL disease in an adult, made this agent especially suitable for preclinical pediatric studies, especially studies on MLL rearranged leukemia in infants as a potential pediatric cancer where obatoclax might have eventual utility in the clinic and substantial impact. In addition, HOXA9 repression of pro-apoptotic BIM (BCL2L11) was shown in a cell line with t(9;11), (Stubbs et€al. 2007) and silencing of the MLL-AF4 fusion oncoprotein in cell lines with t(4;11) using siRNAs not only decreased BCL-XL protein, but also increased activated caspase 3 and increased apoptosis (Thomas et€al. 2005). The rationale to test obatoclax also extended more generally to pediatric cancers and leukemia. In neuroblastoma, the inherent cancer cell survival mechanism derives from anti-apoptotic MCL-1 (Goldsmith et€al. 2006). In contrast, BCL-XL is believed to be the direct transcriptional target of the PAX3-FKHR translocation, the primary molecular aberration in alveolar rhabdomyosarcoma (Xia et€al. 2002). In childhood ALL, pro-apoptotic BIM expression predicted early response to treatment in the high-risk patients, (Bhojwani et€ al. 2008) and decreased expression of the extrinsic apoptosis effector caspase CASP8 predicted MRD during and following induction (Flotho et€al. 2007). Studies of apoptosis genes in relation to in€vitro drug responses in childhood ALL uncovered the importance of MCL-1 in prednisolone sensitivity and of the new BCL-2 family members BCL2L13 and HRK in L-asparaginase resistance (Holleman et€ al. 2006). Importantly, preclinical in€ vitro studies suggested that obatoclax restored the glucocorticoid response in steroid resistant childhood precursor B-cell and T-cell ALL (Bonapace et€ al. 2007). That not all of the emerging BH3 mimetics have high MCL-1 affinity, (van Delft et€ al. 2006) also made obatoclax an attractive agent to advance in pediatrics.
42
P.A. Brown and C.A. Felix
The principal pediatric preclinical studies leading to the ADVL0816 obatoclax Phase I trial in relapsed/refractory solid tumors and leukemia were performed on MLL rearranged leukemia in infants. Quantitative real-time PCR suggested that not only anti- but also several pro-apoptotic BCL-2 family mRNAs, apoptosis execution, and various mitochondrial maintenance mRNAs were abundant in MLL rearranged infant ALL (Zhang et€al. 2008a). Since basal expression levels of cell death and cell survival factors are relevant to finding novel therapeutic targets, these studies formed the basis to evaluate the potential of obatoclax to silence the activity of anti-apoptotic BCL-2 family protein targets in primary MLL leukemia cases (Zhang et€al. 2008b). MTT assays of MLL rearranged infant ALL and bilineal acute leukemias showed variable single-agent sensitivities, many within clinically achievable range (Zhang et€ al. 2008b). In a primary ALL with the MLL-AF4 rearrangement, pharmacostatistical response surface modeling indicated synergy between obatoclax and many of the common antileukemia cytotoxic chemotherapeutic drugs (cytosine arabinoside, doxorubicin, etoposide, dexamethasone, L-asparaginase, vincristine) (Zhang et€ al. 2008b). Interestingly, apoptosis and the alternative cell death pathway of autophagy, in which cellular components are degraded through the lysosomal machinery, are partially overlapping in downstream cell death mediators (Bonapace et€ al. 2007; Klionsky et€ al. 2008; Maiuri et€ al. 2007). Accumulating evidence from standard studies of autophagy (Klionsky et€al. 2008) and autophagy factors is suggesting that the obatoclax cell death mechanism in MLL leukemia involves the autophagy cell death pathway (Zhang et€al. 2008b). Biomarkers in the autophagy pathway will be further studied as the Phase I clinical trial of obatoclax ADVL0816 advances. Beyond the demonstration of preclinical in€vitro activity, obatoclax also is attractive to advance for leukemia in infants because pharmacometric modeling strategies and a pharmacokinetic study carried out in a diseased NOD-scid-IL-2Rgnull (NOG) immunodeficient xenograft model that recapitulates the hyperleukocytosis and extramedullary involvement of MLL leukemias, (Zhang et€al. 2008c) have demonstrated that obatoclax has excellent biodistribution to sites of involvement by MLL leukemia in infants, as indicated by tissue:plasma concentration ratios, including excellent CNS penetration (Zhang et€al. 2008c, 2007). The ADVL0816 Phase I clinical study of obatoclax summarized in Fig.€3 was approved by Clinical Trials Evaluation Program (CTEP) and opened for enrollment in COG Phase I Consortium Institutions in 2009. After an initial single agent dose, obatoclax will be administered in combination with vincristine, doxorubicin, and dexrazoxane. Patients will be enrolled in three separate strata. Patients with relapsed/refractory solid tumors (Stratum 1) will be enrolled in a dose escalation trial with a starting obatoclax dose of approximately 70% of the adult MTD. As a result of the striking preclinical obatoclax activity, patients with relapsed/refractory MLL-rearranged leukemia (Stratum 2) will be eligible for enrollment concurrently with the dose finding phase in solid tumor stratum. Once the dose finding phase has been completed, a patient cohort with relapsed/refractory non-MLL leukemia will be enrolled in Stratum 3.
Molecularly Targeted Therapy for Infant ALL
43
Fig.€3╅ COG study ADVL0816 treatment schema
The foregoing studies have built a foundation for overcoming drug resistance from abnormal cell death regulation in leukemia in infants where current treatments offer little hope due to the inherent drug resistance and unique vulnerabilities of infants to toxicities from intensive treatments. However, that MLL leukemias exhibit variable preclinical obatoclax sensitivities, (Zhang et€ al. 2008b) although promising, also underscores that no one agent targeting this pathway is expected to be active in all cases. The small molecule inhibitor ABT737, which targets anti-apoptotic BCL-2, BCL-XL, and BCL-w, (van Delft et€al. 2006) is another candidate to evaluate for cell death pathway modulation in leukemia in infants. ABT-737 disrupts BCL-2 family protein interactions by targeting the BH3 binding pocket of select anti-apoptotic BCL-2 family members (van Delft et€ al. 2006). Inhibition of interactions of anti-apoptotic BCL-2 family proteins with proapoptotic BCL-2 family proteins by ABT-737 and activity of ABT-737 alone and in combination with various cytotoxic drugs has been demonstrated in several preclinical studies of leukemia (Kang et€ al. 2007; Trudel et€ al. 2007b; Chauhan et€al. 2007; Konopleva et€al. 2006; Kline et€al. 2007; Kojima et€al. 2006) and ABT263, which has identical pharmacologic properties, is advancing to adult clinic trials. This compound demonstrates mechanism-based single agent activity in lymphomas (Oltersdorf et€al. 2005). Of particular interest, in€vitro activity of ABT737 was recently studied in seven different ALL cell lines [T-cell ALL: COG-LL-317, CEM, MOLT-3, MOLT-4; pre-B ALL: COG-LL-317, Nalm-6, RS4:11], (Kang et€ al. 2007) among which the RS4:11 cell line was established from a case of adult ALL with t(4;11); however, no cases of infant leukemia have been tested. Strongly synergistic or synergistic cytotoxicity as defined by combination indices £0.3 or £0.7, respectively, of ABT-737 combined with cytotoxics, especially L-asparaginase or vincristine, was shown in all of the cell lines, even when there was no single agent sensitivity; the strongest synergy was with Lasparaginase. Further study of the pro-Â�apoptotic effects of the ABT-737/L- asparaginase combination in RS4:11 and COG-LL-317 cells showed evidence of
44
P.A. Brown and C.A. Felix
mitochondrial apoptosis pathway �activation including an increase in the truncated form of BID, tBID, increased hypophosphorylated BAD, an increase in the activated form of BAX, BAX(s), an increase in the cleaved form of CASP8, which cleaves BID to tBID, and cleavage of CASP9 and CASP3. The ABT-737/L- asparaginase combination in COG-LL-317 cells resulted in increased Annexin V/FITC staining and increased JC-1 probe staining, and Western blot analysis showed increased mitochondrial cytochrome c release. Importantly, ABT-737 did not increase apoptosis in normal PBMCs. ABT-737 showed synergistic cytotoxicity when combined with vincristine, L- asparaginase and dexamethasone in 5 of the 7 cell lines tested, including RS4:11. ABT-737 combined with vincristine, L- asparaginase, and dexamethasone delayed the progression of established xenografts in two NOD/SCID models of relapsed primary ALL and increased EFS compared to chemotherapy alone or ABT-737 alone in one of the two xenografts. As described in detail above, infant acute leukemia exhibits deregulated FLT3 receptor tyrosine kinase signaling as a separate resistance mechanism (Armstrong et€ al. 2003, 2004b; Brown et€ al. 2006b; Taketani et€ al. 2004b; Brown and Small 2004b; Brown et€al. 2005b; Stam et€al. 2005; Stubbs and Armstrong 2007) in addition to the imbalanced expression of cell death regulatory factors (Zhang et€ al. 2008a). Therefore, it may be of substantial interest that Abbott Pharmaceuticals has developed small molecule inhibitors to both pathways. ABT-869 is an ATPcompetitive small molecule inhibitor with specificity for the VEGF and PDGF families of receptor tyrosine kinases (RTKs), thereby targeting multiple resistance mechanisms (Rodila et€al. 2006). There is rationale to test ABT-869 (Carlson et€al. 2005; Dai et€al. 2007; Shankar et€al. 2007; Albert et€al. 2006) alone and combined with the pro-apoptotic agent ABT-737 in leukemias in this population, since FLT3 is a member of the PDGF family of RTKs (Dai et€al. 2007). The activity of ABT-869 is dependent on mutant kinases (Albert et€al. 2006). In the AML cell lines MV4-11 and Molm13 with MLL translocations and FLT3ITD mutations, ABT-869 already has exhibited in€vitro activity in cell proliferation/ viability assays, and downregulation of FLT3, STAT5, and ERK phosphorylation and the expression of the STAT5 target gene Pim-1 (Shankar et€al. 2007). ABT-869 prevented the formation of subcutaneous MV4-11 xenografts, caused the regression of established subcutaneous MV4-11 xenografts, slowed the progression of subcutaneous Molm13 xenografts and prolonged the survival of mice with marrow engraftment of MV4-11 cells (Shankar et€al. 2007). Even though FLT3 tyrosine inhibitors may effectively decrease FLT3 phosphorylation, persistent phosphorylation of downstream signaling proteins (e.g., AKT, MAPK), activating mutations in parallel signaling pathways (e.g., RAS) or overexpression of other RTKs may be avenues to resistance, as has been shown in cell lines with MLL translocations (Piloto et€al. 2007). Importantly, however, the ABT737 pro-apoptotic drug can neutralize resistance to FLT3 tyrosine kinase inhibition (Kohl et€ al. 2007). Studies of ABT-737 and ABT-869 together in leukemia in infants would enable more detailed characterization of additional pathways to deregulated cell death mechanisms and RTK signaling and how these pathways can be targeted.
Molecularly Targeted Therapy for Infant ALL
45
Targeting MLL Fusion Transcripts The plethora of MLL partner genes and heterogeneity in genomic breakpoint locations, which increases transcript heterogeneity even further, would make targeting the fusion transcripts difficult. Nonetheless, it has been possible to silence several of the more common MLL fusion transcripts experimentally using nucleic acid therapeutics (Gewirtz et€al. 1998). Antisense oligodeoxynucleotides have resulted in successful downregulation of translation of MLL fusion proteins in€vitro in cell lines (Kawagoe et€al. 2001; Niitsu et€al. 2001a; Akao et€al. 1998). The MLL-AF9 downregulation in the THP-1 cell line resulted in reduced HOXA7 and HOXA10 gene expression and increased apoptosis (Kawagoe et€ al. 2001). MLL-CBP downregulation using antisense rendered the AML cell line SN-1 sensitive to all-trans retinoic acid or 1a,25-dihydroxyvitamin D3 induced differentiation (Niitsu et€al. 2001a). Importantly, pro-apoptotic effects of targeting the fusion transcript with an antisense oligodeoxynucleotide also were observed in the KOCL33 cell line derived from a case of infant ALL with the MLL-ENL translocation (Akao et€al. 1998). More recently, fusion transcript specific siRNAs were used to deplete the respective fusion oncoproteins resulting from two different MLL-AF4 genomic rearrangements in the SEM-K2 and RS4:11 cell lines (Thomas et€ al. 2005). The in€ vitro effects of the respective siRNAs included decreased colony formation on methylcellulose, decreased proliferation in MTT assays, an increase in G0/G1 phase cells and a decrease in cells in S phase, increased activated caspase 3 and decreased antiapoptotic BCL-XL protein, decreased HOXA7, HOXA9, and MEIS1 gene expression, and decreased expression of the stem cell marker CD133 (Thomas et€ al. 2005). Furthermore, injection of SEM-K2 cells that had been pretreated MLL-AF4 specific siRNA into SCID mice completely suppressed engraftment in a xenotransplantation assay (Thomas et€al. 2005).
PFWT Targeting of MLL Partner Protein Interactions The MLL partner proteins AF4, ENL, AF9, and AF10 interact in a multiprotein complex, which is involved in the regulation of transcriptional elongation and recruitment of the histone H3K79 methyltransferase Dot1 to elongating RNA Pol II (Bitoun et€al. 2007). The PFWT peptide mimetic of the AF9 binding domain of AF4 with a conjugated Penetratin transduction sequence at the amino terminus is in preclinical development for disruption of the AF4-AF9 interaction (Srinivasan et€ al. 2004). The AF-9 interaction domain of AF4, which binds at the AF9 carboxyl terminus, is highly conserved in all LAF4 family proteins (AF4, LAF4, AF5, FMR2) and is retained in the MLL fusion oncoproteins (Srinivasan et€al. 2004; Erfurth et€al. 2004). PFWT disruption of AF4-AF9 complexes was validated in GST pull-down assays and in NIH 3T3 cells, and PFWT decreased cell proliferation of RS4:11 cells, KP-L-RY (MLL-AF5) cells and, to a lesser degree, THP-1 cells with the MLL-AF9
46
P.A. Brown and C.A. Felix
rearrangement (Srinivasan et€ al. 2004). In the MV4-11 and Molm13 (MLL-AF9) myeloid cell lines, PFWT rapidly induces ultrastructural changes of necrosis; abrogation of PFWT cytotoxicity by the serine protease inhibitor TLCK implicated serine proteases in the cell death mechanism (Palermo et€al. 2008). In MV4-11 cells, promising synergistic interactions were suggested between PFWT and cytosine arabinoside, etoposide, the HSP-90 inhibitor 17-allylamino17-demethoxygeldanamycin (17-AAG), or the FLT3 inhibitor 3,4-dimethoxybenzoylamide, in Molm13 cells, between PFWT and cytosine arabinoside or etoposide and, in KOPN-8 (MLL-ENL) cells, though only at half the IC50 concentration, between PFWT and cytosine arabinoside (Bennett et€ al. 2009). Further study of MV4-11 cells showed that, with exception of 17-AAG, the penetratin sequence did not itself augment cytotoxicity (Bennett et€al. 2009). Interestingly, in MV4-11 cells in which synergy was sequence-dependent (cytosine arabinoside or etoposide before PFWT), PFWT alone caused an increase in cells in G1 and a decrease in cells in S phase (Bennett et€al. 2009). Treatment with etoposide and PFWT resulted in EM features of both apoptosis and necrosis and an increase activated caspase 3, indicating that the two cell death mechanisms are not mutually exclusive (Bennett et€al. 2009).
Targeting Glycogen Synthase Kinase 3 The recent basic research discovery demonstrating that the serine/threonine kinase GSK3 is a required, cell type specific MLL leukemia maintenance factor, suggested GSK3 as a potential therapeutic target. The role of GSK3 in MLL leukemia is paradoxical because GSK3 is involved in tumor suppression in other cancer types, and also because GSK3 normally phosphorylates and inactivates b -catenin, MYCN, and JUN, and inhibits pathways involved in self renewal and proliferation, whereas GSK3 inhibition increases stem cell pluripotency and HSC repopulation (Wang et€al. 2008). Targeting GSK3 using small molecule inhibitors (GSK3-IX, SB216763, alsterpaullone) in MLL-AF4 (SEMK2, RS4:11, MV4-11) and MLL-AF5 (KP-L-RY) cell lines inhibited proliferation and increased b-catenin expression, decreased G1–S phase cell cycle progression, and increased sub-G0-G1 DNA consistent with cell death. Culture of murine primary myeloid progenitors expressing transduced MLL-ENL, MLL-LAF4, MLL-AF6, or MLL-GAS7 in the presence of SB216763, decreased clonogenic potential and proliferation; myeloid differentiation and decreased c-Kit expression were suggested with prolonged exposure (Wang et€al. 2008). Transduction of murine Gsk3b−/− cells with MLL fusion genes, or shRNAmediated Gsk3b knockdown in murine myeloid progenitors transduced with MLL fusion genes, increased sensitivity to pharmacological GSK3 inhibition, even though Gsk3b was not required for transformation, whereas Gsk3a depletion alone
Molecularly Targeted Therapy for Infant ALL
47
did not have these effects. Still cooperativity was suggested since the sensitivity to pharmacological inhibition, decreased proliferation and clonogenicity, and increased myeloid differentiation were more profound in a Gsk3b−/− Gsk3aKD MLL fusion gene model in which both Gsk3 isoforms were depleted, and Gsk3b−/− Gsk3aKD MLL-ENL transduced cells did not induce leukemia in mice (Wang et€al. 2008). Additionally, the treatment of MLL-AF4 leukemia bearing mice with the GSK3 inhibitor lithium carbonate prolonged their survival (Wang et€al. 2008). With GSK3 inhibition, the CDKI protein p27Kip1 significantly increased in the MLL-AF5 leukemia cell line KP-L-RY, and in MLL-ENL transduced murine myeloid progenitors, indicating that the effects of GSK3 in MLL leukemias involve p27Kip1 suppression (Wang et€al. 2008). None of these effects were observed in non-MLL leukemia cell lines (Wang et€al. 2008).
HSP90 as a Potential Therapeutic Target The heat shock protein HSP90 is an ATP-dependent chaperone protein that functions in a multi chaperone protein complex with roles in the regulation of conformational maturation and refolding of client proteins, (Wiech et€al. 1992; Stebbins et€al. 1997; Terasawa et€al. 2005) tertiary structure of the proteasome, (Imai et€al. 2003) and steroid receptor functions and interactions (Terasawa et€ al. 2005). Of particular interest to MLL leukemia, FLT3 is another HSP90 client protein (Yao et€ al. 2003). Interestingly, in cancer cells HSP90 stabilizes PI3K (Belova et€ al. 2008; Fujita et€al. 2002) and AKT (Zhang et€al. 2005) and has an extracellular role in invasion and metastasis via regulation of MMP2 protein maturation (Picard 2004; Eustace and Jay 2004; Eustace et€al. 2004). HSP90 is of interest as a potential therapeutic target in infant leukemia. Proteomics-based 2D-DIGE analysis of whole cell proteins identified HSP-90 as a differentially expressed protein in the MLL-AF4 cell lines RS4:11 and MV4-11 compared to normal CD34+ cells (Yocum et€al. 2006). In studies testing HSP90 inhibition, MLL leukemia cell lines exhibited differential 17-AAG sensitivity depending on FLT3 status; RS4:11 cells in which FLT3 is wild type were less sensitive than MV4-11 or Molm13 with the FLT3 ITD (Yao et€al. 2003). In all three of these cell lines, HSP90 inhibition was associated with decreased FLT3, RAF, and AKT expression (Yao et€ al. 2003). In MV4-11 cells, not only did quantitative proteomics indicate that 17-AAG effectively depleted the HSP90a target, but also 17-AAG resulted in decreased viability and proliferation and increased apoptosis; a decrease in nucleoside diphosphate kinase nm23 suggested nm23 as a novel biomarker of HSP90 inhibition (Yocum et€ al. 2006). More recently, when synergy between 17-AAG and etoposide was tested, greater synergy was observed in Molm13 and MV4;11 with the FLT3 ITD than in RS4:11 cells with wild-type FLT3 (Yao et€al. 2007). Synergistic effects of combining
48
P.A. Brown and C.A. Felix
17-AAG with the FLT3 inhibitor GTP14564 in cell lines with FLT3-ITD (MV411) or amplified wild-type FLT3 (SEM-K2) have also been suggested, with reductions in phospho-FLT3 and phospho-STAT5, G0–G1 arrest and increased apoptosis (Yao et€al. 2005).
Epigenetic Strategies In cell lines with the MLL-AF9 translocation (THP-1, Mono-Mac-6, and Molm13), histone deacetylase inhibition with valproic acid has been associated with decreased cell viability, cell cycle arrest in G1, increased apoptosis and induction of p21, and the G cyclin CG2, which is involved in G1 cell cycle arrest (Tonelli et€al. 2006). In the SN-1 AML cell line with the MLL-CBP translocation, various HDAC inhibitors (sodium butyrate, trichostatin A, polyloxyl bytyrate) exhibited synergy with all-trans retinoic acid in inducing differentiation as measured by NBT-reduction (Niitsu et€ al. 2001a). In the SN-1 cell line, and in the KOCL33, KOCL51, and KOCL44 cell lines, all with the MLL-ENL translocation, the demethylating agent 5-aza-2¢-deoxycytidine resulted in p16/INK4A upregulation and sensitivity to differentiation therapy with all-trans retinoic acid and 1a,25-dihydroxyvitamin D3 (Niitsu et€al. 2001b). Taken together these studies may suggest that combinations of epigenetic therapies with differentiation therapies may have utility in MLL disease.
mTOR Inhibition The serine/threonine kinase mTOR (mammalian Target of Rapamycin) is an integrator of several signal transduction pathways, and is involved in the regulation of cell cycle, apoptosis and angiogenesis. The mTOR pathway is aberrantly activated in many tumor types, and the inhibition of mTOR signaling (using sirolimus or one of its analogs, such as temsirolimus or everolimus) has shown antitumor activity in several model systems as well as early phase clinical trials. In fact, temsirolimus was recently FDA-approved for the treatment of advanced renal cell carcinoma (Hudes et€al. 2007). There is some preclinical evidence that inhibition of the serine/ threonine kinase mTOR (mammalian Target of Rapamycin) may be a useful therapeutic strategy in infant leukemia. Sirolimus (and the related mTOR inhibitor temsirolimus) inhibit the growth of precursor B cell ALL cell lines in€vitro and are also active in ALL transgenic and xenograft murine models (Brown et€ al. 2003). In further studies, mTOR inhibitors showed synergistic cytotoxicity with methotrexate against human ALL xenografts (Teachey et€al. 2008). Moreover, sirolimus has been shown to overcome glucocorticoid resistance in ALL cells via anti-apoptotic MCL1 downregulation (Wei et€al. 2006). Although none of these studies were performed on MLL-rearranged leukemias specifically, several groups have suggested that the
Molecularly Targeted Therapy for Infant ALL
49
glucocorticoid resistance in MLL-rearranged infant ALL may be related to the poor prognosis (Pieters et€al. 1998; Palle et€al. 2005). Since mTOR inhibition can reverse glucocorticoid resistance, these studies may provide rationale to consider clinical mTOR inhibition in infant ALL.
Targeting CD33 Cell Surface Antigen Gemtuzumab ozogamicin is a humanized anti-CD33 antibody linked to the antitumor antibiotic calicheamicin. Not only is CD33 positivity observed in a preponderance of the subtypes of pediatric AML including cases with MLL translocations but also this agent may have a role in MLL rearranged ALL because of the frequent coexpression of myeloid cell surface markers (Zwaan et€al. 2003). The results of clinical testing of gemtuzumab ozogamicin for remission induction in conjunction with intensive post induction therapy for relapsed/refractory CD33+ AML have been encouraging (Zwaan et€al. 2003; Arceci et€al. 2005).
Conclusions Several examples of molecularly targeted agents at various stages of development for infant leukemia have been discussed. As the molecular pathobiology of infant leukemia continues to unravel, the characterization of MLL fusion oncoproteins and factors that interact with MLL fusion oncoproteins in multiprotein complexes or cooperate with them in downstream or parallel signaling pathways have led to identification of a number of promising targets and experiments on potentially efficacious molecularly targeted agents that might benefit this patient population. The agents are at various stages of testing along the drug development pipeline from basic to preclinical in€vitro and in€vivo testing, and more recently to testing in clinical trials. In 2010, most still are at the basic and preclinical stages of development. Identification of efficacious agents with a low toxicity profile holds promise to change the outcome of this subset of pediatric cancer that has heretofore been so refractory to more conventional treatments because the mechanisms of action of targeting specific signaling cascades are very different from the more general mechanisms of action of conventional chemotherapy, which result in such profound toxicities in infants in particular. Appropriate choices of molecularly targeted treatments for infant leukemia in the clinic are also expected to be reliant on molecular testing for the relevant target aberrations because of the heterogeneity in MLL partner proteins and the cooperating events in disease pathogenesis within this patient subset. Furthermore, molecular characterization will be critically important to understand mechanisms of response and resistance. As well exemplified by FLT3 tyrosine kinase inhibition, (Piloto et€al. 2007) the future therapeutic opportunities and challenges lie not only in identifying
50
P.A. Brown and C.A. Felix
primary mutations and coexisting mutations, but also exploiting acquired resistance mechanisms as avenues to new treatment in instances of treatment failure. Moreover, it has also been suggested that the eradication of the leukemia stem cell will likely prove essential to cure MLL disease (Somervaille and Cleary 2006). Acknowledgmentâ•… The work is supported by Leukemia & Lymphoma Society SCOR 7372-07.
References Gurney JG, Smith MA, Ross JA. Cancer among infants. In: Ries LAG, Smith MA, Gurney JG, et€al., eds. Cancer Incidence and Survival among children and adolescents: United States SEER program 1975–1995. Bethesda, MD: National Cancer Institute, SEER Program. NIH; 1999. Smith MA, Gloeckler-Ries LA, Gurney JG, Ross JA. Leukemia. In: Ries LAG, Smith MA, Gurney JG, et€al., eds. Cancer incidence and survival among children and adolescents: United States SEER Program 1975–1995. Bethesda, MD: National Cancer Institute, SEER Program, NIH; 1999. Pui CH, Kane JR, Crist WM. Biology and treatment of infant leukemias. Leukemia. 1995;9:762–769. SEER Cancer Statistics Review, 1975–2006; seer.cancer.gov/csr/1975_2006/ Meyer C, Kowarz E, Hofmann J, et€al. New insights to the MLL recombinome of acute leukemias. Leukemia. 2009. Rubnitz JE, Link MP, Shuster JJ, et€al. Frequency and prognostic significance of HRX rearrangements in infant acute lymphoblastic leukemia: a Pediatric Oncology Group study. Blood. 1994;84:570–573. Pui CH, Chessells JM, Camitta B, et€al. Clinical heterogeneity in childhood acute lymphoblastic leukemia with 11q23 rearrangements. Leukemia. 2003;17:700–706. Robinson BW, Devidas M, Carroll AJ, et al. Specific MLL partner genes in infant acute lymphoblastic leukemia (ALL) associated with outcome are linked to age and white blood cell count (WBC) at diagnosis: A report on the Children’s Oncology Group (COG) P9407 trial. Blood (ASH Annual Meeting Abstracts) 2009;114(22):907. Hilden JM, Dinndorf PA, Meerbaum SO, et€al. Analysis of prognostic factors of acute lymphoblastic leukemia in infants: report on CCG 1953 from the Children’s Oncology Group. Blood. 2006;108:441–451. Balgobind BV, Raimondi SC, Harbott J, et€al. Novel prognostic subgroups in childhood 11q23/ MLL-rearranged acute myeloid leukemia: results of an international retrospective study. Blood. 2009;114:2489–2496. Felix CA, Kolaris CP, Osheroff N. Topoisomerase II and the etiology of chromosomal translocations. DNA Repair (Amst). 2006;5:1093–1108. Smith MA, Rubinstein L, Anderson JR, et€al. Secondary leukemia or myelodysplastic syndrome after treatment with epipodophyllotoxins. J Clin Oncol. 1999;17:569–577. Liedtke M, Cleary ML. Therapeutic targeting of MLL. Blood. 2009. Ziemin-van der Poel S, McCabe NR, Gill HJ, et€al. Identification of a gene, MLL, that spans the breakpoint in 11q23 translocations associated with human leukemias. Proc Natl Acad Sci USA. 1991;88:10735–10739. Djabali M, Selleri L, Parry P, Bower M, Young BD, Evans GA. A trithorax-like gene is interrupted by chromosome 11q23 translocations in acute leukaemias. Nat Genet. 1992;2:113–118. Gu Y, Nakamura T, Alder H, et€al. The t(4;11) chromosome translocation of human acute leukemias fuses the ALL-1 gene, related to Drosophila trithorax, to the AF-4 gene. Cell. 1992;71:701–708. Tkachuk DC, Kohler S, Cleary ML. Involvement of a homolog of Drosophila trithorax by 11q23 chromosomal translocations in acute leukemias. Cell. 1992;71:691–700.
Molecularly Targeted Therapy for Infant ALL
51
Ford AM, Ridge SA, Cabrera ME, et€al. In utero rearrangements in the trithorax-related oncogene in infant leukaemias. Nature. 1993;363:358–360. Megonigal MD, Rappaport EF, Jones DH, et€al. t(11;22)(q23;q11.2) In acute myeloid leukemia of infant twins fuses MLL with hCDCrel, a cell division cycle gene in the genomic region of deletion in DiGeorge and velocardiofacial syndromes. Proc Natl Acad Sci U S A. 1998;95:6413–6418. Yokoyama A, Kitabayashi I, Ayton PM, Cleary ML, Ohki M. Leukemia proto-oncoprotein MLL is proteolytically processed into 2 fragments with opposite transcriptional properties. Blood. 2002;100:3710–3718. Hess JL. MLL: a histone methyltransferase disrupted in leukemia. Trends Mol Med. 2004;10:500–507. Takeda S, Chen DY, Westergard TD, et€ al. Proteolysis of MLL family proteins is essential for taspase1-orchestrated cell cycle progression. Genes Dev. 2006;20:2397–2409. Liu H, Cheng EH, Hsieh JJ. Bimodal degradation of MLL by SCFSkp2 and APCCdc20 assures cell cycle execution: a critical regulatory circuit lost in leukemogenic MLL fusions. Genes Dev. 2007;21:2385–2398. Armstrong SA, Staunton JE, Silverman LB, et al. MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nat Genet. 2002;30:41–47. Meyer C, Schneider B, Jakob S, et€ al. The MLL recombinome of acute leukemias. Leukemia. 2006;20:777–784. Ayton PM, Cleary ML. Transformation of myeloid progenitors by MLL oncoproteins is dependent on Hoxa7 and Hoxa9. Genes Dev. 2003;17:2298–2307. Martin ME, Milne TA, Bloyer S, et€ al. Dimerization of MLL fusion proteins immortalizes hematopoietic cells. Cancer Cell. 2003;4:197–207. Wong P, Iwasaki M, Somervaille TC, So CW, Cleary ML. Meis1 is an essential and rate-limiting regulator of MLL leukemia stem cell potential. Genes Dev. 2007. Somervaille TC, Cleary ML. Identification and characterization of leukemia stem cells in murine MLL-AF9 acute myeloid leukemia. Cancer Cell. 2006;10:257–268. Krivtsov AV, Twomey D, Feng Z, et€al. Transformation from committed progenitor to leukaemia stem cell initiated by MLL-AF9. Nature. 2006;442:818–822. Wang J, Iwasaki H, Krivtsov A, et€al. Conditional MLL-CBP targets GMP and models therapyrelated myeloproliferative disease. EMBO J. 2005;24:368–381. FDA. Challenges and opportunity on the critical path to new medical products. 2004. Brown P, Small D. FLT3 inhibitors: a paradigm for the development of targeted therapeutics for paediatric cancer. European journal of cancer (Oxford, England : 1990). 2004;40:707. Levis M, Small D. FLT3: ITDoes matter in leukemia. Leukemia : official journal of the Leukemia Society of America, Leukemia Research Fund, UK. 2003;17:1738. Carow CE, Levenstein M, Kaufmann SH, et€ al. Expression of the hematopoietic growth factor receptor FLT3 (STK-1/Flk2) in human leukemias. Blood. 1996;87:1089. Rosnet O, Buhring HJ, Marchetto S, et€al. Human FLT3/FLK2 receptor tyrosine kinase is expressed at the surface of normal and malignant hematopoietic cells. Leukemia. 1996;10:238–248. Birg F, Courcoul M, Rosnet O, et€al. Expression of the FMS/KIT-like gene FLT3 in human acute leukemias of the myeloid and lymphoid lineages. Blood. 1992;80:2584–2593. Yeoh EJ, Ross ME, Shurtleff SA, et€al. Classification, subtype discovery, and prediction of outcome in pediatric acute lymphoblastic leukemia by gene expression profiling. Cancer Cell. 2002;1:133–143. Ross ME, Zhou X, Song G, et€al. Classification of pediatric acute lymphoblastic leukemia by gene expression profiling. Blood. 2003;102:2951–2959. Zheng R, Levis M, Piloto O, et€al. FLT3 ligand causes autocrine signaling in acute myeloid leukemia cells. Blood. 2004;103:267. Drexler HG. Expression of FLT3 receptor and response to FLT3 ligand by leukemic cells. Leukemia. 1996;10:588–599. Meierhoff G, Dehmel U, Gruss HJ, et€al. Expression of FLT3 receptor and FLT3-ligand in human leukemia-lymphoma cell lines. Leukemia. 1995;9:1368–1372. Brasel K, Escobar S, Anderberg R, de Vries P, Gruss HJ, Lyman SD. Expression of the flt3 receptor and its ligand on hematopoietic cells. Leukemia. 1995;9:1212–1218.
52
P.A. Brown and C.A. Felix
Yamamoto Y, Kiyoi H, Nakano Y, et€al. Activating mutation of D835 within the activation loop of FLT3 in human hematologic malignancies. Blood. 2001;97:2434. Abu-Duhier FM, Goodeve AC, Wilson GA, Care RS, Peake IR, Reilly JT. Identification of novel FLT-3 Asp835 mutations in adult acute myeloid leukaemia. Br J Haematol. 2001;113:983–988. Nakao M, Yokota S, Iwai T, et€ al. Internal tandem duplication of the flt3 gene found in acute myeloid leukemia. Leukemia. 1996;10:1911–1918. Lavagna-Sevenier C, Marchetto S, Birnbaum D, Rosnet O. FLT3 signaling in hematopoietic cells involves CBL, SHC and an unknown P115 as prominent tyrosine-phosphorylated substrates. Leukemia. 1998;12:301–310. Zhang S, Mantel C, Broxmeyer HE. Flt3 signaling involves tyrosyl-phosphorylation of SHP-2 and SHIP and their association with Grb2 and Shc in Baf3/Flt3 cells. J Leukoc Biol. 1999;65:372–380. Zhang S, Broxmeyer HE. Flt3 ligand induces tyrosine phosphorylation of gab1 and gab2 and their association with shp-2, grb2, and PI3 kinase. Biochem Biophys Res Commun. 2000;277:195–199. Zhang S, Fukuda S, Lee Y, et€al. Essential role of signal transducer and activator of transcription (Stat)5a but not Stat5b for Flt3-dependent signaling. J Exp Med. 2000;192:719–728. Kiyoi H, Towatari M, Yokota S, et€al. Internal tandem duplication of the FLT3 gene is a novel modality of elongation mutation which causes constitutive activation of the product. Leukemia. 1998;12:1333–1337. Kiyoi H, Ohno R, Ueda R, Saito H, Naoe T. Mechanism of constitutive activation of FLT3 with internal tandem duplication in the juxtamembrane domain. Oncogene. 2002;21: 2555–2563. Meshinchi S, Woods WG, Stirewalt DL, et€ al. Prevalence and prognostic significance of Flt3 internal tandem duplication in pediatric acute myeloid leukemia. Blood. 2001;97:89. Armstrong SA, Mabon ME, Silverman LB, et€al. FLT3 mutations in childhood acute lymphoblastic leukemia. Blood. 2004;103:3544–3546. Taketani T, Taki T, Sugita K, et€al. FLT3 mutations in the activation loop of tyrosine kinase domain are frequently found in infant ALL with MLL rearrangements and pediatric ALL with hyperdiploidy. Blood. 2004;103:1085–1088. Kelly LM, Kutok JL, Williams IR, et€ al. PML/RARalpha and FLT3-ITD induce an APL-like disease in a mouse model. Proc Natl Acad Sci U S A. 2002;99:8283–8288. Kelly LM, Liu Q, Kutok JL, Williams IR, Boulton CL, Gilliland DG. FLT3 internal tandem duplication mutations associated with human acute myeloid leukemias induce myeloproliferative disease in a murine bone marrow transplant model. Blood. 2002;99:310–318. Li L, Piloto O, Nguyen HB, et€al. Knock-in of an internal tandem duplication mutation into murine FLT3 confers myeloproliferative disease in a mouse model. Blood. 2008;111:3849–3858. Meshinchi S, Alonzo TA, Stirewalt DL, et€al. Clinical implications of FLT3 mutations in pediatric AML. Blood. 2006;108:3654. Levis M, Allebach J, Tse KF, et€al. A FLT3-targeted tyrosine kinase inhibitor is cytotoxic to leukemia cells in€vitro and in€vivo. Blood. 2002;99:3885. Brown P, Meshinchi S, Levis M, et€al. Pediatric AML primary samples with FLT3/ITD mutations are preferentially killed by FLT3 inhibition. Blood. 2004;104:1841. Levis M, Pham R, Smith BD, Small D. In vitro studies of a FLT3 inhibitor combined with chemotherapy: sequence of administration is important to achieve synergistic cytotoxic effects. Blood. 2004;104:1145. Smith BD, Levis M, Beran M, et€al. Single-agent CEP-701, a novel FLT3 inhibitor, shows biologic and clinical activity in patients with relapsed or refractory acute myeloid leukemia. Blood. 2004;103:3669–3676. Stone RM, DeAngelo DJ, Klimek V, et€al. Patients with acute myeloid leukemia and an activating mutation in FLT3 respond to a small-molecule FLT3 tyrosine kinase inhibitor, PKC412. Blood. 2005;105:54–60. Brown P, Levis M, Shurtleff S, Campana D, Downing J, Small D. FLT3 inhibition selectively kills childhood acute lymphoblastic leukemia cells with high levels of FLT3 expression. Blood. 2005;105:812–820.
Molecularly Targeted Therapy for Infant ALL
53
Meshinchi S, Stirewalt DL, Alonzo TA, et€al. Activating mutations of RTK/RAS signal transduction pathway in pediatric acute myeloid leukemia. Blood. 2003;102:1474–1479. Brown P, Levis M, McIntyre E, Griesemer M, Small D. Combinations of the FLT3 inhibitor CEP701 and chemotherapy synergistically kill infant and childhood MLL-rearranged ALL cells in a sequence-dependent manner. Leukemia: official journal of the Leukemia Society of America, Leukemia Research Fund, UK. 2006;20:1368. Piloto O, Nguyen B, Huso D, et€ al. IMC-EB10, an anti-FLT3 monoclonal antibody, prolongs survival and reduces nonobese diabetic/severe combined immunodeficient engraftment of some acute lymphoblastic leukemia cell lines and primary leukemic samples. Cancer research. 2006;66:4843. Piloto O, Levis M, Huso D, et€al. Inhibitory anti-FLT3 antibodies are capable of mediating antibodydependent cell-mediated cytotoxicity and reducing engraftment of acute myelogenous leukemia blasts in nonobese diabetic/severe combined immunodeficient mice. Cancer Res. 2005;65:1514–1522. Levis M, Brown P, Smith BD, et€ al. Plasma inhibitory activity (PIA): a pharmacodynamic assay reveals insights into the basis for cytotoxic response to FLT3 inhibitors. Blood. 2006;108:3477–3483. Levis M, Smith BD, Beran M, et€al. A randomized, open-label study of lestaurtinib (CEP-701), an oral FLT3 inhibitor, administered in sequence with chemotherapy in patients with relapsed AML harboring FLT3 activating mutations: clinical response correlates with successful FLT3 inhibition. ASH Annual Meeting Abstracts. 2005;106:403. Robinson BW, Behling KC, Gupta M, et€al. Abundant anti-apoptotic BCL-2 is a molecular target in leukaemias with t(4;11) translocation. Br J Haematol. 2008;141:827–839. Zhang AY, Robinson BW, Kao K, et€al. Cell death regulatory gene expression correlates with MLL rearrangement status and prognostic clinical covariates in acute leukemia in infants. ASH Annual Meeting Abstracts. 2008;112:2255. Green D. Apoptotic pathways: ten minutes to dead. Cell. 2005;121:671–674. Reed JC. Apoptosis-based therapies. Nat Rev Drug Discov. 2002;1:111–121. Reed JC. Apoptosis-targeted therapies for cancer. Cancer Cell. 2003;3:17–22. Danial NN, Korsmeyer SJ. Cell death: critical control points. Cell. 2004;116:205–219. Nguyen M, Marcellus RC, Roulston A, et€al. Small molecule obatoclax (GX15-070) antagonizes MCL-1 and overcomes MCL-1-mediated resistance to apoptosis. Proc Natl Acad Sci U S A. 2007;104:19512–19517. Gewirtz AM, Sokol DL, Ratajczak MZ. Nucleic acid therapeutics: state of the art and future prospects. Blood. 1998;92:712–736. Letai A, Bassik MC, Walensky LD, Sorcinelli MD, Weiler S, Korsmeyer SJ. Distinct BH3 domains either sensitize or activate mitochondrial apoptosis, serving as prototype cancer therapeutics. Cancer Cell. 2002;2:183–192. Garber K. Targeting mitochondria emerges as therapeutic strategy. J Natl Cancer Inst. 2005;97:1800–1801. Nicholson DW. From bench to clinic with apoptosis based therapeutic agents. Nature. 2000;407:810–816. Marcucci G, Byrd JC, Dai G, et€ al. Phase 1 and pharmacodynamic studies of G3139, a Bcl-2 antisense oligonucleotide, in combination with chemotherapy in refractory or relapsed acute leukemia. Blood. 2003;101:425–432. O’Brien SM, Cunningham CC, Golenkov AK, Turkina AG, Novick SC, Rai KR. Phase I to II multicenter study of oblimersen sodium, a Bcl-2 antisense oligonucleotide, in patients with advanced chronic lymphocytic leukemia. J Clin Oncol. 2005;23:7697–7702. Waters JS, Webb A, Cunningham D, et€al. Phase I clinical and pharmacokinetic study of bcl-2 antisense oligonucleotide therapy in patients with non-Hodgkins lymphoma. J Clin Oncol. 2000;18:1812–1823. Deng X, Gao F, May WS, Jr. Bcl2 retards G1/S cell cycle transition by regulating intracellular ROS. Blood. 2003;102:3179–3185.
54
P.A. Brown and C.A. Felix
Huang S, Okumura K, Sinicrope FA. BH3 mimetic obatoclax enhances TRAIL-mediated apoptosis in human pancreatic cancer cells. Clin Cancer Res. 2009;15:150–159. Williamson NR, Fineran PC, Gristwood T, Chawrai SR, Leeper FJ, Salmond GP. Anticancer and immunosuppressive properties of bacterial prodiginines. Future Microbiol. 2007;2: 605–618. Campas C, Cosialls AM, Barragan M, et€al. Bcl-2 inhibitors induce apoptosis in chronic lymphocytic leukemia cells. Exp Hematol. 2006;34:1663–1669. Galan P, Roue G, Villamor N, Campo E, Colomer D. The Small Molecule Pan-Bcl-2 Inhibitor GX15–070 Induces Apoptosis In Vitro in Mantle Cell Lymphoma (MCL) Cells and Exhibits a Synergistic Effect in Combination with the Proteasome Inhibitor Bortezomib. ASH Annual Meeting Abstracts. 2005;106:1490. Perez-Galan P, Roue G, Villamor N, Campo E, Colomer D. The BH3-mimetic GX15-070 synergizes with bortezomib in mantle cell lymphoma by enhancing Noxa-mediated activation of Bak. Blood. 2007;109:4441–4449. Li J, Viallet J, Haura EB. A small molecule pan-Bcl-2 family inhibitor, GX15-070, induces apoptosis and enhances cisplatin-induced apoptosis in non-small cell lung cancer cells. Cancer Chemother Pharmacol. 2007. Li J, Viallet J, Haura EB. A small molecule pan-Bcl-2 family inhibitor, GX15-070, induces apoptosis and enhances cisplatin-induced apoptosis in non-small cell lung cancer cells. Cancer Chemother Pharmacol. 2008;61:525–534. Martinez-Paniagua MA, Vega MI, Huerta-Yepez S, et€al. Direct and enhanced cytotoxicity of the Bcl-2 family inhibitor GX15-070 on rituximab-sensitive and rituximab-resistant B-NHL clones. ASH Annual Meeting Abstracts. 2007. Trudel S, Li ZH, Rauw J, Tiedemann RE, Wen XY, Stewart AK. Preclinical studies of the pan-Bcl inhibitor obatoclax (GX015-070) in multiple myeloma. Blood. 2007;109:5430–5438. Witters LM, Witkoski A, Planas-Silva MD, Berger M, Viallet J, Lipton A. Synergistic inhibition of breast cancer cell lines with a dual inhibitor of EGFR-HER-2/neu and a Bcl-2 inhibitor. Oncol Rep. 2007;17:465–469. Konopleva M, Watt J, Contractor R, et€al. Mechanisms of antileukemic activity of the novel Bcl-2 homology domain-3 mimetic GX15-070 (obatoclax). Cancer Res. 2008;68:3413–3420. Bebb G, Muzik H, Nguyen S, Morris D, Stewart DA. In Vitro and In Vivo Anti Lymphoma Effect of GX15-070 in Mantle Cell Lymphoma. Blood (ASH Annual Meeting Abstracts). 2006;108:4756. Yazbeck VY, Georgakis GV, Li Y, McConkey D, Andreeff M, Younes A. Inhibition of the Pan-Bcl-2 Family by the Small Molecule GX15-070 Induces Apoptosis in Mantle Cell Lymphoma (MCL) Cells and Enhances the Activity of Two Proteasome Inhibitors (NPI-0052 and Bortezomib), and Doxorubicin Chemotherapy. Blood (ASH Annual Meeting Abstracts). 2006;108:2532. Hernandez-Ilizaliturri FJ, Iqbal A, Alam N, et€al. Targeting BH3-Domain Anti-Apoptotic Proteins with GX15-070 Decreases DNA Synthesis, Induces Cell Death and Sensitizes RituximabSensitive and Resistant Non-Hodgkins Lymphoma Cell Lines to the Anti-Tumor Activity of Chemotherapy Agents. Blood (ASH Annual Meeting Abstracts). 2006;108:2523. Hernandez-Ilizaliturri FJ, Bhat S, Iqbal A, Olejniczak S, Knight J, Czuczman MS. Targeting BH3Domain Anti-Apoptotic Proteins with GX15-070 Significantly Increases Rituximab-Mediated Antibody Dependent Cellular Cytotoxicity (ADCC) and Complement Mediated Cytotoxicity (CMC) Against B-Cell Lymphomas. Blood (ASH Annual Meeting Abstracts). 2006;108:2502. Borthakur G, O’Brien S, Ravandi-Kashani F, et€al. A Phase I Trial of the Small Molecule PanBcl-2 Family Inhibitor Obatoclax Mesylate (GX15-070) Administered by 24 Hour Infusion Every 2 Weeks to Patients with Myeloid Malignancies and Chronic Lymphocytic Leukemia (CLL). ASH Annual Meeting Abstracts. 2006;108:2654. Goy A, Ford P, Feldman T, et€al. A phase 1 trial of the pan Bcl-2 family inhibitor obatoclax mesylate (GX15-070) in combination with bortezomib in patients with relapsed/refractory mantle cell lymphoma. ASH Annual Meeting Abstracts. 2007. O’Brien SM, Claxton DF, Crump M, et€al. Phase I study of obatoclax mesylate (GX15-070), a small molecule pan-Bcl-2 family antagonist, in patients with advanced chronic lymphocytic leukemia. Blood. 2009;113:299–305.
Molecularly Targeted Therapy for Infant ALL
55
Schimmer AD, Brandwein J, O’Brien SM, et€al. A phase I trial of the small molecule pan-Bcl-2 family inhibitor obatoclax mesylate (GX15-070) administered by continuous Infusion for up to four days to patients with hematological malignancies. ASH Annual Meeting Abstracts. 2007. Verstovsek S, Raza A, Schimmer AD, Viallet J, Kantarjian H. A Phase II trial of the small molecule Pan-Bcl-2 family inhibitor obatoclax mesylate (GX15-070) administered by a 24-h continuous infusion every 2 weeks to patients with chronic idiopathic myelofibrosis (CIMF). ASH Annual Meeting Abstracts. 2007. Goy A, Ford P, Feldman T, et€al. A Phase 1 Trial of the Pan Bcl-2 Family Inhibitor Obatoclax Mesylate (GX15-070) in Combination with Bortezomib in Patients with Relapsed/Refractory Mantle Cell Lymphoma. ASH Annual Meeting Abstracts. 2007;110:2569. Schimmer AD, Brandwein J, O’Brien SM, et€al. A Phase I Trial of the Small Molecule Pan-Bcl-2 Family Inhibitor Obatoclax Mesylate (GX15-070) Administered by Continuous Infusion for up to Four Days to Patients with Hematological Malignancies. ASH Annual Meeting Abstracts. 2007;110:892. Stubbs MC, Faber J, Kung AL, Cameron S, Armstrong SA. HOXA9 represses bim expression in MLL rearranged leukemia: implications for drug therapy. ASH Annual Meeting Abstracts. 2007. Thomas M, Gessner A, Vornlocher HP, Hadwiger P, Greil J, Heidenreich O. Targeting MLL-AF4 with short interfering RNAs inhibits clonogenicity and engraftment of t(4;11)-positive human leukemic cells. Blood. 2005;106:3559–3566. Goldsmith KC, Liu X, Dam V, et€al. BH3 peptidomimetics potently activate apoptosis and demonstrate single agent efficacy in neuroblastoma. Oncogene. 2006;25:4525–4533. Xia SJ, Pressey JG, Barr FG. Molecular pathogenesis of rhabdomyosarcoma. Cancer Biol Ther. 2002;1:97–104. Bhojwani D, Kang H, Menezes RX, et€al. Gene expression signatures predictive of early response and outcome in high-risk childhood acute lymphoblastic leukemia: A Children’s Oncology Group Study [corrected]. J Clin Oncol. 2008;26:4376–4384. Flotho C, Coustan-Smith E, Pei D, et€al. A set of genes that regulate cell proliferation predicts treatment outcome in childhood acute lymphoblastic leukemia. Blood. 2007;110:1271–1277. Holleman A, den Boer ML, de Menezes RX, et€al. The expression of 70 apoptosis genes in relation to lineage, genetic subtype, cellular drug resistance, and outcome in childhood acute lymphoblastic leukemia. Blood. 2006;107:769–776. Bonapace L, Bornhauser BC, Cario G, et€al. The BH3-Mimetic Obatoclax Restores the Response to Dexamethasone in Glucocorticoid-Resistant ALL through Induction of Autophagy. ASH Annual Meeting Abstracts. 2007;110:806. van Delft MF, Wei AH, Mason KD, et€ al. The BH3 mimetic ABT-737 targets selective Bcl-2 proteins and efficiently induces apoptosis via Bak/Bax if Mcl-1 is neutralized. Cancer Cell. 2006;10:389–399. Zhang AY, Robinson BW, Wang L-S, et€ al. Pan-Anti-Apoptotic BCL-2 Family Inhibitor, Obatoclax, Activates Autophagic Cell Death Pathway and Has Potent Cytotoxicity in Infant and Pediatric MLL-Rearranged Leukemias. ASH Annual Meeting Abstracts. 2008;112:2647. Klionsky DJ, Abeliovich H, Agostinis P, et€al. Guidelines for the use and interpretation of assays for monitoring autophagy in higher eukaryotes. Autophagy. 2008;4:151–175. Maiuri MC, Le Toumelin G, Criollo A, et€al. Functional and physical interaction between BclX(L) and a BH3-like domain in Beclin-1. Embo J. 2007;26:2527–2539. Zhang AY, Barrett JS, Danet-Desnoyers G, et€ al. Obatoclax Biodistribution in MLL leukemia NOG Mouse Model is predicted by modeling and simulation and shows high tissue penetration at clinically important sites. J Clin Pharmacol. 2008;48:1130. Zhang AY, Barrett JS, Beauparlant P, et€ al. Modeling and Simulation Approach to Advance Molecularly-Targeted Pro-Apoptotic Agents for Infant Leukemias. J Clin Pharmacol. 2007;47:1209. Kang MH, Kang YH, Szymanska B, et€ al. Activity of vincristine, L-ASP, and dexamethasone against acute lymphoblastic leukemia is enhanced by the BH3-mimetic ABT-737 in€vitro and in€vivo. Blood. 2007.
56
P.A. Brown and C.A. Felix
Trudel S, Stewart AK, Li Z, et€al. The Bcl-2 family protein inhibitor, ABT-737, has substantial antimyeloma activity and shows synergistic effect with dexamethasone and melphalan. Clin Cancer Res. 2007;13:621–629. Chauhan D, Velankar M, Brahmandam M, et€al. A novel Bcl-2/Bcl-X(L)/Bcl-w inhibitor ABT737 as therapy in multiple myeloma. Oncogene. 2007;26:2374–2380. Konopleva M, Contractor R, Tsao T, et€al. Mechanisms of apoptosis sensitivity and resistance to the BH3 mimetic ABT-737 in acute myeloid leukemia. Cancer Cell. 2006;10:375–388. Kline MP, Rajkumar SV, Timm MM, et€al. ABT-737, an inhibitor of Bcl-2 family proteins, is a potent inducer of apoptosis in multiple myeloma cells. Leukemia. 2007;21:1549–1560. Kojima K, Konopleva M, Samudio IJ, Schober WD, Bornmann WG, Andreeff M. Concomitant inhibition of MDM2 and Bcl-2 protein function synergistically induce mitochondrial apoptosis in AML. Cell Cycle. 2006;5:2778–2786. Oltersdorf T, Elmore SW, Shoemaker AR, et€ al. An inhibitor of Bcl-2 family proteins induces regression of solid tumours. Nature. 2005;435:677–681. Armstrong SA, Kung AL, Mabon ME, et€al. Inhibition of FLT3 in MLL. Validation of a therapeutic target identified by gene expression based classification. Cancer Cell. 2003;3:173–183. Brown P, Levis M, McIntyre E, Griesemer M, Small D. Combinations of the FLT3 inhibitor CEP701 and chemotherapy synergistically kill infant and childhood MLL-rearranged ALL cells in a sequence-dependent manner. Leukemia. 2006;20:1368–1376. Brown P, Small D. FLT3 inhibitors: a paradigm for the development of targeted therapeutics for paediatric cancer. Eur J Cancer. 2004;40:707–721, discussion 722–704. Stam RW, den Boer ML, Schneider P, et€ al. Targeting FLT3 in primary MLL gene rearranged infant acute lymphoblastic leukemia. Blood. 2005. Stubbs MC, Armstrong SA. FLT3 as a therapeutic target in childhood acute leukemia. Curr Drug Targets. 2007;8:703–714. Rodila RC, Kim JC, Ji QC, El-Shourbagy TA. A high-throughput, fully automated liquid/liquid extraction liquid chromatography/mass spectrometry method for the quantitation of a new investigational drug ABT-869 and its metabolite A-849529 in human plasma samples. Rapid Commun Mass Spectrom. 2006;20:3067–3075. Carlson DM, Steinberg JL, Gordon G. Targeting the unmet medical need: the Abbott Laboratories oncology approach. Clin Adv Hematol Oncol. 2005;3:703–710. Dai Y, Hartandi K, Ji Z, et€al. Discovery of N-(4-(3-amino-1H-indazol-4-yl)phenyl)-N¢-(2-fluoro5-methylphenyl)urea (ABT-869), a 3-aminoindazole-based orally active multitargeted receptor tyrosine kinase inhibitor. J Med Chem. 2007;50:1584–1597. Shankar DB, Li J, Tapang P, et€al. ABT-869, a multitargeted receptor tyrosine kinase inhibitor: inhibition of FLT3 phosphorylation and signaling in acute myeloid leukemia. Blood. 2007;109: 3400–3408. Albert DH, Tapang P, Magoc TJ, et€al. Preclinical activity of ABT-869, a multitargeted receptor tyrosine kinase inhibitor. Mol Cancer Ther. 2006;5:995–1006. Piloto O, Wright M, Brown P, Kim KT, Levis M, Small D. Prolonged exposure to FLT3 inhibitors leads to resistance via activation of parallel signaling pathways. Blood. 2007;109:1643–1652. Kohl TM, Hellinger C, Ahmed F, et€ al. BH3 mimetic ABT-737 neutralizes resistance to FLT3 inhibitor treatment mediated by FLT3-independent expression of BCL2 in primary AML blasts. Leukemia. 2007. Kawagoe H, Kawagoe R, Sano K. Targeted down-regulation of MLL-AF9 with antisense oligodeoxyribonucleotide reduces the expression of the HOXA7 and -A10 genes and induces apoptosis in a human leukemia cell line, THP-1. Leukemia. 2001;15:1743–1749. Niitsu N, Hayashi Y, Honma Y. Downregulation of MLL-CBP fusion gene expression is associated with differentiation of SN-1 cells with t(11;16)(q23;p13). Oncogene. 2001;20:375–384. Akao Y, Mizoguchi H, Misiura K, et€al. Antisense oligodeoxyribonucleotide against the MLL-LTG19 chimeric transcript inhibits cell growth and induces apoptosis in cells of an infantile leukemia cell line carrying the t(11;19) chromosomal translocation. Cancer Res. 1998;58:3773–3776. Bitoun E, Oliver PL, Davies KE. The mixed-lineage leukemia fusion partner AF4 stimulates RNA polymerase II transcriptional elongation and mediates coordinated chromatin remodeling. Hum Mol Genet. 2007;16:92–106.
Molecularly Targeted Therapy for Infant ALL
57
Srinivasan RS, Nesbit JB, Marrero L, Erfurth F, LaRussa VF, Hemenway CS. The synthetic peptide PFWT disrupts AF4-AF9 protein complexes and induces apoptosis in t(4;11) leukemia cells. Leukemia. 2004;18:1364–1372. Erfurth F, Hemenway CS, de Erkenez AC, Domer PH. MLL fusion partners AF4 and AF9 interact at subnuclear foci. Leukemia. 2004;18:92–102. Palermo CM, Bennett CA, Winters AC, Hemenway CS. The AF4-mimetic peptide, PFWT, induces necrotic cell death in MV4-11 leukemia cells. Leuk Res. 2008;32:633–642. Bennett CA, Winters AC, Barretto NN, Hemenway CS. Molecular targeting of MLL-rearranged leukemia cell lines with the synthetic peptide PFWT synergistically enhances the cytotoxic effect of established chemotherapeutic agents. Leuk Res. 2009;33:937–947. Wang Z, Smith KS, Murphy M, Piloto O, Somervaille TC, Cleary ML. Glycogen synthase kinase 3 in MLL leukaemia maintenance and targeted therapy. Nature. 2008;455:1205–1209. Wiech H, Buchner J, Zimmermann R, Jakob U. Hsp90 chaperones protein folding in€vitro. Nature. 1992;358:169–170. Stebbins CE, Russo AA, Schneider C, Rosen N, Hartl FU, Pavletich NP. Crystal structure of an Hsp90-geldanamycin complex: targeting of a protein chaperone by an antitumor agent. Cell. 1997;89:239–250. Terasawa K, Minami M, Minami Y. Constantly updated knowledge of Hsp90. J Biochem. 2005;137:443–447. Imai J, Maruya M, Yashiroda H, Yahara I, Tanaka K. The molecular chaperone Hsp90 plays a role in the assembly and maintenance of the 26S proteasome. EMBO J. 2003;22:3557–3567. Yao Q, Nishiuchi R, Li Q, Kumar AR, Hudson WA, Kersey JH. FLT3 expressing leukemias are selectively sensitive to inhibitors of the molecular chaperone heat shock protein 90 through destabilization of signal transduction-associated kinases. Clin Cancer Res. 2003;9:4483–4493. Belova L, Brickley DR, Ky B, Sharma SK, Conzen SD. Hsp90 regulates the phosphorylation and activity of serum- and glucocorticoid-regulated kinase-1. J Biol Chem. 2008;283:18821–18831. Fujita N, Sato S, Ishida A, Tsuruo T. Involvement of Hsp90 in signaling and stability of 3-phosphoinositide-dependent kinase-1. J Biol Chem. 2002;277:10346–10353. Zhang R, Luo D, Miao R, et€al. Hsp90-Akt phosphorylates ASK1 and inhibits ASK1-mediated apoptosis. Oncogene. 2005;24:3954–3963. Picard D. Hsp90 invades the outside. Nat Cell Biol. 2004;6:479–480. Eustace BK, Jay DG. Extracellular roles for the molecular chaperone, hsp90. Cell Cycle. 2004;3:1098–1100. Eustace BK, Sakurai T, Stewart JK, et€al. Functional proteomic screens reveal an essential extracellular role for hsp90 alpha in cancer cell invasiveness. Nat Cell Biol. 2004;6:507–514. Yocum AK, Busch CM, Felix CA, Blair IA. Proteomics-based strategy to identify biomarkers and pharmacological targets in leukemias with t(4;11) translocations. J Proteome Res. 2006;5:2743–2753. Yao Q, Weigel B, Kersey J. Synergism between etoposide and 17-AAG in leukemia cells: critical roles for Hsp90, FLT3, topoisomerase II, Chk1, and Rad51. Clin Cancer Res. 2007;13:1591–1600. Yao Q, Nishiuchi R, Kitamura T, Kersey JH. Human leukemias with mutated FLT3 kinase are synergistically sensitive to FLT3 and Hsp90 inhibitors: the key role of the STAT5 signal transduction pathway. Leukemia. 2005;19:1605–1612. Tonelli R, Sartini R, Fronza R, et€ al. G1 cell-cycle arrest and apoptosis by histone deacetylase inhibition in MLL-AF9 acute myeloid leukemia cells is p21 dependent and MLL-AF9 independent. Leukemia. 2006;20:1307–1310. Niitsu N, Hayashi Y, Sugita K, Honma Y. Sensitization by 5-aza-2¢-deoxycytidine of leukaemia cells with MLL abnormalities to induction of differentiation by all-trans retinoic acid and 1alpha,25-dihydroxyvitamin D3. Br J Haematol. 2001;112:315–326. Hudes G, Carducci M, Tomczak P, et€al. Temsirolimus, interferon alfa, or both for advanced renalcell carcinoma. N Engl J Med. 2007;356:2271–2281. Brown VI, Fang J, Alcorn K, et€al. Rapamycin is active against B-precursor leukemia in€vitro and in€vivo, an effect that is modulated by IL-7-mediated signaling. Proc Natl Acad Sci U S A. 2003;100:15113–15118. Teachey DT, Sheen C, Hall J, et€al. mTOR inhibitors are synergistic with methotrexate: an effective combination to treat acute lymphoblastic leukemia. Blood. 2008;112:2020–2023.
58
P.A. Brown and C.A. Felix
Wei G, Twomey D, Lamb J, et€al. Gene expression-based chemical genomics identifies rapamycin as a modulator of MCL1 and glucocorticoid resistance. Cancer Cell. 2006;10:331–342. Pieters R, den Boer ML, Durian M, et€al. Relation between age, immunophenotype and in€vitro drug resistance in 395 children with acute lymphoblastic leukemia--implications for treatment of infants. Leukemia. 1998;12:1344–1348. Palle J, Frost BM, Forestier E, et€al. Cellular drug sensitivity in MLL-rearranged childhood acute leukaemia is correlated to partner genes and cell lineage. Br J Haematol. 2005;129:189–198. Zwaan CM, Reinhardt D, Corbacioglu S, et€al. Gemtuzumab ozogamicin: first clinical experiences in children with relapsed/refractory acute myeloid leukemia treated on compassionate-use basis. Blood. 2003;101:3868–3871. Arceci RJ, Sande J, Lange B, et€al. Safety and efficacy of gemtuzumab ozogamicin in pediatric patients with advanced CD33+ acute myeloid leukemia. Blood. 2005;106:1183–1188.
Targeted Therapeutic Approaches for AML Robert J. Arceci and Donald Small
Introduction Acute myeloid leukemia comprises about 20% of the acute leukemias in children, but it is responsible for more than half of leukemic deaths due to leukemia. Compared to the tremendous success in the treatment of acute lymphocytic leukemia in the last three decades, resulting in more than 80% cure rate, improvements in AML therapy have been more limited with only about half of patients with AML being cured. Risk-adapted therapy has been the cornerstone of ALL therapy. One of the reasons for the success of this approach in ALL is that standard ALL induction and consolidation have been able to be intensified without causing significant morbidity and mortality. In contrast, the leukemic stem cell in most AML subtypes is inherently more drug resistant requiring significantly intensified courses of near myeloablative combinations of chemotherapeutic agents. This has resulted in a plateau in survival at approximately 50% along with significant morbidity and mortality. Current AML therapy is based on the use of multi-agent combinations of �noncross-resistant chemotherapeutic agents, dose intensification, risk-adapted use of allogeneic HSCT as well as aggressive, pre-emptive use of supportive care interventions. Such therapeutic approaches are also relatively nonselective and associated� with significant treatment-related toxicities. Advances in the molecular basis for AML have provided improvements in subtype classification, a better understanding of risk stratification, and the introduction of molecularly targeted therapies. This increased the understanding of the molecular basis of AML in combination with a growing list of host genetic risk factors provides for the possibility of truly individualized therapy.
R.J. Arceci (*) The Sidney Kimmel Comprehensive Cancer Center at Johns Hopkins, Pediatric Oncology, Johns Hopkins University, Baltimore, MD 21231, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_4, © Springer Science+Business Media, LLC 2010
59
60
R.J. Arceci and D. Small
Mechanisms of Leukemogenesis: Implications for Targeted Therapy Development The explosion of new discoveries relating to our understanding of the cellular and molecular basis of AML in adults and children has significant implications for the development of less toxic and more effective therapies but also is associated with challenges in terms of validating potential therapeutic targets both in preclinical models and in clinical trials. The genetic changes that lead to AML can be both inherited as well as somatically acquired. Many of these molecular changes have been largely derived from abnormal chromosomal rearrangements that in turn have led to the identification of gene translocations and resulting fusion products that play roles in the development of AML. The genes targeted by these abnormal chromosome changes often include transcription factors that are important in regulating myeloid differentiation as well as partner genes that may relocate the DNA binding or regulatory region of the transcription factor to altered sites within the cell or along DNA. For example, although the cytogenetic abnormalities observed in AML are quite heterogeneous, they may result in the altered targeting of the same molecular pathways. For example, all the t(8;21), t(3;21), t(16;16), t(16;21), and inv(16) chromosomal changes observed in AML alter the function of core binding factor complexes that regulate chromatin structure and function. Another example includes translocations that involve PML and RAR alpha in APL. Such genetic abnormalities result in a change in the normal function of these core binding factors leading to transcriptional repression of target genes regulating proliferation, cell survival, and differentiation. Furthermore, the convergence of these altered differentiation regulatory pathways has important implications for therapeutic targeting. For example, instead of developing therapeutic inhibitors directed toward each fusion product, it may be possible to target the critical and shared transcriptional pathways that contribute to this step in AML development (Ichikawa et€al. 2006). However, preclinical murine models designed to test the contribution of important translocations in AML have determined that the presence of only the translocation fusion product is usually insufficient for leukemic transformation. This observation led to a search for cooperating genetic pathways, which, in turn, led to the recognition of the importance of activating mutations of tyrosine kinase receptors such as FLT3 and c-KIT as regulators of leukemia cell survival and proliferation. The presence of activating mutations of specific tyrosine kinase receptors has also been shown to contribute to chemotherapeutic resistance through their activation of downstream resistance pathways as well as increased proliferation and survival of leukemic blasts. Thus, the pathways affected by these mutated receptors and signal transduction molecules also represent additional therapeutic targets.
Targeted Therapeutic Approaches for AML
61
Refining the Definition and Relevance of Targeted Therapy The considerable overlap in the phenotypes of AML and normal hematopoietic precursor cells has presented significant challenges in terms of distinguishing them for targeted therapies. The development of fluorescence activated cell sorting in conjunction with monoclonal antibodies directed to lineage specific markers has provided an approach to isolate primitive self-repopulating leukemic stem cells apart from normal progenitors. The AML, self-repopulating primitive cells are quite rare within the population of leukemic blasts, with estimates of their �frequency being in the 0.2 to 200 per 106 mononuclear cells range. These data have led to the conclusion that in many instances, the subtype of AML observed clinically depends on the types of initiating and subsequent molecular events that occur in an early stem cell population of hematopoietic precursors rather than from transformation occurring in a later stage, lineage-determined, precursor cell (Jordan 2007; Barabe et€al. 2007; Hope et€al. 2004). The complex hierarchy of the self-initiating AML stem cell creates important challenges for targeted therapy development. Thus, for a pathway to represent a potentially curative target and not simply a cytoreductive one, the altered pathway should have a specific, or at least a highly selective, effect on the leukemia selfinitiating stem cell. Thus, a potentially curative target should (1) be expressed in the relevant cell population�, (2) survival of the relevant cell type should be dependent on the expression of the target, and (3) modulation of the targeted pathways should result in the demise or permanent inactivity of the relevant cell type. As noted above, the relevant cell type in AML should include the self-replicating leukemic stem cell. Clinical trials to assess the relevance of agents directed toward enumerating and eradicating this �self-repopulating AML in order to test the clinical relevance of these concepts are needed.
Development of Targeted Therapies for AML Although many therapies are particularly effective at cytoreducing leukemic cell burden in patients with AML, very few have been selectively developed to target the self-repopulating leukemic stem cell population. In order to make this type of targeting a reality, a significant amount of investigation needs to be done into what distinguishes AML stem cells from their normal counterparts. Little information has thus far been reported on this issue, although some reports are starting to reveal potentially important differences between leukemic and the normal stem cell counterparts (Andersson et€al. 2005a,b; Ross et€al. 2004; Oshima et€al. 2003). In addition, while most cytoreductive agents produce rapid antileukemic responses, agents targeted at the relatively rare but self-repopulating leukemic stem cell population would be expected to result in slow clinical responses. Most clinical trials do not �usually provide the opportunity to observe such slow responses, especially in
62
R.J. Arceci and D. Small
Cancer Self-Renewing Stem Cell
Non-Self-Replicating Cancer Cell
Replication, Proliferation and Differentiation
Lineage Specific Antigens Drug Transporter Inhibitors Cytokine Receptors Signal Transduction Pathways Chromatin/Transcription
Replication or Dormancy Immunotherapy
Proliferation and Differentiation Microenvironment
Fig.€1╅ Schematic of possible cellular and molecular targets in AML (adapted from Arceci and Cripe (2002))
patients with acute leukemia. Thus, if a targeted agent is not an effective cytoreductive agent, it may still provide important antileukemic effects, particularly when used in �vcombination with more traditional cytoreductive chemotherapies. Quantitation of the self-repopulating AML stem cells in such combination trials may be an approach to evaluate the effect of such stem-cell directed treatments. Several pathways and targets are being currently tested in pediatric AML that are directed alone or in combination toward reversing the abnormalities in leukemia cell survival/drug resistance, proliferation, and differentiation. Promising new approaches to more selective treatments include inhibition of proliferation and survival pathways such as FLT3-ITD and c-KIT receptors and their downstream targets, such as RAS, as well as transcriptional or chromatin-based strategies and immunotargeted therapies with monoclonal antibodies and vaccines (Fig.€1).
Specific Examples of Targeting AML Targeting Pathways that Alter Leukemia Cell Proliferation and Survival Tyrosine Kinases Laboratory studies demonstrating the importance of cytokines (survival and differentiation factors) and their receptors along with the induced downstream �signaling
Targeted Therapeutic Approaches for AML
63
pathways that direct normal hematopoiesis were fundamental in elucidating the role of these same pathways in the etiology and pathophysiology of AML. Several of these cytokine receptors, such as FLT3, c-KIT, c-FMS, PDGF-R, and VEGF-R, have been shown to be expressed at high levels or have activating mutations that lead to increased proliferative, survival, and drug resistance in AML (Kiyoi et€al. 1997, 1998; Meshinchi et€al. 2001, 2003a; Yamamoto et€al. 2001). These changes in leukemic cell behavior have furthermore been shown in several instances to changes in clinical outcome of patients. For example, pediatric and adult patients with FLT3/ITD mutations, particularly in the group of patients with normal karyotypes, have a poor outcome. Such observations have provided the rationale to target these mutant receptors and their downstream signaling pathways. Mutations effecting c-KIT occur more frequently in the subset of AML characterized by alterations in core binding factors and have been reported to portend a poor outcome in an otherwise favorable group of patients (Schnittger et€al. 2006, 2007; Shimada et€al. 2006). In both the examples of FLT3/ITD and c-KIT, the mutant receptors result in distinct changes in signal transduction compared to activation of the respective wild type receptors, leading to cytokine independent survival, increased proliferation, and greater resistance to chemotherapeutic drugs (Tse et€ al. 2000; Ning et€ al. 2001a,b,c). Another important observation regarding FLT3/ITD is that at least some cases of AML appear to have the mutation in a leukemia repopulating subset of cells as well as their progeny, thus in part addressing the issues raised above concerning the targeting of the leukemia initiating stem cell as well as the right molecular target (Levis et€al. 2005a; Pollard et€ al. 2006). However, in some patients with FLT3/ITD positive AML, relapsed samples lacked the FLT3/ITD, suggesting that in these patients, FLT3/ITD is more likely to be a secondary mutational event rather than arising in a leukemia selfrepopulating cell; an alternative explanation would be that more than one leukemia initiating cell was initially present and differentially selected for during treatment (Kottaridis et€al. 2002; Cloos et€al. 2006). Several inhibitors of FLT3/ITD, such as CEP-701 (Cephalon) (Brown et€al. 2006; Levis et€al. 2002, 2004; Smith et€al. 2004), MLN518 (Millenium) (DeAngelo et€al. 2006) and PKC412 (Novartis) (Cools et€ al. 2003; Stone et€ al. 2005), have been developed and tested in preclinical models and early clinical trials. One of these inhibitors, CEP-701 (lestaurtinib), showed excellent activity in a preclinical model (Levis et€al. 2002) followed by clinical trials as a single agent (Smith et€al. 2004; Levis et€ al. 2006). In these clinical trials, decreases in peripheral blood leukemic blasts were observed but no significant marrow responses leading to complete remissions. Such results have also been observed with other inhibitors of FLT3 (DeAngelo et€al. 2006; Stone et€al. 2005; Knapper et€al. 2006). However, an important aspect of the single agent clinical trials with CEP-701 was that methods were developed and validated which showed that sufficiently high plasma levels of the inhibitor could be obtained to inhibit the activated, mutant receptor (Levis et€al. 2006). Additional preclinical work demonstrated increased efficacy of combining CEP701 with cytosine arabinoside (Levis et€al. 2004) which subsequently led to encouraging early results in a randomized phase II trial in adults with relapsed AML expressing FLT3/ITD receptors (Levis et€al. 2005b). In the preliminary data reported
64
R.J. Arceci and D. Small
from this trial, 76% (13 of 17) of patients showed plasma levels of drug that inhibited FLT3 activity by at least 85%. Further, true clinical bone marrow responses were observed primarily in the group of patients whose AML was sensitive and who had sufficient, inhibitory plasma levels of CEP-701 (Levis et€al. 2005b). COG AAML06P1 in combination with chemotherapy (idarubicin and cytosine arabinoside) for pediatric patients with relapsed AML harboring FLT3/ITD mutations. In addition, the increased expression of wild type FLT3 in AML, particularly in infant leukemia, has led to a COG clinical trial combining CEP-701 with chemotherapy in this very high risk leukemia in infants (Brown et€al. 2006; Armstrong et€al. 2002). There are plans to test the addition of FLT3 inhibitors into the COG phase III trial for patients with newly diagnosed AML expressing the FLT3/ITD mutant receptor. The c-KIT tyrosine kinase receptor has also been associated with poor outcome in the subtype of otherwise good prognostic core binding factor AML (Schnittger et€al. 2006; Shimada et€al. 2006) although a recent report from COG was not able to confirm a worse prognosis associated with c-KIT mutations (Poland et al. 2009). Mutations of c-KIT involve single base pair changes leading to amino acid changes in the kinase domain and resulting in an auto-activating receptor. Less common in AML, some c-KIT mutations involve the cytoplasmic, juxtamembrane portion of the receptor; such mutations have different kinase inhibitor characteristics than the kinase domain mutations (Roberts et€al. 2007). c-KIT mutations are observed in only about 3 to 5% of AML, but in 20 to 30% of AML characterized by changes in core binding factors (Schnittger et€al. 2006; Shimada et€al. 2006; Meshinchi et€al. 2003b). Thus, kinase inhibitors with activity against such c-KIT mutations would be expected to be potentially useful in this subset of AML. Imatinib, first developed for the inhibition of the BCR-ABL fusion kinase and the treatment of CML, also has significant activity against c-KIT as well as the platelet derived growth factor receptor (PDGFR) (Heinrich et€al. 2000; Fernandez et€al. 2007). In vitro results have demonstrated that imatinib selectively inhibits wild type and mutant c-KIT activity as well as its activation of downstream signaling molecules, such as MAPK and AKT (Heinrich et€al. 2000). An initial phase II study of imatinib in patients with refractory or recurrent AML showed disappointing results with no significant clinical responses (Kindler et€al. 2004). However, a subsequent clinical trial using imatinib in 21 patients with refractory AML expressing c-KIT demonstrated 5 out of 21 patients with hematologic responses, two of which were complete hematologic remissions in patients who had relatively low blast counts in the bone marrow or in peripheral blood (Ito et€al. 2005; Jentsch-Ullrich et€al. 2004; Pompetti et€al. 2007). Imatinib and low dose cytosine arabinoside has been tested in older adults in a phase II trial based on in€vitro data that showed synergistic cytotoxicity with the combination compared to either agent used alone (Heidel et€al. 2007). However, the clinical response rate was only 11% and did not appear significantly better than AraC alone based on comparison with historical controls (Heidel et€al. 2007). There are several potential problems with such studies. First is that imatinib may not be a particularly potent or specific inhibitor of mutant c-KIT receptors; second, the inhibition of the c-KIT receptor
Targeted Therapeutic Approaches for AML
65
kinase activity and/or downstream signaling pathways were not assessed in leukemic blasts from patients, thus raising the possibility that the kinase was not sufficiently inhibited. Although c-KIT remains a potentially important therapeutic target in a subset of patients with AML, more selective and potent inhibitors need to develop along with synergistic combinations conventional chemotherapy or other targeting agents. Although their name suggests the unlikely possibility that the IGF-I and VEGF receptors would play a role in AML, there are data to the contrary. IGF-IR expression has been demonstrated on AML cells although activating mutations have not been reported (Hizuka et€al. 1987). In addition, IGF-I has been shown to contribute along with other cytokines to the survival and proliferation of AML cells as well as their response to AraC (Abe et€al. 2006; Frostad and Bruserud 1999). In vitro studies have further shown that blockade of the IGF-IR was able to induce apoptosis in AraC resistant AML cells (Abe et€al. 2006). As IGF-IR inhibitors, whether small molecules or monoclonal antibodies, are further studied in the laboratory and in clinical trials (Cosaceanu et€al. 2007; Jerome et€al. 2006), a potential role of these approaches alone or in combination with other targeted therapies, may prove worthwhile in AML. Several lines of evidence have established a role for VEGF and its receptor in hematopoiesis and in AML. Laboratory studies have demonstrated that a high percentage of AML cells express VEGF as well as VEGFR (Perl and Carroll 2007). Furthermore, the secretion of VEGF by AML blasts can stimulate bone marrow stromal fibroblasts to secrete hematopoietic cytokines such as G-CSF and GM-CSF, which, in turn, may contribute to further myeloid progenitor expansion (Aguayo 2004; Aguayo et€al. 1999, 2003; Hussong et€al. 2000; Kessler et€al. 2007; Schuch et€ al. 2002). In addition, increased bone marrow angiogenesis, observed in both MDS and AML, may be in part a result of VEGF secretion. Thus, VEGF may play both an autocrine and paracrine role in MDS and AML. Leukemic cell proliferation, survival, and resistance to chemotherapy have also been directly linked to signaling through VEGF-C and its activation of VEGFR-3 (Dias et€ al. 2002; Liersch et€ al. 2007). Inhibitory monoclonal antibodies that bind and inactivate VEGF (e.g., Avastin or Bevacizumab) (Karp et€al. 2004), or small molecule inhibitors of VEGFR (e.g., SU5416 or AZD2171) (Fiedler et€al. 2003; Maris et€al. 2008), have been developed for therapeutic anti-angiogenic treatment approaches. Karp et€al. (2004) tested Bevacizumab post chemotherapy treatment in adults with AML and showed clearance of blasts and decreased bone marrow microvasculature. Bevacizumab is also being tested in adults with newly diagnosed AML in combination with idarubicin and AraC. Initial clinical studies with SU5416 reported a complete remission in a patient with AML, although subsequent studies have not shown significant response rates to this agent when used alone (Perl and Carroll 2007; Fiedler et€al. 2003). It is possible that the selective preference of SU5416 may be too narrow or mechanisms of resistance are activated to prevent significant leukemic cell killing. Current attempts to combine such agents with chemotherapy or to use tyrosine kinase inhibitors with a broader spectrum of targets are being developed (Perl and Carroll 2007).
66
R.J. Arceci and D. Small
The above examples demonstrate several key principles for targeted therapy in AML using TK inhibitors. First, unlike in CML, such approaches using single agents are unlikely to be highly effective in AML. However, these TK inhibitors appear to share the characteristic of sensitizing AML cells to conventional chemotherapeutic drugs, without necessarily resulting in significant overlapping toxicities. Nevertheless, because some of the activated TKs signal through common, downstream pathways, investigators are attempting to determine whether inhibition of these shared, convergence points can be therapeutically exploited (Fig.€2). However, similar to imatinib, resistance to TK inhibitors has been demonstrated, although the pathways leading to resistance may be different (Piloto et€al. 2007).
Inhibition of Signaling Pathways Downstream of TK Receptors Not long after its initial identification as an important dominant oncogene in malignant transformation, RAS mutations were documented in about 25% of AML/MDS samples (Farr et€al. 1991). Further, in the absence of RAS mutations, RAS activation through its conversion from GDP to GTP and farnesylated tethering to the inner cell membrane is known to play a critical role in the development of AML as well as the myeloproliferative syndrome, juvenile myelomonocytic leukemia (JMML) (Loh et€al. 2004). While specific inhibitors of GTP activation of RAS have not yet been successfully reported, farnesyl transferase inhibitors (FTI) have been tested in patients with AML (Karp 2001, 2005; Emanuel et€al. 2000; Gotlib 2005). Phase I studies in adults with AML reported significant clinical responses that did not appear to be dose dependent (Karp 2001). A subsequent multi-institutional phase II study reported a 5% complete remission rate using tipifarnib as a single agent in adult patients with relapsed and/or refractory disease (Harousseau et€ al. 2007). Tipifarnib as a single agent in patients aged 65 or older with high risk AML showed a response rate of 10 to 20% (Lancet et€al. 2007). Based on such results, a randomized phase III trial comparing tipifarnib to supportive care in patients aged 70 years or more with AML is being conducted in Europe. A pediatric randomized trial of tipifarnib versus no drug in the postallogeneic HSCT setting by the COG in patients with relapsed AML. While tipifarnib has shown a low, but real level of activity in AML as a single agent, future use and testing will likely be in combination with other targeted agents and/or conventional chemotherapy. Of note, however, is that the mechanism of action of such FTI in AML and whether they are mediating their effect through the RAS pathway remain unclear. Downstream of RAS are the RAF, MEK, ERK, and STAT pathways, all of which are being considered as potential therapeutic targets in AML (Fig.€2). For example, small molecule inhibitors of the MAPK pathway, such as PD98059 and PD184352, show significant and selective induction of cell cycle arrest and apoptosis in AML cells (Milella et€al. 2001, 2002, 2003). MAPK inhibition can also result in increased sensitization to chemotherapy of AML cells, particularly from AML expressing the
Fig.€2╅ Schematic of FLT3 signaling pathways as an example potential downstream molecular targets as a result of activating mutations of tyrosine kinases. From Doepfner et€al. (2007)
Targeted Therapeutic Approaches for AML 67
68
R.J. Arceci and D. Small
stem cell marker, CD34 (James et€ al. 2003). RAF-1 kinase inhibitors, such as sorafenib, are also being clinically developed. Of note, sorafenib also has significant activity against c-KIT (Tong et€al. 2006). The mTOR (mammalian Target Of Rapamycin) serine/threonine kinase plays a primary role as a regulator of cell cycle progression, proliferation, apoptosis, and angiogenesis in both normal and neoplastic cells. mTOR functions as a type of �utility player of signal transduction and in that it is a central regulator of a variety of signaling pathways originating from TK receptor activation, including FLT3, c-KIT, PDGFR, c-FMS, and VEGF, as well as through RAS, various phosphatases, such as PTEN and SHP-1/2 (Weisberg et€al. 2008). Coactivation of the PI3K/AKT pathway is also usually observed with mTOR activation (Tamburini et€al. 2008; Xu et€al. 2003, 2005). In each of these examples, a role of mTOR has been demonstrated in the proliferation, survival, and drug resistant phenotype of approximately 70% of AML cases examined (Xu et€al. 2003; Recher et€al. 2005; Min et€al. 2003). Several highly selective inhibitors of mTOR, including rapamycin, CCI-779, RAD001, AP23573, and MK-8669 have been developed. A clinical trial in adults with refractory/relapsed or secondary AML reported that four of nine patients had partial responses to rapamycin when used as a single agent (Recher et€al. 2005). A phase I/II study using RAD001 (everolimus) as a single agent in adults with refractory and/or relapsed hematologic malignancies showed that a 10€mg daily dose was well tolerated and two of five patients with MDS had a major and minor platelet responses. No patients with AML were included on this trial (Yee et€ al. 2006). Because of its central role and activation in a majority of patients with AML, mTOR inhibition as a therapeutic approach, is being planned for clinical trial testing in pediatric patients with relapsed/ refractory AML. Combining mTOR inhibitors with conventional chemotherapy, particularly ARAC containing regimens, or with other signal transduction inhibitors will likely be the best approach for definitive clinical testing.
Drug Resistance Mechanisms and Leukemic Cell Survival As discussed above, in addition to conveying a proliferative advantage, mutant or highly expressed tyrosine kinase receptors and their downstream signaling pathways can also lead to increased drug resistance on the part of AML blasts. Other molecular pathways may also lead to de€novo or acquired drug resistance as well. The identification of the drug efflux transporter P-glycoprotein (Pgp 170 or MDR1 P-glycoprotein) as an important mechanism of multidrug resistance in cancer cells provided a potential therapeutic target. To this end, a variety of competitive inhibitors, including cyclosporine A and its nonimmunosuppressive D isomer, PSC833, as well as a number of other small molecules, were developed to block the drug efflux capacity of this ATP-dependent transporter. Inhibition of the MDR1 transporter results in decreased efflux of multiple chemotherapeutic, natural product drugs, such as anthracyclines and vinca alkaloids, leading to their increased intracellular concentration and subsequent cell death. The observations that many tumor types that are resistant to
Targeted Therapeutic Approaches for AML
69
such chemotherapeutic agents also express high levels of the MDR1 P-glycoprotein and could be resensitized to chemotherapy in the presence of an MDR1 inhibitor provided the experimental foundation to clinically test this approach. Of note, the expression of MDR1 P-glycoprotein was found to be high in adult AML and, furthermore, showed even higher expression at the time of relapse (Baer et€al. 2002; Leith et€al. 1995, 1999; Willman 1996). Such high levels of MDR1 P-glycoprotein were correlated with poor outcome (Willman 1997). Expression in childhood AML was reported to be considerably lower (Sievers et€al. 1995). After the completion of phase I/II trials with MDR1 P-glycoprotein inhibitors, it became clear that significant dose adjustments of the chemotherapeutic drugs needed to be made to correct for decreased drug elimination secondary to inhibition of cellular efflux in normal cells of the liver, kidney, and intestine (List et€al. 1993; Dahl et€ al. 2000). Most randomized phase III trials in adults with AML tested the impact on outcome of inclusion of an MDR1 inhibitor, e.g., PSC833, in combination with chemotherapy, did not show an improvement compared to chemotherapy along (Greenberg et€al. 2004). One exception was a randomized comparison of cyclosporine A and no cyclosporine in combination with daunorubicin and ARAC during consolidation for adults with poor-risk AML (List et€al. 2001). Although the rate of complete remission was not different in the two groups of patients (39 vs. 33%, p╛=╛0.14), relapse free survival and overall survival were significantly improved in the group receiving cyclosporine (RFS 34 vs. 9% at 2 years, p╛=╛0.031 and OS 22 vs. 12%, p╛=╛0.46, respectively). Currently, the Eastern Cooperative Oncology group is testing the nonpharmacologically active inhibitor, zosuquidar and the Southwest Oncology Group is testing cyclosporine A in randomized, phase III trials. Phase I and II studies in pediatric patients with relapsed/refractory AML established the dose of cyclosporine A and chemotherapy when used in combination (Dahl et€al. 2000). Based on these results, the POG 9421 study randomized patients to receive in consolidation cyclosporine or no cyclosporine in combination with mitoxantrone and ARAC during the consolidation phase of treatment (Becton et€al. 2006). In addition, this study randomized high dose to standard dose ARAC during induction therapy. The results showed that the best outcome was associated receiving high dose ARAC during induction and cyclosporine plus mitoxantrone and ARAC during consolidation. However, no statistically significant differences in outcome were observed for either the high dose ARAC or the addition of cyclosporine when considered as independent variables (Becton et€al. 2006). The lack of a definitive benefit for MDR1 P-glycoprotein inhibition may be related to the emergence of other mechanisms of resistance in the presence of MDR1 inhibition, including the possible role of one or more of the dozens of other transport pumps expressed in cells, increased expression of the MDR1 P-glycoprotein as well as ineffective blockade and increased toxicity. While further MDR transporter inhibition trials in pediatrics are unlikely in the near future, one exception might be in conjunction with anti-CD33-calicheamicin immunotargeted therapy because a known mechanism of resistance to calicheamicin is MDR1 P-glycoprotein efflux (Arceci et€al. 2005; Linenberger et€al. 2001).
70
R.J. Arceci and D. Small
A second pathway to chemotherapy-induced resistance is the upregulation of bcl-2. Extensive work on the role of bcl-2 in drug resistance has established an important role in AML. The ability of inhibitors of bcl-2 to sensitize AML cells to chemotherapeutic agents has led to the development of clinical trials using bcl-2 inhibitors (e.g., antisense oligonucleotide oblimersen or Genasense) (Marcucci et€ al. 2003, 2005). While this approach is being tested in adults with AML, no immediate clinical trials are being planned in pediatric AML (Moore et€al. 2006). An exciting new approach of therapeutically targeting protein degradation mechanisms in cancer cells as a means to sensitize them to chemotherapy has arisen out of experimental studies demonstrating the extreme dependency of tumor cells on such pathways (Zavrski et€al. 2007). In part because of the increased metabolic rate, the increased concentration of reactive oxygen species and the expression of multiple, mutated or aberrant proteins, malignant cells heavily depend on proteasome degradation. By blocking this system, proteasome inhibitors have been shown to increase the sensitivity of tumor cells to chemotherapeutic agents as well as induce apoptosis. Thus, this “proteasome addiction” of tumor cells, similar to their “kinase addiction” may make them more susceptible to inhibition of these pathways than normal cells. The observations that AML stem cells express high levels of the transcription and survival promoting factor, NF-KB, along with the expression of the NF-KB inhibitor, IKB, provides a potentially important link between survival factors and proteasome inhibition (Guzman et€al. 2001, 2002, 2007). For example, the use of the proteasome inhibitor, bortezomib (PS-341 or Velcade), has been shown to decrease the degradation of IKB and thereby reducing the amount of translocation of NF-KB to the nucleus, resulting in increased apoptosis of AML self-repopulating or stem cells (Guzman et€al. 2001, 2002). Further, the apoptotic effects of an anthracycline, for example, idarubicin, were greatly enhanced in AML stem cells in the presence of proteasome inhibition (Guzman et€al. 2001, 2002). The induction of the proapoptotic factors, BIM and Bax, have also been observed to be increased in AML cells exposed to idarubicin and proteasome inhibition (Pigneux et€al. 2007). Based on such data, the COG is planning to conduct a clinical trial with bortezemib in combination with chemotherapy in children with relapsed and/or refractory AML. This type of approach should help resolve the issue of the ability to selectively target AML stem cells in a clinical setting.
Epigenetic and Chromatin Remodeling-Directed Targets The modifications of DNA that help regulate its expression, whether cytosine methylation, histone type and/or modification or changes in other DNA-associated proteins, are considered epigenetic events (Holliday 2006). Epigenetic modifications of the genome have been causally linked to altered gene expression, often involving gene silencing of tumor suppressor genes and the development of various types of cancer. Several of the chromosomal translocations that characterize
Targeted Therapeutic Approaches for AML
71
various subtypes of AML are also known to generate fusion proteins of gene products that profoundly influence DNA and/or chromatin structure and function. For instance, several of these fusion proteins, including the t(15;17)/PML/RAR alpha in APL, those involving core binding factors such as AML1/ETO resulting from t(8;21), or those involving the MLL gene product, have been demonstrated to recruit repression-associated chromatin remodeling protein complexes to specific genomic regions (Wood et€al. 2005; Nie et€al. 2003). In a significant percentage of AML, silencing of expression coincides with promoter hypermethylation of the tumor suppressor gene, p15INK4b or the mismatch repair gene, hMLH1, is observed (Preisler et€al. 2001; Herman et€al. 1997; Seedhouse et€al. 2003). Other chromatin marks of the repression of gene expression are decreased histone H3 acetylation as well as trimethylation of lysine 27 of histone H3 (Bernstein et€al. 2007; Ting et€al. 2006). Rearrangements of the MLL gene along with its fusion to other partner genes results in both activation and repression of a subset of genes involved in AML, particularly the upregulation of HOX gene expression (Nie et€ al. 2003; So et€ al. 2003, 2004). Although the epigenetic changes that characterize different subtypes of AML are complex, many of them may converge on several critical changes, similar to the convergence of some downstream signaling pathways following activation of different TK receptors. Thus, strategies to reprogram the aberrant epigenetic changes observed in cancer and, in particular, MDS and AML are being pursued. Small molecule inhibitors or modulators of both DNA methylation and histone deacetylation have been developed and are being tested alone and in combination with the intent of inducing the reactivation of expression of genes which may promote leukemia cell maturation, senescence, or apoptosis (Ghoshal and Bai 2007; Santini et€al. 2007). Studies on MDS in adults led to the approval by the US Food and Drug Administration for the DNA methylation inhibitor 5¢-azacytidine (Garcia-Manero et€ al. 2007; Gattermann et€ al. 2007; Plimack et€ al. 2007; Kaminskas et€ al. 2005; Silverman 2001). However, in€ vitro and subsequently in€vivo data have demonstrated synergy in terms of the induction of gene expression with the use of combinations of inhibitors of DNA methylation and HDACs (Griffiths and Gore 2008; Qin et€al. 2007). Clinical trials testing the DNA methyltransferase inhibitors, 5¢-azacytidine or the deoxy-form, decitabine, alone and in combination with HDAC inhibitors, such as suberoylanilide hydroxamic acid (SAHA, Vorinostat or Zolinza), the benzamide derivative, MS275, depsipeptide, and valproic acid, are being intensely pursued in adults with MDS and high risk AML (Griffiths and Gore 2008; Gore et€al. 2006). Some of the results from these trials have demonstrated clinical responses that are comparable to conventional chemotherapy in older adults with high risk MDS/AML (Garcia-Manero et€ al. 2006; Issa et€ al. 2004; Soriano et€ al. 2007; Kantarjian et€ al. 2007). Trials using DNA methyltransferase and HDAC inhibitors alone and in combination are either ongoing or being planned in pediatric patients. Several challenges exist for this type of therapeutic targeting. The first is how specific or global the effects of the inhibitors should be in terms of effecting epigenetic change. An unwanted result would be if DNA demethylation, for example,
72
R.J. Arceci and D. Small
induced genes that would prevent apoptosis or induce drug resistance; there is some precedent for this occurring (Li et€al. 1999). Another challenge is how best to use epigenetically directed therapies in combination with each other, with other targeting agents or with conventional chemotherapy. For example, DNA replication is required for the effects of DNA methyltransferase inhibitors to be most effective. HDAC inhibitors have been known to cause cell cycle arrest. Thus, the sequencing of agents with these characteristics will be a significant issue to solve. A third challenge is to determine how such epigenetic therapies will impact the AML stem cell population. Increasing evidence suggests that the cancer stem cells share many epigenetic marks with embryonic stem cells. How such complex genomic signatures can be therapeutically manipulated remains an immense challenge.
Targeted Immunotherapy and Immunostimulatory Therapy Immense potential of the discovery of how to generate antigen specific, monoclonal antibodies from B cell/myeloma hybridomas by Kohler and Milstein has only started to be realized in the last decade. An outcome of such work in target AML has been thus far most using monoclonal antibodies directed against the differentiation antigen, CD33, a sialic acid-dependent cell adhesion molecule that is differentially expressed during myeloid differentiation. Gemtuzumab ozogamicin (GO; Mylotarg™) is a humanized, IgG4 subtype, monoclonal antibody conjugated to calicheamicin and directed against CD33; GO has demonstrated significant activity in AML and was the first toxin/monoclonal antibody fusion approved by the US FDA (Larson et€al. 2005). Initial clinical trials in adults and children with relapsed or primary refractory disease AML showed response rates of approximately 30 to 35% when GO is used as a single agent (Arceci et€ al. 2005; Larson et€ al. 2005). Of particular interest, the response in children with primary refractory disease was the same as that for patients with relapsed disease who had achieved a prior remission, suggesting that some conventional resistance mechanisms can be circumvented with this agent (Arceci et€al. 2005). The maximal tolerated dose (MTD) for children was determined to be 6€mg/m2 on the pediatric phase I study. The main toxicity was myelosuppression. However, there was a 24% (7 of 29 patients) overall incidence of veno-occlusive disease (VOD) with 6 of 13 (40%) developing VOD during a subsequent HSCT. VOD was most frequently observed in patients undergoing allogeneic SCT in less than 3.5 months from receiving GO. A COG phase I trial using GO in combination with either mitoxantrone plus high dose ARAC (regimen A) or with Capizzi II chemotherapy (regimen B) showed a maximum tolerated dose of GO to be 3 and 2€ mg/m2, respectively (Aplenc et€ al. 2008). Significant toxicities included those typically observed in patients with AML. DLTs for the Capizzi II included grade III transaminitis, pancreatitis and hyperbilirubinemia requiring a dose de-escalation to 2€mg/m2. The starting dose of 3€mg/m2 of GO in combination with mitoxantrone plus high dose ARAC was well tolerated. Responses included 52 and 40% CR
Targeted Therapeutic Approaches for AML
73
or CRp for regimens A and B, respectively (Aplenc et€ al. 2008). The MRC has performed a randomized trial of GO plus induction chemotherapy with ARAC, daunomycin, and etoposide (ADE) as well as with the second intensification course of mitoxantrone plus ARAC (Kell et€al. 2003). Results from this trial have been reported in abstract form and an advantage of EFS and OS for good risk patients who received GO (Burnett et€al. 2006). The current COG trial is test in a randomized fashion whether the addition of GO to induction and intensification courses of treatment results in an improved outcome for pediatric patients with newly diagnosed AML. In addition to the potential role of GO in improving the outcome of patients with AML, several additional questions remain. The question of why many patients are resistant to GO therapy has been partially addressed by the findings that calicheamicin is a substrate for the MDR1 P-glycoprotein drug transporter (Arceci et€al. 2005; Linenberger et€al. 2001). Thus, AML cells that express high levels of this transporter are more likely to be resistant to GO. These observations have led to the consideration of a clinical trial in pediatric patients with relapsed or refractory AML that combines GO with an inhibitor of the MDR1 transporter. A second question is whether GO will augment killing of AML repopulating stem cells. There is now definitive evidence of expression of CD33 on the surface of some AML stem cells, although the expression does not appear to be uniform across all AML (Pollard et€al. 2006; Taussig et€al. 2005). Phase III randomized trials will help answer the question of the utility of adding GO to conventional chemotherapy, but trials developed to look more specifically at the effects of such combination on AML stem cells should also be performed. The potential for targeting AML with monoclonal antibodies remains an important avenue for future research. Defining antigens that are more selectively or specifically expressed on AML stem cells remains an important challenge. CLL-1 has been reported to be such a specific AML stem cell antigen although its utility as a therapeutic target will need to be definitively tested in clinical trials (van Rhenen et€al. 2007). The hope that the inherent ability of the immune system to recognized foreign antigens and cells could be directed toward antitumor therapy has been a challenge for many decades. Several preclinical models for vaccine-mediated treatment of AML have shown excellent antileukemic effects, including cures (Arceci 1998; DunussiJoannopoulos et€al. 1996, 1997, 1998). Most of these preclinical models have involved upon the transduction of costimulatory receptors into AML cells or the use of cytokines to induce immune costimulation (Cheuk and Guinn 2008; Chan et€al. 2005). Clinical trials in adults with AML have involved similar approaches but remain in very early stages of development (Houtenbos et€al. 2006). Another approach has been to expose patients with potentially immunogenic antigens that are preferentially expressed on AML blasts (Greiner et€ al. 2004, 2005). For example, WT1 peptides are being tested in an attempt to augment anti-AML responses (Greiner et€al. 2005, 2006; Gaiger et€al. 2000). Attempts to stimulate anti-AML immune responses in pediatric patients have been tested in the randomized CCG-2961 trial, which used a limited course of IL-2 in the postremission period (Lange et€al. 2008). No significant difference in out-
74
R.J. Arceci and D. Small
come was observed between the patients who received IL-2 and those who did not (Lange et€al. 2008). This result could certainly have been due to the relatively short course or dosing of IL-2. Based in part on the empiric evidence that significant graft versus host disease is correlated with a lower relapse rate for patients with AML undergoing allogeneic HSCT, the use of donor lymphocytes and attempts to modulate graft versus tumor responses with cytokines and or early withdrawal of immunosuppressive drugs have been tested but with limited success. Another possibility, supported by preclinical models, is that this approach simply may not be sufficiently robust to elicit significant antitumor responses. Attempts to augment anti-AML immune responses in the allogeneic HSCT setting are also being tested in a COG trial attempting to exploit HLA mismatching and blockade of natural killer cell inhibition.
Future Challenges A major hope for improving the outcome of patients with AML is that more molecularly targeted therapies that will increase antileukemic activity without adding significant toxicity will be developed. Some of these agents, such as antibody-directed therapies and tyrosine kinase inhibitors, are currently in phase II and III clinical trials. The results of such trials should determine whether these approaches will be integrated into future regimens, establishing a new standard backbone for AML therapy. Although the thought of eliminating the need to use currently available cytotoxic chemotherapeutic agents with their short- and long-term associated adverse sequelae may seem improbable at this time, this should be our goal. Subtypes of AML with distinct genomic, gene expression or proteomic signatures or epigenetic patterns as well as with mutations of genes encoding critical molecules affecting signaling pathways should be prime targets for this type of therapy. In addition, essential to future strategies will be the important consideration of targeting leukemic self-repopulating progenitors in clinical trials. There is also a significant need for more predictive in€vitro and animal models for the many subtypes of AML. Such truly predictive in€ vitro or in€ vivo models would help greatly in the prioritization of new agents. However, at the current time and in the conceivable future, the definitive proof that a drug will improve outcomes in patients is to demonstrate this in the context of a well-designed and statistically powered clinical trial. The increasing emphasis on subgroup analysis, prognostic factor stratification and patient characteristics should lead to more individualized therapeutic approaches. The historical approach of using one or two hammers to hit a heterogeneous group of nails will be challenged by the biologically based delineation of increasingly smaller subgroups of patients for testing targeted therapies. Thus, the accrual to phase III trials will also continue to be a challenge with diminishing numbers of patients characterized by specific subgroups. The use of smaller, albeit less statistically robust, trials, such as randomized phase II or trials comparing an experimental treatment with a comparable historical
Targeted Therapeutic Approaches for AML
75
control group, may provide approaches to allow for both the testing of novel agents and decision-making concerning how to incorporate them into future regimens. We should also not forget that about 85% of children with cancer reside in �developing countries. And although significant obstacles may need to be overcome, it is becoming increasingly clear that approaches for effectively performing international trials need to be established. In addition, the predicted reduction in toxicity to be realized with molecularly targeted therapies is likely to be of significant benefit to patients in developing parts of the world where the complications associated with intensive chemotherapy regimens are often difficult to manage.
References Ichikawa H, Tanabe K, Mizushima H, et€al. Common gene expression signatures in t(8;21)- and inv(16)-acute myeloid leukaemia. Br J Haematol. 2006;135:336–347. Jordan CT. The leukemic stem cell. Best Pract Res Clin Haematol. 2007;20:13–18. Barabe F, Kennedy JA, Hope KJ, Dick JE. Modeling the initiation and progression of human acute leukemia in mice. Science. 2007;316:600–604. Hope KJ, Jin L, Dick JE. Acute myeloid leukemia originates from a hierarchy of leukemic stem cell classes that differ in self-renewal capacity. Nat Immunol. 2004;5:738–743. Andersson A, Eden P, Lindgren D, et€al. Gene expression profiling of leukemic cell lines reveals conserved molecular signatures among subtypes with specific genetic aberrations. Leukemia. 2005;19:1042–1050. Andersson A, Olofsson T, Lindgren D, et€al. Molecular signatures in childhood acute leukemia and their correlations to expression patterns in normal hematopoietic subpopulations. Proc Natl Acad Sci U S A. 2005;102:19069–19074. Ross ME, Mahfouz R, Onciu M, et€al. Gene expression profiling of pediatric acute myelogenous leukemia. Blood. 2004;104:3679–3687. Oshima Y, Ueda M, Yamashita Y, et€al. DNA microarray analysis of hematopoietic stem cell-like fractions from individuals with the M2 subtype of acute myeloid leukemia. Leukemia. 2003;17:1990–1997. Kiyoi H, Naoe T, Yokota S, et€al. Internal tandem duplication of FLT3 associated with leukocytosis in acute promyelocytic leukemia. Leukemia Study Group of the Ministry of Health and Welfare (Kohseisho). Leukemia. 1997;11:1447–1452. Kiyoi H, Towatari M, Yokota S, et€al. Internal tandem duplication of the FLT3 gene is a novel modality of elongation mutation which causes constitutive activation of the product. Leukemia. 1998;12:1333–1337. Meshinchi S, Alonzo TA, Gerbing R, Lange B, Radich JP. FLT3 internal tandem duplication is a prognostic factor for poor outcome in pediatric AML; a CCG-1961 study [abstract]. Blood. 2003;102:335a. Meshinchi S, Woods WG, Stirewalt DL, et€ al. Prevalence and prognostic significance of Flt3 internal tandem duplication in pediatric acute myeloid leukemia. Blood. 2001;97:89–94. Yamamoto Y, Kiyoi H, Nakano Y, et€al. Activating mutation of D835 within the activation loop of FLT3 in human hematologic malignancies. Blood. 2001;97:2434–2439. Schnittger S, Bacher U, Haferlach C, Kern W, Haferlach T. Rare CBFB-MYH11 fusion transcripts in AML with inv(16)/t(16;16) are associated with therapy-related AML M4eo, atypical cytomorphology, atypical immunophenotype, atypical additional chromosomal rearrangements and low white blood cell count: a study on 162 patients. Leukemia. 2007;21:725–731. Schnittger S, Kohl TM, Haferlach T, et€al. KIT-D816 mutations in AML1-ETO-positive AML are associated with impaired event-free and overall survival. Blood. 2006;107:1791–1799.
76
R.J. Arceci and D. Small
Shimada A, Taki T, Tabuchi K, et€al. KIT mutations, and not FLT3 internal tandem duplication, are strongly associated with a poor prognosis in pediatric acute myeloid leukemia with t(8;21): a study of the Japanese Childhood AML Cooperative Study Group. Blood. 2006;107:1806–1809. Tse KF, Mukherjee G, Small D. Constitutive activation of FLT3 stimulates multiple intracellular signal transducers and results in transformation. Leukemia. 2000;14:1766–1776. Ning ZQ, Li J, Arceci RJ. Activating mutations of c-kit at codon 816 confer drug resistance in human leukemia cells. Leuk Lymphoma. 2001;41:513–522. Ning ZQ, Li J, Arceci RJ. Signal transducer and activator of transcription 3 activation is required for Asp(816) mutant c-Kit-mediated cytokine-independent survival and proliferation in human leukemia cells. Blood. 2001;97:3559–3567. Ning ZQ, Li J, McGuinness M, Arceci RJ. STAT3 activation is required for Asp(816) mutant c-Kit induced tumorigenicity. Oncogene. 2001;20:4528–4536. Levis M, Murphy KM, Pham R, et€al. Internal tandem duplications of the FLT3 gene are present in leukemia stem cells. Blood. 2005;106:673–680. Pollard JA, Alonzo TA, Gerbing RB, et€al. FLT3 internal tandem duplication in CD34+/CD33− precursors predicts poor outcome in acute myeloid leukemia. Blood. 2006;108:2764–2769. Kottaridis PD, Gale RE, Langabeer SE, Frew ME, Bowen DT, Linch DC. Studies of FLT3 mutations in paired presentation and relapse samples from patients with acute myeloid leukemia: implications for the role of FLT3 mutations in leukemogenesis, minimal residual disease detection, and possible therapy with FLT3 inhibitors. Blood. 2002;100:2393–2398. Cloos J, Goemans BF, Hess CJ, et€ al. Stability and prognostic influence of FLT3 mutations in paired initial and relapsed AML samples. Leukemia. 2006;20:1217–1220. Brown P, Levis M, McIntyre E, Griesemer M, Small D. Combinations of the FLT3 inhibitor CEP701 and chemotherapy synergistically kill infant and childhood MLL-rearranged ALL cells in a sequence-dependent manner. Leukemia. 2006;20:1368–1376. Levis M, Allebach J, Tse KF, et€al. A FLT3-targeted tyrosine kinase inhibitor is cytotoxic to leukemia cells in€vitro and in€vivo. Blood. 2002;99:3885–3891. Levis M, Pham R, Smith BD, Small D. In vitro studies of a FLT3 inhibitor combined with chemotherapy: sequence of administration is important to achieve synergistic cytotoxic effects. Blood. 2004;104:1145–1150. Smith BD, Levis M, Beran M, et€al. Single-agent CEP-701, a novel FLT3 inhibitor, shows biologic and clinical activity in patients with relapsed or refractory acute myeloid leukemia. Blood. 2004;103:3669–3676. DeAngelo DJ, Stone RM, Heaney ML, et€al. Phase 1 clinical results with tandutinib (MLN518), a novel FLT3 antagonist, in patients with acute myelogenous leukemia or high-risk myelodysplastic syndrome: safety, pharmacokinetics, and pharmacodynamics. Blood. 2006;108:3674–3681. Cools J, Stover EH, Boulton CL, et€ al. PKC412 overcomes resistance to imatinib in a murine model of FIP1L1-PDGFRalpha-induced myeloproliferative disease. Cancer Cell. 2003;3:459–469. Stone RM, DeAngelo DJ, Klimek V, et€al. Patients with acute myeloid leukemia and an activating mutation in FLT3 respond to a small-molecule FLT3 tyrosine kinase inhibitor, PKC412. Blood. 2005;105:54–60. Levis M, Brown P, Smith BD, et€al. Plasma inhibitory activity (PIA): a pharmacodynamic assay reveals insights into the basis for cytotoxic response to FLT3 inhibitors. Blood. 2006;108:3477–3483. Knapper S, Burnett AK, Littlewood T, et€ al. A phase 2 trial of the FLT3 inhibitor lestaurtinib (CEP701) as first-line treatment for older patients with acute myeloid leukemia not considered fit for intensive chemotherapy. Blood. 2006;108:3262–3270. Levis M, Smith BD, Beran M, et€al. A randomized, open-label study of lestaurtinib (CEP-701), an oral FLT3 inhibitor, administered in sequence with chemotherapy in patients with relapsed AML harboring FLT3 activating mutations: clinical response correlates with successful FLT3 inhibition. Blood. 2005;106:403a.
Targeted Therapeutic Approaches for AML
77
Armstrong SA, Staunton JE, Silverman LB, et€ al. MLL translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nat Genet. 2002;30:41–47. Roberts KG, Odell AF, Byrnes EM, et€ al. Resistance to c-KIT kinase inhibitors conferred by V654A mutation. Mol Cancer Ther. 2007;6:1159–1166. Meshinchi S, Stirewalt DL, Alonzo TA, et€al. Activating mutations of RTK/RAS signal transduction pathway in pediatric acute myeloid leukemia. Blood. 2003;102:1474–1479. Heinrich MC, Griffith DJ, Druker BJ, Wait CL, Ott KA, Zigler AJ. Inhibition of c-kit receptor tyrosine kinase activity by STI 571, a selective tyrosine kinase inhibitor. Blood. 2000;96:925–932. Fernandez A, Sanguino A, Peng Z, et€al. An anticancer c-kit kinase inhibitor is reengineered to make it more active and less cardiotoxic. J Clin Invest. 2007;117:4044–4054. Kindler T, Breitenbuecher F, Marx A, et€al. Efficacy and safety of imatinib in adult patients with c-kit-positive acute myeloid leukemia. Blood. 2004;103:3644–3654. Ito K, Tominaga K, Suzuki T, Jinnai I, Bessho M. Successful treatment with imatinib mesylate in a case of minor BCR–ABL-positive acute myelogenous leukemia. Int J Hematol. 2005;81:242–245. Jentsch-Ullrich K, Pelz AF, Braun H, et€ al. Complete molecular remission in a patient with Philadelphia-chromosome positive acute myeloid leukemia after conventional therapy and imatinib. Haematologica. 2004;89:ECR15. Pompetti F, Spadano A, Sau A, et€al. Long-term remission in BCR/ABL-positive AML-M6 patient treated with imatinib mesylate. Leuk Res. 2007;31:563–567. Heidel F, Cortes J, Rucker FG, et€al. Results of a multicenter phase II trial for older patients with c-kit-positive acute myeloid leukemia (AML) and high-risk myelodysplastic syndrome (HR-MDS) using low-dose Ara-C and imatinib. Cancer. 2007;109:907–914. Hizuka N, Sukegawa I, Takano K, et€al. Characterization of insulin-like growth factor I receptors on human erythroleukemia cell line (K-562 cells). Endocrinol Jpn. 1987;34:81–88. Abe S, Funato T, Takahashi S, et€al. Increased expression of insulin-like growth factor i is associated with Ara-C resistance in leukemia. Tohoku J Exp Med. 2006;209:217–228. Frostad S, Bruserud O. In vitro effects of insulin-like growth factor-1 (IGF-1) on proliferation and constitutive cytokine secretion by acute myelogenous leukemia blasts. Eur J Haematol. 1999;62:191–198. Cosaceanu D, Carapancea M, Alexandru O, et€al. Comparison of three approaches for inhibiting insulin-like growth factor I receptor and their effects on NSCLC cell lines in€ vitro. Growth Factors. 2007;25:1–8. Jerome L, Alami N, Belanger S, et€ al. Recombinant human insulin-like growth factor binding protein 3 inhibits growth of human epidermal growth factor receptor-2-overexpressing breast tumors and potentiates herceptin activity in€vivo. Cancer Res. 2006;66:7245–7252. Perl AE, Carroll M. Exploiting signal transduction pathways in acute myelogenous leukemia. Curr Treat Options Oncol. 2007;8:265–276. Aguayo A. The role of angiogenesis in the biology and therapy of myelodysplastic syndromes. Curr Hematol Rep. 2004;3:184–191. Aguayo A, Estey E, Kantarjian H, et€al. Cellular vascular endothelial growth factor is a predictor of outcome in patients with acute myeloid leukemia. Blood. 1999;94:3717–3721. Aguayo A, Giles F, Albitar M. Vascularity, angiogenesis and angiogenic factors in leukemias and myelodysplastic syndromes. Leuk Lymphoma. 2003;44:213–222. Hussong JW, Rodgers GM, Shami PJ. Evidence of increased angiogenesis in patients with acute myeloid leukemia. Blood. 2000;95:309–313. Kessler T, Fehrmann F, Bieker R, Berdel WE, Mesters RM. Vascular endothelial growth factor and its receptor as drug targets in hematological malignancies. Curr Drug Targets. 2007;8:257–268. Schuch G, Machluf M, Bartsch G, Jr., et€al. In vivo administration of vascular endothelial growth factor (VEGF) and its antagonist, soluble neuropilin-1, predicts a role of VEGF in the progression of acute myeloid leukemia in€vivo. Blood. 2002;100:4622–4628.
78
R.J. Arceci and D. Small
Dias S, Choy M, Alitalo K, Rafii S. Vascular endothelial growth factor (VEGF)-C signaling through FLT-4 (VEGFR-3) mediates leukemic cell proliferation, survival, and resistance to chemotherapy. Blood. 2002;99:2179–2184. Liersch R, Schliemann C, Bieker R, et€al. Expression of VEGF-C and its receptor VEGFR-3 in the bone marrow of patients with acute myeloid leukaemia. Leuk Res. 2008;32:954–961. Karp JE, Gojo I, Pili R, et€ al. Targeting vascular endothelial growth factor for relapsed and Â�refractory adult acute myelogenous leukemias: therapy with sequential 1-beta-d-arabinofuranosylcytosine, mitoxantrone, and bevacizumab. Clin Cancer Res. 2004;10:3577–3585. Fiedler W, Mesters R, Tinnefeld H, et€ al. A phase 2 clinical study of SU5416 in patients with refractory acute myeloid leukemia. Blood. 2003;102:2763–2767. Maris JM, Courtright J, Houghton PJ, et€al. Initial testing of the VEGFR inhibitor AZD2171 by the pediatric preclinical testing program. Pediatr Blood Cancer. 2008;50:581–587. Piloto O, Wright M, Brown P, Kim KT, Levis M, Small D. Prolonged exposure to FLT3 inhibitors leads to resistance via activation of parallel signaling pathways. Blood. 2007;109:1643–1652. Farr C, Gill R, Katz F, Gibbons B, Marshall CJ. Analysis of RAS gene mutations in childhood myeloid leukaemia. Br J Haematol. 1991;77:323–327. Loh ML, Vattikuti S, Schubbert S, et€al. Mutations in PTPN11 implicate the SHP-2 phosphatase in leukemogenesis. Blood. 2004;103:2325–2331. Karp JE. Farnesyl protein transferase inhibitors as targeted therapies for hematologic malignancies. Semin Hematol. 2001;38:16–23. Karp JE. Farnesyl transferase inhibition in hematologic malignancies. J Natl Compr Canc Netw. 2005;3 Suppl 1:S37–S40. Emanuel PD, Snyder RC, Wiley T, Gopurala B, Castleberry RP. Inhibition of juvenile myelomonocytic leukemia cell growth in€ vitro by farnesyltransferase inhibitors. Blood. 2000;95: 639–645. Gotlib J. Farnesyltransferase inhibitor therapy in acute myelogenous leukemia. Curr Hematol Rep. 2005;4:77–84. Harousseau JL, Lancet JE, Reiffers J, et€al. A phase 2 study of the oral farnesyltransferase inhibitor tipifarnib in patients with refractory or relapsed acute myeloid leukemia. Blood. 2007;109: 5151–5156. Lancet JE, Gojo I, Gotlib J, et€al. A phase 2 study of the farnesyltransferase inhibitor tipifarnib in poor-risk and elderly patients with previously untreated acute myelogenous leukemia. Blood. 2007;109:1387–1394. Milella M, Estrov Z, Kornblau SM, et€ al. Synergistic induction of apoptosis by simultaneous disruption of the Bcl-2 and MEK/MAPK pathways in acute myelogenous leukemia. Blood. 2002;99:3461–3464. Milella M, Kornblau SM, Andreeff M. The mitogen-activated protein kinase signaling module as a therapeutic target in hematologic malignancies. Rev Clin Exp Hematol. 2003;7:160–190. Milella M, Kornblau SM, Estrov Z, et€al. Therapeutic targeting of the MEK/MAPK signal transduction module in acute myeloid leukemia. J Clin Invest. 2001;108:851–859. James JA, Smith MA, Court EL, et€al. An investigation of the effects of the MEK inhibitor U0126 on apoptosis in acute leukemia. Hematol J. 2003;4:427–432. Tong FK, Chow S, Hedley D. Pharmacodynamic monitoring of BAY 43-9006 (Sorafenib) in phase I clinical trials involving solid tumor and AML/MDS patients, using flow cytometry to monitor activation of the ERK pathway in peripheral blood cells. Cytometry B Clin Cytom. 2006;70:107–114. Weisberg E, Banerji L, Wright RD, et€al. Potentiation of anti-leukemic therapies by the dual PI3K/ PDK-1 inhibitor, BAG956: effects on BCR–ABL- and mutant FLT3-expressing cells. Blood. 2008. Tamburini J, Chapuis N, Bardet V, et€ al. Mammalian target of rapamycin (mTOR) inhibition activates phosphatidylinositol 3-kinase/Akt by up-regulating insulin-like growth factor-1 receptor signaling in acute myeloid leukemia: rationale for therapeutic inhibition of both pathways. Blood. 2008;111:379–382. Xu Q, Simpson SE, Scialla TJ, Bagg A, Carroll M. Survival of acute myeloid leukemia cells requires PI3 kinase activation. Blood. 2003;102:972–980.
Targeted Therapeutic Approaches for AML
79
Xu Q, Thompson JE, Carroll M. mTOR regulates cell survival after etoposide treatment in primary AML cells. Blood. 2005;106:4261–4268. Recher C, Beyne-Rauzy O, Demur C, et€al. Antileukemic activity of rapamycin in acute myeloid leukemia. Blood. 2005;105:2527–2534. Min YH, Eom JI, Cheong JW, et€al. Constitutive phosphorylation of Akt/PKB protein in acute myeloid leukemia: its significance as a prognostic variable. Leukemia. 2003;17:995–997. Yee KW, Zeng Z, Konopleva M, et€ al. Phase I/II study of the mammalian target of rapamycin inhibitor everolimus (RAD001) in patients with relapsed or refractory hematologic malignancies. Clin Cancer Res. 2006;12:5165–5173. Baer MR, George SL, Dodge RK, et€al. Phase 3 study of the multidrug resistance modulator PSC833 in previously untreated patients 60 years of age and older with acute myeloid leukemia: Cancer and Leukemia Group B Study 9720. Blood. 2002;100:1224–1232. Leith CP, Chen IM, Kopecky KJ, et€ al. Correlation of multidrug resistance (MDR1) protein expression with functional dye/drug efflux in acute myeloid leukemia by multiparameter flow cytometry: identification of discordant MDR−/efflux+ and MDR1+/efflux− cases. Blood. 1995;86:2329–2342. Leith CP, Kopecky KJ, Chen IM, et€al. Frequency and clinical significance of the expression of the multidrug resistance proteins MDR1/P-glycoprotein, MRP1, and LRP in acute myeloid leukemia: a Southwest Oncology Group Study. Blood. 1999;94:1086–1099. Willman CL. Immunophenotyping and cytogenetics in older adults with acute myeloid leukemia: significance of expression of the multidrug resistance gene-1 (MDR1). Leukemia. 1996;10 Suppl 1:S33–S35. Willman CL. The prognostic significance of the expression and function of multidrug resistance transporter proteins in acute myeloid leukemia: studies of the Southwest Oncology Group Leukemia Research Program. Semin Hematol. 1997;34:25–33. Sievers EL, Smith FO, Woods WG, et€ al. Cell surface expression of the multidrug resistance P-glycoprotein (P-170) as detected by monoclonal antibody MRK-16 in pediatric acute myeloid leukemia fails to define a poor prognostic group: a report from the Childrens Cancer Group. Leukemia. 1995;9:2042–2048. List AF, Spier C, Greer J, et€al. Phase I/II trial of cyclosporine as a chemotherapy-resistance modifier in acute leukemia. J Clin Oncol. 1993;11:1652–1660. Dahl GV, Lacayo NJ, Brophy N, et€al. Mitoxantrone, etoposide, and cyclosporine therapy in pediatric patients with recurrent or refractory acute myeloid leukemia. J Clin Oncol. 2000;18:1867–1875. Greenberg PL, Lee SJ, Advani R, et€al. Mitoxantrone, etoposide, and cytarabine with or without valspodar in patients with relapsed or refractory acute myeloid leukemia and high-risk myelodysplastic syndrome: a phase III trial (E2995). J Clin Oncol. 2004;22:1078–1086. List AF, Kopecky KJ, Willman CL, et€al. Benefit of cyclosporine modulation of drug resistance in patients with poor-risk acute myeloid leukemia: a Southwest Oncology Group study. Blood. 2001;98:3212–3220. Becton D, Dahl GV, Ravindranath Y, et€al. Randomized use of cyclosporin A (CsA) to modulate P-glycoprotein in children with AML in remission: Pediatric Oncology Group Study 9421. Blood. 2006;107:1315–1324. Arceci RJ, Sande J, Lange B, et€al. Safety and efficacy of gemtuzumab ozogamicin in pediatric patients with advanced CD33+ acute myeloid leukemia. Blood. 2005;106:1183–1188. Linenberger ML, Hong T, Flowers D, et€al. Multidrug-resistance phenotype and clinical responses to gemtuzumab ozogamicin. Blood. 2001;98:988–994. Marcucci G, Byrd JC, Dai G, et€ al. Phase 1 and pharmacodynamic studies of G3139, a Bcl-2 antisense oligonucleotide, in combination with chemotherapy in refractory or relapsed acute leukemia. Blood. 2003;101:425–432. Marcucci G, Stock W, Dai G, et€al. Phase I study of oblimersen sodium, an antisense to Bcl-2, in untreated older patients with acute myeloid leukemia: pharmacokinetics, pharmacodynamics, and clinical activity. J Clin Oncol. 2005;23:3404–3411. Moore J, Seiter K, Kolitz J, et€al. A phase II study of Bcl-2 antisense (oblimersen sodium) combined with gemtuzumab ozogamicin in older patients with acute myeloid leukemia in first relapse. Leuk Res. 2006;30:777–783.
80
R.J. Arceci and D. Small
Zavrski I, Jakob C, Kaiser M, Fleissner C, Heider U, Sezer O. Molecular and clinical aspects of proteasome inhibition in the treatment of cancer. Recent Results Cancer Res. 2007;176: 165–176. Guzman ML, Neering SJ, Upchurch D, et€al. Nuclear factor-kappaB is constitutively activated in primitive human acute myelogenous leukemia cells. Blood. 2001;98:2301–2307. Guzman ML, Swiderski CF, Howard DS, et€ al. Preferential induction of apoptosis for primary human leukemic stem cells. Proc Natl Acad Sci U S A. 2002;99:16220–16225. Guzman ML, Rossi RM, Neelakantan S, et€al. An orally bioavailable parthenolide analog selectively eradicates acute myelogenous leukemia stem and progenitor cells. Blood. 2007;110:4427–4435. Pigneux A, Mahon FX, Moreau-Gaudry F, et€ al. Proteasome inhibition specifically sensitizes leukemic cells to anthracyclin-induced apoptosis through the accumulation of Bim and Bax pro-apoptotic proteins. Cancer Biol Ther. 2007;6:603–611. Holliday R. Epigenetics: a historical overview. Epigenetics. 2006;1:76–80. Wood A, Schneider J, Shilatifard A. Cross-talking histones: implications for the regulation of gene expression and DNA repair. Biochem Cell Biol. 2005;83:460–467. Nie Z, Yan Z, Chen EH, et€al. Novel SWI/SNF chromatin-remodeling complexes contain a mixedlineage leukemia chromosomal translocation partner. Mol Cell Biol. 2003;23:2942–2952. Preisler HD, Li B, Chen H, et€al. P15INK4B gene methylation and expression in normal, myelodysplastic, and acute myelogenous leukemia cells and in the marrow cells of cured lymphoma patients. Leukemia. 2001;15:1589–1595. Herman JG, Civin CI, Issa JP, Collector MI, Sharkis SJ, Baylin SB. Distinct patterns of inactivation of p15INK4B and p16INK4A characterize the major types of hematological malignancies. Cancer Res. 1997;57:837–841. Seedhouse CH, Das-Gupta EP, Russell NH. Methylation of the hMLH1 promoter and its association with microsatellite instability in acute myeloid leukemia. Leukemia. 2003;17:83–88. Bernstein BE, Meissner A, Lander ES. The mammalian epigenome. Cell. 2007;128:669–681. Ting AH, McGarvey KM, Baylin SB. The cancer epigenome – components and functional correlates. Genes Dev. 2006;20:3215–3231. So CW, Karsunky H, Wong P, Weissman IL, Cleary ML. Leukemic transformation of hematopoietic progenitors by MLL-GAS7 in the absence of Hoxa7 or Hoxa9. Blood. 2004; 103:3192–3199. So CW, Lin M, Ayton PM, Chen EH, Cleary ML. Dimerization contributes to oncogenic activation of MLL chimeras in acute leukemias. Cancer Cell. 2003;4:99–110. Ghoshal K, Bai S. DNA methyltransferases as targets for cancer therapy. Drugs Today (Barc). 2007;43:395–422. Santini V, Gozzini A, Ferrari G. Histone deacetylase inhibitors: molecular and biological activity as a premise to clinical application. Curr Drug Metab. 2007;8:383–393. Garcia-Manero G, Yang AS, Jagasia M. Evaluating new treatment options for MDS. Clin Adv Hematol Oncol. 2007;5:1–9 Gattermann N, Kundgen A, Germing U. Treatment of patients with high-risk myelodysplastic syndromes. Cancer Treat Rev. 2007;33 Suppl 1:S64–S68. Plimack ER, Kantarjian HM, Issa JP. Decitabine and its role in the treatment of hematopoietic malignancies. Leuk Lymphoma. 2007;48:1472–1481. Kaminskas E, Farrell AT, Wang YC, Sridhara R, Pazdur R. FDA drug approval summary: azacitidine (5-azacytidine, Vidaza) for injectable suspension. Oncologist. 2005;10:176–182. Silverman LR. Targeting hypomethylation of DNA to achieve cellular differentiation in myelodysplastic syndromes (MDS). Oncologist. 2001;6 Suppl 5:8–14. Griffiths EA, Gore SD. DNA methyltransferase and histone deacetylase inhibitors in the treatment of myelodysplastic syndromes. Semin Hematol. 2008;45:23–30. Qin T, Youssef EM, Jelinek J, et€al. Effect of cytarabine and decitabine in combination in human leukemic cell lines. Clin Cancer Res. 2007;13:4225–4232. Gore SD, Baylin S, Sugar E, et€ al. Combined DNA methyltransferase and histone deacetylase inhibition in the treatment of myeloid neoplasms. Cancer Res. 2006;66:6361–6369.
Targeted Therapeutic Approaches for AML
81
Garcia-Manero G, Kantarjian HM, Sanchez-Gonzalez B, et€al. Phase 1/2 study of the combination of 5-aza-2¢-deoxycytidine with valproic acid in patients with leukemia. Blood. 2006;108:3271–3279. Issa JP, Garcia-Manero G, Giles FJ, et€al. Phase 1 study of low-dose prolonged exposure schedules of the hypomethylating agent 5-aza-2¢-deoxycytidine (decitabine) in hematopoietic malignancies. Blood. 2004;103:1635–1640. Soriano AO, Yang H, Faderl S, et€al. Safety and clinical activity of the combination of 5-azacytidine, valproic acid, and all-trans retinoic acid in acute myeloid leukemia and myelodysplastic syndrome. Blood. 2007;110:2302–2308. Kantarjian HM, O’Brien S, Huang X, et€al. Survival advantage with decitabine versus intensive chemotherapy in patients with higher risk myelodysplastic syndrome: comparison with historical experience. Cancer. 2007;109:1133–1137. Li Q, Kopecky KJ, Mohan A, et€al. Estrogen receptor methylation is associated with improved survival in adult acute myeloid leukemia. Clin Cancer Res. 1999;5:1077–1084. Larson RA, Sievers EL, Stadtmauer EA, et€al. Final report of the efficacy and safety of gemtuzumab ozogamicin (Mylotarg) in patients with CD33-positive acute myeloid leukemia in first recurrence. Cancer. 2005;104:1442–1452. Aplenc R, Alonzo T, Gerbing A, Robert B, et€al. Safety and efficacy of gemtuzumab ozogamicin in combination with chemotherapy for pediatric acute myeloid leukemia. J Clin Oncol. 2008;26:2390–3295. Kell WJ, Burnett AK, Chopra R, et€al. A feasibility study of simultaneous administration of gemtuzumab ozogamicin with intensive chemotherapy in induction and consolidation in younger patients with acute myeloid leukemia. Blood. 2003;102:4277–4283. Burnett AK, Kell WJ, Goldstone AH, et€al. The addition of gemtuzumab ozogamicin to induction chemotherapy for AML improves disease free survival without extra toxicity: preliminary analysis of 1115 patients in the MRC AML15 trial. Blood. 2006;108:8a. Taussig DC, Pearce DJ, Simpson C, et€ al. Hematopoietic stem cells express multiple myeloid markers: implications for the origin and targeted therapy of acute myeloid leukemia. Blood. 2005;106:4086–4092. van Rhenen A, van Dongen GA, Kelder A, et€ al. The novel AML stem cell associated antigen CLL-1 aids in discrimination between normal and leukemic stem cells. Blood. 2007;110: 2659–2666. Arceci RJ. The potential for antitumor vaccination in acute myelogenous leukemia. J Mol Med. 1998;76:80–93. Dunussi-Joannopoulos K, Dranoff G, Weinstein HJ, Ferrara JL, Bierer BE, Croop JM. Gene immunotherapy in murine acute myeloid leukemia: granulocyte-macrophage colony-stimulating factor tumor cell vaccines elicit more potent antitumor immunity compared with B7 family and other cytokine vaccines. Blood. 1998;91:222–230. Dunussi-Joannopoulos K, Krenger W, Weinstein HJ, Ferrara JL, Croop JM. CD8+ T cells activated during the course of murine acute myelogenous leukemia elicit therapeutic responses to late B7 vaccines after cytoreductive treatment. Blood. 1997;89:2915–2924. Dunussi-Joannopoulos K, Weinstein HJ, Nickerson PW, et€al. Irradiated B7-1 transduced primary acute myelogenous leukemia (AML) cells can be used as therapeutic vaccines in murine AML. Blood. 1996;87:2938–2946. Cheuk AT, Guinn BA. Immunotherapy of acute myeloid leukaemia: development of a whole cell vaccine. Front Biosci. 2008;13:2022–2029. Chan L, Hardwick N, Darling D, et€al. IL-2/B7.1 (CD80) fusagene transduction of AML blasts by a self-inactivating lentiviral vector stimulates T cell responses in€vitro: a strategy to generate whole cell vaccines for AML. Mol Ther. 2005;11:120–131. Houtenbos I, Westers TM, Ossenkoppele GJ, van de Loosdrecht AA. Feasibility of clinical dendritic cell vaccination in acute myeloid leukemia. Immunobiology. 2006;211:677–685. Greiner J, Li L, Ringhoffer M, et€al. Identification and characterization of epitopes of the receptor for hyaluronic acid-mediated motility (RHAMM/CD168) recognized by CD8+ T cells of HLA-A2-positive patients with acute myeloid leukemia. Blood. 2005;106:938–945.
82
R.J. Arceci and D. Small
Greiner J, Ringhoffer M, Taniguchi M, et€al. mRNA expression of leukemia-associated antigens in patients with acute myeloid leukemia for the development of specific immunotherapies. Int J Cancer. 2004;108:704–711. Gaiger A, Reese V, Disis ML, Cheever MA. Immunity to WT1 in the animal model and in patients with acute myeloid leukemia. Blood. 2000;96:1480–1489. Greiner J, Dohner H, Schmitt M. Cancer vaccines for patients with acute myeloid leukemia – definition of leukemia-associated antigens and current clinical protocols targeting these antigens. Haematologica. 2006;91:1653–1661. Lange B, Smith FO, Feusner J, et€al. Outcomes in CC-2961: a Children’s Cancer Group Phase 3 Trial for untreated pediatric acute myeloid leukemia (AML). Blood. 2008;111:1044–1053. Arceci RJ, Cripe TP. Emerging cancer-targeted therapies. Pediatr Clin North Am. 2002;49: 1339–1368, vii–viii Doepfner KT, Boller D, Arcaro A. Targeting receptor tyrosine kinase signaling in acute myeloid leukemia. Crit Rev Oncol Hematol. 2007;63:215–230. Poland KS, Shardy DL, Azim M, et al. Overexpression of ZNF 342 by juxtaposition with MPO Promoter/enhancer in the novel translocation t(17;19) (q23; q13.32) in pediatric acute myeloid leukemia and analysis of ZNF 342 expression in leukemia. Genes Chromosomes Cancer. 2009;48:480–489.
Acute Promyelocytic Leukaemia Andrea Biondi, Anna Maria Testi, and Brenda E.S. Gibson
Introduction Since the time of its first description (Degos 2003), Acute Promyelocytic Leukaemia (APL) has drawn the attention of clinicians and scientists for its clinical and biological features (Degos 2003). APL is notable for distinctive clinical and biological characteristics, including its particular morphological features, the presence of potentially devastating hemorrhagic syndrome that is related to disseminated intravascular coagulopathy and abnormal fibrinolysis, and finally, the sensitivity to anthracyclinecontaining chemotherapy (Bernard et€ al. 1973). The availability of differentation therapy with All-Trans Retinoic Acid (ATRA), and more recently, the discovery of the beneficial effect of arsenic trioxide (ATO), together with the molecular characterization of the t(15;17) specific translocation (Grignani et€al. 1994) have produced a remarkable improvement in patient outcome in the last decade. Most patients with APL are now cured (Sanz 2006). APL is a paradigm of a malignant disease that can be treated by cell modulation, using agents that act specifically on oncogenic, molecular events. This chapter reviews its impact in children APL.
Demographic Features APL represents approximately 4 to 8% of paediatric AML (Kaspers and Creutzig 2005). The median age at presentation is probably similar to that of other AML subtypes (7 to 9€years), but APL has rarely been reported in the first year of life. A remarkable epidemiologic feature of APL is its high incidence in certain ethnic groups (Ribeiro and Rego 2006). Douer described the increased incidence of APL
A. Biondiâ•›(*) Centro M. Tettamanti, Clinica Pediatrica Università Milano-Bicocca, Ospedale San Gerardo,Via Pergolesi, 33, 20052 Monza, Italy e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_5, © Springer Science+Business Media, LLC 2010
83
84
A. Biondi et al.
among patients of Latin American descent (Douer 2003). Several series from hospitalbased registries in Italy (Biondi et€al. 1994), Spain (Tomas and Fernandez-Ranada 1996), Mexico (Ruiz-Arguelles 1997), Perù (Otero et€ al. 1996), and Nicaragua (Malta Corea et€al. 1993) have noted a higher-than expected frequency of APL. It is still unknown whether such differences may suggest a genetic predisposition to APL and/or exposure to distinct environmental factor(s).
Pathogenesis of APL Acute promyelocytic leukaemia (APL) is characterized by rearrangements of the retinoic acid receptor-a(RARA) gene on chromosome 17q21 reviewed by Melnick and Licht (1999). To date, six different partner genes have been identified, with the vast majority of cases having the majority of the RAR gene fused to the PML gene on chromosome 15q21 reviewed by Grignani et€al. (1994). Determination of the underlying molecular lesion is critical for appropriate management of APL, with the presence of an underlying PML-RARA fusion gene predicting a favorable response to molecularly targeted therapies in the form of ATRA and ATO. Two rare subtypes of retinoid-sensitive APL have been reported in the paediatric population, that is, the NPM1-RARA and NuMA-RARA fusions resulting from t(5;17)(q34-35;q21) and t(11;17)(q13;q21), respectively (Redner et€al. 1996; Wells et€al. 1997; Grimwade et€al. 2000). The response of these rare subtypes of APL to ATO has not been established in patients. PLZF-RARA fusion, associated with t(11;17)(q23;q21) has only been reported in adults with the disease but is clinically important since it is resistant to ATRA and ATO reviewed by Mistry et€al. (2003). The STAT5b-RARA fusion resulting from an interstitial deletion on chromosome 17 is also considered to be retinoid-resistant (Arnould et€al. 1999). Fusion of the PRKAR1A to RARA was described as a variant APL. (Catalano et€al. 2007). The presence of the PML-RARA fusion gene is responsible for the peculiar response to ATRA and ATO. RARA (there are three homologous RAR proteins, called a, b and g) is a ligand-dependent transcriptional activator that binds through its zinc (Zn) finger domain, as heterodimer with members of the RXR family of nuclear receptors (Grignani et€al. 1994), to specific DNA sequences (called retinoic acid responsive elements, RARE) found in the promoters of retinoic acid-responsive genes (Umesono et€al. 1991). This process is accomplished in concert with a coactivator complex that includes proteins such as p300 and pCAF. As shown schematically in Fig.1, in the absence of retinoic acid, a co-repressor complex is recruited comprising either Nuclear Receptor Co-Repressor (N-CoR) or Silencing Mediator of Retinoid and Thyroid Receptors (SMRT), which binds to another protein called Sin3 and then to HDAC1 (Umesono et€ al. 1991; Alland et€ al. 1997; Laherty et€al. 1997; Heinzel et€al. 1997; Redner et€al. 1999). HDAC1 is a histone deacetylase protein, which epigenetically alters histones to keep DNA in an untranscribable form. The expression of retinoic-acid responsive genes is essential for
Acute Promyelocytic Leukaemia
85
Differentiation RAR (wild-type) PML-RARa
Cofactors
in3
N-CoR
mS
RAR
N-CoR
3
in mS
HDAC
HDAC
RAR
PML
Myelocyte
All-transretinoic acid RAR
X Transcription arrest
APL blast cell
PML
PML (dispersed)
Metamyelocyte PML-RARα degradation
PML in PODs
Granulocyte
Fig.€1╅ A schematic representation of the effects of PML-RAR and ATRA in the blast cells of acute promyelocytic leukaemia. DNA-bound PML-RARA interacts with N-CoR (or SMRT) and recruits the m-Sin3-HD complex, decreasing hystone acetylation and producing repressive chromatin organisation and transcriptional repression. Retinoic acid interaction with PML/RAR leads to rapid degradation of this fusion protein and assembly of the wild-type PML protein into normal nuclear structures called PODs (PML oncogenic domains). Moreover, it induces dissociation of the N-CoR-mSin3-HD complex, recruitment of coactivators with acetyltransferase activity, increased levels of hystone acetylation, chromatin remodelling, and transcriptional activation
normal myeloid development (Tsai and Collins 1993). In APL, the presence of the PML-RARA fusion gene generates a chimeric protein that is capable of DNA binding, but it recruits the N- CoR/Sin3/HDAC1 co-repressor complex and prevents transcription of RAR target genes (Collins 1998) and myeloid cell differentiation (Tsai and Collins 1993) under normal conditions. Additionally, the PML-RARA forms a heterodimer with the wild type PML protein and disrupts the PML nuclear body/PML oncogenic domain (Koken et€al. 1994). ATRA functions by binding to the RAR and causing degradation of the PML-RAR protein through both the ubiquitin-proteosome and caspase systems, thus allowing for the terminal differentiation of leukemic promyelocytes (Nervi et€ al. 1998; Zhu et€ al. 1999; Liu et€ al. 2000). This binding also results in N-CoR being dissociated from the PML-RAR fusion protein, and subsequently, the nuclear co-activator complex is recruited to reverse histone deacetylase-mediated repression Fig.€1. Finally, PML nuclear bodies are restored to their normal structure.
86
A. Biondi et al.
Diagnosis of APL Compared with diagnosis of other AML subtypes, the identification of APL conveys unique therapeutic and prognostic implications. In fact, (1) this leukaemia is a medical emergency, and up to 10% of early hemorrhagic deaths are currently recorded even in patients receiving modern state-of-the-art treatments; (2) the optimised front-line approach (ATRA plus chemotherapy) is different from that used in other AMLs and is effective also in controlling the life-threatening coagulopathy. Figureâ•›2 summarises the different tools available for the proper diagnosis of APL. According to FAB classification, the term M3-AML was assigned to hypergranular promyelocytic leukaemia characterised by blast cells with heavy azurophilic granules, bundles of Auer rods, and a reniform or bilobed nucleus (Fig.€2 panel A). Although the vast majority of M3 cases fit the description of hypergranular or classical M3, a cytological hypogranular or microgranular variant form, M3v, has been identified (Fig.€2 panel B). It is commonly associated with hyperleukocytosis, accounts for 15 to 20% of APL cases, and shares the same t(15;17). Albeit less frequently observed, M3v occurs even in paediatric patients (Rovelli et€al. 1992). As compared with control cases with the classical t(15;17) translocation, in the cases with PLZF/RARA rearrangements, the majority of blasts had (1) a regular nucleus, and, (2) an abundant cytoplasm with either coarse granules or, less frequently, with fine or no granules and Chediak-like granules are rarely detectable in such cases Sainty et€al. (2000). The presence of a regular nuclear outline is a key feature of t(11;17)-associated APL. The few cases so far described of APL expressing NPM/RARA (Redner et€al. 1996) could be classified as M3 and presented with hyperleukocytosis the remaining case had an M3v morphology. As for morphology and molecular genetics, the immunophenotype of acute promyelocytic leukaemia is also very distinctive. Both in adults and children, APL blasts show a typical surface marker expression characterised by positivity for CD33, CD13, CD9, absence of HLA-DR and rare expression of CD10, CD7, and CD11b (Guglielmi et€al. 1998). The t(15;17) is the diagnostic hallmark of APL and initially had been considered to be present in all patients with this condition reviewed by Grimwade et€al. (2000). Conventional cytogenetics or Fluorescence in situ hybridization (FISH) enable the identification of both derivatives chromosome 15q and 17q (Fig.€ 2, panel C). In addition to the classic form of the t(15;17) translocation, the existence of cryptyic translocations or microinsertion have been reported (Grimwade et€al. 2000). The study of the PML distribution pattern in leukemic cells provides a rapid, specific, low-cost, and relatively simple diagnostic approach (Dyck et€ al. 1995; Falini et€al. 1997). Different from the wild-type (speckled) staining, which corresponds to the localization of PML into 5 to 20 discrete nuclear particles (so-called “nuclear bodies”), APL cells show a characteristic and easily distinguishable nuclear PML positivity known as “microspeckled,” resulting from the disruption of the nuclear bodies and redistribution of the protein into greater than 50 small
87
Fig.€2â•… Diagnosis of acute promyelocytic leukaemia. Representative morphology of M3-hypergranular (Panel a) and M3v (Panel b). Bone marrow smears were stained with May-Grumwald-Giemsa. ×1,000-fold magnification. Panel c: FISH analysis of an APL case. The result shown has been obtained by using the Vysis probe set, which is designed to detect only the PML-RARA fusion gene. It comprises a mixture of directly labelled probes: a PML probe, which begins in intron 7 and extends toward the centromere for 180€kb, and a RARA probe, which begins in intron 4 and extends toward the telomere for 400€kb, Panel d: Immunolabeling APL blast with thePG-M3MoAb shows a microgranular distribution of the PML/RARa protein within the nucleus. Panel e: Results of the RT-PCR analysis of the PML-RARA fusion gene. Ethidium bromide-stained agarose gel showing bcr1+ and bcr3+ nested PCR products according to the procedures in van Dongen et al. (1999). Pt.pos: positive samples, Pt:neg: negative samples; Ctrl neg: negative controls, MK: markers
88
A. Biondi et al.
granules/per cell. Either immunocytochemistry or immunofluorescence have been successfully used as detection systems. The use of Reverse-Transcription PCR (RT-PCR) for the detection of the PMLRARA and RARA/PML fusion genes is the only technique that defines the PML breakpoint type and that allows the definition of a correct strategy for subsequent MRD monitoring. The advantage of routinely using this assay at diagnosis to better address treatment (Miller et€ al. 1992; Biondi et€ al. 1992) has been subsequently validated in prospective multicenter reviewed by Grimwade (2002). According to most investigators, high-quality RNA and efficient RT are the crucial determinants for successful RT-PCR of PML-RARA (van Dongen et€ al. 1999). As shown in Fig.€2, panel E, a proper set of primers has been identified that allow the detection of all the different PML-RARA junctions, generated by the existence of different PML breakpoint regions in the PML locus and the presence of alternative PML splicing of PML transcripts. Moreover, the alternative usage of two alternative usage of RARA sites generates extra PML-RARA transcripts of different size (van Dongen et€al. 1999).
Treatment of APL Despite areas of controversy in the management of APL, there is reasonable consensus in both adults and children for the approach of induction therapy with simultaneous ATRA and anthracycline, two or three blocks of anthracycline-based consolidation with simultaneous ATRA, with or without cytarabine or nonintercalating agents, and 1 or 2€ years of maintenance therapy with intermittent ATRA, with or without 6-mercaptopurine and methotrexate. Consolidation may be risk adapted and usually based on the presenting white cell count (WCC). Many frontline protocols incorporate serial molecular monitoring during treatment and at three month intervals for 2€years following completion of therapy to detect persistent or recurrent molecular disease, which would be an indication for further treatment. There is no role for hematopoietic stem cell transplantation (HSCT) in frontline therapy, except for the small number of patients with persistent minimal residual disease (MRD) at the end of consolidation who may proceed to either allogenic or autologous transplantation based on molecular status after further treatment. While the incorporation of ATO in frontline studies appears encouraging, there are limited data regarding children and none based on a randomized comparison of ATO-based therapy against ATRA and anthracycline based therapy. Pediatric APL trials, which include ATRA with anthracycline-based chemotherapy, report a 5€year overall survival (OS) of 87 to 90%, disease-free survival (DFS) of 78 to 82% and event-free survival (EFS) of 71 to 77% (Mann et€al. 2001; de Botton et€ al. 2004; Ortega et€ al. 2005; Testi et€ al. 2005) (Table 1). However, ATRA and anthracycline-based chemotherapy is only appropriate for children confirmed positive for a fusion gene associated with ATRA sensitivity and should not be used for children with APL who lack these fusion genes, and who should receive standard intensive AML type therapy with anthracyclines and cytarabine.
Acute Promyelocytic Leukaemia
89
Table€1â•… Four studies of children with newly diagnosed APL treated with combinations of ATRA and anthracycline-based chemotherapya Study results German-Austrian- European Trial Swiss BFM APL 93 GIMEMA PETHEMA No. of children 22 31 124 66 Study period 1994–2000 1993–1998 1993–2000 1996–2004 Treatment: ATRA dose (mg/m2/d) 25(15) 45(7) 45 25 25 Anthracycline dose DNR 180 or IDR DNR 495 IDR 80â•›+â•›MTZ IDR 80/100â•›+â•›MTZ (mg/m2) 36â•›+â•›ADR 120 50 50 ATRA toxicity (%): APL DS 14 13 7.5 19.5 Definite 2 4.5 Indeterminate 6 15 Pseudotumour cerebri 5 16 9 6 Headache 27 39 13 30 Results: CR Rate % 95 97 96 92 Deaths in CR 0 0 0 0 NA NA 3 3 Molecular persistence at the end of consolidation 5-year EFS, % 76 71 76 77 5-year DFS, % NA 78 82 5-year OS, % 87 90 89 87 GIMEMA Gruppo Italiano Malattie Ematologiche Maligne dell’Adulto group; PETHEMA Programa de Estudio y Tratamiento de las Hemopatias Malignas Group; NA not available; DNR daunorubicin; IDR idarubicin; ADR adriamycin; MTZ mitoxantrone; CR complete remission; APL DS APL differentiation syndrome; EFS event-free survival; OS overall survival; DFS disease free survival a Adapted from Ortega et€al. (2005)
Induction Therapy The simultaneous use of ATRA and anthracycline monotherapy is considered optimal induction chemotherapy (Fenaux et€al. 1999; Sanz et€al. 2003). Primary resistance to ATRA and idarubicin is extremely rare and there may be no advantage for the addition of cytarabine or other chemotherapeutic agents to ATRA and anthracyclines in induction. A study, which did not include ATRA, found no advantage in CR or EFS for dual therapy with idarubicin and cytarabine compared to monotherapy with idarubicin (Avvisati et€al. 2002), although a larger anthracycline dose may have conferred an advantage in the monotherapy arm. However, an increase in the relapse rate, but no difference in CR, was reported when cytarabine was omitted from a schedule employing daunorubicin as the anthracycline (Ades et€al. 2006). Daunorubicin and idarubicin have not been randomly compared in APL although idarubicin is often favored because of its relatively long half-life and CNS penetration.
90
A. Biondi et al.
In two separate trials employing simultaneous anthracycline (idarubicin) and ATRA as induction therapy, the CR and EFS for children were reported at 92 and 96% and 77 and 76%, respectively (Ortega et€al. 2005; Testi et€al. 2005). While ATRA followed sequentially by chemotherapy is superior in terms of CR and RR to single agent ATRA or chemotherapy alone (Ortega et€al. 2005; Fenaux et€ al. 2000; Tallman et€ al. 1997; Kanamaru et€ al. 1995), the simultaneous use of chemotherapy in combination with ATRA is superior to the sequential use (de Botton et€ al. 2004; Fenaux et€ al. 1999; Burnett et€ al. 1999; Mandelli et€ al. 1997; Tallman et€al. 2002). The European APL 93 study (Fenaux et€al. 1999) randomized patients to simultaneous induction treatment (ATRA, with chemotherapy added on day 3) or sequential treatment (ATRA followed by chemotherapy after CR to a maximum of 90 days) in patients with a WCCâ•›<â•›5â•›×â•›109/L and allocated simultaneous treatment with ATRA and chemotherapy from day 1 to patients with a WCCâ•›>â•›5â•›×â•›109/L. Fifty-five percent of patients randomized to sequential treatment had chemotherapy added before CR achievement because of a rising WCC. The CR rate was similar for simultaneous and sequential treatment at 94 and 95% (pâ•›=â•›0.79) respectively, but the estimated RR at 2€ years was in favor of simultaneous treatment (pâ•›=â•›0.04). This implies that the advantage of simultaneous over sequential ATRA is due to a reduction in relapse risk rather than a significant increase in CR or reduction in early deaths. The number of children less than 18€years of age and treated on this trial was too small (31; 5%) to compare simultaneous with sequential treatment, but overall, there was no difference in outcome between adults and children, except for a significantly better survival in children after adjustment for WCC (Ortega et€al. 2005). Within the pediatric age group, there are limited data on the advantage of ATRA and anthracycline monotherapy over ATRA and polychemotherapy and on simultaneous over sequential ATRA with strategies extrapolated from adult studies. AML- BFM 93 reported a CR rate of 95% with a single induction death (4.5%) in a cohort of 22 children who received 3 days of ATRA followed by simultaneous multi-agent chemotherapy, compared to a CR rate of 64% with seven induction deaths (32%) in a historical cohort of 22 children who were treated with similar chemotherapy alone (Mann et€al. 2001) and no ATRA (pâ•›<â•›0.04). The GIMEMA-AIEOP (Testi et€al. 2005) and PETHEMA (Ortega et€ al. 2005) groups both delivered simultaneous chemotherapy with idarubicin and ATRA and reported CR rates of 96 and 92% and induction death rates of 4 and 7.5%, respectively. European APL 93 study using a combination of sequential and simultaneous treatment reported a CR rate of 97% with a single induction death (3%) (de Botton et€al. 2004). There are therefore clear advantages for the combined use of ATRA and chemotherapy with excellent CR rates being achieved and low induction death rates.
Complications of Induction Therapy Primary resistance of APL to simultaneous ATRA and anthracycline-based chemotherapy is exceptionally rare and the most common cause of failure to achieve CR is death during induction therapy from hemorrhage. A bleeding diathesis, particularly
Acute Promyelocytic Leukaemia
91
intracranial hemorrhage is more common in APL than in other AML subtypes, and most commonly involves children with a high presenting WCC (>10â•›×â•›109/L) and in those with M3 variant. The severe coagulopathy initiated or exacerbated by chemotherapy, can result in both hemorrhagic and thrombotic complications. While ATRA significantly reduces the incidence of intracranial hemorrhage, this complication remains a cause of early death in 3 to 4% of children despite increased awareness, preemptive therapy, and the use of ATRA (Mann et€al. 2001; de Botton et€al. 2004; Ortega et€al. 2005; Testi et€al. 2005). The risk of early hemorrhagic death, often prior to, or in the first week of treatment, classifies APL as a medical emergency, and ATRA, which improves the coagulopathy and may reduce the risk period, should be started immediately on morphological suspicion (Sanz et€ al. 1999; Falanga 2003; Huang et€al. 1988; Chomienne et€al. 1990; Tallman et€al. 2004a) and not delayed until the diagnosis is molecularly confirmed. The rapid initiation of combined ATRA and anthracycline-based chemotherapy is particularly important in patients with a high WCC, who should not undergo leucopheresis as this can exacerbate the coagulopathy (Tallman et€al. 2004b). Coagulation parameters remain abnormal (low fibrinogen, elevated d-dimers) until hypocellularity of the bone marrow is achieved; a median of about 9 days. The coagulation screen and platelet count should be checked twice daily during induction therapy. Fibrinogen levels should be maintained above 1.5 to 2€g/L with fresh frozen plasma (FFP) or cryoprecipitate and the platelet count above 50â•›×â•›109/L until the coagulopathy resolves (Sanz et€al. 1999; Chomienne et€al. 1990; Tallman et€al. 2004a, b; Milligan et€al. 2006). There is no proven benefit for the use of heparin, Tranexamic acid, or antifibrinolytic drugs and these drugs should not be used routinely. Indeed, anti-fibrinolytic agents, when combined with ATRA, may increase the risk of thrombosis. Anti-fibrinolytic agents and NOVOSEVEN may be considered in life-threatening hemorrhage unresponsive to platelets, FFP, and cryoprecipitate. The APL differentation syndrome (APL DS) is characterized by fluid retention and capillary leak and is associated with a high mortality if not promptly treated (Sanz et€al. 2005a; Vahdat et€al. 1994; Frankel et€al. 1992; De Botton et€al. 1998). It is thought to be related to surface adhesion molecule modulation and cytokine release following differentiation of APL cells and can occur at any time from day 1 to day 35 after the start of induction therapy with a mean of day 7 (Frankel et€al. 1992). Patients with a high WCC (>10â•›×â•›109/L) or a rising WCC are at greatest risk and should receive simultaneous chemotherapy and ATRA, which is reported to reduce the incidente (Frankel et€al. 1992; De Botton et€al. 1998). Early recognition and prompt treatment with dexamethasone reduces the associated mortalità (Milligan et€al. 2006; De Botton et€al. 1998). At the first sign or symptom suggesting APL differentation syndrome, dexamethasone should be promptly initiated at a dose of 5€mg/m2â•›bd (max single dose 10€mg) iv, until disappearance of symptoms and signs, and for a minimum of 3 days. ATRA should be stopped if symptoms are severe or unresponsive to steroid (Milligan et€ al. 2006; de Botton et€ al. 2003). ATRA can be cautiously reintroduced because the occurrence of APL differentiation syndrome during induction, linked to the differentiation of APL blasts, is not a contraindication to its subsequent use. Although the prophylactic use of steroids
92
A. Biondi et al.
with induction therapy is practised, its benefit is uncertain (Firkin et€ al. 1999; Wiley and Firkin 1995). APL differentiation syndrome appears to occur with the same incidence in children as in adults. In adults receiving combined ATRA and chemotherapy and the early use of dexamethasone, the incidence is reported at 15%, but with a low associated mortality of 1.2% (De Botton et€ al. 1998). The incidence in similarly treated children of definite APL differentiation syndrome is reported at 3 to 5% (de Botton et€al. 2004; Ortega et€al. 2005; Testi et€al. 2005) with an additional 5 to 15% having indeterminate APL differentiation syndrome and a mortality rate of less than 1% (Mann et€ al. 2001; de Botton et€ al. 2004; Ortega et€al. 2005; Testi et€al. 2005). Children have a higher incidence of ATRA-associated headache and pseudotumor cerebri than adults (Mahmoud et€ al. 1993). Symptoms are reversible with discontinuation of the drug, which may be tolerated when reinstituted at a lower dose. The incidence of headaches may be dose-dependent and age-dependent (<10€years). Doses as low as 25€mg/m2 may be effective in children (Mann et€al. 2001; Ortega et€al. 2005; Testi et€al. 2005; Castaigne et€al. 1993). The incidence of headache and pseudotumor cerebri has been reported between 13 to 30 and 5 to 9% respectively, for children receiving ATRA at a dose of 25€mg/m2 (Ortega et€al. 2005; Testi et€al. 2005) and 39 and 16%, respectively, for those receiving 45€mg/m2. There are no randomized studies comparing 25€mg/m2 with 45€mg/m2 in terms of relapse reduction but similar outcomes have been reported (Castaigne et€al. 1993). Three studies using an ATRA dose of 25€mg/m2 report EFS of 76 and 77% and OS of 87 and 89%, suggesting efficacy. Failure of ATRA may reflect a reduction in plasma drug levels due to induction of cytochrome p450 or due to cellular retinoic acid binding proteins (CRABPs) (Warrell 1993), which cannot be overcome by dose escalation. Care should be taken with the concomitant use of azoles and ATRA because ATRA is metabolized by cytochrome P450, which is inhibited by azoles. Resistance to retinoids may be due to mutations in the PML-RAR alpha fusion gene (Ding et€al. 1998). ATRA capsules can be difficult for small children to swallow. The pharmokinetics of formulations prepared from capsules and of such preparations given by nasogastric tube are poorly defined. Liposomal ATRA has the advantage of being an intravenous preparation.
Consolidation Therapy The controversial issues in consolidation therapy are the benefits of simultaneous ATRA with chemotherapy, of cytarabine or any chemotherapy additional to anthracyclines and of a risk-stratified approach. Morphological assessment at the end of induction therapy is uninformative. About 50% of patients have molecular evidence of residual disease at the end of induction, while more than 95% (Testi et€ al. 2005; Mandelli et€ al. 1997) are in molecular remission at the end of consolidation following two or three intensive anthracycline-based courses. There is no correlation between molecular positivity
Acute Promyelocytic Leukaemia
93
at the end of induction, but molecular positivity at the end of consolidation is highly predictive of relapse. ATRA may have a synergistic effect when given simultaneously with chemotherapy in consolidation, but this has not been demonstrated in a randomized trial. The use of ATRA for 28 days during induction and with two cycles of daunorubicin and cytarabine consolidation without subsequent maintenance reported comparable outcomes (5€year OS and LFS of 82 and 78%, respectively) to studies employing prolonged maintenance with ATRA with or without low dose chemotherapy, suggesting that ATRA in consolidation may allow shortening of the overall duration of APL treatment (Gupta et€ al. 2005). The PETHEMA group, employing anthracycline monotherapy in consolidation, reported a reduction in the relapse rate in high and intermediate risk patients from 20.1% to 8.7% (p╛=╛0.004) when ATRA was added to consolidation. The relapse rate fell from 14 to 2.5% (p╛=╛0.006) in intermediate risk patients (Sanz et€al. 2004). The role of cytarabine in consolidation therapy is questioned comparable results reported for anthracycline monotherapy and combination chemotherapy, commonly an anthracycline combined with cytarabine. However, some groups report benefit for high dose cytarabine in consolidation (Schlenk et€al. 2005), while others report encouraging results for anthracycline monotherapy (Tallman et€ al. 2002; Sanz et€al. 2004). Fifty-one percent of patients receiving anthracycline monotherapy in induction and consolidation were PCR negative after induction, and 93%, after consolidation. The OS, DFS, and cumulative RR at 5€years were 87, 82 and 17%, respectively (Ortega et€al. 2005) for the 66 children who received anthracycline monotherapy in consolidation with which ATRA was subsequently added for intermediate and high-risk patients defined by their presenting WCC and platelet counts. Combined analysis of the GIMEMA and PETHEMA APL trials, which differ only in the inclusion or exclusion of drugs other than anthracyclines during consolidation showed that the omission of nonanthracycline drugs was not associated with an inferior antileukaemia effect (Sanz et€al. 2000). The view that cytarabine plays a minor role in consolidation therapy has been challenged by the European APL Group who found benefit for cytarabine in combination with ATRA and daunorubicin in newly diagnosed patients with APL (Avvisati et€al. 2002). Any advantage for cytarabine may be protocol-dependent, and particularly, may relate to the anthracycline, its cumulative dose, and the content of consolidation and maintenance courses. Cytarabine or polytherapy consolidation may play a particular role in patients with a high WCC, a more common situation in childhood APL (Lo-Coco et€al. 2004a; Diverio et€al. 1998). Heterogeneous therapeutic regimens appear to produce similar results, but all deliver high cumulative doses of anthracycline, which may be very important in APL. While anthracycline monotherapy in consolidation has the advantage of reduced morbidity and mortality, there were no remission deaths in four reported pediatric studies (Mann et€al. 2001; de Botton et€al. 2004; Ortega et€al. 2005; Testi et€al. 2005) using variable consolidation courses. The optimal number of consolidation courses has not been identified by randomized trial. WCC at presentation is a strong predictive indicator of outcome
94
A. Biondi et al.
and the number and intensity of the blocks of consolidation may be tailored to the individual patient. Children with a low presenting WCC (<10â•›×â•›109/L) may require less treatment than those with a high presenting WCC (>10â•›×â•›109/L) and treatment may be further modified by response as assessed by MRD monitoring. In the GIMEMA-AIEOP study, the EFS was 83% versus 59% at 10€years for children with a low versus high WCC, respectively (Testi et€al. 2005). The WCC is not as predictive as MRD.
Maintenance Treatment A number of groups have reported that maintenance therapy, given for 2€ years, reduces the relapse risk, with the best results being obtained with intermittent ATRA (15 days every 3€months) in combination with 6-mercaptopurine and methotrexate (Testi et€al. 2005; Fenaux et€al. 1999; Tallman et€al. 1997). The intermittent administration of ATRA recognizes its pharmacological properties; particularly that continuous use leads to a fall in plasma levels secondary to hypercatabolism of the drug. The European APL study (Fenaux et€al. 1999) reported improved OS in patients who received maintenance chemotherapy (p╛=╛0.01) and a trend in favor of ATRA (p╛=╛0.22). The best results were obtained in patients who received both ATRA and low-dose maintenance chemotherapy, particularly patients with a high WCC, and therefore, the combined use of ATRA, 6MP, and MTX may have particular benefit for this group (Fenaux et€al. 1999). However, the GIMEMA study found no benefit for maintenance in patients who are in molecular remission at the end of consolidation (Avvisati et€ al. 2003), questioning its value for this cohort. Similarly, it is unclear if maintenance retinoid confers a survival advantage in patients who receive prolonged ATRA as a component of induction and consolidation chemotherapy. The benefit of maintenance may depend on the intensity of previous induction and consolidation therapy.
Molecular Monitoring in APL MRD Detection Is an Independent Predictor of Outcome in APL Presenting leucocyte count has been shown to be a key prognostic factor in clinical trials involving pediatric and adult patients with APL, with WCCâ•›>â•›10â•›×â•›109/L predicting a significantly increased risk of induction death and relapse risk reviewed by Grimwade (2002). Relapse risk among patients treated with ATRA and anthracycline-based chemotherapy with WBCâ•›<â•›10 is typically approximately 10%, as opposed to about 30% in those who are being overtreated with current protocols leading to unnecessary morbidity, while patients at higher risk of relapse could benefit from additional therapy. There is evidence to suggest that more precise tailoring of therapy may be achieved through
Acute Promyelocytic Leukaemia
95
MRD monitoring, which has been shown to provide an independent risk factor for relapse reviewed by Lo-Coco and Ammatuna (2006). Using conventional end-point PCR assays, which achieve a sensitivity of approximately 1 in 104, Italian studies undertaken in children and adults in conjunction with the AIDA protocol have established that patients with PML-RARA transcripts still detectable at the end of consolidation (who account for less than 10% of cases overall) or those subject to a later recurrence of PCR positivity (molecular relapse) are destined to undergo subsequent hematological relapse, which may however in both instances, be averted by additional therapy (Diverio et€al. 1998). Indeed, preliminary data from the Italian trials suggest that preemptive treatment at the time of molecular relapse may lead to improved survival in comparison to patients who are re-treated in frank hematological relapse (Lo Coco et€al. 1999). In addition, analysis of children and adults treated in the UK Medical Research Council (MRC) ATRA trial showed that MRD monitoring can distinguish subgroups of patients at differing risk of subsequent relapse relatively early during the treatment course. Patients with detectable fusion transcripts following three courses of treatment who accounted for 10% of patients had a significantly increased risk of subsequent relapse (57% vs. 26% at 5€years, p╛=╛0.008) associated with poorer overall survival (57% vs. 82%, p╛=╛0.03) compared to those testing PCR negative at the same stage (Burnett et€al. 1999). These findings have recently been supported by a study from the German AML Cooperative Group using real-time quantitative RT-PCR (RQ-PCR), in which patients who failed to achieve a 3-log reduction in PML-RARA transcript level within the first 3 to 4€ months of therapy, had an increased risk of early relapse (Schnittger et€al. 2003).
Real-Time Quantitative RT-PCR Assays Enhance MRD Detection in APL While MRD monitoring using conventional nested RT-PCR has provided valuable prognostic information, its clinical utility has been somewhat compromised by failure to detect residual disease in a significant proportion of APL patients who ultimately relapse reviewed by Grimwade (2002). This may be a reflection of the relatively limited sensitivity of conventional assays and/or variation in RNA quality/ quantity and efficiency of the reverse transcription (RT) step. Quantitative PCR approaches using hydrolysis (Taqman) or hybridization probe technology afford a number of advantages in comparison to conventional end-point assays. In particular, quantitation of fusion gene and endogenous control gene transcripts enables more reliable determination of kinetics of molecular remission achievement or relapse and readily identifies poor quality samples that could potentially give rise to falsenegative results. Moreover, RQ-PCR assays are rapid, facilitate high throughput sample analysis, are highly reproducible, and readily standardized, thereby lending themselves to MRD assessment in multicenter clinical trials. Optimized RQ-PCR protocols for detection of the PML-RARA fusion gene and ubiquitously expressed control genes have been established by the Europe Against Cancer (EAC) Group
96
A. Biondi et al.
(Gabert et€al. 2003). These assays have subsequently been validated in relation to conventional nested RT-PCR (Gallagher et€al. 2003). However, despite the higher sensitivity of RQ-PCR, a significant number of patients who ultimately relapsed tested PCR negative at the end of consolidation, underlining the importance of serial monitoring to increase the predictive value of this approach.
Anthracycline Cumulative Dose and Cardiotoxicity Idarubicin and daunorubicin have not been randomly compared in APL. Idarubicin is more often used as monotherapy and daunorubicin in combination therapy. The optimum cumulative dose of anthracycline in children is unknown. A high cumulative dosage of anthracyclines (650€ mg/m2) was delivered to children in the GIMEMAAIEOP AIDA protocol with no serious cardio-toxicity observed in the 101 long-term survivors, but late cardiotoxicity may occur (Testi et€al. 2005). While anthracyclines are very effective in APL, the favorable outcome increases the importance of minimizing late effects and clearly the lowest effective dose should be identified. The role of liposomal anthracycline and cardioprotectants have not been evaluated in a clinical trial.
Stem Cell Transplantation Because of the favorable results achieved with chemotherapy and ATRA, Hematopoietic stem cell transplantation (HSCT), allogenic or autologous, are not recommended in first remission (Burnett et€al. 1998, 2002). Patients with persistent molecular disease at the end of consolidation may be candidates for HSCT after further chemotherapy.
Arsenic Trioxide ATO, like ATRA, is a differentiating agent, which offers the potential for a non or low chemotherapy approach to the treatment of APL. While the data on ATO use in children are limited, it may be the most effective single agent in APL. Preliminary data from adult studies suggest at least comparability between ATO and single agent ATRA in terms of the achievement of CR (>90%) and that combined ATO and ATRA is superior to ATO or ATRA alone (Niu et€al. 1999; Soignet et€al. 2001; Shen et€al. 2004). ATRA and ATO may act in part synergistically in triggering down regulation of telomerase causing telomerase shortening and subsequent cell death (Tarkanyi et€al. 2005). A small single center pediatric study using monotherapy with ATO reported a haematological and molecular remission rate of 91% with a RFS and OS of 81 and 91%, respectively at 30€months (George et€al. 2004). The addition of ATO to consolidation following remission induction with ATRA and chemotherapy has been reported to significantly improve the EFS and OS in adults with newly
Acute Promyelocytic Leukaemia
97
diagnosed APL (Powell 2007). While ATO appears to be excellent at inducing remission, its long-term benefit and comparability with ATRA-based chemotherapy is less clearly defined in frontline therapy, especially in children, where there are many unanswered questions. ATO may have a role in both induction and consolidation, combined with either ATRA or chemotherapy and used with the aim of minimizing or eliminating chemotherapy. ATO is usually well tolerated, although its use is associated with serious adverse events including hyperleukocytosis, the APL differentiation syndrome, and prolongation of QT interval. Although severe neuropathy has been rarely observed, mild peripheral neuropathy, which usually resolves after ATO discontinuation, has been reported in approximately 40% of patients reviewed by Sanz et€al. (2005b). Little is known of late ATO cardio and neurological long-term toxicity. There is a need to define toxicity in children and to assess the role and best strategy for the use of ATO within a pediatric study.
What to Do When Treatment Fails? Failure to treatment currently constitutes an infrequent event in APL, involving only one-fourth of patients, owing to the high efficacy of frontline therapy. Failure is represented by disease recurrence at hematological and/or molecular level (relapse rate 20 to 25%) and by the persistence of molecular disease at recovery from 2 to 3 consolidation cycles (about 3 to 5%). Relapse occurs more frequently in those patients with a WCC countâ•›>â•›10â•›×â•›109/L at presentation (Tallman et€al. 2002; Sanz et€al. 2004; Tallman 2007). Approximately 3 to 5% of APL patients develop extramedullary (EM) sites of relapse, predominating in the central nervous system (CNS) and the skin, followed by other sites (testes, sites of vascular access, external ear, and auditory canal). It has been suggested that EM relapse also may be associated with initial adverse prognostic features such as high WCC count, and bcr3 PML-RARA isoform, but the risk factors for this kind of relapse have not been precisely defined (Tallman 2007; de Botton et€al. 2006; Specchia et€al. 2001; Breccia et€al. 2003). Nonetheless, the adequate management of relapsed APL patients is still unclear, and few studies have analyzed the outcome of these patients in a prospective fashion. In particular, very limited information is available in paediatric age, with only few sporadic cases, mainly included in adult series, described in literature. In any case, the prognosis following hematological relapse is usually dismal; on the contrary, institution of therapy at the first evidence of molecular disease results in more favorable outcome, reinforcing the rationale of PCR-triggered therapy in APL (Lo Coco et€al. 2003). Reinduction with ATRA alone in patients did not provide long-term benefit, since second complete remission (CR) achieved (approximately 50%) in this manner, lacked durability and required consolidation by further intensive chemotherapy (Castagnola et€al. 1998). Better results have been achieved in relapsed APL with salvage therapy consisting of ATRA and chemotherapy for induction, generally containing high-dose cytarabine, followed by further chemotherapy and/or hematopoietic stem cell transplant (HSCT) (CR rate 87%; disease-free-survival 50 to 60
98
A. Biondi et al.
(99, 100). In these studies, the choice of transplant modality was mainly based on PCR status achieved with chemotherapy: autologous HSCT was the preferred option in patients without detectable MRD, while allogeneic HSCT was chosen for patients failing to achieve a second molecular remission. However, the outcome of patients treated in molecular relapse compared favorably to those who received the treatment for hematological relapse, with longer survival and lower relapse risk (Specchia et€ al. 2001). Additionally, molecular response to salvage therapy predicted longer leukaemia-free survival (LFS), independently by the therapeutic consolidation choice (Castagnola et€al. 1998; Esteve et€al. 2007). Currently, given the high antileukemic efficacy observed with ATO in APL patients relapsing after ATRA-containing regimen, this agent is regarded as the best option in this context. ATO has emerged as the single most active agent in patients with relapsed APL (Tallman 2007; Lo Coco et€al. 2007; Ghavamzadeh et€al. 2006). In the several published trials with relatively small numbers of patients, the CR rate ranges from approximately 80 to 90%, accompanied by molecular remission after two cycles in the majority of cases. ATO is administered at the dosage of 0.15€mg/ kg/day by 3€hours intravenous infusion, until CR achievement, and for a maximum of 50 days. Combining ATO with other agents may also be useful: in a report from China, the combination of ATO and anthracycline resulted in a superior outcome compared to ATO alone. However, the addition of ATRA with ATO, providing high CR rate in newly diagnosed patients, does not appear to be beneficial in the setting of relapse disease. In the randomised French study, ATO alone was compared to ATO plus ATRA in 20 patients with relapsed APL. Haematological and molecular response (8/10 patients in each treatment group), time necessary to reach CR (42 days in the two groups) and outcome (2-year overall survival 59% in the two groups) were comparable in both treatment groups (Tallman 2007; Lo-Coco et€al. 2004b, 2007; Ghavamzadeh et€al. 2006; Raffoux et€al. 2003). Also in these experiences, best results are achieved in patients treated in molecular relapse; in these cases molecular and durable remission can be obtained after ATO monotherapy with less risk of ATO-related toxicity. The quality of CR obtained after ATO salvage treatment remains an open question. However, the best consolidation strategy after ATO-induced second remission is unknown. Several options are available, including repeated cycles of ATO (0.15€ mg/kg/day, ×5 days ×5€ weeks; 1 to 2 courses), combination with standard chemotherapy, HSCT, and recently, antiCD33. Gemtuzumab ozogamicin combines an anti-CD33 antibody with calicheamicin a cytotoxic agent with similarities to anthracyclines, and therefore, this targeted therapy might be expected to have activity in APL. CD33 is expressed in virtually 100% of APL cases with a highly homogenous expression pattern and calicheamicin belongs to the anthracycline family. While it has been used mainly in relapsed disease, it has shown efficacy in adult frontline studies (Lo-Coco et€al. 2004b; Takeshita et€al. 2005). CR rates comparable to ATRA and anthracycline have been reported for ATRA and Gemtuzumab (Takeshita et€al. 2005). Nonetheless, the precise role of this agent in the management of relapsed APL remains unsettled. In a limited series, molecular response was obtained in 9/11 (91%) patients tested by RT-PCR after two doses and in all 13 (100%) patients tested after the third dose. GO resulted in mild
Acute Promyelocytic Leukaemia
99
myelosuppression and little extrahematologic toxicity, even in heavily pretreated patients (Zheng et€al. 2007). Furthermore, it is presumable that combining GO with simultaneous ATRA or other agents, such as ATO, may result in improved outcome also in more advanced APL (Takeshita et€al. 2005; Zheng et€al. 2007). The selection of one of the above-mentioned options for consolidation therapy, as well as the modality of HSCT, should take into account various variables that may influence the outcome. Recent data published from the EBMT Group indicate that HSCT has continued to be part of treatment strategy for patients with second CR (CR2) (Sanz et€al. 2007). These results show that a high proportion of patients in CR2 achieve a long-term overall survival (OS) after auto- and allo-HSCT, and both procedures present valid therapeutic options in this setting. The choice of one or other procedures will depend on the molecular status, age, the availability of an HLA identical donor and the time from diagnosis to transplant. For the 195 autografted patients in CR2, the 5-year cumulative incidence of LFS was 51% and the transplant-related mortality (TRM) 16%; for the 137 patients allografted in CR2, the 5-year LFS and TRM were 59% and 24%, respectively. The use of peripheral blood stem cells was associated with decreased TRM as compared to bone marrow (12 and 31%). In the near future the results of all consolidation procedures in similar series of CR2 patients need to be compared for better understanding how to improve the prognosis in this group of patients.
Extramedullary Relapse Because of the rarity of this event in APL, the outcome of these patients remains undetermined and only few studies have been reported in literature. However, since the introduction of ATRA in the treatment of this disease, cases of EM relapse have been increasingly described in the last years (3-year cumulative incidence from 1 to 5 to 12% of the first relapse) (Tallman 2007; de Botton et€al. 2006; Specchia et€al. 2001; Breccia et€ al. 2003). The reason for the potential increase in EM relapse observed with ATRA treatment are unknown. One possibility is that the longer survival of patients may increase the number of patients at risk of relapse. The role of ATRA in mediating increased expression of adhesion molecules has been implicated as possible mechanism in the infiltration of EM sites. EM relapses largely predominate in the CNS and are frequently associated with overt or molecular relapse (about 80% of cases). Some reports found, in patients with EM relapse, a high incidence of presenting features such as microgranular M3 variant, the bcr3 PML-RARalpha isoform, both features being correlated to high WBC counts and younger age. Finally, it has been suggested that occurrence of APL differentation syndrome during induction, might be associated with an increased risk of EM relapse (Tallman 2007; de Botton et€al. 2006; Specchia et€al. 2001; Breccia et€al. 2003; Knipp et€al. 2007). The available data suggest that EM relapse in patients with APL should be regarded as systemic disease. CNS-directed therapy with intrathecal methotrexate and cytarabine and cranial radiotherapy, has to be associated to systemic treatment
100
A. Biondi et al.
including drugs administered at high doses and HSCT. Heterogeneous protocols have been used for the treatment of EM relapse; second CR is achieved in almost 90% of the patients. Despite intensive systemic therapy followed by autologous or allogeneic HSCT in most cases, patients with EM relapse have a poor outcome and the overall survival is comparable to those patients who relapsed at hematological level only. The new drugs, such as ATO, alone or in combination with other chemotherapeutic agents, were recently employed in the treatment of EM relapse; ATO crosses the blood-brain-barrier but the concentrations in Cerebral Spinal Fluid (CSF) achieved by intravenous infusion, are probably insufficient for the treatment of meningeal leukemia (Knipp et€ al. 2007). Also, the described ATO neurological toxicity might be increased by the previous or subsequent intrathecal and radiotherapy, necessary for the whole treatment of CNS relapse. For these reasons and because of the small number of patients, it is difficult to give more exact indications about the treatment of CNS relapse in APL. In response to the above considerations, many institutions have adopted CNS intrathecal prophylaxis in the frontline protocols (starting from consolidation phase), only for those patients with unfavorable presenting features, at higher risk of CNS relapse.
New Drugs Recent novel therapeutic agents potentially effective in APL are currently under investigation. In particular, there is great promise for the FLT-3 inhibitors in APL. FLT-3 aberrations, in the form of an internal tandem duplication (ITD) or mutation at the activation loop position 835, D835, have been reported in 20 to 45% of APL cases (Arrigoni et€al. 2003; Au et€al. 2004; Sohal et€al. 2003). The FLT-3 ITD has also frequently associated with high WBC, microgranular morphology, and PML breakpoint at bcr3. The FLT-3 inhibitors have been shown to be effective with ATRA in a APL mouse model (Fazi et€al. 2005). These raise the enticing possibility that APL may be more effectively managed in the future using orally bioavailable drugs that molecular target both classes of mutation that contribute to the pathogenesis of APL: FLT-3 inhibitors for FLT-3 ITD and ATRA that targets the PML/ RARalpha fusion. The upfront combination of ATRA and FLT-3 antagonists could be highly effective for the treating of APL patients possessing these adverse factors. On the contrary, FLT-3 aberrations might only exist in subclones or represent secondary changes in relapse, and finally might be lost at relapse. The use of FLT-3 antagonists in APL patients in relapse might not be warranted. These data should be addressed in future large prospective trials. A potential role of histone deacetylase (HDAC) inhibitors has been also suggested in the setting of APL. The addition of HDAC inhibitor to ATRA, may restore the sensitivity to retinoids in ATRA-resistant relapse. The effects of a novel mixed retinoic/butyric hyaluronan ester (HBR) were explored in€vitro and in€ vivo on a retinoic acid (RA)-sensitive human myeloid cell line and on its RA-resistant subclone. The results showed that HBR blocks cell growth in both RA-sensitive and RA-resistant cell lines and induces terminal granulocytic
Acute Promyelocytic Leukaemia
101
differentiation in RA-sensitive cells and apoptosis in RA-resistant cells. The antiproliferative effect observed in€vitro is confirmed by in€vivo significant prolongation of the lifetime expectancy in treated immunodeficient mice. Also, HBR increased the host life spam similarly to maximum tolerated doses of RA alone. HBR, due to its strong affinity for the CD44 membrane receptors could be a promising antineoplastic agent for APL patients regardless of their possible resistance to RA. Other retinoid/HDCA inhibitor associations are currently under investigation with promising results. Finally, a possible role for alemtuzumab, an anti-CD52 monoclonal antibody, has also been suggested in APL. ATRA treatment of a patient’s APL cells in€vitro, induced high level of CD52 expression on the leukemic cells. This antigen may become a new target for antibody therapy in patients with APL.
Conclusions Much of the data on which strategies for the treatment of children with APL are based have been extrapolated from adult trials. Questions which require to be addressed exclusively in children include both the optimal ATRA and cumulative anthracycline dose, the role of lipsomal anthracyclines and cardioprotectants, the optimal risk-adapted therapy (as children often have a higher WCC at presentation compared to adults), optimal maintenance and its duration, the role of GO and ATO in combination with traditional first line therapies, the ideal frequency of molecular monitoring and best preemptive therapy, and finally, the role that novel therapies will play in this disease.
References Degos L: The history of acute promyelocytic leukaemia. Br J Haematol. 2003; 122(4):539–553. Bernard J, Weil M, Boiron M, Jacquillat C, Flandrin G, Gemon MF: Acute promyelocytic leukemia: results of treatment by daunorubicin. Blood 1973: 41:489–496. Grignani F, Fagioli M, Alcalay M, Longo L, Pandolfi PP, Donti E, Biondi A, Lo Coco F, Pelicci PG: Acute promyelocytic leukemia: from genetics to treatment. Blood 1994: 83:10–25. Sanz MA: Treatment of acute promyelocytic leukemia. Hematology Am Soc Hematol Educ Program. 2006:147–55. Review Kaspers GJ, Creutzig U. Pediatric acute myeloid leukemia: international progress and future directions. Leukemia. 2005; 19(12):2025–2029. Review Ribeiro RC and Rego E: Management of APL in developing countries: epidemiology, challenges and opportunities for international collaboration.Hematology Am Soc Hematol Educ Program. 2006:162–8. Review. Douer D. The epidemiology of acute promyelocytic leu-kaemia. Best Pract Res Clin Haematol. 2003;16:357–367. Biondi A, Rovelli A, Cantu-Rajnoldi A, Fenu S, Basso G, Luciano A, Rondelli R, Mandelli F, Masera G, Testi AM: Acute promyelocytic leukemia in children: experience of the Italian Pediatric Hematology and Oncology Group (AIEOP). Leukemia 1994: 8:1264–1268. Tomas JF, Fernandez-Ranada JM. About the increased frequency of acute promyelocytic leukemia among Latinos: the experience from a center in Spain. Blood. 1996;88:2357– 2358.
102
A. Biondi et al.
Ruiz-Arguelles GJ. Promyelocytic leukemia in Mexican Mestizos. Blood. 1997;89:348–349. Otero JC, Santillana S, Fereyros G. High frequency of acute promyelocytic leukemia among Latinos with acute myeloid leukemia. Blood. 1996;88:377. Malta Corea A, Pacheco Espinoza C, Cantu Rajnoldi A, Conter V, Lietti G, Masera G, Sessa C, Cavalli F, Biondi A, Rovelli A: Childhood acute promyelocytic leukemia in Nicaragua. Ann Oncol 1993: 4:892–894. Melnick A, Licht JD: Deconstructing a disease: RARalpha, its fusion partners, and their roles in the pathogenesis of acute promyelocytic leukemia.Blood. 1999, 93(10):3167–3215. Review. No abstract available. Redner RL, Rush EA, Faas S, Rudert WA, Corey SJ: The t(5;17) variant of acute promyelocytic leukemia expresses a nucleophosmin-retinoic acid receptor fusion. Blood. 1996; 87(3):882–886. Wells RA, Catzavelos C, Kamel-Reid S: Fusion of retinoic acid receptor alpha to NuMA, the nuclear mitotic apparatus protein, by a variant translocation in acute promyelocytic leukaemia. Nat Genet. 1997; 17(1):109–113. Grimwade D, Biondi A, Mozziconacci MJ, Hagemeijer A, Berger R, Neat M, Howe K, Dastugue N, Jansen J, Radford-Weiss I, Lo Coco F, Lessard M, Hernandez JM, Delabesse E, Head D, Liso V, Sainty D, Flandrin G, Solomon E, Birg F, Lafage-Pochitaloff M: Characterization of acute promyelocytic leukemia cases lacking the classic t(15;17): results of the European Working Party. Groupe Francais de Cytogenetique Hematologique, Groupe de Francais d’Hematologie Cellulaire, UK Cancer Cytogenetics Group and BIOMED 1 European Community-Concerted Action “Molecular Cytogenetic Diagnosis in Haematological Malignancies”. Blood. 2000 Aug 15;96(4):1297–1308. Mistry AR, Pedersen EW, Solomon E, Grimwade D. The molecular pathogenesis of acute promyelocytic leukaemia: implications for the clinical management of the disease. Blood Rev. 2003;17(2):71–97. Review. Arnould C, Philippe C, Bourdon V, Gr goire MJ, Berger R, Jonveaux P: The signal transducer and activator of transcription STAT5b gene is a new partner of retinoic acid receptor alpha in acute promyelocytic-like leukaemia. Hum Mol Genet. 1999;8(9):1741–1749. Catalano A, Dawson MA, Somana K, Opat S, Schwarer A, Campbell LJ, Iland H. The PRKAR1A gene is fused to RARA in a new variant acute promyelocytic leukemia. Blood. 2007 [Epub ahead of print]. Umesono K, Murakami KK, Thompson CC, Evans RM: A direct repeat in the cellular retinolbinding protein type II gene confers differential regulation by RXR and RAR. Cell. 1991;66(3):555–561. Alland, L., Muhle, R., Hou, H., Jr., Potes, J., Chin, L., Schreiber-Agus, N. and DePinho, R. A.: Role for N-CoR and histone deacetylase in Sin3-mediated transcriptional repression. Nature. 1997;387(6628):49–55. Laherty, C. D., Yang, W. M., Sun, J. M., Davie, J. R., Seto, E. and Eisenman, R. N: Histone deacetylases associated with the mSin3 corepressor mediate mad transcriptional repression. Cell. 1997;89(3):349–356. Heinzel, T., Lavinsky, R. M., Mullen, T. M., Soderstrom, M., Laherty, C. D., Torchia, J., Yang, W. M., Brard, G., Ngo, S. D., Davie, J. R., Seto, E., Eisenman, R. N., Rose, D. W., Glass, C. K. and Rosenfeld, M. G: A complex containing N-CoR, mSin3 and histone deacetylase mediates transcriptional repression. Nature. 1997;387(6628):43–48. Redner, R. L., Wang, J. and Liu, J. M.: Chromatin remodeling and leukemia: new therapeutic paradigms. Blood. 1999;94(2):417–428. Review. No abstract available. Tsai, S. and Collins, S. J.: A dominant negative retinoic acid receptor blocks neutrophil differentiation at the promyelocyte stage. Proc Natl Acad Sci U S A. 1993;90(15):7153–7157. Collins, S. J.: Acute promyelocytic leukemia: relieving repression induces remission.Blood. 1998 Apr 15;91(8):2631–2633. Review. No abstract available. Koken, M. H., Puvion-Dutilleul, F., Guillemin, M. C., Viron, A., Linares-Cruz, G., Stuurman, N., de Jong, L., Szostecki, C., Calvo, F., Chomienne, C. and et€ al.: The t(15;17) translocation alters a nuclear body in a retinoic acid-reversible fashion.EMBO J. 1994 Mar 1;13(5): 1073–1083.
Acute Promyelocytic Leukaemia
103
Nervi, C., Ferrara, F. F., Fanelli, M., Rippo, M. R., Tomassini, B., Ferrucci, P. F., Ruthardt, M., Gelmetti, V., Gambacorti-Passerini, C., Diverio, D., Grignani, F., Pelicci, P. G. and Testi, R.: Caspases mediate retinoic acid-induced degradation of the acute promyelocytic leukemia PML/RARalpha fusion protein. Blood. 1998;92(7):2244–2251. Zhu, J., Gianni, M., Kopf, E., Honore, N., Chelbi-Alix, M., Koken, M., Quignon, F., RochetteEgly, C. and de The, H.: Retinoic acid induces proteasome-dependent degradation of retinoic acid receptor alpha (RARalpha) and oncogenic RARalpha fusion proteins. Proc Natl Acad Sci U S A. 1999;96(26):14807–14812. Liu, T. X., Zhang, J. W., Tao, J., Zhang, R. B., Zhang, Q. H., Zhao, C. J., Tong, J. H., Lanotte, M., Waxman, S., Chen, S. J., Mao, M., Hu, G. X., Zhu, L. and Chen, Z: Gene expression networks underlying retinoic acid-induced differentiation of acute promyelocytic leukemia cells. Blood. 2000;96(4):1496–1504. Rovelli A, Biondi A, Cantu Rajnoldi A, Conter V, Giudici G, Jankovic M, Locasciulli A, Rizzari C, Romitti L, Rossi MR, et€ al.: Microgranular variant of acute promyelocytic leukemia in children.J Clin Oncol. 1992 Sep;10(9):1413–1418. Sainty D, Liso V, Cantu-Rajnoldi A, Head D, Mozziconacci MJ, Arnoulet C, Benattar L, Fenu S, Mancini M, Duchayne E, Mahon FX, Gutierrez N, Birg F, Biondi A, Grimwade D, Lafage-Pochitaloff M, Hagemeijer A, Flandrin G: A new morphologic classification system for acute promyelocytic leukemia distinguishes cases with underlying PLZF/RARA gene rearrangements. Group Francais de Cytogenetique Hematologique, UK Cancer Cytogenetics Group and BIOMED 1 European Coomunity-Concerted Acion “Molecular Cytogenetic Diagnosis in Haematological Malignancies”. Blood. 2000 Aug 15;96(4): 1287–1296. Guglielmi C, Martelli MP, Diverio D, Fenu S, Vegna ML, Cantu-Rajnoldi A, Biondi A, Cocito MG, Del Vecchio L, Tabilio A, Avvisati G, Basso G, Lo Coco F: Immunophenotype of adult and childhood acute promyelocytic leukaemia: correlation with morphology, type of PML gene breakpoint and clinical outcome. A cooperative Italian study on 196 cases. Br J Haematol. 1998 Sep;102(4):1035–1041. Dyck JA, Warrell RP, Evans RM, Miller WH: Rapid diagnosis of acute promyelocytic leukemia by immunohistochemical localization of PML/RAR-alpha protein. Blood 1995 86:862–7. Falini B, Flenghi L, Fagioli M, Coco FL, Cordone I, Diverio D, Pasqualucci L, Biondi A, Riganelli D, Orleth A, Liso A, Martelli MF, Pelicci PG, Pileri S: Immunocytochemical diagnosis of acute promyelocytic leukemia (M3) with the monoclonal antibody PG-M3 (antiPML). Blood 1997; 90:4046–4053. Miller WH, Kakizuka A, Frankel SR, Warrell RP, DeBlasio A, Levine K, Evans RM, Dmitrovsky E: Reverse transcription polymerase chain reaction for the rearranged retinoic acid receptor alpha clarifies diagnosis and detects minimal residual disease in acute promyelocytic leukemia. Proc Natl Acad Sci U S A 1992; 89:2694–2698. Biondi A, Rambaldi A, Pandolfi PP, Rossi V, Giudici G, Alcalay M, Lo Coco F, Diverio D, Pogliani EM, Lanzi EM, et€al.: Molecular monitoring of the myl/retinoic acid receptor-alpha fusion gene in acute promyelocytic leukemia by polymerase chain reaction. Blood 1992; 80:492–497. Grimwade D, Lo Coco F. Acute promyelocytic leukemia: a model for the role of molecular diagnosis and residual disease monitoring in directing treatment approach in acute myeloid leukemia. Leukemia. 2002;16:1959–1973. van Dongen JJ, Macintyre EA, Gabert JA, Delabesse E, Rossi V, Saglio G, Gottardi E, Rambaldi A, Dotti G, Griesinger F, Parreira A, Gameiro P, Diaz MG, Malec M, Langerak AW, San Miguel JF, Biondi A: Standardized RT-PCR analysis of fusion gene transcripts from chromosome aberrations in acute leukemia for detection of minimal residual disease. Report of the BIOMED-1 Concerted Action: investigation of minimal residual disease in acute leukemia. Leukemia 1999; 13:1901–1928. Mann G, Reinhardt D, Ritter J, Hermann J, Schmitt K, Gadner H, and Creutzig U. Treatment with all-trans retinoic acid in acute promyelocytic leukemia reduces early deaths in children. Annals of Hematology 2001; 80: 417–422.
104
A. Biondi et al.
de Botton S, Coiteux V, Chevret S, Raynon C, Vilmer E, Sanz M, de La Serna J, Philippe N, Baruchel A, Leverger G, Robert A, San Miguel J, Conde E, Sotto JJ, Bordessoule D, Fegueux N, Fey M, Parry A, Chomienne C, Degos L, Fenaux P. Outcome of childhood acute promyelocytic leukemia with all-trans-retinoic acid and chemotherapy J Clin Oncol 2004; 22:1404–1412. Ortega JJ, Madero L, Martin G, Verdeguer A, Garcia P, Parody R, Fuster J, Molines A, Novo A, Debén G, Rodriguez A, Conde E, de la Serna J, Allegue MJ, Capote FJ, González JD, Bolufer P, González M, Sanz MA. Treatment with all-trans retinoic acid and anthracycline monochemotherapy for children with acute promyelocytic leukemia: a multicenter study by the PETHEMA Group. J Clin Oncol 2005; 23: 7632–7640. Testi AM, Biondi A, Lo-Coco F, Moleti ML, Giona F, Vignetti M, Menna G, Locatelli F, Pession A, Barisone E, De Rossi G, Diverio D, Micalizzi C, Aricò M, Basso G, Foa R, Mandelli F. GIMEMA-AIEOP AIDA protocol for the treatment of newly diagnosed acute promyelocytic leukemia (APL) in children. Blood 2005; 106: 447–453. Fenaux P, Chastang C, Chevret S, Sanz M, Dombret H, Archimbaud E, Fey M, Rayon C, Huguet F, Sotto JJ, Gardin C, Makhoul PC, Travade P, Solary E, Fegueux N, Bordessoule D, Miguel JS, Link H, Desablens B, Stamatoullas A, Deconinck E, Maloisel F, Castaigne S, Preudhomme C & Degos L. A randomized comparison of all trans-retinoic acid (ATRA) followed by chemotherapy and ATRA plus chemotherapy and the role of maintenance therapy in newly diagnosed acute promyelocytic leukemia. Blood 1999; 94: 1192–1200. Sanz MA, Martin G and Lo Coco F. Choice of chemotherapy in induction,consolidation and maintenance in acute promyelocytic leukaemia. Best Pract Res Clin Haematol 2003: 16(3): 443–451. Avvisati G, Petti MC, Lo Coco F, Vegna ML, Amadori S, Baccarani M, Cantore N, Di Bona E, Ferrara F, Fioritoni G, Gallo E, Invernizzi R, Lazzarino M, Liso V, Mariani G, Ricciuti F, Selleri C, Sica S, Veneri D, Mendelli F, GIMEMA Italian Coopertive Group. Induction therapy with idarubicin alone significantly influences event-free survival duration in patients with newly diagnosed hypergranular acute promyelocytic leukaemia: final results of the GIMEMA randomized study LAP 0389 with 7 years of minimum follow-up. Blood 2002; 100: 3141–6. Ades L, Chevret S, Raffoux E, de Botton S, Guerci A, Pigneux A, Stoppa AM, Lamy T, RigalHuget F, Vekhoff A, Meyer-Monard S, Maloisei F, Deconinck E, Ferrant A, Thomas X, Fegueux N, Chomienne C, Dombret H, Degos L, Fenaux P, European Acute Promyelocytic Group. Is cytarabine useful in the treatment of acute promyelocytic leukaemia? Results of a randomized trial from the European Acute Promyelocytic Leukaemia Group. J Clin Oncol 2006; 24(36): 5703–5710. Fenaux P, Chevret S, Guerci A, Fegueux N, Dombret H, Thomas X, Sans M, Link H, Maloisel F, Gardin C, Bordessoule D, Stoppa AM, Sadoun A, Muus P, Wandt H, Mineur P, Whittaker JA, Fey M, Daniel MT, Castaigne S, Degos L. Long-term follow-up confirms the benefit of alltrans retinoic acid in acute promyelocytic leukaemia. European APL Group. Leukaemia 2000; 14; 1371–1377. Tallman MS, Andersen JW, Schiffer CA, Appelbaum FR, Feusner JH, Ogden A, Shepherd L, Willman C, Bloomfield CD, Rowe JM & Wiernik PH All-trans-retinoic acid in acute promyelocytic leukemia. New England Journal of Medicine 1997; 337: 1021–1028. Kanamaru A, Takemoto Y, Tanimoto M, Marakami H, Asou N, Kobayashi T, Kuriyama K, Ohmoto E, Shakamaki H, Tsubaki K. All-trans retinoic acid for the treatment of newly diagnosed acute promyelocytic leukaemia Blood 1995: 85:1202–1206. Burnett A.K., Grimwade D, Solomon E, Wheatley K, Goldstone A.H on behalf of the MRC adult leukaemia working party Presenting white cell blood count and kinetics of molecular remission predict prognosis in acute promyelocytic leukaemia treated with all-trans retinoic acid: result of the randomised MRC trial. Blood 1999; 93, 4131–4143. Mandelli F, Diverio D, Avvisati G, Luciano A, Barbui T, Bernasconi C, Broccia G, Cerri R, Falda M, Fioritoni G, Leoni F, Liso V, Concetta Petti, M, Rodeghiero F, Sagilo G, Luce Vegna M, Visani G, Jehn U, Willemze R, Muus P, Pelicci PG, Biondi A, and Lo Coco F. for the Gruppo Italiano Malattie Ematologiche Maligne dell’ Adulto and Associazione Italiana di Ematologica
Acute Promyelocytic Leukaemia
105
ed Oncologia Pediatrica Cooperative groups Molecular remission in PML/RARAa positive acute promyelocytic leukemia by combined all-trans retinoic acid and idarubicin (AIDA) therapy. Blood 1997; 90:1014–1021. Tallman MS, Nabhan CH, Feusner JH, Rowe JM. Acute promyelocytic leukemia: evolving therapeutic strategies. Blood 2002; 99:759–767. Sanz MA, Martin G, Rayon C, Esteve J, Gonzalez M, Diaz-Mediavilla J, Bolufer P, Barragan E, Terol MJ, Gonzalez JD, Colomer D, Chillon C, Rivas C, Gomez T, Ribera JM, Bornstein R, Roman J, Calasanz MJ, Arias J, Alvarez C, Ramos F, and Deben G for the PETHEMA Group. A modified AIDA protocol with anthracycline-based consolidation results in high anti leukemic efficacy and reduced toxicity in newly diagnosed PML/RARa -positive acute promyelocytic leukemia. Blood 1999; 94:3015–3021. Falanga A, & Rickles FR Pathogenesis and management of the bleeding diathesis in acute promyelocytic leukemia. Best Practice & Research Clinical Hematology 2003; 16: 463–482. Huang ME, Ye YC, Chen SR, Chai JR, Lu JX, Zhoa L, Gu LJ & Wang ZY. Use of all-trans retinoic acid in the treatment of acute promyelocytic leukemia. Blood 1988; 72: 567–572. Chomienne C, Ballerini P, Balitrans N, Daniel MT, Fenaux P, Castaigne S, Degos L. All-trans retinoic acid in acute promyelocytic leukemias. II. In vitro studies: structure-function relationship. Blood 1990; 76: 1710–1717. Tallman MS, Lefebvre P, Baine RM, Shoji M, Cohen I, Green D, Kwaan HC, Peitta E & Rickles FR Effects of all-trans retinoic acid or chemotherapy on the molecular regulation of systemic blood coagulation and fibrinolysis in patients with acute promyelocytic leukemia. Journal of Thrombosis and Haemostatis 2004, 2, 1341–1350. Tallman MS, Brenner B, Serna Jde L, Dombret H, Falanga A, Kwaan HC, Liebman H, Raffoux E, Rickles FR. promyelocytic leukemia-associated coagulopathy, 21 January 2004, London, United Kingdom. Leuk Res 2005; 29: 347–351. Milligan DW, Grimwade D, Cullis J O, Bond L, Swirsky D, Craddock C, Kell J, Homewood J, Campbell K, McGinley S, Wheatley K, Jackson G. Guidelines on the management of acute myeloid leukemia in adults. British Committee for Standards in Haematology (BCSH). Br J Haematol 2006; 135(4):450–474. Sanz MA, Tallman MS & Lo Coco F. The tricks of the trade for the appropriate management of newly diagnosed acute promyelocytic leukemia. Blood 2005; 105: 3019–3025. Vahdat L, Maslak P, Miller WH, Eardkey A, Heller G, Scheinberg DA, Warrell RP. Early mortality and the retinoic acid syndrome in acute promyelocytic leukaemia: impact of leucocytosis, low – dose chemotherapy, PML-RAR(alpha) isoform, and CD 13 expression in patients treated with all-trans retinoic acid. Blood 1994; 84: 3843–3849. Frankel S.R, Eardley A, Lauwers G, Weiss M, Warrell R.P. The “retinoic acid syndrome” in acute promyelocytic leukaemia. Annals of Internal Medicine 1992; 117: 292–296. De Botton S, Dombret H, Sanz M, San Miguel J, Caillot D, Zittoun R, Gardembas M, Stamatoulas A, Conde E, Guerci A, Gardin C, Gelser K, Cony Makhoul D, Reman O, de la Serna J, Lefrere F, Chomienne C, Chastang C, Degos L, Fenaux P and the European APL Group. Incidence, clinical features, and outcome of all trans-retinoic acid syndrome in 413 cases of newly diagnosed acute promyelocytic leukemia. The European APL Group. Blood 1998; 92: 2712–2718. de Botton S, Chevret S, Coiteux V, Dombret H, Sanz M, San Miguel J, Calliot D, Vekhoff A, Gardembas M, Stamatoulas A, Conde E, Guerci A, Gardin C, Fey M, Cony Makhoul D, Reman O, de la Serna J, Lefrere F, Chomienne C, Degos L, Fenaux P, European APL group. Early onset of chemotherapy can reduce the incidence of ATRA syndrome in newly diagnosed acute promyelocytic leukaemia (APL) with low white blood cell counts: results from APL 93 trial. Leukemia 2003; 17 (2): 339–342. Firkin F, Matthews J, Bradstock K & Wiley JS A phase II study of all-trans retinoic acid (ATRA) with Prednisolone prophylaxis in the treatment of acute promyelocytic leukemia (APL). Blood 1999; 94 (Suppl. 2): 228B (abstract). Wiley JS & Firkin FC. Reduction of pulmonary toxicity by Prednisolone prophylaxis during alltrans retinoic acid treatment of acute promyelocytic leukemia. Leukemia 1995; 9: 774–778.
106
A. Biondi et al.
Mahmoud HH, Hurwitz CA, Roberts WM, Santana VM, Ribeiro RC, Krance RA. Tretinoin toxicity in children with acute promyelocytic leukaemia. Lancet 1993; 342: 1394–1395. Castaigne S, Lefebvre P, Chomienne C, Suc E, Rigal-Huguet F, Gardin C, Delmer A, Archimbaud E, Tilly H, Janvier M, Isnard F, Travade P, Montfort L, Delannoy A, Rapp MJ, Christian B, Montastruc M, Weh Hans, Fenaux P, Dombret H, Gourmel B and Degos L. Effectiveness and pharmacokinetics of low –dose all-trans retinoic acid (25mg/m2) in acute promyelocytic leukaemia. Blood 1993; 82: 3560–3563. Warrell R.P. Retinoid resistance in acute promyelocytic leukemia: new mechanisms, strategies and implications. Blood 1993; 82: 1949–1953. Ding W, Li Y.-P., Nobile L.M. Grills G, Carrera I. Paietta E, Tallman M.S, Wiernik P.H., Gallagher R.E. Leukemic cellular retinoic acid resistance and missense mutations in the PML-RAR alpha fusion gene after relapse of acute promyelocytic leukemia from treatment with all-trans retinoic acid and intensive chemotherapy. Blood 1998; 92: 1172–1183. Gupta V, Yi QL, Branwein J, Lipton JH, Messner HA, Schuch AC, Wells RA, Minden MD. Role of all-trans retinoic acid (ATRA) in the consolidation therapy of acute promyelocytic leukaemia (APL). Leuk Res 2005; 29(1): 113–114. Sanz MA, Martin G, Gonzalez M, Leon A, Rayon C, Rivas C, Colomer D, Amutio E, Capote FJ, Milone GA, de la Serna J, Roman J, Barragan E, Bergua J, Escoda L, Parody R, Negri S, Calasanz MJ, and Bolufer P. Risk-adapted treatment of acute promyelocytic leukemia with all-trans retinoic acid and anthracycline monochemotherapy: a multicenter study by the PETHEMA Group. Blood 2004; 103:1237–1243. Schlenk RF, Germing U, Hartmann F, Glasmacher A, Fischer JT, del Valle Y, Fuentes F, Götze K, Pralle H, Nerl C, Salwender H, Grimminger W, Petzer A, Hensel M, Brenner A, Zick L, Döhner K, Fröhling S, Döhner H, AML Study Group.High-dose cytarabine and mitoxantrone in consolidation therapy for acute promyelocytic leukaemia. Leukemia 2005; 19( 6): 978–983. Sanz MA, Lo Coco F, Martin G, Avvisati G, Rayón C, Barbui T, Diaz-Mediavilla, Fioritoni G, Gonález J D,Liso V, Esteve J, Ferrara F, Bolufer P, Bernasconi C, Gonalez M, Rodeghiero F, Colomer D, Petti M C, Ribera JM, Mandelli F for the Spanish PETHEMA and the Italian GIMEMA Cooperative Groups. Definition of relapse risk and role of nonanthracycline drugs for consolidation in patients with acute promylecytic leukaemia: a joint study of the PETHEMA and GIMEMA cooperative groups. Blood 2000; 96(4): 1247–1253. Lo-Coco F, Avvisati G, Vignetti M. Front-line treatment of acute promyelocytic leukemia with AIDA induction followed by risk-adapted consolidation: results of the AIDA-2000 trial of the Italian GIMEMA group. Blood 2004; 104:392a. Diverio D, Rossi V, Avvisati G, De Santis S, Pistilli A, Pane F, Sagilio G, Martinelli G, Petti MC, Santoro A, Pelicci PG, Mandelli F, Biondi A & Lo Coco F. Early detection of relapse by prospective reverse transcriptase-polymerase chain reaction analysis of the PML/RAR[alpha] fusion gene in patients with acute promyelocytic leukemia enrolled in the GIMEMA-AIEOP multicenter “AIDA” trial. Blood 1998; 92: 784–789. Avvisati G, Petti MC, Lo Coco F, Testi AM, Fazi P, Specchia G, Malagola M, Di Bona E, Recchia A, Marmont F, Buelli M, Lazzarino M, Di Raimondo F, Leoni F, Kropp MG, Vineri D, Miccolis I, Rossi G, Venditti A & Mandelli F AIDA: the Italian way of treating acute promyelocytic leukemia (APL), final act. Blood 2003; 102 (Suppl. 1): 487 (abstract). Lo-Coco F, Ammatuna E.The biology of acute promyelocytic leukemia and its impact on diagnosis and treatment.Hematology Am Soc Hematol Educ Program. 2006:156–61, 514. Review. Lo Coco F, Diverio D, Avvisati G, Petti MC, Meloni G, Pogliani EM, Biondi A, Rossi G, CarloStella C, Selleri C, Martino B, Specchia G, Mandelli F. Therapy of molecular relapse in acute promyelocytic leukemia. Blood 1999; 94: 2225–2229. Schnittger S, Weisser M, Schoch C, Hiddemann W, Haferlach T, Kern W. New score predicting for prognosis in PML-RARA+, AML1-ETO+, or CBFBMYH11+ acute myeloid leukemia based on quantification of fusion transcripts. Blood. 2003;102:2746–2755. Gabert J, Beillard E, van der Velden VH, Bi W, Grimwade D, Pallisgaard N, Barbany G, Cazzaniga G, Cayuela JM, Cavé H, Pane F, Aerts JL, De Micheli D, Thirion X, Pradel V,
Acute Promyelocytic Leukaemia
107
González M, Viehmann S, Malec M, Saglio G, van Dongen JJ. Standardization and quality control studies of ‘real-time’ quantitative reverse transcriptase polymerase chain reaction of fusion gene transcripts for residual disease detection in leukemia - a Europe Against Cancer program. Leukemia. 2003;17(12):2318–2357. Review. Gallagher RE, Yeap BY, Bi W, et€al. Quantitative real-time RT- PCR analysis of PMLRAR alpha mRNA in acute promyelocytic leukemia: assessment of prognostic significance in adult patients from intergroup protocol 0129. Blood. 2003;101:2521–2528. Burnett AK, Wheatley K, Goldstone AH, Stevens RF, Hann IM, Rees JH, Harrison G. The value of allogeneic bone marrow transplant on patients with acute myeloid leukaemia at differing risk of relapse; results of the UK MRC AML 10 trial. British Journal of Haematology 2002: 118: 385-400. Burnett A.K, Goldstone A.H., Stevens R.M.F., Hann I.M., Rees J.K.H. Gray R.G.,Wheatley K. Randomised comparison of addition of autologous bone-marrow transplantation to intensive chemotherapy for acute promyelocytic leukaemia in first remission – results of MRC AML 10-trial. Lancet 1998; 351: 700–708. Niu C, Yan H, Yu T, Sun HP, Liu JX, Li XS, Wu W, Zhang FQ, Chen Y, Zhou L, Li JM, Zeng XY, Yang RR, Yuan MM, Ren MY, Gu FY, Cao Q, Gu BW, Su XY, Chen GQ, Xiong SM, Zhang TD, Waxman S, Wang ZY, Chen Z, Hu J, Shen ZX, Chen SJ. Studies on treatment of acute promyelocytic leukemia with arsenic trioxide: remission induction, follow-up, and molecular monitoring in 11 newly diagnosed and 47 relapsed acute promyelocytic leukemia patients. Blood 1999; 94: 3315–3324. Soignet SL, Frankel SR, Douer D, Tallman MS, Kantarjian H, Calleja E, Stone RM, Kalaycio M, Scheinberg DA, Steinherz P, Sievers EL, Coutré S, Dahlberg S, Ellison R, Warrell RP Jr. United States Multicenter Study of arsenic trioxide in relapsed acute promyelocytic leukemia. J Clin Oncol 2001; 19: 3852–3860. Shen ZX, Shi ZZ, Fang J, Gu BW, Li JM, Zhu YM. All-trans retinoic acid/As2O3 combination yields a high quality remission and survival in newly diagnosed acute promyelocytic leukaemia. Proc Natl Acad Sci USA 2004; 101: 5328–5335. Tarkanyi I, Dudognon C, Hillion J, Pendino F, Lanotte M, Aradi J, Sigal L, Bendirdjian E. Retinoid/arsenic combination therapy of promyelocytic leukaemia: induction of telomerasedependent cell death. Leukemia 2005; 19(10):1806–1811. George B,Mathews V, Poonkuzhali B, Shaji RV, Srivastava A,Chandy M. Treatment of children with newly diagnosed acute promyelocytic leukemia with arsenic trioxide: a single centre experience. Leukemia 2004; 18(10): 1587–1590. Powell BL. Effect of consolidation with arsenic troxide (As2O3) on event-free survival (EFS) and overall survival (OS) among patients with newly diagnosed acute promyelocytic leukaemia (APL): North American Intergroup Protocol C9710. Journal of Clinical Oncology 2007; 25(18S): 2 Sanz MA, Fenaux P, Lo Coco F. Arsenic trioxide in the treatment of acute promyelocytic leukemia. A review of current evidence. Haematologica 2005; 90: 1231–1235. Tallman MS. Treatment of relapsed refractory acute promyelocytic leukemia. Best Pract res Clin Haematol 2007; 20: 57–65. de Botton S, Sanz MA, Chevret S, Dombret H, Martin G, Thomas X, Mediavilla JD, Recher C, Ades L, Quesnel B, Brault P, Fey M, Wandt H, Machover D, Guerci A, Maloisel F, Stoppa AM, Rayon C, Ribera JM, Chomienne C, Degos L, Fenaux P; European APL Group; PETHEMA Group. Extramedullary relapse in acute Promyelocytic leukemia treated with alltrans retinoic acid and chemotherapy. Leukemia 2006; 20: 35–41. Specchia G, Lo Coco F, Vignetti M, Avvisati G, Fazi P, Albano F, Di Raimondo F, Martino B, Ferrara F, Selleri C, Liso V, Mandelli F. Extramedullary involvement at relapse in acute promyelocytic leukemia patients treated or not with all-trans retinoic acid: a report by the Gruppo Italiano Malattie Ematologiche dell’Adulto. J Clin Oncol 2001; 19: 4023–4028. Breccia M, Carmosino I, Diverio D, De Santis S, De Propris MS, Romano A, Petti MC, Mandelli F, Lo-Coco F. Early detection of meningeal localization in acute promyelocytic leukemia patients with high presenting leucocyte count. British J Haematol 2003; 120: 266–270.
108
A. Biondi et al.
Lo Coco F, Breccia M, Diverio D. The importance of molecular monitoring in acute promyelocytic leukemia. Best Pract Res Clin Haematol 2003; 16: 503–520. Castagnola C, Lunghi M, Corso A, Tajana M, Zappasodi P, Dabusti M, Lazzarino M, Bernasconi C. Management of acute promyelocytic leukemia relapse in the ATRA era. Haematologica 1998; 83: 714–717. Esteve J, Escoda L, Martín G, Rubio V, Díaz-Mediavilla J, González M, Rivas C, Alvarez C, González San Miguel JD, Brunet S, Tomás JF, Tormo M, Sayas MJ, Sánchez Godoy P, Colomer D, Bolufer P, Sanz MA; the Spanish Cooperative Group PETHEMA. Outcome of patients with acute Promyelocytic leukemia failing to front-line treatment with all-trans retinoic acid and anthracycline-based chemotherapy (PETHEMA protocols LPA96 and LPA99): benefit of an early intervention. Leukemia 2007; 21: 446–452. Lo Coco F, Ammatuna E, Sanz MA. Current treatment of acute promyelocytic leukemia. Haematologica 2007; 92: 289–290. Ghavamzadeh A, Alimoghaddam K, Ghaffari SH, Rostami S, Jahani M, Hosseini R, Mossavi A, Baybordi E, Khodabadeh A, Iravani M, Bahar B, Mortazavi Y, Totonchi M, Aghdami N. Treatment of acute promyelocytic leukemia with arsenic trioxide without ATRA and/or chemotherapy. Ann Oncol 2006; 17: 131–134. Raffoux E, Rousselot P, Poupon J, Daniel MT, Cassinat B, Delarue R, Taksin AL, Réa D, Buzyn A, Tibi A, Lebbé G, Cimerman P, Chomienne C, Fermand JP, de Thé H, Degos L, Hermine O, Dombret H.Combined tretament with arsenic trioxide and all-trans retinoic acid in patients with relapsed acute Promyelocytic leukemia. J Clin Oncol 2003; 21: 2326–2334. Lo-Coco F, Cimino G, Breccia M, Noguera NI, Diverio D, Finolezzi E, Pogliani EM, Di Bona E, Micalizzi C, Kropp M, Venditti A, Tafuri A, Mandelli F.Gentuzumab ozogamicin (Mylotarg) as a single agent for molecularly relapsed acute Promyelocytic leukemia. Blood 2004; 104: 1995–1999. Takeshita A, Shinjo K, Naito K, Matsui H, Sahara N, Shigeno K, Horii T, Shirai N, Maekawa M, Ohnishi K, Naoe T, Ohno R.Efficacy of gentuzumab ozogamicin on ATRA- and arsenicresistant acute promyelocytic leukemia (APL) cells. Leukemia 2005; 19: 1306–1311. Zheng X, Seshire A, Rüster B, Bug G, Beissert T, Puccetti E, Hoelzer D, Henschler R, Ruthardt M.Arsenic but not all-trans retinoic acid overcomes the aberrant stem cell capacity of PML/ RARa-positive leukemic cells. Haematologica 2007; 92: 324–331. Sanz MA, Labopin M, Gorin NC, de la Rubia J, Arcese W, Meloni G, Bacigalupo A, Alessandrino P, Carreras E, Iriondo A, Novitzky N, Jacobs P, Bandini G, Lo-Coco F, Frassoni F, Rocha V; Acute Leukemia Working Party (ALWP) of European Cooperative Group for Blood and Marrow Transplantation (EBMT). Hematopoietic stem cell transplantation for adults with acute promyelocytic leukemia in the ATRA era: a survey of the European Cooperative Group for Blood and Marrow Transplantation. Bone Marrow Transplantation 2007; 39: 461–469. Knipp S, Gattermann N, Schapira M, Käferstein H, Germing U.Arsenic in the cerebrospinal fluid of a patient receiving arsenic trioxide for relapsed acute promyelocytic leukemia with CNS involvement. Leuk Res. 2007;31(11):1585–1587. Epub 2007 Apr 9. Arrigoni P, Beretta C, Silvestri D, Rossi V, Rizzari C, Valsecchi MG, Cazzaniga G, Biondi A. FLT3 internal tandem duplication in childhood acute myeloid leukaemia: association with hyperleucocytosis in acute promyelocytic leukaemia. Brit J Haematol. 2003;120(1):89–92. Au WY, Fung A, Chim CS, Lie AK, Liang R, Ma ES, Chan CH, Wong KF, Kwong YL. FLT-3 aberrations in acute promyelocytic leukaemia: clinicopathological associations and prognostic impact. Brit J Haematol 2004; 125: 463–469. Sohal J, Phan VT, Chan PV, Davis EM, Patel B, Kelly LM, Abrams TJ, O’Farrell AM, Gilliland DG, Le Beau MM, Kogan SC. A model of APL with FLT3 mutation is responsive to retinoic acid and a receptor tyrosine kinase inhibitor, SUII657. Blood 2003; 101: 3188–3197. Fazi F, Travaglini L, Carotti D, Palitti F, Diverio D, Alcalay M, McNamara S, Miller WH Jr, Lo Coco F, Pelicci PG, Nervi C. Retinoic acid targets DNA-methyltransferases and histone deacetylases during APL blast differentiation in€vitro and in€vivo. Oncogene 2005; 24: 1820–1830.
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity to Chemotherapy Jeffrey W. Taub, Yubin Ge, and Yaddanapudi Ravindranath
Introduction Following the first description of leukemia in a Down syndrome (DS) child in 1930 (Brewster and Cannon 1930), a national survey in 1957 confirmed that DS individuals had an increased risk of developing leukemia (Krivit and Good 1957). It has been estimated that DS children have a 10 to 20-fold increased risk of developing both acute lymphoblastic leukemia (ALL) and acute myeloid leukemia (AML) compared to nonDS children (Taub 2001). A Danish population-based study reported that the cumulative risk of developing leukemia in DS children by the age of 5 was 2.1%, and there was an approximately fourfold higher standardized incidence ratio of AML compared to ALL (Hasle et€al. 2000). The Nordic Society of Pediatric Hematology and Oncology (NOPHO) cooperative group reported that DS children with ALL and AML comprised 2.1% and 14%, respectively, of total childhood leukemia cases (Zeller et€al. 2005). In the Children’s Cancer Group (CCG) 2891 study, 15% of the AML patients had DS, indicating that DS children comprise one of the largest subgroup of AML patients. Acute megakaryocytic leukemia (AMkL; M7) is the most common FrenchAmerican-British (FAB) subtype of DS AML patients, as reported by the Pediatric Oncology Group (POG), CCG, Berlin-Frankfort-Münster (BFM)-AML, NOPHO, Medical Research Council (MRC), Toronto and Japanese Childhood AML
J.W. Taub (*) Division of Hematology/Oncology, Children’s Hospital of Michigan, 3901 Beaubien Boulevard, Detroit, MI 48201, USA and Department of Pediatrics, Wayne State University School of Medicine, Detroit, MI 48201, USA and Developmental and Therapeutics Program, Barbara Ann Karmanos Cancer Institute, 4100 John R Street, Detroit, MI 48201, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_6, © Springer Science+Business Media, LLC 2010
109
110
J.W. Taub et al.
�cooperative groups (Zeller et€al. 2005; Ravindranath et€al. 1992, 1996; Gamis et€al. 2003; Creutzig et€ al. 2005; Al-Ahmari et€ al. 2006; Kudo et€ al. 2005; Rao et€ al. 2006) (Table€1). In contrast, AMkL is estimated to represent approximately 10% of pediatric AML cases and 1 to 2% of adult AML cases. The reported frequencies of the AMkL phenotype in DS AML cases have ranged from 39% to 100% and the discrepancy among the reported frequencies is likely due to variations in identifying� megakaryoblasts by either morphology or expression of platelet-associated membrane antigens (glycoprotein IIb/IIIa) using CD41/61 antibodies. Zipursky has estimated that DS children have a 500-fold increased risk of developing AMkL compared to nonDS children (Zipursky et€al. 1994). Prior to the early 1990s, the prognosis for DS AML patients was considered to be extremely poor. In 1992, 12 DS AML patients treated on the POG 8498 study, had a 100% event-free survival (EFS) rate compared to 28% for nonDS AML patients treated with the identical therapies (Ravindranath et€ al. 1992). Subsequent studies from multiple pediatric oncology cooperative groups have confirmed this observation (Table€1). The improved survival rates highlighted by studies from these groups coincided with the utilization of high-dose cytosine arabinoside (ara-C)-based therapy for DS AML patients (Ravindranath et€ al. 1992, 1996; Lange et€ al. 1998; Gamis et€ al. 2003). The high EFS rates of DS AML patients (particularly patients with AMkL) contrasts with the extremely poor complete remission (CR) rates and EFS rates for nonDS pediatric and adult patients with AMkL (Athale et€ al. 2001; Barnard et€ al. 2007; Reinhardt et€ al. 2005; Ravindranath et€ al. 2005; Lie et€ al. 2005; Ruiz-Arguelles et€ al. 1992; Duchayne et€al. 2003; Tallman et€al. 2000; Oki et€al. 2006; Pagano et€al. 2002) (Table€2). The EFS rates of nonDS AMkL patients are also markedly lower compared
Table€1╅ Treatment outcome of Down syndrome children with AML/MDS Induction failure/ Rx relapse deaths Protocol N M6/7 EFS References POG 8498 12 58% 100% 0% 0% Ravindranath et€al. (1992) POG 8821 34 47% 68% 15% 18% Ravindranath et€al. (1996) POG 9421 62 91% 80% 11% 11% CCG-2891 161 55% 77% 13% 4% Gamis et€al. (2003) NOPHO 38 66% 83% 11% 0% Zeller et€al. (2005) BFM AML98 58 97% 89% 6% 6% Creutzig et€al. (2005) Japan-AML72 90% 83% 9% 9% Kudo et€al. (2005) Down MRC AML10/12 46 39% 74% 3% 27% Rao et€al. (2006) Toronto (low 18 100% 72% 28% 0% Al-Ahmari et€al. dose ara-C) (2006) Toronto (standard 16 100% 75% 25% 0% Al-Ahmari et€al. dose ara-C) (2006)
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity Table€2╅ Treatment results of non-Down syndrome AMkL cases Group N CR rate EFS St. Jude 35 (12.7%) 60.5% 14% De novo 20% Secondary CCG 2891 52 (5.9%) 66% 22.5% BFM 89 (7.1%) 75% 34%
111
References Athale et€al. (2001) Barnard et€al. (2007) Reinhardt et€al. (2005) Ravindranath et€al. (2005) Lie et€al. (2005) Ruiz-Argüelles et€al. (1992) Duchayne et€al. (2003)
POG 8821
32 (6.5%)
Not reported
25%
NOPHO-AML93 Mexico
16 (6.6 %) 10 children 35 adults 23 children
Not reported 90% 74% 77%
35% 30% (OS) 9% (OS) 26%
23 adult 20 (1.2%) 37 (2%)
33% 50% 43%
0% 10% 0%
Tallman et€al. (2000) Oki et€al. (2006)
24 (0.6%)
50%
10% (OS)
Pagano et€al. (2002)
Groupe Francais De Cytogenetique Hematologique ECOG (adult) MD Anderson (adult) Gimema (adult)
to current EFS rates for nonDS pediatric AML patients, overall. This indicates that the AMkL group is one of the highest risk AML subgroups, despite the use of intensive chemotherapy protocols.
Chemotherapy Sensitivity and Down Syndrome What is the basis for the significantly higher EFS rates of DS AML patients compared to other subtypes of AML? In the 3-(4,5-Dimethylthiazol-2-yl)-2,5diphenyltetrazolium bromide (MTT) in€vitro drug sensitivity assays, DS megakaryoblasts (nâ•›=â•›22) were 4.5-fold more sensitive to ara-C (median IC50: 77.5€ nm) and 12-fold more sensitive to daunorubicin (median IC50: 5.8€nm) compared to a large sample (nâ•›=â•›362) of nonDS AML blasts (median IC50’s: 350.9€nm and 71.2€nm, respectively) (Taub and Ge 2005). Further, DS megakaryoblasts generated significantly higher levels (4.4-fold) of the active intracellular metabolite, ara-CTP, following in€vitro incubations with 3H-ara-C, indicating that the metabolism of ara-C in DS cells is altered (Taub et€al. 1999). Increased in€vitro sensitivities of DS AML blasts to ara-C and anthracyclines have been confirmed by other groups (Frost et€al. 2000; Zwaan et€al. 2002). These studies suggest that chromosome 21-localized genes, which may be overexpressed in trisomy 21 containing cells, contribute to enhanced drug sensitivities of DS AML blasts. Do the increased chemotherapy sensitivities of DS megakaryoblasts translate into increased treatment related toxicity of DS children with AML therapy? For the CCG-2891 AML study (standard timed induction), there were no significant Â�differences in Grade III/IV toxicity at any site between DS and nonDS patients
112
J.W. Taub et al.
treated with the high-dose ara-C (Capizzi II) cycle (Gamis et€ al. 2003). NonDS patients experienced significantly greater hepatic toxicity (7.4% vs. 2%; Pâ•›=â•›0.036) compared to DS patients. In contrast, DS patients experienced greater mucositis (9.6% vs. 3.7%; Pâ•›=â•›0.023) and skin toxicity (3.7% vs. 0.4%; 0.024) compared to nonDS patients. However, other studies have reported a high frequency of treatment-related toxicity and mortality of DS patients. The intensive-timed DCTER regimen used in the CCG 2861/2891 AML protocol was associated with a 32% mortality rate for DS patients (Lange et€ al. 1998). Similarly, the MRC AML 10/12 studies reported a 27% treatment related mortality rate for DS patients, which used no modification in chemotherapy dosing for DS compared to nonDS AML patients (Rao et€ al. 2006). Infectious complications have also had a negative impact on the outcome of DS AML patients in the BFM studies (Lehrnbecher et€ al. 2004). A study from Argentina reported seven treatment-related deaths among 11 DS AML patients (Zubizarreta et€al. 1998). Studies from the POG have suggested that DS children have an increased risk of anthracycline-induced cardiac toxicity. A multivariate analysis of 6,493 children treated on POG protocols from 1974 to 1990 with anthracycline chemotherapy found that DS children had a relative risk of 3.4 to develop cardiac toxicity (Krischer et€al. 1997). On the POG 9421 AML study (total cumulative anthracycline dose: daunorubicin: 135€ mg/m2; mitoxantrone: 80€ mg/m2), symptomatic Â�cardiomyopathy developed in 10 (17.5%) DS patients during or soon after completion of treatment, with three patients dying of congestive heart failure (O’Brien et€al. 2006). The treatment outcome of DS ALL patients have been distinctly different Â�compared to DS AML patients. Prior studies have reported inferior survival rates of DS ALL patients compared to nonDS ALL patients (Whitlock 2006) and Â�frequently DS patients have experienced treatment-related toxicity, particularly with methotrexate (Peeters and Poon 1987). In vitro drug sensitivity studies have also demonstrated that DS lymphoblasts do not have the same patterns of increased in€vitro chemotherapy sensitivities as in DS AML cases (Frost et€al. 2000; Zwaan et€ al. 2002). This suggests that DS ALL and DS AML cells differ biologically, Â�possibly reflecting the lack of GATA1 mutations (discussed below) in DS ALL and potential differences in expression of chromosome 21-localized genes.
The Role of Chromosome 21-Localized Genes and Chemotherapy Sensitivity in Down Syndrome AMkL Several chromosome 21-localized genes may be involved in the increased Â�sensitivity of DS megakaryoblasts to chemotherapy agents. The transsulfuration pathway enzyme, cystathionine-ß-synthase (CBS; localized to chromosome 21q22.3), Â�catalyzes the condensation of serine and homocysteine to form Â�cystathionine, an intermediate step in the synthesis of cysteine. Increased CBS
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity
113
activity is associated with significantly lower homocysteine, methionine, and S-adenosylmethionine (AdoMet) levels in DS individuals, which may have downstream effects on reduced folate metabolism. Increased CBS activity may also indirectly impact metabolism of nucleotides (including ara-C) and their antileukemic activity, analogus to the established synergism of sequential methotrexate and ara-C therapy (Newman et€al. 1990). We previously hypothesized that increased CBS expression in DS leukemia cells leads to decreased allosteric regulation of 5,10-methylenetetrahydrofolate reductase (MTHFR) by AdoMet, which diverts increased levels of 5,10-methylenetetrahydrofolate to 5-methyltetrahydrofolate. The lower 5,10-methylenetetrahydrofolate pools could result in decreased dTMP and dTTP, which would relieve inhibition of deoxycytidylate deaminase (DCD) and result in lower dCTP pools. The decreased dCTP pools, in turn, would result in increased deoxycytidine kinase (dCK) catalytic activity and greater activation of ara-C to the active intracellular metabolite, araCTP, and decreased competition with ara-CTP for incorporation into DNA (Taub et€al. 1999, 1996) (Fig.€1). Portions of this hypothesis have been validated experimentally in both clinical leukemia samples and leukemia cell line models. (1) The CBS gene was expressed at significantly higher levels in DS leukemia cells (median 12-fold higher) �compared to nonDS leukemia cells and correlated with both in€ vitro ara-CTP �generation and ara-C sensitivities (Taub et€al. 1999). Interestingly, CBS transcripts were not detected in DS and nonDS ALL blasts. (2) Transfection of the CBS-null CCRF-CEM leukemia cell line with the CBS coding cDNA resulted in significantly increased in€vitro and in€vivo ara-C sensitivities compared to wild-type CEM cells (Taub et€al. 2000). (3) In AMkL cell lines with gene expression patterns similar to clinical AML samples, high CBS transcripts in DS CMK cells were accompanied by tenfold greater ara-C sensitivities and 2.4-fold higher levels of ara-CTP �generation compared to the CBS-null, nonDS AMkL cell line, CMS (Ge et€al. 2003). A relationship may also exist between the presence of the CBS 844ins68 polymorphism and in€vitro ara-C sensitivities of DS and nonDS AML blast cells (Ge et€al. 2002). Copper/zinc superoxide dismutase (SOD1, gene localized to 21q22) catalyzes the reaction from superoxide anions to hydrogen peroxide, resulting in the generation of hydroxyl free radicals. Imbalances of oxygen radicals secondary to increased SOD activity are known to be involved in the pathophysiology of DS, reflecting increased generation of oxygen free radicals and increased susceptibility of cells to undergo apoptosis (Busciglio and Yankner 1995; de Haann et€ al. 1997). The increased susceptibility of DS cells to undergo apoptosis via oxygen radicals, may contribute to enhanced chemotherapy sensitivity, particularly following daunorubicin therapy, which mediates its cytotoxicity via the generation of oxygen radicals (Chien et€al. 2004). Consistent with this notion, SOD transcripts were significantly higher (median fourfold) in DS megakaryoblasts compared to nonDS AML cells with a trend toward a correlation with in€vitro daunorubicin sensitivities (Taub et€al. 1999). Altered oxygen radical metabolism in DS may also contribute to the increased risk of anthracycline-induced cardiotoxicity.
114
J.W. Taub et al.
a
Down syndrome Megakaryoblast
B O N E
GATA1s (40-kDa)
ARA-U
E C
J
CDA
A
ARA-C
I
F
ARA-C dCTP
H
B
D
BST2
M A R R O W S T R O M A
ARA-CTP G
Fig.€1╅ The Role of GATA1 in cytidine deaminase expression and cytarabine metabolism in AML cells. After facilitated intracellular transport via the nucleoside transporter, HENT1 (a), cytarabine (ara-C) undergoes three sequential phosphorylation steps to form the active metabolite ara-CTP. The first phosphorylation step is rate limiting and is catalyzed by deoxycytidine kinase (dCK) (b), which is negatively regulated by dCTP. Cytidine deaminase (CDA) deaminates ara-C to the inactive metabolite, uridine arabinoside (ara-U) (c). Ara-CTP competes with dCTP for incorporation into DNA and inhibits DNA polymerase (d). Down syndrome (DS) acute megakaryocytic leukemia (AMkL) blasts are characterized by the presence of somatic mutations in exon 2 of the X-linked transcription factor gene, GATA1, resulting in the sythesis of a truncated GATA1 protein, GATA1s (40-kDa), which has reduced transactivation activity compared to the wild-type 50-kDa GATA1 protein (e). The CDA gene consists of a CDAsf (short-form) intronic promoter that acts as an enhancer to the CDAlf (long-form) promoter, which may enhance the transcription of the CDA gene. Therefore, lower CDA expression in DS megakaryoblasts compared to nonDS AMkL, may be due to the synthesis of the GATA1s protein, secondary to GATA1 mutations (f). Decreased CDA in DS megakaryoblasts results in greater ara-C phosphorylation and higher ara-CTP levels compared to nonDS megakaryoblast cells (g). Increased CBS expression in DS AMkL blasts results in altered intracellular reduced folate pools and decreased dCTP pools. This leads to: (1) reduced feedback inhibition of dCK resulting in greater phosphorylation of ara-C to form ara-CTP and (2) greater ara-CTP incorporation into DNA due to reduced competition from the lower dCTP pools (h). Bone marrow stromal cell antigen 2 (BST2), is a cell surface membrane protein expressed in normal tissues and malignant cells and supports the stromal cell-dependent growth of other cells. The transcription of BST2 gene is regulated by GATA1. Stimulation of BST2 promoter activity by the GATA1s protein is substantially reduced compared to GATA1, resulting in lower BST2 expression in DS megakaryoblasts. This results in reduced leukemia cell protection from ara-C-induced cytotoxicity modulated by bone marrow stromal cells (i). Other potential genes regulated by GATA1, may modulate the cytotoxicity of ara-C and other chemotherapy drugs via apoptosis pathways (e.g., HSP70, bcl-2) (j). In nonDS megakaryoblasts, higher CDA expression results in greater ara-C deamination to ara-U (k), decreased generation of ara-CTP and reduced ara-C cytotoxicity. Increased BST2 expression leads to greater leukemic blast protection from ara-C-induced apoptosis in the presence of bone marrow stromal cells (l)
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity
b
115
Non-Down Syndrome Megakaryoblast
B O N E
GATA1 (50-kDa) K
E
ARA-U
L C
ARA-C
A
F
J
CDA
BST2
M A R R O W
ARA-C B
dCTP
H
D
S T R O M A
ARA-CTP G
Fig.€1╅ (continued)
Relationship Between GATA1 and Chemotherapy Drug Sensitivities A seminal discovery in understanding the biology of AML in DS children was initially reported in 2002, (Wechsler et€ al. 2002) and subsequently confirmed by other groups (Hitzler et€al. 2003; Rainis et€al. 2003; Ahmed et€al. 2004). Acquired somatic mutations of the transcription factor gene, GATA1 (localized to Xp11.23), have been detected with nearly 100% penetrance in DS AMkL and transient myeloproliferative disorder (TMD) cases, while mutations have not been detected in nonDS AML and nonAMkL DS leukemia cases. Hence, GATA1 mutations appear to be synonymous with the DS AMkL phenotype and are not detected in remission bone marrows. GATA1 encodes a zinc finger transcription factor that binds to the WGATAR motif and is essential for normal erythroid and megakaryocytic differentiation. Reported sequence alterations in the region encoding the N-terminal activation domain of GATA1 include insertion, deletion, missense, nonsense, and splice site mutations at the exon 2/intron boundary (Muntean et€al. 2006). The net effect of GATA1 mutations is the introduction of stop codons and synthesis of a shorter GATA1 (designated GATA1s) protein (40-kDa), initiated from a downstream initiation site and distinguishable from the wild-type GATA1 protein (50-kDa) (Wechsler et€al. 2002). Both GATA1s and the wild-type GATA1 proteins show similar DNA binding abilities and interact with FOG1, though the GATA1s protein exhibits altered transactivation capacity due to the loss of the N-terminal activation domain.
116
J.W. Taub et al.
Thus, mutations at the GATA1 locus in DS are believed to result in the accumulationÂ� of poorly differentiated megakaryocytic precursors and represent initiating or early genetic hits in a multistep process of leukemogenesis. The uniform detection of somatic mutations in the GATA1 gene in DS AMkL cases suggested the possibility that the mutant GATA1 protein in DS AMkL may contribute to the high EFS rates. A relationship between GATA1 and AML outcome has been suggested by a Japanese study in which nonDS AML patients with lower GATA1 expression had the highest complete remission rates (Shimamoto et€al. 1995). It is conceivable that GATA1 may regulate the expression of several differentially expressed genes (based on the localization of GATA1 binding sites in their promoters), leading to altered activity of chemotherapy drugs. One likely target is cytidine deaminase (CDA; gene localized to chromosome 1p), which deaminates ara-C to the inactive metabolite, uridine arabinoside (ara-U). CDA transcripts were a median 5.1-fold lower in DS megakaryoblasts compared to nonDS AML blast cells (Ge et€al. 2005). The CDA transcript is transcribed from a CDA “long form” promoter (CDAlf) while a “short form” promoter (CDAsf) acts as an enhancer for the CDAlf promoter (Ge et€ al. 2004). Thus, mutations of the GATA1 gene in DS AMkL blasts that generate a functionally altered GATA1 protein, CDA enhancer activity and decreased overall CDA expression, decreased net conversion of ara-C to ara-U (Fig.€1). To assess the relationship between ara-C and GATA1, the DS AMkL cell line, CMK, was stably transfected with the full-length GATA1 cDNA (Ge et€al. 2005). The CMK transfectants expressing the full-length GATA1 protein were approximately 8 to 17-fold less sensitive to ara-C compared to wild-type and mock-transfected CMK cells. The GATA1-transfected cells showed slightly decreased daunorubicin (approximately 2-fold) sensitivity compared with the wild-type and mock-Â�transfected CMK cells, though the relationship between GATA1 and genes involved in daunorubicin metabolism/activity is unknown.
Differential Gene Expression Studies and Down Syndrome AMkL Two studies have used microarrays to identify differentially expressed genes between DS and nonDS AMkL cases, (Ge et€al. 2006; Bourquin et€al. 2006), which may account for the biological differences between the two groups of AMkL and identify genes linked to the increased chemotherapy responses of DS AMkL cases and/or chemotherapy resistance of nonDS AMkL cases. In the study by Ge et€al. (2006), 551 differentially expressed genes (105 overexpressed in the DS group) with a minimum twofold change were identified between DS and nonDS AMkL cases. Among the genes overexpressed in the nonDS AMkL group was bone �marrow stromal cell antigen 2 (BST2; localized to 19p13.2), a surface membrane protein expressed in normal tissues, malignant cells (e.g., nonDS AMkL cell line, Mo7, multiple myeloma cells). BST2 has been reported to support the stromal
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity
117
� cell-dependent growth of the DW34 cell line (Ishikawa et€al. 1995; Ohtomo et€al. 1999). Coincubation of leukemia cells with bone marrow stromal cells protects the former from chemotherapy-induced apoptosis, while megakaryocytic differentiation of progenitor cells is inhibited by direct contact with bone marrow stromal cells (Garrido et€al. 2001; Konopleva et€al. 2002; Zweegman et€al. 1999). The BST2 promoter region contains numerous putative cis elements including GATA1. Using gel shift and chromatin immunoprecipitation (ChIP) assays, BST2 was confirmed to be a GATA1 target gene with transcript levels significantly higher in nonDS AMkL blasts compared to DS megakaryoblasts (Ge et€ al. 2006). Interestingly, stable transfection of the BST2 cDNA into the DS AMkL cell line, CMK, resulted in reduced ara-C-induced apoptosis of the CMK transfectant cells in the presence of bone marrow stromal cells compared to a mock transfectant (Ge et€ al. 2006). These results suggest that bone marrow stromal cells protect leukemia cells through cell interaction mediated by cell surface antigens (e.g., BST2) and receptors. In addition to bone marrow stromal cells, bone marrow derived mesenchymal cells have also been found to protect leukemia cells from asparaginase cytotoxicity (Iwamoto et€al. 2007). Additional genes overexpressed in the nonDS AMkL group including bcl-2 and HSP70, (Ge et€al. 2006) have known antiapoptosis activity, suggesting a potential common mechanism of resistance to chemotherapy drugs in nonDS megakaryoblasts. Interestingly, the chromosome 21-localized gene, AML1 (also known as RUNX1) is underexpressed in the DS AMkL group compared to the nonDS AMkL group despite the gene dosage effect of trisomy 21 (Bourquin et€al. 2006). Knockdown expression of AML1 in a cell line model was associated with increased ara-C sensitivities (Edwards et al. 2009). Savasan et€al. (2006) used CD 36 expression by flow cytometry as a marker for megakaryocytic blast maturation in DS and nonDS AMkL blasts. Interestingly, in a small sample of nonDS megakaryoblasts expressing CD36 at similar levels as DS AMkL blasts (high CD 36 expression is universal in DS AMkL), exhibited the same level of in€vitro sensitivities as DS megakaryoblasts and the corresponding patients had markedly superior outcome compared to nonDS AMkL patients with low CD36 expression (Savasan et€al. 2006).
Future Challenges The optimal therapy for DS AML patients has not been clearly defined, balancing curative therapy against the potential risk of significant morbidity and treatmentrelated mortality. The current COG AAML041 clinical trial “Treatment of Down Syndrome Children with AML and MDS Under the Age of 4€Years” is designed to determine whether intensification of ara-C therapy during the second cycle of induction therapy, a reduction of the total cumulative anthracycline dose by 25% and a reduction in the number of intrathecal chemotherapy treatments, can maintain the high EFS rates of DS patients. Correlative pharmacologic and genetic studies may also
118
J.W. Taub et al.
identify DS patients who could be treated with further reduced intensive therapies, such as low dose ara-C containing regimens, as pioneered by Zipursky (Al-Ahmari et€ al. 2006) and/or identify “high risk” DS AML patients. In an analysis of DS patients treated on the POG 9421 and CCG-2891 (standard timing arm) AML studies, 11% (24/218) of the patients relapsed with an overall survival of only 12%, despite intensive salvage therapies including bone marrow transplantation (Loew et€al. 2004). In addition, DS AML patients >4€years of age (which comprise 5% of total DS AML cases), have EFS rates of <35%, (Gamis et€al. 2003) Â�indicating potential biological differences in AML blasts of DS children <4€years compared to >4€years of age. This suggests that a small subset of DS AML cases have an inherent resistance to chemotherapy, which may require intensification of therapy at diagnosis. No studies to date have analyzed GATA1 mutations in the group of DS AML patients >4€years of age. It is conceivable that these cases may represent true de€novo AMLs rather than the classic DS AMkL phenotype associated with GATA1 mutations.
Conclusion Can the mechanisms of enhanced chemotherapy sensitivities in DS AMkL be translated to the treatment of nonDS children with AMkL, and possibly AML overall? From the studies to date, it appears that the high cure rates of DS AML patients, and in particular, patients with the AMkL phenotype, are multifactorial based upon the trisomy 21 phenotype (overexpression of the chromosome 21 localized genes, CBS and potentially SOD1) and the unique defining genetic abnormality in DS megakaryoblasts (mutations in the GATA1 gene leading to the synthesis of the 40-kDa GATA1s protein). Using gene targeting methods to knock down GATA1 expression in nonDS AMkL may be one potential strategy, though difficulties in targeting genes in€ vivo exist. Alternative approaches include identifying downstream GATA1 target genes (e.g., genes in the apoptotic pathways), which mediate chemotherapy sensitivity/resistance. Ultimately, a universal mechanism linking the generation of GATA1 mutations, leukemogenesis, and enhanced chemotherapy sensitivities in DS AMkL may be identified. Acknowledgments╅ Supported by grants RO1 CA92308 and CA120772 from the National Cancer Institute, the Leukemia and Lymphoma Society, The Elana Fund, The Ring Screw Textron Chair in Pediatric Cancer Research and The Georgie Ginopolis Chair for Pediatric Cancer and Hematology.
References Ahmed, M., Sternberg, A., Hall, G., Thomas, A., Smith, O., O’Marcaigh, A., Wynn, R., Stevens, R., Addison, M., King, D., Stewart, B., Gibson, B., Roberts, I., Vyas, P. 2004. Natural history of GATA1 mutations in Down Syndrome. Blood 103:2480–2489.
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity
119
Al-Ahmari, A., Shah, N., Sung, L., Zipursky, A., Hitzler, J. 2006. Long-term results of an ultra low-dose cytarabine-based regimen for the treatment of acute megakaryoblastic leukaemia in children with Down syndrome. Br J Haematol 133:646–648. Athale, U.H., Razzouk, B.I., Raimondi, S.C., Tong, X., Behm, F.G., Head, D.R., Srivastava, D.K., Rubnitz, J.E., Bowman, L., Pui, C.H., Ribeiro, R.C. 2001. Biology and outcome of childhood acute megakaryoblastic leukemia: a single institution’s experience. Blood 97:3727–3732. Barnard, D.R., Alonzo, T.A., Gerbing, R.B., Lange, B., Woods, W.G. 2007. Comparison of childhood myelodysplastic syndrome, AML FAB M6 or M7, CCG 2891: report from the Children’s Oncology Group. Pediatr Blood Cancer 49:17–22. Bourquin, J.P., Subramanian, A., Langebrake, C., Reinhardt, D., Bernard, O., Ballerini, P., Baruchel, A., Cave, H., Dastugue, N., Hasle, H., Kaspers, G.L., Lessard, M., Michaux, L., Vyas, P., van Wering, E., Zwaan, C.M., Golub, T.R., Orkin, S.H. 2006. Identification of distinct molecular phenotypes in acute megakaryoblastic leukemia by gene expression profiling. Proc Natl Acad Sci USA 103:3339–3344. Brewster, H.F., Cannon, H.E. 1930. Acute lymphatic leukemia: report of a case in an eleventh month mongolian idiot. New Orleans Med Surg J 82:872–873. Busciglio, J., Yankner, B.A. 1995. Apoptosis and increased generation of reactive oxygen species in Down’s syndrome neurons in€vitro. Nature 378:776–779. Chien, M.T., Buck, S., Johnson, R.M., Stout, M., Ravindranath, Y. 2004. Increased generation of reactive oxygen species correlates with cytotoxicity in Acute Myeloid Leukemia (AML) of Down Syndrome and is augmented by cytotoxic agents affecting the mitochondrial electron transport chain. Pediatr Res 55: (abstract 1681). Creutzig, U., Reinhardt, D., Diekamp, S., Dworzak, M., Stary, J., Zimmermann, M. 2005. AML patients with Down syndrome have a high cure rate with AML-BFM therapy with reduced dose intensity. Leukemia 19:1355–1360. de Haann, J.B., Wolvetang, E.J., Cristiano, F., Iannello, R., Bladier, C., Kelner, M.J., Kola, I. 1997. Reactive oxygen species and their contribution to pathology in Down syndrome. Adv Pharmacol 38:379–402. Duchayne, E., Fenneteau, O., Pages, M.P., Sainty, D., Arnoulet, C., Dastugue, N., Garand, R., Flandrin, G., Groupe Francais d’Hematologie Cellulaire, Groupe Francais de Cytogenetique Hematologique. 2003. Acute megakaryoblastic leukaemia: a national clinical and biological study of 53 adult and childhood cases by the Groupe Francais d’Hematologie Cellulaire (GFHC). Leuk Lymphoma 44:49–58. Edwards, H., Xie, C., LaFiura, K. M., Dombkowski, A., Buck, S., Boerner, J., Taub, J. W., Matherly, L. H., Ge, Y. 2009. RUNX1 regulates phosphoinositide 3-kinase/AKT pathway: role in chemotherapy sensitivity in acute megakaryocytic leukemia. Blood 114:2744–2752. Frost, B.M., Gustafsson, G., Larsson, R., Nygren, P., Lonnerholm, G. 2000. Cellular cytotoxic drug sensitivity in children with acute leukemia and Down’s syndrome: an explanation to differences in clinical outcome? Leukemia 14:943–944. Gamis, A.S., Woods, W.G., Alonzo, T.A., Buxton, A., Lange, B., Barnard, D.R., Gold, S., Smith, F.O. Children’s Cancer Group Study 2891. 2003. Increased age at diagnosis has a significantly negative effect on outcome in children with Down syndrome and acute myeloid leukemia: a report from the Children’s Cancer Group Study 2891. J Clin Oncol 21:3415–3422. Garrido, S.M., Appelbaum, F.R., Willman, C.L., Banker, D.E. 2001. Acute myeloid leukemia cells are protected from spontaneous and drug-induced apoptosis by direct contact with a human bone marrow stromal cell line (HS-5). Exp Hematol 29:448–457. Ge, Y., Jensen, T., James, S.J., Becton, D.L., Massey, G.V., Weinstein, H.J., Ravindranath, Y., Matherly, L.H., Taub, J.W. 2002. High frequency of the 844€ins 68 cystathionine-b-synthase gene variant in Down Syndrome children with acute myeloid leukemia. Leukemia 16:2339–2341. Ge, Y., Jensen, T.L., Matherly, L.H., Taub, J.W. 2003. Transcriptional regulation of the cystathionine-ß-synthase gene in Down syndrome and non-Down syndrome megakaryocytic leukemia cell lines. Blood 101:1551–1557.
120
J.W. Taub et al.
Ge, Y., Jensen, T.L., Stout, M.L., Flatley, R.M., Grohar, P.J., Ravindranath, Y., Matherly, L.H., Taub, J.W. 2004. The role of cytidine deaminase and GATA1 mutations in the increased cytosine arabinoside sensitivity of Down syndrome myeloblasts and leukemia cell lines. Cancer Res 64:728–735. Ge, Y., Stout, M.L., Tatman, D.A., Jensen, T.L., Buck, S., Thomas, R.L., Ravindranath, Y., Matherly, L.H., Taub, J.W. 2005. GATA1, cytidine deaminase and the high cure rate of Down syndrome children with acute megakaryocytic leukemia. J Natl Cancer Institute 97:226–231. Ge, Y., Dombkowski, A.A., LaFiura, K.M., Tatman, D., Yedidi, R.S., Stout, M.L., Buck, S.A., Massey, G., Becton, D.L., Weinstein, H.J., Ravindranath, Y., Matherly, L.H., Taub, J.W. 2006. Differential gene expression, GATA1 target genes and the chemotherapy sensitivity of Down syndrome megakaryocytic leukemia. Blood 107:1570–1581. Hasle, H., Clemmensen, I.H., Mikkelsen, M. 2000. Risks of leukaemia and solid tumours individuals with Down’s syndrome. Lancet 355:165–169. Hitzler, J.K., Cheung, J., Li, Y., Scherer, S.W., Zipursky, A. 2003. GATA1 mutations in transient leukemia and acute megakaryoblastic leukemia of Down syndrome. Blood 101:4301–4304. Ishikawa, J., Kaisho, T., Tomizawa, H., Lee, B.O., Kobune, Y., Inazawa, J., Oritani, K., Itoh, M., Ochi, T., Ishihara, K., Hirano, T. 1995. Molecular cloning and chromosomal mapping of a bone marrow stromal cell surface gene, BST2, that may be involved in pre-B-cell growth. Genomics 26:527–534. Iwamoto, S., Mihara, K., Downing, J.R., Pui, C.H., Campana, D. 2007. Mesenchymal cells regulate the response of acute lymphoblastic leukemia cells to asparaginase. J Clin Invest 117:1049–1057. Konopleva, M., Konoplew, S., Hu, W., Zaritskey, A.Y., Afanasiev, B.V., Andreeff, M. 2002. Stromal cells prevent apoptosis of AML cells by up-regulation of anti-apoptotic proteins. Leukemia 16:1713–1724. Krischer, J.P., Epstein, S., Cuthbertson, D.D., Goorin, A.M., Epstein, M.L., Lipshultz, S.E. 1997. Clinical cardiotoxicity following anthracycline treatment for childhood cancer: the Pediatric Oncology Group experience. J Clin Oncol 15:1544–1552. Krivit, W., Good, R.A. 1957. Simultaneous occurrence of mongolism and leukemia. Am J Dis Child 94:289–293. Kudo, K., Kojima, S., Tabuchi, K., Ishii, E., Yabe, H., Tawa, A., Imaizumi, M., Hanada, R., Hamamoto, K., Kobayashi, R., Morimoto, A., Nakayama, H., Tsuchiya, S., Tsuchida, M., Horibe, K. 2005. Results of the Japanese Childhood Acute Myeloid Leukemia 99 protocol for Down Syndrome and Acute Myeloid Leukemia. Blood 106: abstract 276. Lange, B.J., Kobrinsky, N., Barnard, D.R., Arthur, D.C., Buckley, J.D., Howells, W.B., Gold, S., Sanders, J., Neudorf, S., Smith, F.O., Woods, W.G. 1998. Distinctive demography, biology, and outcome of acute myeloid leukemia and myelodysplastic syndrome in children with Down syndrome: Children’s Cancer Group Studies 2861 and 2891. Blood 91:608–615. Lehrnbecher, T., Varwig, D., Kaiser, J., Reinhardt, D., Klinebiel, T., Creutzig, U. 2004. Infectious complications in pediatric acute myeloid leukemia: analysis of the prospective multi-institutional clinical trial AML-BFM 93. Leukemia 18:72–77. Lie, S.O., Abrahamsson, J., Clausen, N., Forestier, E., Hasle, H., Hovi, L., Jonmundsson, G., Mellander, L., Siimes, M.A., Yssing, M., Zeller, B., Gustafsson, G.; Nordic Society of Pediatric Hematology and Oncology (NOPHO), AML Study Group. 2005. Long-term results in children with AML: NOPHO-AML Study Group – report of three consecutive trials. Leukemia 19:2090–2100. Loew, T.W., Gamis, A., Smith, F.O., Massey, G.V., Woods, W.G., Hilden, J.M., Lange, B., Alonzo, T.A., Chang, M., Gerbing, R.B. 2004. Down syndrome patients with relapsed acute myelogenous leukemia. Blood 104: (abstract 4526). Muntean, A.G., Ge, Y., Taub, J.W., Crispino, J.D. 2006. Transcription factor GATA-1 and Down syndrome leukemogenesis. Leuk Lymphoma 47:986–997. Newman, E.M., Villacorte, D.G., Testi, A.M., Krance, R.A., Harris, M.B., Ravindranath, Y., Pinkel, D. 1990. Biochemical interactions between methotrexate and 1-beta-D-arabinofura-
Down Syndrome and Acute Myeloid Leukemia: An Unique Genetic Sensitivity
121
nosylcytosine in hematopoietic cells of children: a Pediatric Oncology Group study. Cancer Chemother Pharmacol 27:60–66. O’Brien, M., Taub, J., Stine, K., Chang, M., Ravindranath, Y., Becton, D., Dahl, G.V. 2006. Excessive cardiotoxicity despite excellent leukemia-free survival for pediatric patients with Down syndrome and acute myeloid leukemia: results from POG (Pediatric Oncology Group) protocol 9421. Blood 108: (abstract 559). Ohtomo, T., Sugamata, Y., Ozaki, Y., Ono, K., Yoshimura, Y., Kawai, S., Koishihara, Y., Ozaki, S., Kosaka, M., Hirano, T., Tsuchiya, M. 1999. Molecular cloning and characterization of a surface antigen preferentially overexpressed on multiple myeloma cells. Biochem Biophys Res Commun 258:583–591. Oki, Y., Kantarjian, H.M., Zhou, X., Cortes, J., Faderl, S., Verstovsek, S., O’Brien, S., Koller, C., Beran, M., Bekele, B.N., Pierce, S., Thomas, D., Ravandi, F., Wierda, W.G., Giles, F., Ferrajoli, A., Jabbour, E., Keating, M.J., Bueso-Ramos, C.E., Estey, E., Garcia-Manero, G. 2006. Adult acute megakaryocytic leukemia: an analysis of 37 patients treated at M.D. Anderson Cancer Center. Blood 107:880–884. Pagano, L., Pulsoni, A., Vignetti, M., Mele, L., Fianchi, L., Petti, M.C., Mirto, S., Falcucci, P., Fazi, P., Broccia, G., Specchia, G., Di Raimondo, F., Pacilli, L., Leoni, P., Ladogana, S., Gallo, E., Venditti, A., Avanzi, G., Camera, A., Liso, V., Leone, G., Mandelli, F. 2002. Acute megakaryoblastic leukemia: experience of GIMEMA trials. Leukemia 16:1622–1626. Peeters, M., Poon, A. 1987. Down syndrome and leukemia: unusual clinical aspects and unexpected methotrexate sensitivity. Eur J Pediatr 146:416–422. Rainis, L., Bercovich, D., Strehl, S., Teigler-Schlegel, A., Stark, B., Trka, J., Amariglio, N., Biondi, A., Muler, I., Rechavi, G., Kempski, H., Haas, O.A., Izraeli, S. 2003. Mutations in exon 2 of GATA1 are early events in megakaryocytic malignancies associated with trisomy 21. Blood 102:981–986. Rao, A., Hills, R.K., Stiller, C., Gibson, B.E., Graaf, S.S., Hann, I.M., O’Marcaigh, A., Wheatley, K., Webb, D.K. 2006. Treatment for myeloid leukaemia of Down syndrome: population-based experience in the UK and results from the Medical Research Council AML 10 and AML 12 trials. Br J Haematol 132:576–583. Ravindranath, Y., Abella, E., Krischer, J., Wiley, J., Inoue, S., Harris, M., Chauvenet, A., Alvarado, C.S., Dubowy, R., Ritchey, A.K., Land, V., Steuber, C.P., Weinstein, H.J. 1992. Acute myeloid leukemia (AML) in Down’s syndrome is highly responsive to chemotherapy: experience on Pediatric Oncology Group AML study 8498. Blood 80:2210–2214. Ravindranath, Y., Yeager, A.M., Chang, M.N., Steuber, C.P., Krischer, J., Graham-Pole, J., Carroll, A., Inoue, S., Camitta, B., Weinstein, J.H. 1996. Acute myeloid leukemia in children: a randomized comparative study of purged autologous bone marrow transplantation versus intensive multiagent consolidation chemotherapy in first remission (Pediatric Oncology Group Study-POG 8821). N Engl J Med 334:1428–1434. Ravindranath, Y., Chang, M., Steuber, C.P., Becton, D., Dahl, G., Civin, C., Camitta, B., Carroll, A., Raimondi, S.C., Weinstein, H.J., Pediatric Oncology Group. 2005. Pediatric Oncology Group (POG) studies of acute myeloid leukemia (AML): a review of four consecutive childhood AML trials conducted between 1981 and 2000. Leukemia 19:2101–2116. Reinhardt, D., Diekamp, S., Langebrake, C., Ritter, J., Stary, J., Dworzak, M., Schrauder, A., Zimmermann, M., Fleischhack, G., Ludwig, W.D., Harbott, J., Creutzig, U. 2005. Acute megakaryoblastic leukemia in children and adolescents, excluding Down’s syndrome: improved outcome with intensified induction treatment. Leukemia 19:1495–1496. Ruiz-Argüelles, G.J., Lobato-Mendizábal, E., San-Miguel, J.F., González, M., Caballero, M.D., Ruiz-Argüelles, A., Orfao, A., Gómez-Almaguer, D., Vidriales, B., Ruiz-Reyes, G., López-Borrasca, L. 1992. Long-term treatment results for acute megakaryoblastic leukaemia patients: a multicentre study. Br J Haematol 82:671–675. Savasan, S., Buck, S., Raimondi, S.C., Becton, D.L., Weinstein, H., Chang, M., Ravindranath, Y. 2006. CD36 (thrombospondin receptor) expression in childhood acute megakaryoblastic leukemia: in€vitro drug sensitivity and outcome. Leuk Lymphoma 47:2076–2083. Shimamoto, T., Ohyashiki, K., Ohyashiki, J.H., Kawakubo, K., Fujimura, T., Iwama, H., Nakazawa, S., Toyama, K. 1995. The expression pattern of erythrocyte/megakaryocyte-
122
J.W. Taub et al.
related transcription factors GATA-1 and the stem cell leukemia gene correlates with hematopoietic differentiation and is associated with outcome of acute myeloid leukemia. Blood 86:3173–3180. Tallman, M.S., Neuberg, D., Bennett, J.M., Francois, C.J., Paietta, E., Wiernik, P.H., Dewald, G., Cassileth, P.A., Oken, M.M., Rowe, J.M. 2000. Acute megakaryocytic leukemia: the Eastern Cooperative Oncology Group experience. Blood 96:2405–2411. Taub, J.W. 2001. Relationship of chromosome 21 and acute leukemia in children with Down syndrome. J Pediatr Hematol Oncol 23:175–178. Taub, J.W., Ge, Y. 2005. Down syndrome, drug metabolism and chromosome 21. Pediatr Blood Cancer 44:33–39. Taub, J.W., Matherly, L.H., Stout, M.L., Buck, S.A., Gurney, J.G., Ravindranath, Y. 1996. Enhanced metabolism of 1-ß-D-arabinofuranosylcytosine in Down syndrome cells: a contributing factor to the superior event free survival of Down syndrome children with acute myeloid leukemia. Blood 87:3395–3403. Taub, J.W., Huang, X., Matherly, L.H., Stout, M.L., Buck, S.A., Massey, G.V., Becton, D.L., Chang, M.N., Weinstein, H.J., Ravindranath, Y. 1999. Expression of chromosome 21-localized genes in acute myeloid leukemia: differences between Down syndrome and nonDown syndrome blast cells and relationship to in€vitro sensitivity to cytosine arabinoside and daunorubicin. Blood 94:1393–1400. Taub, J.W., Huang, X., Ge, Y., Dutcher, J.A., Stout, M.L., Mohammad, R.M., Ravindranath, Y., Matherly, L.H. 2000. Cystathionine-ß-synthase cDNA transfection alters the sensitivity and metabolism of cytosine arabinoside in CCRF-CEM leukemia cells in€vitro and in€vivo: a model of leukemia in Down syndrome. Cancer Res 60:6421–6426. Wechsler, J., Greene, M., McDevitt, M.A., Anastasi, J., Karp, J.E., LeBeau, M.M., Crispino, J.D. 2002. Acquired mutations in GATA1 in the megakaryoblastic leukemia of Down syndrome. Nat Genet 32:148–152. Whitlock, J.A. 2006. Down syndrome and acute lymphoblastic leukaemia. Br J Haematol 135:595–602. Zeller, B., Gustafsson, G., Forestier, E., Abrahamsson, J., Clausen, N., Heldrup, J., Hovi, L., Jonmundsson, G., Lie, S.O., Glomstein, A., Hasle, H. 2005. Acute leukaemia in children with Down syndrome: a population-based Nordic study. Br J Haematol 128:797–804. Zipursky, A., Thorner, P., De Harven, E., Christensen, H., Doyle, J. 1994. Myelodysplasia and acute megakaryoblastic leukemia in Down’s syndrome. Leuk Res 18:163–171. Zubizarreta, P., Felice, M.S., Alfaro, E., Fraquelli, L., Casak, S., Quinteros, R., Cygler, A., Gallego, M., Perez, L.E., Sackmann-Muriel, F. 1998. Acute myelogenous leukemia in Down’s syndrome: report of a single pediatric institution using a BFM treatment strategy. Leuk Res 22:465–472. Zwaan, C.M., Kaspers, G.J., Pieters, R., Hahlen, K., Janka-Schaub, G.E., van Zantwijk, C.H., Huismans, D.R., de Vries, E., Rots, M.G., Peters, G.J., Jansen, G., Creutzig, U., Veerman, A.J. 2002. Different drug sensitivity profiles of acute myeloid and lymphoblastic leukemia and normal peripheral blood mononuclear cells in children with and without Down syndrome. Blood 99:245–251. Zweegman, S., Veenhof, M.A., Debili, N., Schuurhuis, G.J., Huijgens, P.C., Drager, A.M. 1999. Megakaryocytic differentiation of human progenitor cells is negatively influenced by direct contact with stroma. Leukemia 13:935–943.
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML) Jennifer O’Hara Lauchle and Benjamin S. Braun
Juvenile myelomonocytic leukemia (JMML) is an aggressive, clonal �myeloproliferative disorder (MPD) of childhood characterized by the overproduction of myelomonocytic cells that infiltrate the spleen, lung, and gastrointestinal tract (Arico et€al. 1997; Emanuel et€al. 1996). Children frequently present with anemia, thrombocytopenia, splenomegaly, and failure to thrive. The median age of diagnosis is 2€years of age. Without definitive treatment, the median survival of JMML patients is less than 1€year. Allogeneic hematopoietic stem cell transplantation (HSCT) is the only curative therapy with a probability of event-free survival at 5€ years of 50% (Liu et€ al. 2004). The main cause of treatment failure continues to be leukemia relapse with death due to organ infiltration, infection, or transformation to acute myeloid leukemia. Stem cell transplantation is also associated with significant acute and chronic morbidity in young children. Therefore, new approaches to therapy are needed for children with newly diagnosed and relapsed JMML. Hematopoietic progenitors isolated from the peripheral blood and bone marrow of JMML patients are hypersensitive to multiple cytokines, particularly granulocytemacrophage colony-stimulating factor (GM-CSF) (Emanuel et€ al. 1991). This is demonstrated by increased growth of granulocyte-macrophage colony forming units (CFU-GM) at low doses of GM-CSF compared to progenitor cells from healthy children. The molecular genetics of JMML implicates hyperactive RAS as an essential initiating event, as 85% of patients harbor mutations in either NRAS, KRAS, NF1, PTPN11 or CBL (reviewed in Lauchle et€al. 2006; Loh et al. 2009). Due to deregulation of the RAS pathway in many malignancies, a tremendous effort has been expended developing novel inhibitors targeted at RAS signaling for cancer therapies. We will summarize recent experience with preclinical and clinical evaluation of targeted therapeutics for JMML and discuss opportunities to study agents currently in �development. With molecular insights into pathogenesis, new assays to evaluate molecular response and preclinical testing systems in murine
J. O’Hara Lauchle (*) University of California, Helen Diller Family Cancer Research Building (Optional), 1450 3rd Street, Room 264 (Lauchle) and 265 (Braun), San Francisco, CA 94158, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_7, © Springer Science+Business Media, LLC 2010
123
124
J. O’Hara Lauchle and B.S. Braun
models and human samples, JMML researchers are uniquely poised to evaluate novel mechanism-based therapeutics for RAS pathway malignancies.
GM-CSF and Deregulated RAS Signaling in Pathogenesis of JMML RAS Signaling is a critical determinant of proliferative cell output in response to extracellular stimuli including GM-CSF and other hematopoietic growth factors. Upon ligand binding, the GM-CSF receptor activates a receptor associated tyrosine kinase, JAK2. JAK2 phosphorylates the b-chain of the GM-CSF receptor and other substrates including Signal Transducer and Activator of Transcription (STAT) proteins. The phosphorylated GM-CSF receptor also recruits adaptor and signaling molecules such as SHP-2 and results in activation of RAS signaling pathways. RAS Proteins are molecular switches that cycle between an inactive GDP-bound state and an active GTP-bound state after receptor activation. Guanine nucleotide exchange factors (GEFs), such as son of sevenless homolog (SOS1), increase the proportion of RAS bound to GTP by disassociating guanine nucleotides from RAS and favoring passive binding to GTP, which is abundant in the cytosol. GTP bound RAS interacts with effector molecules including phosphatidylinositol 3-kinase (PI3K), RAF, and Ral-GDS to activate downstream kinase cascades. Fig.€1 shows a simplified schematic of RAS pathway effector cascades, although both cell context specific interactions between effectors and higher order network regulation are likely to be relevant. RAS signaling is terminated by hydrolysis of bound GTP to GDP by a slow intrinsic GTPase, which is markedly accelerated by the binding of GTPase activating proteins (GAPs), including p120GAP and neurofibromin. RAS Proteins are encoded by the closely related KRAS, NRAS, and HRAS genes, which undergo somatic point mutations in approximately one-third of human cancers. NRAS and KRAS mutations have been documented in 25% of JMML cases, though HRAS mutations are rare (Flotho et€al. 1999; Miles et€al. 1996). These mutations introduce amino acid substitutions that lead to accumulation of RAS in the GTP-bound conformation due to defective intrinsic GTPase activity and insensitivity to GAPs. Children with neurofibromatosis type 1 (NF1) are predisposed to developing several distinct malignancies, especially JMML, brain tumors, and malignant peripheral nerve sheath tumors. The NF1 gene encodes neurofibromin, which is a GAP that negatively regulates RAS signaling by accelerating the conversion of RAS-GTP to RAS-GDP (reviewed in refs. Cichowski and Jacks 2001; Donovan et€al. 2002). Genetically, NF1 functions as a classic tumor suppressor gene. Patients with NF1 demonstrate germline heterozygosity for NF1 and develop malignancies that have undergone biallelic inactivation of NF1 (Schubbert et€al. 2005; Shannon et€al. 1994). In JMML, homozygous inactivation of NF1 has been shown to result in reduction in neurofibromin-specific GAP activity and elevated levels of RAS-GTP (Bollag et€al. 1996). Recent studies have demonstrated that there is acquired isodisomy of the mutant NF1 allele in these patients, a commonly observed phenomenon in human cancer (Flotho et€al. 2007; Michaelson et€al. 2005).
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
125
GF
JAK2 XL019 lestaurtinib INCB018424
P
SHP-2
P
Shc
Gab2 Grb2 SOS
RAS GDP
GAP NF1
RAS GTP
tipifarnib salirasib
CBL
IC87114 GDC0941 PX866 XL147 BGT226
PI3K
sorafenib
Raf
RalGDS
PDK
PD0325901 AZD6244 XL518
MEK
JNKK
ERK
JNK
NVP-BEZ235 XL765 Akt rapamycin everolimus temsirolimus
mTOR
Fig.€1â•… RAS Pathway signaling and inhibitors in clinical development. A simplified schematic of the RAS signal transduction network is depicted. RAS cycles between active (GDP-bound) and inactive (GTP-bound) states upon activation of growth factor (GF) receptors, and in turn stimulates multiple cytoplasmic signal transduction cascades (see text). JMML-associated mutations in KRAS, NRAS, NF1, CBL, or PTPN11 (SHP-2) activate this system. Selected agents in clinical development that might be used to treat JMML are shown with their most relevant targets. Note that, in some cases, “off-target” proteins not shown in this figure are inhibited to an equal or greater degree. As reviewed in the text, initial efforts were focused on inhibition of RAS. Current efforts are more focused on targeting downstream RAS effectors, singly or in combination. An alternative approach is to inhibit upstream biochemical events, such as JAK2 activation
Germline PTPN11 mutations are present in 40 to 50% cases of Noonan syndrome, which shares clinical features with NF1 (Sun et€al. 2003). Similar to NF1, the occurrence of JMML and transient myeloproliferative disorders in children with Noonan syndrome led investigators to screen nonsyndromic patients with de€novo JMML for somatic mutations in PTPN11. Somatic mutations in PTPN11 are found in 35% of nonsyndromic JMML cases (Locatelli et€ al. 2005; Tartaglia et€ al. 2001). PTPN11 encodes the nonreceptor protein tyrosine phosphatase (PTP) SHP-2. SHP-2 contains two src homology 2 (SH2) domains and a PTP domain. In its basal state, the PTP activity is blocked by an interaction with the N-SH2 domain. Ligand binding disrupts this auto-inhibitory interaction resulting in PTP activation and enhanced RAS/ERK pathway activation. Most PTPN11 mutations in JMML alter residues in the interface between the N-SH2 and PTP domains thereby preventing basal auto-inhibition (Keilhack et€al. 2005). Leukemia associated alleles encode SHP-2 variants, such as D61Y and E76K, that tend to have higher PTP activity than those found in germline mutations found in NS (Nakamura et€al. 2005; Rotblat et€al. 2008).
126
J. O’Hara Lauchle and B.S. Braun
More recently, mutations in CBL have been found in approximately 10% of JMML cases (Loh et al. 2009). The Cbl protein has multiple roles in signal transduction. A ubiquitin ligase function targets various growth factor receptors for degradation to terminate signaling, whereas other domains promote signal transduction by formation of signaling complexes. Leukemia associated point mutations appear to disrupt ubiquitin transfer activity but leave adapter functions intact, leading to a net activation of signal transduction by aberrant CBL proteins (Sanada et al. 2009). Complete loss of Cbl ubiquitin ligase activity may be important, because the normal allele is usually lost (specifically, by acquired uniparental isodisomy). In one child with JMML, the mutant allele was found to be heterozygous in buccal DNA, suggesting possible germline inheritance. In this scenario, patients that inherit a heterozygous point mutation would be at risk for developing JMML upon acquired somatic homozygosity, in a manner similar to patients heterozygous for NF1. Mutations in PTPN11, RAS and NF1 are rarely found in the same patient, supporting the hypothesis that hyperactive RAS signaling is essential to the pathogenesis of JMML. The molecular events in 25% of JMML cases are unknown, yet they are likely to also dysregulate the RAS pathway. Analysis of JMML samples has not yet uncovered mutations in GM-CSF, SOS1, SHC1, GRB2, GAB1, BRAF, MEK1, or MEK2 (de Vries et€al. 2007a; Kratz et€al. 2007). In addition, FLT3 and JAK2 mutations, relatively common in other myeloid leukemias and myeloproliferative disorders, are rarely detected in JMML (de Vries et€al. 2007b; Winter-Vann and Casey 2005). In addition to providing insights into pathogenesis, detection of mutations in NF1, RAS, PTPN11, or CBL will be incorporated into revised diagnostic criteria for JMML, and may have prognostic implications. For therapeutic studies, mutation detection provides a way to evaluate molecular response to treatment. Specific quantitative assays for JMML-associated mutant alleles have been developed and used to detect relapse following stem cell transplant with high sensitivity (Archambeault et€al. 2008). Importantly, this assay could yield objective response criteria for therapies given prior to transplantation or in the context of minimal residual disease. Because the morphology of neoplastic cells in JMML is nearly normal, quantitative assessment of the mutant clone by DNA analysis provides the only realistic way to measure disease burden in these settings. More recently, a population of cells with increased STAT5 phosphorylation in response to GM-CSF was identified in diagnostic JMML samples with sensitivity of 91% and a specificity of 95% (Kotecha et€al. 2008). Presence of this population correlated with clinical disease status in several patients for whom diagnostic, remission, and relapse samples were evaluated (Kotecha et€al. 2008). Application of this assay to evaluate therapeutic response and disease status may be especially important in cases for which PCR assays are not available such as those patients without a known mutation or with a diagnosis of NF1. Beyond its potential use as a diagnostic tool, intracellular phosphoprotein analysis of JMML samples may provide opportunities to identify new disease targets, facilitate preclinical drug testing, and evaluate clinical responses. For example, in€ vitro exposure of JMML samples to a JAK2 inhibitor abrogated both the phospho-STAT5 and phospho-ERK response to GM-CSF (Kotecha et€al. 2008). This work follows prior efforts to develop assays using patient samples for drug �development but will
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
127
need to be linked to leukemia growth in€ vitro, and most importantly, clinical responses. The GM-CSF hypersensitive myeloid progenitor colony growth that aids in diagnosis has also been applied as an in€vitro assay to screen new therapeutic compounds (Emanuel et€ al. 2000; Loh et€ al. 2004). Inhibition of spontaneous versus GM-CSF stimulated colonies may be assessed, and therapeutic thresholds can be defined when colony growth from normal bone marrow progenitors is evaluated concurrently. Engraftment of JMML samples into NOD/SCID mice provides another method to test new agents on patient-derived samples (Mohi et€al. 2005).
Mouse Models Genetically engineered mouse models have substantial potential to address several important issues in drug development. Clinical research in patients is limited by high cost, long duration, and the need to protect the safety of subjects. The rarity of JMML is also a major barrier, as even cooperative groups can expect to enroll fewer than 50 patients each year. Hopefully, mouse models that recapitulate the genetic, biochemical, and cell biologic features of JMML will have useful positive and negative predictive value for screening candidate therapeutics. Studies in these animal models can be performed at far lower cost, complexity, and time in comparison to conventional human trials. Several strategies have been used to model JMML in the mouse. The first used gene targeting to disrupt Nf1, and heterozygous mice indeed developed a JMML-like disease associated with somatic loss of the wild type allele (Bollag et€al. 1996; Jacks et€al. 1994). This occurred only in old mice, however, so homozygous loss of Nf1 was accomplished by breeding Nf1−/− mice. Because of late-term embryonic lethality, hematopoietic cells were harvested from Nf1−/− fetal livers and transplanted into irradiated recipients (Largaespada et€al. 1996). This method reliably produced mice with MPD but also required significant effort and introduced potentially confounding effects of radiation. To create a more tractable and physiologic model, a conditional Nf1flox allele was constructed by placing loxP recombination sites around exons 31 and 32 (Zhu et€al. 2001). When the Cre recombinase protein is expressed, it performs a splicing event between the loxP sites that removes the intervening DNA. This results in the deletion of these two exons and inactivation of Nf1. Importantly, this lesion is similar to mutations found in NF1 patients, who frequently have deletion, nonsense, or frameshift mutations in this region. The inducible Mx1-Cre transgene allows Nf1 disruption to be induced in hematopoietic stem cells after birth by treating mice with interferon or substances that elicit interferon production such as polyinosinic-polycytidilic acid (pIpC). Nf1flox/flox, Mx1-Cre mice that are treated with pIpC at 3–5€days of age uniformly develop a monocytic MPD that closely resembles JMML (Le et€al. 2004). This represents a major technical advance over the earlier systems, and the conditional system is much more amenable to preclinical studies. Similar conditional alleles have also been used to model effects of dominant oncogenes such as Kras and Ptpn11 (Braun et€al. 2004; Chan et€al. 2004, 2009). In this case, an oncogenic mutation is introduced in the germ line, but it is initially
128
J. O’Hara Lauchle and B.S. Braun
silenced by a repressor cassette (“LoxP-Stop-LoxP” or LSL). The LSL is then excised by Cre in juvenile mice. In contrast to the indolent MPD seen in Nf1flox/flox and LSL-Ptpn11D61Y mice, which survive for 6–12€months, the LSL-KrasG12D model is characterized by aggressive disease leading to death in less than 3€months. The LSL-KrasG12D model is also the most readily transplanted into naïve recipients. Interestingly, despite broad phenotypic similarities among all these models, there remain distinctions between them with respect to erythroid maturation, progenitor distribution, signaling biochemistry, and stem cell function. It is yet unknown as to which of these features will reflect genotype-specific findings in JMML patients and which will be unique to mice. Nonetheless, it is a testament to the efficiency of modern genetic engineering that three of the 5 known JMML oncogenes have now been modeled using conditional targeted alleles in mice. Such systems have significant advantages, most notably (a) ease of use, (b) preservation of normal cis-acting regulatory control of the targeted locus, including normal copy number, and (c) the ability to study the effects of somatic mutation in animals that have undergone normal development. In principle, these mouse cancer models will allow researchers to conduct well-controlled experiments aimed at both discovering fundamental mechanisms of leukemogenesis and evaluating potential therapies prior to human testing. However, it is prudent to note that no in€vitro or in€vivo model system has yet been used to identify a clinically important therapy for JMML.
Targeted Therapeutics Considering its role in many cancers, inhibition of activated RAS is a logical therapeutic strategy. However, there are several obstacles to this approach. The importance of RAS signaling in diverse physiologic processes in normal cells raises the concern of systemic toxicities from any therapies in which normal RAS proteins or effectors are inhibited. In addition, oncogenic mutations in RAS lead to decreased GTPase activity, and developing agents that help restore rather than inhibit enzyme function has not been achieved to date. Competitive GTP inhibitors would need to confront the very high affinity of RAS for GTP, as well as the high intracellular concentration of GTP. With this in mind, researchers turned their focus to inhibiting posttranslational RAS processing at the C-terminal CAAX sequence. The multistep process of CAAX protein modification renders proteins hydrophobic at their C-termini, allowing for membrane association. The rate-limiting step in normal and mutant RAS protein modification is prenylation, which covalently attaches an isoprenoid lipid at the CAAX motif at the carboxyl terminus. Based on the terminal amino acid (X) of the CAAX motif, proteins are prenlyated by farnesyltransferase (FTase), which transfers a 15 carbon farnesyl isoprenoid, or geranylgeranyltransferase (GGtase), which in turn, transfers a 20 carbon geranylgeranyl isoprenoid. As prenylation of H-RAS, N-RAS, and K-RAS is catalyzed by FTase, farnesyltransferase inhibitors (FTIs) were developed to prevent RAS localization to the plasma membrane and inhibit RAS signaling. Colony formation of JMML cells was reduced by 25 to 50% at �concentrations of the FTI L739,749 that only modestly reduced colony growth
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
129
of control hematopoietic progenitors (Emanuel et€al. 2000). Treatment of Nf1 deficient murine models of JMML with a similar FTI, L744,832, did not inhibit Nf1−/− hematopoietic cell growth or splenic infiltration with myeloid progenitors (Mahgoub et€ al. 1999). FTI treatment was associated with a decrease in farnesylation of H-RAS but not N-RAS (Mahgoub et€al. 1999). Response rate and acute toxicities of a nonpeptidic CAAX-competitive selective inhibitor of FTase tipifarnib ( R115777 ) was studied in a Children’s Oncology Group (COG) Phase II treatment window in newly diagnosed JMML (Castleberry et€ al. 2005). The majority of JMML patients treated had partial clinical responses defined by decreased white blood count and organomegaly prior to transplantation; however, event-free survival was not impacted by FTI treatment. Similar to other clinical trials with tipifarnib, clinical response did not correlate with the degree of inhibition of FTase activity as measured by farnesylation of the heat shock molecule, HDJ2 (Castleberry et€al. 2005). Residual FTase activity at clinical doses, alternative prenylation of N-RAS and K-RAS by GGTase, as well as a lack of correlation between RAS mutation status and response in other tumor models underscore that the authentic target(s) of FTIs are unknown. The cellular effects of FTIs may require inhibition of any of the many cellular proteins that require farnesylation such as the Rho family GTPases. As N-RAS and K-RAS have been shown to be prenylated by GGTase during FTase inhibition, GGTase inhibitors have been developed and their clinical applications are under investigation (Park et€al. 2008; Side et€al. 1997). Prenylation is one step in a larger metabolic pathway that targets RAS to plasma membranes in both physiologic signaling and transformation. After prenyl transfer by FTase or GGTase, cleavage of the AAX residues occurs by RAS converting enzyme 1 (Rce1). This is followed by methylation catalyzed by isoprenylcystine carboxyl methyltransferase (ICMT). These steps increase efficiency of membrane association of the RAS proteins. A few small molecule inhibitors of Rce1 protease and ICMT have recently been described and remain interesting for preclinical evaluation (reviewed in Wilhelm et€al. 2004). Analysis of Rce1 and Icmt null cells suggests that farnesylated RAS and other farnesylated proteins are mislocalized to a greater degree than geranylated proteins (Mahgoub et€ al. 1999). Disruption of Icmt or Rce1 results in embryonic lethality, suggesting an important physiologic role for this pathway and potential for systemic toxicities from its inhibition (Bergo et€al. 2001; Kim et€al. 1999). In the Kras MPD model, conditional deletion of Icmt in the malignant populations reduced severity of the disease but did not correct anemia or prolong survival (Wahlstrom et€ al. 2007). Paradoxically, conditional deletion of Rce1 caused acceleration of MPD, perhaps due to effects of Rce1 on CAAX-containing proteins that inhibit RAS signaling (Tartaglia et€al. 2003). Overall, investigation of agents targeting RAS processing has revealed complex and poorly predicted effects on growth of normal and malignant cells. These problems can be attributed both to ineffective targeting of mutant RAS, in the case of FTIs that spare GGTase modification of N-RAS and K-RAS, and to a large class of “off-target” substrates other than RAS with both positive and negative effects on cell growth. This experience underscores the need to investigate the therapeutic targets, mechanism, and selectivity of these drugs if we are to understand which diseases and patients might benefit from their clinical application.
130
J. O’Hara Lauchle and B.S. Braun
Direct targeting of RAS by strategies that do not inhibit prenylation is also being evaluated. Farnesylthiosalicylic acid (FTS, Salirasib) mimics the carboxy-terminal farnesyl cysteine carboxymethyl ester of RAS and competes with RAS-GTP for binding sites on cellular membrane proteins, galectin 1 (Gal-1) and galectin 3 (Gal-3) (Ashery et€al. 2006; Rinehart et€al. 2004). It is postulated that disruption of galectin binding displaces all isoforms of RAS-GTP from the plasma membrane interfering with effector interactions and increasing the rate of inactivation to GDP (Ashery et€al. 2006; Rinehart et€al. 2004). FTS was recently shown to inhibit growth of NF1 mutant MPNST cell lines in a dose-dependent fashion that correlated with basal RAS-GTP levels (Barkan et€al. 2006). FTS treatment shortened duration of RAS-GTP induction following serum stimulation and was associated with lower levels of phosphorylated ERK, AKT, and Ral-GTP (Barkan et€al. 2006). Phase I and II trials of Salirasib in pancreatic and nonsmall cell lung cancer are currently enrolling patients and will be important in evaluating the pharmacodynamics and toxicity of this compound in malignancies associated with activated RAS signaling. Finally, N-RAS needs to be both palmitoylated and prenylated for stable membrane association, and inihibition of the palmitoylation step has been proposed as a therapy for N-RAS mutant leukemias (Cuiffo and Ren 2010). However, the K-RAS 4B isoform is not palmitoylated and therefore such a strategy would not apply to cases with K-RAS mutations. As targeting mutated RAS, SHP-2, and neurofibromin is difficult, inhibiting effectors of RAS signaling has emerged as an alternative therapeutic strategy. Nonmutated effector proteins would be inhibited to the same degree in neoplastic and normal cells, and thus, a therapeutic response would arise only if cancer cells respond differently to pathway inhibition. Cellular adaptation to oncogenic signals can alter the dependence on specific signals and determine the outcome of pathway inhibition, a process referred to as “oncogene addiction” (Hingorani and Tuveson 2003; Watanabe et€ al. 1996). One potential advantage of this strategy is that inhibiting a critical cellular signal that is altered by diverse molecular mechanisms has broader clinical applicability. Given the complexity of normal and malignant signaling networks, it is hard to predict a priori whether inhibition of one or more signaling proteins will be selectively toxic to cancer cells. The two RAS effector pathways most highly associated with cancer are the Raf/MEK/ERK and the PI3K/AKT/mTOR networks, though their individual contributions to malignant phenotypes are not clear. A growing number of compounds targeting various components of these systems are now available for preclinical and clinical investigations. Studying cellular responses and signaling in the presence of specific inhibitors will inform the pathobiology of cancer and hopefully yield therapeutic discoveries.
RAF/MEK/ERK RAS-GTP associates with and activates the Raf kinases (A-Raf, B-Raf, and c-Raf /Raf-1). Raf phosphorylates and activates mitogen-activated protein kinase kinase (MEK1 and MEK2). MEK, in turn, activates ERK, which phosphorylates diverse
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
131
nuclear and cytosolic substrates. RAF gene mutations that encode constitutively active proteins are found in several cancers, and therefore, it is reasonable to hypothesize that inhibiting the Raf/MEK/ERK cascade has therapeutic potential in cancers associated with elevated RAS-GTP. Degradation of Raf-1 with a DNA enzyme resulted in a reduction of spontaneous colony formation and GM-CSF hypersensitivity in JMML samples without effect on colony formation in normal bone marrow cells (Iversen et€ al. 2002). Reduced levels of GM-CSF mRNA and protein were observed in a panel of treated JMML samples (Iversen et€al. 2002). Immunodeficient mice transplanted with primary JMML cells and treated for 4€weeks with an Raf-1 degrading enzyme experienced a greater than 50% reduction in JMML bone marrow cell mass compared to control mice (Iversen et€al. 2002). Several Raf inhibitors have been developed that vary in their specificity for Raf isoforms and mutant (V600E) B-Raf commonly found in melanoma (reviewed in Caraglia et€al. 2006). To date, the largest clinical experience with Raf inhibitors is with Sorafenib (BAY 43-9006), which is approved by the FDA for treatment of advanced renal cell carcinoma and hepatocellular carcinoma. Sorafenib is a novel bi-aryl urea that has been shown to inhibit Raf-1 (c-Raf) as well as B-Raf, and reduces tumor proliferation in several human tumor xenograft models (Weinstein 2002). In preclinical testing, antitumor activity was noted to correlate with a reduction in MEK and ERK phosphorylation in the presence of Sorafenib in diverse tumor cell lines and murine models (Weinstein 2002). However, Sorafenib is also active against a variety of other kinases, including VEGFR-2 and 3, PDGFR-b, FGFR-1, FLT3 and c-KIT (Weinstein 2002). RAS and BRAF mutation status failed to correlate with growth inhibition in xenograft models suggesting the importance of inhibiting angiogenesis in these responses. Early phase testing is currently being conducted in pediatric solid tumor patients. A future study of this FDA-approved multikinase inhibitor in relapsed JMML is feasible, yet the preclinical activity in JMML cells is unknown. The broad kinase inhibition of Sorafenib may have more disadvantages than benefits in hematopoietic malignancies, as inhibition of angiogenesis may cause toxicities and is unlikely to have therapeutic benefit. These potential “off-target” toxicities may be avoided with more specific inhibitors of Raf or MEK. Although not frequently mutated in cancers, MEK activation is implicated in tumor development based on the frequency of RAS and RAF mutations in tumors. MEK is an attractive target because the only known substrate is ERK. Highly selective inhibition of MEK can be achieved with non competitive inhibitors that bind a novel region and lock the kinase in a catalytically inactive state (Niihori et€ al. 2005). The addition of U0126, a MEK inhibitor, to myeloid colony assays selectively inhibited the growth of Ptpn11E76K expressing cells when compared with wild-type cells (Miyauchi et€ al. 1994). In contrast, the MEK inhibitor CI-1040 inhibited myeloid colony growth in Nf1 mutant bone marrow only at high concentrations that equivalently inhibited myeloid colonies in wild-type bone marrow (Lauchle et€al. 2007). Treatment with CI-1040 twice a day at the maximal tolerated dose (MTD) is known to inhibit phosphorylation of ERK for 6 to 8€ hours
132
J. O’Hara Lauchle and B.S. Braun
(Peterson et€al. 2006). Four weeks of treatment with this regimen failed to decrease WBC or �splenomegaly in the Nf1flox/flox MPD model (Lauchle et€al. 2007). There may be a role for MEK inhibition for patients with refractory JMML, especially in those patients that transform to acute myeloid leukemia. This is suggested by the inhibition of Nf1 mutant acute myeloid leukemic colony growth at low doses of CI-1040 and prolonged survival of mice transplanted with Nf1 mutant AMLs when treated with CI-1040 (Lauchle et€al. 2007). Recently developed MEK inhibitors (PD0325901 and AZD 6244) that have more favorable pharmacodynamic profiles are currently in clinical trials, including a Phase II trial of AZD 6244 in adults with acute myeloid leukemia (Adjei et€al. 2008).
PI3K/AKT/mTOR The PI3K/AKT/mTOR pathway is activated in human cancers by diverse genetic events including RAS mutations, activating PI3K mutations, and inactivation of PTEN, a negative regulator of PI3K. The largest clinical experience is with PI3K/ AKT/mTOR pathway inhibitors that target the more distal component of the pathway, mTOR. Rapamycin and its analogs, temsirolimus (CCI-779) and everolimus (RAD001), bind to the FK506 binding protein, FKBP-12, which then binds to and inhibits mTOR. Initially discovered in 1975, rapamycin is FDA approved as an immunosuppressant and is frequently used after HSCT to prevent graft versus host disease. In JMML, antileukemic activity of low dose rapamycin (1 to 10€nM) has been observed in Ptpn11E76K mutant transformed cells and in JMML patient samples (Le et€al. 2004; Miyauchi et€al. 1994). In Ptpn11 mutant cells, the addition of a MEK inhibitor to rapamycin resulted in a synergistic decrease in colony formation (Miyauchi et€al. 1994). Inhibition of mTOR by rapamycin results in AKT activation in some cell systems. This, combined with the potential of other effectors downstream of PI3K/ AKT activation to contribute to the malignant phenotype, has increased interest in chemical inhibitors of more proximal pathway components. The first generation PI3K inhibitors, wortmannin and LY294002, both target the p110 catalytic subunit of PI3K, yet they also inhibit other kinases. They are associated with broad cellular toxicities that limit therapeutic evaluation of PI3K pathway inhibition. Recently, development of more specific inhibitors of PI3K (PX-866, GDC-0941, IC87114, XL147, BGT226) has revitalized the study of the PI3K pathway in malignancies, and several agents are under clinical investigation. In addition, several dual inhibitors of PI3K and mTOR (NVP-BEZ235, XL765 and PI-103) are under active preclinical and clinical study. Antileukemic activity in AML has recently been reported for PI-103, a dual p110 and mTOR inhibitor, which was not induced either by selective PI3K inhibition with a p110 inhibitor, IC87714, or by mTOR inhibition with everolimus (Pardanani 2008). In AML samples, these authors noted both AKT activation during mTOR inhibition as well as mTOR activation during PI3K inhibition, supporting a rationale for combined inhibition of pathway components
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
133
(Pardanani 2008). AKT inhibitors with diverse mechanisms including ATP �competition, prevention of membrane localization, and allosteric inhibition provide another approach to inhibit this complex pathway.
PTPN11/SHP-2 As SHP-2 is the most commonly mutated protein in sporadic JMML and promotes signal transduction through RAS pathway, inhibiting the nonreceptor protein tyrosine phosphatase (PTP) is a rational therapeutic strategy for JMML. By screening the NCI diversity set chemical library, NSC-87877 was identified as a specific inhibitor of SHP-2 and SHP-1 (Chen et€ al. 2006). Site directed mutagenesis and modeling suggested that NSC-87877 inhibits SHP-2 PTP activity by binding to the catalytic cleft (Chen et€al. 2006). NCS-87877 may preferentially inhibit activated SHP-2, as 50% inhibition of PTP activity in fetal liver cells transduced with fulllength wild-type SHP-2 occurred at a NSC-87877 dose 4.5-fold higher than the dose required to achieve the same inhibition of the E76K mutant SHP-2 or the wildtype PTP domain lacking the autoinhibitory N-SH2 domain (Chen et€ al. 2006). Cells exposed to inhibitor had lower RAS-GTP and phospho-ERK levels following cytokine stimulation (Chen et€al. 2006). The cytotoxic effects of NSC-87877 on a human breast cancer cell line were not significantly different than those seen with a MEK inhibitor or a PI3K inhibitor (LY294002) alone; however, greater cell death was noted with the combination of NSC-878777 and LY294002 (Chen et€al. 2006). A second compound, phenylhydrazonopyrazolone (PHSP1) was identified in a screen modeling binding interactions with the SHP-2 catalytic site (Hellmuth et€al. 2008). PHSP1 decreased growth factor stimulated SHP-2 PTPase activity, phosphorylation of ERK, and SHP-2 dependent morphologic changes in epithelial cells (Hellmuth et€al. 2008). PHSP1 inhibited growth and colony formation in a panel of human tumors in a SHP-2 dependent manner, with high expression of mutant SHP-2 predicting strong inhibition (Hellmuth et€al. 2008). A reasonable next step is to study these compounds in clinical and murine JMML samples to evaluate therapeutic potential of SHP-2 inhibitors as they are currently unavailable for clinical administration.
JAK2-STAT5 Cellular responses to GM-CSF are primarily regulated by the JAK2 kinase (Wahlstrom et€al. 2008). With the recent discovery of STAT5 phosphorylation in response to low dose GM-CSF in JMML cells and a prominent role of activating JAK2 mutations in myeloproliferative disorders, it is logical to evaluate the functional consequences of inhibiting JAK signaling (Kotecha et€al. 2008). Based on the gain of function mutation JAK2 V617F, identified in the majority of patients with
134
J. O’Hara Lauchle and B.S. Braun
polycythemia vera, essential thrombocythemia, and primary myelofibrosis, a number of inhibitors (XL019, INCB018424, and TG101348) that inhibit wild-type and mutant JAK2 are in various stages of preclinical and clinical trial testing (Ohren et€al. 2004). In addition, several agents designed to inhibit Aurora kinases (VX680, and AT9283) or FLT3 (CEP-701, lestaurtinib) have significant JAK2 kinase inhibition and Phase I and II clinical trials in MPDs are being conducted (Hexner et€al. 2008; Ohren et€ al. 2004). Therefore, these agents may be among the first signal transduction inhibitors used in clinical trials for JMML. In a preclinical setting, the orally available JAK inhibitor (XL019) abrogated the phosphorylation of both STAT5 and ERK in response to GM-CSF (Kotecha et€al. 2008). Further correlating an inhibitor’s effect on RAS signaling proteins and colony formation may be used to evaluate the role of JAK/STAT5 signaling in JMML disease pathogenesis.
Conclusion Therapy with selective small molecule inhibitors has revolutionized cancer treatment for several hematopoietic malignancies, most notably CML. Advances in the molecular understanding of JMML have arrived concurrently with development of many inhibitors directed at RAS signaling. This provides an unprecedented opportunity to evaluate the impact of targeted inhibitors alone and in combination in JMML treatment. At a minimum, the drugs reviewed in this chapter will serve as chemical probes to explore the critical biochemical alterations in JMML cells and inform future drug development. Ultimately, the successful identification of a new therapy for JMML will require collaborative research to generate and analyze preclinical data necessary to prioritize the growing number of drugs available for clinical trial. Trial design must give careful consideration to incorporation of novel agents and biologic correlates into current clinical trial design to evaluate each target thoroughly. The options for JMML therapy are expanding rapidly and hold promise for improved clinical outcomes.
References Adjei AA, Cohen RB, Franklin W, Morris C, Wilson D, Molina JR, Hanson LJ, Gore L, Chow L, Leong S, Maloney L, Gordon G, Simmons H, Marlow A, Litwiler K, Brown S, Poch G, Kane K, Haney J, Eckhardt SG (2008) Phase I pharmacokinetic and pharmacodynamic study of the oral, small-molecule mitogen-activated protein kinase kinase 1/2 inhibitor AZD6244 (ARRY-142886) in patients with advanced cancers. J Clin Oncol 26:2139–46 Archambeault S, Flores NJ, Yoshimi A, Kratz CP, Reising M, Fischer A, Noellke P, Locatelli F, Sedlacek P, Flotho C, Zecca M, Emanuel PD, Castleberry RP, Niemeyer CM, Bader P, Loh ML (2008) Development of an allele-specific minimal residual disease assay for patients with juvenile myelomonocytic leukemia. Blood 111:1124–7 Arico M, Biondi A, Pui CH (1997) Juvenile myelomonocytic leukemia. Blood 90:479–88 Ashery U, Yizhar O, Rotblat B, Elad-Sfadia G, Barkan B, Haklai R, Kloog Y (2006) Spatiotemporal organization of signaling: rasosomes and the galectin switch. Cell Mol Neurobiol 26:471–95
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
135
Barkan B, Starinsky S, Friedman E, Stein R, Kloog Y (2006) The inhibitor Â�farnesylthiosalicylic acid as a potential therapy for neurofibromatosis type 1. Clin Cancer Res 12:5533–42 Bergo MO, Leung GK, Ambroziak P, Otto JC, Casey PJ, Gomes AQ, Seabra MC, Young SG (2001) Isoprenylcysteine carboxyl methyltransferase deficiency in mice. J Biol Chem 276:5841–5 Bollag G, Clapp DW, Shih S, Adler F, Zhang Y, Thompson P, Lange BJ, Freedman MH, McCormick F, Jacks T, Shannon K (1996) Loss of NF1 results in activation of the signaling pathway and leads to aberrant growth in murine and human hematopoietic cells. Nat Genet 12:144–48 Braun BS, Tuveson DA, Kong N, Le DT, Kogan SC, Rozmus J, Le Beau MM, Jacks TE, Shannon KM (2004) Somatic activation of oncogenic Kras in hematopoietic cells initiates a rapidly fatal myeloproliferative disorder. Proc Natl Acad Sci USA 101:597–602 Caraglia M, Tassone P, Marra M, Budillon A, Venuta S, Tagliaferri P (2006) Targeting Raf-kinase: molecular rationales and translational issues. Ann Oncol 17 Suppl. 7:vii124–7 Castleberry RP, Loh ML, Jayaprakash N, Peterson A, Casey V, Chang M, Widemann B, Emanuel PD (2005) Phase II WIndow Study of R115777(Zarnesta) in Untreated Juvenile Myelomonocytic Leukemia (JMML): a Children’s Oncology Group Study. Blood 106:727a Chan G, Kalaitzidis D, Usenko T, Kutok JL, Yang W, Mohi MG, Neel BG (2009) Leukemogenic Ptpn11 causes fatal myeloproliferative disorder via cell-autonomous effects on multiple stages of hematopoiesis. Blood 113:4414–24 Chan IT, Kutok JL, Williams IR, Cohen S, Kelly L, Shigematsu H, Johnson L, Akashi K, Tuveson DA, Jacks T, Gilliland DG (2004) Conditional expression of oncogenic K-ras from its endogenous promoter induces a myeloproliferative disease. J Clin Invest 113:528–38 Chen L, Sung SS, Yip ML, Lawrence HR, Ren Y, Guida WC, Sebti SM, Lawrence NJ, Wu J (2006) Discovery of a novel shp2 protein tyrosine phosphatase inhibitor. Mol Pharmacol 70:562–70 Cichowski K, Jacks T (2001) NF1 tumor suppressor gene function: narrowing the GAP. Cell 104:593–604 Cuiffo B, Ren R (2010) Palmitoylation of oncogenic N is essential for leukemogenesis. Blood 115:3598–605 de Vries AC, Stam RW, Kratz CP, Zenker M, Niemeyer CM, van den Heuvel-Eibrink MM (2007a) Mutation analysis of the BRAF oncogene in juvenile myelomonocytic leukemia. Haematologica 92:1574–5 de Vries AC, Stam RW, Schneider P, Niemeyer CM, van Wering ER, Haas OA, Kratz CP, den Boer ML, Pieters R, van den Heuvel-Eibrink MM (2007b) Role of mutation independent constitutive activation of FLT3 in juvenile myelomonocytic leukemia. Haematologica 92:1557–60 Donovan S, Shannon KM, Bollag G (2002) GTPase activating poteins: critical regulators of intracellular signaling. BBA Rev Cancer 1602:23–45 Emanuel PD, Bates LJ, Castleberry RP, Gualtieri RJ, Zuckerman KS (1991) Seletive hypersensitivity to granulocyte-macrophage colony stimulating factor by juvenile chronic myeloid leukemia hematopoietic progenitors. Blood 77:925–9 Emanuel PD, Shannon KM, Castleberry RP (1996) Juvenile myelomonocytic leukemia: molecular understanding and prospects for therapy. Mol Med Today 2:468–75 Emanuel PD, Snyder RC, Wiley T, Gopurala B, Castleberry RP (2000) Inhibition of juvenile myelomonocytic leukemia cell growth in€vitro by farnesyltransferase inhibitors. Blood 95:639–45 Flotho C, Steinemann D, Mullighan CG, Neale G, Mayer K, Kratz CP, Schlegelberger B, Downing JR, Niemeyer CM (2007) Genome-wide single-nucleotide polymorphism analysis in juvenile myelomonocytic leukemia identifies uniparental disomy surrounding the NF1 locus in cases associated with neurofibromatosis but not in cases with mutant or PTPN11. Oncogene 26:5816–21 Flotho C, Valcamonica S, Mach-Pascual S, Schmahl G, Corral L, Ritterbach J, Hasle H, Arico M, Biondi A, Niemeyer CM (1999) mutations and clonality analysis in children with juvenile myelomonocytic leukemia (JMML). Leukemia 13:32–7 Hellmuth K, Grosskopf S, Lum CT, Wurtele M, Roder N, von Kries JP, Rosario M, Rademann J, Birchmeier W (2008) Specific inhibitors of the protein tyrosine phosphatase Shp2 identified by high-throughput docking. Proc Natl Acad Sci USA 105:7275–80
136
J. O’Hara Lauchle and B.S. Braun
Hexner EO, Serdikoff C, Jan M, Swider CR, Robinson C, Yang S, Angeles T, Emerson SG, Carroll M, Ruggeri B, Dobrzanski P (2008) Lestaurtinib (CEP701) is a JAK2 inhibitor that suppresses JAK2/STAT5 signaling and the proliferation of primary erythroid cells from patients with myeloproliferative disorders. Blood 111:5663–71 Hingorani SR, Tuveson DA (2003) Targeting oncogene dependence and resistance. Cancer Cell 3:414–7 Iversen PO, Emanuel PD, Sioud M (2002) Targeting Raf-1 gene expression by a DNA enzyme inhibits juvenile myelomonocytic leukemia cell growth. Blood 99:4147–53 Jacks T, Shih S, Schmitt EM, Bronson RT, Bernards A, Weinberg RA (1994) Tumorigenic and developmental consequences of a targeted Nf1 mutation in the mouse. Nat Genet 7:353–61 Keilhack H, David FS, McGregor M, Cantley LC, Neel BG (2005) Diverse biochemical properties of Shp2 mutants. Implications for disease phenotypes. J Biol Chem 280:30984–93 Kim E, Ambroziak P, Otto JC, Taylor B, Ashby M, Shannon K, Casey PJ, Young SG (1999) Disruption of the mouse Rce1 gene results in defective processing and mislocalization of within cells. J Biol Chem 274:8383–90 Kotecha N, Flores NJ, Irish JM, Simonds EF, Sakai DS, Archambeault S, Diaz-Flores E, Coram M, Shannon KM, Nolan GP, Loh ML (2008) Single-cell profiling identifies aberrant STAT5 activation in myeloid malignancies with specific clinical and biologic correlates. Cancer Cell 14:335–43 Kratz CP, Niemeyer CM, Thomas C, Bauhuber S, Matejas V, Bergstrasser E, Flotho C, Flores NJ, Haas O, Hasle H, van den Heuvel-Eibrink MM, Kucherlapati RS, Lang P, Roberts AE, Stary J, Strahm B, Swanson KD, Trebo M, Zecca M, Neel B, Locatelli F, Loh ML, Zenker M (2007) Mutation analysis of Son of Sevenless in juvenile myelomonocytic leukemia. Leukemia 21:1108–9 Largaespada DA, Brannan CI, Jenkins NA, Copeland NG (1996) Nf1 deficiency causes mediated granulocyte/macrophage colony stimulating factor hypersensitivity and chronic myeloid leukaemia. Nat Genet 12:137–43 Lauchle JO, Braun BS, Loh ML, Shannon K (2006) Inherited predispositions and hyperactive in myeloid leukemogenesis. Pediatr Blood Cancer 46:579–85 Lauchle JO, Le DT, Kim D, Akagi K, Gorman MF, Tran M, Sebolt-Leopold J, Wolff L, Parada LF, Jenkins N, Copeland N, Shannon KM (2007) Mutations that cooperate with Nf1 inactivation in leukemogenesis influence therapeutic response to MEK inhibition. Blood 110:594 Kong N, Zhu Y, Lauchle JO, Aiyigari A, Braun BS, Wang E, Kogan SC, Lebeau MM, Parada L, Shannon KM (2004) Somatic inactivation of Nf1 in hematopoietic cells results in a progressive myeloproliferative disorder. Blood 103:4243–50 Liu YL, Castleberry RP, Emanuel PD (2004) Rapamycin a potential mechanistically targeted therapeutic for juvenile myelomonocytic leukemia. Blood 104:653a Locatelli F, Nollke P, Zecca M, Korthof E, Lanino E, Peters C, Pession A, Kabisch H, Uderzo C, Bonfim CS, Bader P, Dilloo D, Stary J, Fischer A, Revesz T, Fuhrer M, Hasle H, Trebo M, van den Heuvel-Eibrink MM, Fenu S, Strahm B, Giorgiani G, Bonora MR, Duffner U, Niemeyer CM (2005) Hematopoietic stem cell transplantation (HSCT) in children with juvenile myelomonocytic leukemia (JMML): results of the EWOG-MDS/EBMT trial. Blood 105:410–9 Loh ML, Vattikuti S, Schubbert S, Reynolds MG, Carlson E, Lieuw KH, Cheng JW, Lee CM, Stokoe D, Bonifas JM, Curtiss NP, Gotlib J, Meshinchi S, Le Beau MM, Emanuel PD, Shannon KM (2004) Mutations in PTPN11 implicate the SHP-2 phosphatase in leukemogenesis. Blood 103:2325–31 Loh ML, Sakai DS, Flotho C, Kang M, Fliegauf M, Archambeault S, Mullighan CG, Chen L, Bergstraesser E, Bueso-Ramos CE, Emanuel PD, Hasle H, Issa JP, van den Heuvel-Eibrink MM, Locatelli F, Stary J, Trebo M, Wlodarski M, Zecca M, Shannon KM, Niemeyer CM (2009) Mutations in CBL occur frequently in juvenile myelomonocytic leukemia. Blood 114:1859–63 Mahgoub N, Taylor BR, Gratiot M, Kohl NE, Gibbs JB, Jacks T, Shannon KM (1999) In vitro and In vivo effects of a farnesyltransferase inhibitor on Nf1- deficient hematopoietic cells. Blood 94:2469–76
Targeting RAS Signaling Pathways in Juvenile Myelomonocytic Leukemia (JMML)
137
Michaelson D, Ali W, Chiu VK, Bergo M, Silletti J, Wright L, Young SG, Philips M (2005) Postprenylation CAAX processing is required for proper localization of but not Rho GTPases. Mol Biol Cell 16:1606–16 Miles DK, Freedman MH, Stephens K, Pallavicini M, Sievers EL, Weaver M, Grunberger T, Thompson P, Shannon KM (1996) Patterns of hematopoietic lineage involvement in children with neurofibromatosis type 1 and malignant myeloid disorders. Blood 88:4314–20 Miyauchi J, Asada M, Sasaki M, Tsunematsu Y, Kojima S, Mizutani S (1994) Mutations of the N-ras gene in juvenile chronic myelogenous leukemia. Blood 83:2248–54 Mohi MG, Williams IR, Dearolf CR, Chan G, Kutok JL, Cohen S, Morgan K, Boulton C, Shigematsu H, Keilhack H, Akashi K, Gilliland DG, Neel BG (2005) Prognostic, therapeutic, and mechanistic implications of a mouse model of leukemia evoked by Shp2 (PTPN11) mutations. Cancer Cell 7:179–91 Nakamura Y, Ito M, Yamamoto T, Yan XY, Yagasaki H, Kamachi Y, Kudo K, Kojima S (2005) Engraftment of NOD/SCID/gammac(null) mice with multilineage neoplastic cells from patients with juvenile myelomonocytic leukaemia. Br J Haematol 130:51–7 Niihori T, Aoki Y, Ohashi H, Kurosawa K, Kondoh T, Ishikiriyama S, Kawame H, Kamasaki H, Yamanaka T, Takada F, Nishio K, Sakurai M, Tamai H, Nagashima T, Suzuki Y, Kure S, Fujii K, Imaizumi M, Matsubara Y (2005) Functional analysis of PTPN11/SHP-2 mutants identified in Noonan syndrome and childhood leukemia. J Hum Genet 50:192–202 Ohren JF, Chen H, Pavlovsky A, Whitehead C, Zhang E, Kuffa P, Yan C, McConnell P, Spessard C, Banotai C, Mueller WT, Delaney A, Omer C, Sebolt-Leopold J, Dudley DT, Leung IK, Flamme C, Warmus J, Kaufman M, Barrett S, Tecle H, Hasemann CA (2004) Structures of human MAP kinase kinase 1 (MEK1) and MEK2 describe novel noncompetitive kinase inhibition. Nat Struct Mol Biol 11:1192–7 Pardanani A (2008) JAK2 inhibitor therapy in myeloproliferative disorders: rationale, preclinical studies and ongoing clinical trials. Leukemia 22:23–30 Park S, Chapuis N, Bardet V, Tamburini J, Gallay N, Willems L, Knight ZA, Shokat KM, Azar N, Viguie F, Ifrah N, Dreyfus F, Mayeux P, Lacombe C, Bouscary D (2008) PI-103, a dual inhibitor of Class IA phosphatidylinositide 3-kinase and mTOR, has antileukemic activity in AML. Leukemia 22:1698–706 Peterson YK, Kelly P, Weinbaum CA, Casey PJ (2006) A novel protein geranylgeranyltransferase-I inhibitor with high potency, selectivity, and cellular activity. J Biol Chem 281:12445–50 Rinehart J, Adjei AA, Lorusso PM, Waterhouse D, Hecht JR, Natale RB, Hamid O, Varterasian M, Asbury P, Kaldjian EP, Gulyas S, Mitchell DY, Herrera R, Sebolt-Leopold JS, Meyer MB (2004) Multicenter phase II study of the oral MEK inhibitor, CI-1040, in patients with advanced non-small-cell lung, breast, colon, and pancreatic cancer. J Clin Oncol 22:4456–62 Rotblat B, Ehrlich M, Haklai R, Kloog Y (2008) The inhibitor farnesylthiosalicylic acid (Salirasib) disrupts the spatiotemporal localization of active: a potential treatment for cancer. Methods Enzymol 439:467–89 Sanada M, Suzuki T, Shih LY, Otsu M, Kato M, Yamazaki S, Tamura A, Honda H, SakataYanagimoto M, Kumano K, Oda H, Yamagata T, Takita J, Gotoh N, Nakazaki K, Kawamata N, Onodera M, Nobuyoshi M, Hayashi Y, Harada H, Kurokawa M, Chiba S, Mori H, Ozawa K, Omine M, Hirai H, Nakauchi H, Koeffler HP, Ogawa S (2009) Gain-of-function of mutated C-CBL tumour suppressor in myeloid neoplasms. Nature 460:904–8 Schubbert S, Lieuw K, Rowe SL, Lee CM, Li X, Loh ML, Clapp DW, Shannon KM (2005) Functional analysis of leukemia-associated PTPN11 mutations in primary hematopoietic cells. Blood 106:311–7 Shannon KM, O’Connell P, Martin GA, Paderanga D, Olson K, Dinndorf P, McCormick F (1994) Loss of the normal NF1 allele from the bone marrow of children with type 1 neurofibromatosis and malignant myeloid disorders. N Engl J Med 330:597–601 Side L, Taylor B, Cayouette M, Connor E, Thompson P, Luce M, Shannon K (1997) Homozygous inactivation of the NF1 gene in bone marrow cells from children with neurofibromatosis type 1 and malignant myeloid disorders. N Engl J Med 336:1713–20
138
J. O’Hara Lauchle and B.S. Braun
Sun J, Ohkanda J, Coppola D, Yin H, Kothare M, Busciglio B, Hamilton AD, Sebti SM (2003) Geranylgeranyltransferase I inhibitor GGTI-2154 induces breast carcinoma apoptosis and tumor regression in H transgenic mice. Cancer Res 63:8922–9 Tartaglia M, Mehler EL, Goldberg R, Zampino G, Brunner HG, Kremer H, van der Burgt I, Crosby AH, Ion A, Jeffery S, Kalidas K, Patton MA, Kucherlapati RS, Gelb BD (2001) Mutations in PTPN11, encoding the protein tyrosine phosphatase SHP-2, cause Noonan syndrome. Nat Genet 29:465–8 Tartaglia M, Niemeyer CM, Fragale A, Song X, Buechner J, Jung A, Hahlen K, Hasle H, Licht JD, Gelb BD (2003) Somatic mutations in PTPN11 in juvenile myelomonocytic leukemia, myelodysplastic syndromes and acute myeloid leukemia. Nat Genet 34:148–50 Wahlstrom AM, Cutts BA, Karlsson C, Andersson KM, Liu M, Sjogren AK, Swolin B, Young SG, Bergo MO (2007) Rce1 deficiency accelerates the development of K-induced myeloproliferative disease. Blood 109:763–8 Wahlstrom AM, Cutts BA, Liu M, Lindskog A, Karlsson C, Sjogren AK, Andersson KM, Young SG, Bergo MO (2008) Inactivating Icmt ameliorates K-induced myeloproliferative disease. Blood 112:1357–65 Watanabe S, Itoh T, Arai K (1996) Roles of JAK kinases in human GM-CSF receptor signal transduction. J Allergy Clin Immunol 98:S183–91 Weinstein IB (2002) Cancer. Addiction to oncogenes – the Achilles heal of cancer. Science 297:63–4 Wilhelm SM, Carter C, Tang L, Wilkie D, McNabola A, Rong H, Chen C, Zhang X, Vincent P, McHugh M, Cao Y, Shujath J, Gawlak S, Eveleigh D, Rowley B, Liu L, Adnane L, Lynch M, Auclair D, Taylor I, Gedrich R, Voznesensky A, Riedl B, Post LE, Bollag G, Trail PA (2004) BAY 43-9006 exhibits broad spectrum oral antitumor activity and targets the RAF/MEK/ERK pathway and receptor tyrosine kinases involved in tumor progression and angiogenesis. Cancer Res 64:7099–109 Winter-Vann AM, Casey PJ (2005) Post-prenylation-processing enzymes as new targets in oncogenesis. Nat Rev Cancer 5:405–12 Zhu Y, Romero MI, Ghosh P, Ye Z, Charnay P, Rushing EJ, Marth JD, Parada LF (2001) Ablation of NF1 function in neurons induces abnormal development of cerebral cortex and reactive gliosis in the brain. Genes Dev 15:859–76
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics Seth J. Corey and Jorge Cortes
Chronic Myeloid Leukemia (CML) is a clonal myeloproliferative disorder Â�characterized by a translocation between the long arms of chromosomes 9 and 22. Often referred to as the Philadelphia (Ph) chromosome, this genetic rearrangement results in an oncogenic tyrosine kinase, Bcr-Abl. The disease may smolder for years before terminating in a blast crisis. Once curable, mostly with bone marrow transplantation and only seldom with interferon alpha-based therapy, its management has been radically changed with the introduction of imatinib mesylate, an orally available Abl kinase inhibitor, in 1998. More common in adults, only an estimated 50 pediatric cases occur annually in North America. The treatment of CML will remain in flux as second-generation Bcr-Abl kinase inhibitors are studied clinically. Because of its rarity in pediatrics, clinical insights must come from Â�experience with adult CML. CML serves as the paradigm of how molecular understanding of a cancer’s pathophysiology can produce a revolutionary form of Â�targeted therapy, providing hope for those pediatric cancers with a well-defined genetic lesion.
Pathophysiology of Bcr-Abl in CML More than 95% of all adult and pediatric CML possess the Ph chromosome (designated as Ph+ or Ph1) (Melo and Barnes 2007; Sherbenou and Druker 2007). This results from a reciprocal and balanced translocation in hematopoietic stem cells, which may be more precisely described t(9;22)(q34;q11.2) (Nowell and Hungerford 1960; Rowley 1973). Consequently, the Abl gene on chromosome 9 fuses to the Bcr
S.J. Corey (*) Departments of Pediatrics and Cell & Molecular Biology, Children’s Memorial Hospital and the Robert H. Lurie Comprehensive Cancer Center, Northwestern University, Chicago, IL e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_8, © Springer Science+Business Media, LLC 2010
139
140
S.J. Corey and J. Cortes
gene on chromosome 22 (Ben-Neriah et€al. 1986; Shtivelman et€al. 1985). The fusion mRNA transcript encodes a protein, Bcr-Abl, with constitutive tyrosine kinase activity. Bcr stands for breakpoint cluster region, because there are variable breakpoints at which the translocation arises. The two most common breakpoints produce transcripts designated as e13a2 and e14a2 (also known as b2a2 and b3a2) typically found in CML. Mapping specific sites of chromosomal translocations aids in the design of quantitative PCR, which is commercially available, to monitor disease activity. The fusion of the partial genes for Bcr and Abl produce transcripts encoding either a 210 or approximately 190€kDa protein (Shtivelman et€al. 1985). The larger (p210 Bcr-Abl) is found predominantly in CML, while the smaller form (p190 Bcr-Abl) occurs most frequently in Ph+ acute lymphoblastic leukemia (ALL) (Chan et€al. 1987). Discussion of therapeutically targeting the p190 Bcr-Abl in ALL is discussed in Chap. 20. Bcr itself is a serine/threonine kinase and contains regulatory domains that affect the activities of RAS-related proteins Rac and Cdc42 (Diekmann et€al. 1991; Maru and Witte 1991). Named after a viral transforming gene (v-Abl, Abelson murine leukemia virus) that Herb Abelson discovered as pediatric hematology fellow (Abelson and Rabstein 1970; Wang et€al. 1984), cellular Abl (c-Abl) is a tyrosine kinase (Groffen et€ al. 1983). Weakly active in its proto-oncogenic form, Abl becomes a much more potent intracellular tyrosine kinase that is constitutively activated upon fusion with Bcr. (Fig.€1). Bcr-Abl kinase drives cellular proliferation and inhibits apoptosis of myeloid progenitor and precursor cells. Ultimately, differentiation to neutrophils is impaired and blasts accumulate, leading to an acute leukemia-like presentation known as blast phase (BP). As the disease evolves, additional chromosomal abnormalities and genetic mutations may arise. Corresponding to the progression are three clinically defined phases: chronic, accelerated, and blast (discussed below). Murine retroviral transplantation experiments and the clinical success of imatinib confirm that Bcr-Abl is necessary and sufficient for CML. Bone marrow mononuclear cells harvested from healthy mice can be infected with retrovirus containing the Bcr-Abl gene (Daley et€al. 1990). These modified cells, now expressing Bcr-Abl kinase, can be transplanted into irradiated, genetically related mice. Within several weeks, the transplant engrafts and a fatal myeloproliferative disease develops. At the same time that the mouse model for CML was being developed, organic chemists were synthesizing an array of tyrosine kinase inhibitors (Lydon and Druker 2004). Tyrosine kinases may be found in one of about a dozen families of receptors, such as those for epidermal growth factor, insulin, and �platelet-derived growth factor. At least eight families of non-receptor (or cytosolic) tyrosine kinases exist: Csk, Src, Btk/Itk, Syk/Zap-70, Fes, FAK/Pyk2, Jak, and Abl/Arg. Structural similarities among the several hundred receptor and cytosolic tyrosine kinases exist, which result in only relative specificity for most inhibitors. One lead compound (CGP57148B, later renamed STI571, and now known as imatinib mesylate) was discovered by scientists at Ciba-Geigy and showed potency against Abl and PDGF Receptors (Buchdunger et€ al. 1996). Administration of that Abl tyrosine kinase inhibitor, eventually known as imatinib mesylate �(commercially known as
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
141
Fig.€1â•… Structure and signaling function of Bcr-Abl. Upper panel, structure of Bcr and Abl. Arrows point to sites of where the two proteins fuse forming either a p190 or, more commonly, p210 fusion. For Bcr, portions of the N-terminal Ser/Thr kinase is preserved. For Abl, most of the domains, except for SH3, are retained. Lower panel, the Bcr-Abl fusion kinase recruits multiple signaling pathways. A few of the more important ones are shown here. When phosphorylated, Grb2 binds to Tyr177. This recruits two pathways, Sos-RAS-Raf-MAP Kinase and Gab2-PI 3¢kinase-Akt. Additional signaling come through activation of the Src kinase Lyn and the transcription factor STAT5. Mitochondrial associated apoptosis is deregulated through effects on Bad and Bcl-XL by Akt and STAT5, respectively. Altogether, these signaling pathways affect proliferation, survival, altered adhesiveness, and loss of differentiation
Gleevec™ or Glivec™), to these retrovirally transduced cell lines (Druker et€ al. 1996) or mice (le Coutre et€ al. 1999) or patients (Druker et€ al. 2001) prevented development of the fatal disease. Additional genetic lesions contribute to or drive the disease through its stages. However, Bcr-Abl remains a critical target in disease progression (Shah et€al. 2007). The structure and signaling function of Bcr-Abl has been thoroughly investigated, and still, new insights into how this fusion tyrosine kinase causes CML arise (Melo and Barnes 2007). Bcr-Abl tyrosine kinase’s activity is absolutely required for leukemogenesis. Increased genomic instability arising from the abnormal kinase likely contributes to disease progression with accumulation of additional genetic mutations. The constitutive tyrosine kinase activity of Bcr-Abl Â�phosphorylates a number of substrates, which activate a variety of signaling pathways (Fig.€ 1). Abl’s kinase activity autophosphorylates tyrosine 177 of Bcr-Abl, which serves as
142
S.J. Corey and J. Cortes
a docking site for the SH2 domain of Grb2 (Pendergast et€al. 1993). In turn, Grb2 recruits SOS, a GTP exchange factor for RAS. RAS activates both PI 3¢kinase and Raf (and subsequently MAP Kinases). Grb2 also binds another scaffolding protein, Gab2, which also activates PI 3¢kinase (Sattler et€al. 2002). PI 3¢kinase phosphorylates phosphatidylinositol at the 3¢ position of inositol. This leads to several 3¢OH-phosphoinositides. One major effect is to activate the serine/threonine kinase Akt (Skorski et€al. 1997). Akt phosphorylates a variety of substrates involved in the regulation of metabolism and apoptosis control. One physiologically important substrate is Bad, which when phosphorylated, can no longer promote mitochondrial-associated apoptosis. Through less understood mechanisms, Bcr-Abl activates Src family kinases, such as Lyn and Hck (Danhauser-Riedl et€al. 1996). These may serve to recruit additional or salvage growth-promoting pathways. Bcr-Abl also leads to activation of the transcription factor STAT5 (Carlesso et€al. 1996; Ilaria and Van Etten 1996). STAT5 promotes the expression of the anti-apoptotic factor, Bcl-XL. Another group of Bcr-Abl substrates are cytoskeletal proteins, such as CrkL, paxillin, talin, and focal adhesion kinase, which affect cellular adhesiveness (Salgia et€al. 1997). As a result of fusion of Bcr with Abl, normal cellular functions of c-Abl (and Bcr) are lost. Since c-Abl binds to DNA and is involved in mediating genotoxic repair, genetic instability results. Altogether, these signaling events promote cell cycle division, survival, altered adhesiveness, and genetic instability (Skorski 2002). Ultimately, a loss of differentiation occurs (Schuster et€al. 2003).
Clinical Phases Pediatric CML is a rare disease with an annual incidence of 1 case per million children (the incidence in adults is approximately ten times greater) (Linet et€ al. 1999). The clinical characteristics do not differ from those seen in the adult population, with the exception of age at presentation, and perhaps, white blood cell count (Millot et€al. 2005). There is a slight preponderance of boys affected (ratio of 1.5:1) with a median age at diagnosis of 12.5 years. A third of the children will be diagnosed within the first decade of life. Almost all children with CML (approximately 95%) are diagnosed during the chronic phase (CP). Common presenting symptoms are: weakness, abdominal discomfort or fullness, weight loss, bleeding, fever, and bone pain. More than 2/3 of the children will have a palpable spleen on examination; most will not have a palpable liver or enlarged lymph nodes. Hematologic findings consist of elevated white blood cell count, with almost two-thirds having a count greater than 100â•›×â•›109/L. Two-thirds will also have an anemia (hemoglobin less than 12€g/dL) and thrombocytosis (greater than 450â•›×â•›109/L). The CP of CML presents as a myeloproliferative disease, with an increasing number of normal appearing and functioning granulocytes and their precursors (metamyelocytes and promyelocytes). Blasts count remains below 15% in both the peripheral circulation and the bone marrow. On examination, the spleen becomes increasingly enlarged. The CP typically lasts between 3 and 5 years. Before the disease terminates in a BP, which may be either myeloid (predominantly) or
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
143
Â� lymphoid, there is frequently an intermediary accelerated phase (AP). Almost all of the children never go through AP, or it goes unrecognized. There are several criteria the different stages of CML (Cortes et€al. 2006). As defined by the World Health Organization, criteria for AP include any of the following: 10 to 19% blasts in either the periphery or bone marrow, greater than 20% basophils in blood or marrow, platelet count either less than 100â•›×â•›109â•›L or greater than 1,000â•›×â•›109â•›L, new cytogenetic abnormalities, or increasing splenomegaly or white blood count unresponsive to therapy. However, these criteria have never been used prospectively, and several characteristics do not appear to remain as predictive of outcome in patients treated with tyrosine kinase inhibitors. The duration of AP is short, lasting 1 to 2 years, and usually terminates with a BP. Though there may not be an antecedent AP, BP occurs when there are at least 30% blasts (20% according to the WHO classification) in either the periphery or bone marrow or evidence of a chloroma or large clusters of blasts in the bone marrow. Blasts most frequently display myeloid markers, but a few will be lymphoid, and some undifferentiated.
Targeted Therapies Imatinib is a small molecule tyrosine kinase inhibitor, a 2-phenylaminopyrimidine compound originally identified in a drug screen for activity against the PDGF Receptor (Lydon and Druker 2004; Buchdunger et€al. 1996). Pre-clinical evaluation of the compound (known then as CGP57148B and later as STI571) demonstrated its exquisite sensitivity against Bcr-Abl-transfected cell lines, cells derived from patients with CML, and mouse retroviral transduction models (Druker et€al. 2001). Active also against stem cell factor receptor (c-Kit) and Abl, imatinib was tested first in patients with interferon-resistant CML, and subsequently, in patients with gastrointestinal stromal cell tumor (GIST) where activating mutations of c-Kit are commonly found (Demetri et€al. 2002). Of note, imatinib interferes with the ATP binding domain of the inactive form of Bcr-Abl. For both diseases, responses were swift and dramatic. Complete cytogenetic remission can now be achieved in more than 80% of patients with CML in CP. The initial phase I trial of imatinib in patients with CP CML resistant to interferon-a was conducted in 1998 (Druker et€al. 2001). Though a maximally tolerated dose was not estAblished, even at doses up to 1,000€ mg daily, an oral dose of 400€mg daily for adults was selected for phase II trials. Responses in CML may be defined as hematologic, cytogenetic, or molecular (Table€1). A phase III International Randomized Study of Interferon and STI571 (IRIS) cross-over study compared imatinib 400€mg/day with combination of interferon-a and cytarabine in untreated CML patients in CP (O’Brien et€al. 2003). As reported in 2003, complete hematologic response was seen in 96% of patients treated with imatinib (76% with Â�complete cytogenetic responses). With 5 years of additional follow-up, projected rates of complete hematologic responses were 98% and complete cytogenetic responses were 87%. Responses are mostly durable, with 5-year event-free survival rates of 83%, with 93% alive and free from transformation to accelerated or BP.
144 Table€1╅ Response criteria Hematologic ╅ Complete
â•… Partial
Cytogenetic â•… Complete â•… Major â•… Minor Molecular â•… Complete
S.J. Corey and J. Cortes
Complete normalization of peripheral blood counts Leukocyte count <10â•›×â•›109/L Platelet count <450â•›×â•›109/L No immature cells in peripheral blood Disappearance of palpable spleen Presence of immature cells (myelocytes, promyelocytes, blasts) Platelet count <50% of pre-treatment count, but >450â•›×â•›109/L Persistent splenomegaly No Ph+ metaphases 1–35% Ph+ metaphases >35% Ph+ metaphases
Bcr-Abl undetectable by qPCR 5-Log reduction in Bcr-Abl transcript levels â•… Major Bcr-Abl/Abl <0.05% 3-Log reduction in Bcr-Abl transcript levels Cytogenetic analysis requires at least 20 metaphases qPCR quantitative PCR
Five-year survival was 90% (92% when censoring patients at the time of transplant). Importantly, the response achieved at 12 months correlated with the longterm outcome. Patients who achieved a complete cytogenetic response and at least a 3-log reduction in Bcr-Abl transcript levels (i.e., major molecular response) after 12 months of therapy had a significantly better probability of being alive and free from transformation to accelerated or CP at 5 years compared to those with complete cytogenetic response but less than a 3-log reduction in transcript levels, and those without a complete cytogenetic response (100% vs. 95% vs. 88%, respectively; pâ•›<â•›0.001) (Druker et€al. 2006). With these results, achieving a major molecular response has become the main goal of therapy in CML, and adequate molecular monitoring has become mandatory. Clinical experience based on treatment of over 60,000 patients has shown that imatinib is well-tolerated with once-daily administration (Baccarani et€ al. 2006). The drug is cleared through the liver, and thus, the ingestion of grapefruit juice or other known cytochrome P450 inhibitors is to be discouraged (Peng et€ al. 2005). These drugs include dexamethasone, phenobarbital, carbamezipine, and St. John’s Wort. Administering the drug with meals reduces gastrointestinal distress; however, an occasional patient may need anti-emetics such as prochlorperazine or ondanestron (Deininger et€al. 2003). Diarrhea may also occur but it is frequently mild, not needing therapy; when it does, imodium adequately controls it in most instances. Approximately 60% will show signs of peripheral edema (e.g., peripheral edema, periorbital swelling) (Cohen et€al. 2002). This tends to subside, and can be treated with diuretics (Esmaeli et€al. 2002). Dyspnea can occur in approximately 12%, usually Â�representing fluid retention; it may be treated with diuretics and/or brief pulse
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
145
of prednisone. Myelosuppression occurs frequently, particularly in the first 2 to 3 months of therapy (Cohen et€al. 2002). If grade 3 to 4 neutropenia (<109/L) and/or thrombocytopenia (<50â•›×â•›109/L) occur, imatinib may be discontinued until the neutrophils and/or platelets return to above these levels, and then resumed. The dose may be reduced if the recovery takes longer than 2 weeks. A recent report of severe congestive heart failure in ten patients treated with imatinib and without known prior history of cardiovascular disease has not been confirmed by subsequent monitoring (Kerkela et€al. 2006; Verweij et€al. 2007; Atallah et€al. 2007a) The incidence of heart failure among patients treated with imatinib appears not to be different than the risk in the general population. The incidence of congestive heart failure is rare and is associated with advanced age with pre-existing heart disease (Atallah et€al. 2007b). Very little is known about long-term effects of imatinib in children (Bond 2007; Millot et€al. 2006a; Millot et€al. 2006b; Champagne et€al. 2004). Although imatinib uncommonly affects bone metabolism through hypophosphatemia (Berman et€ al. 2006), stunting of growth or osteoporosis has not been noted in pediatrics.
Imatinib in Children A Children’s Oncology Group Phase I study reported that imatinib was well tolerated (Champagne et€al. 2004). The study group of 31 patients included 14 patients with CP, 6 with myeloid BP, and 9 with Ph+ ALL. CML patients were resistant to interferon-a. All 14 patients with CP had a hematologic response, with 10 having a major cytogenetic response. There was a 55% partial response rate among those with advanced disease. No maximally tolerated dose was defined, with 570€mg/m2 as the largest dose studied (comparable to 800€mg adult high-dose). Pharmacokinetic data showed steady state levels and area under the curve concentrations similar to those found in adults given 400€mg/m2/day. Also, as seen in adult patients, there was interpatient variability, perhaps due to CYP3A activity (Peng et€al. 2005; Peng et€al. 2004). A multi-center phase II study of imatinib conducted in children with AP (either after transplantation or with interferon resistance) showed a 75 to 80% complete hematologic response (Millot et€ al. 2006a). One-year survival in this high-risk group was 95% for those with CP and 75% with AP. Since imatinib is the only currently FDA-approved drug for pediatric CML, we propose in Fig.€2, the following imatinib-based management of a child or adolescent with CML.
Monitoring With improved results with imatinib, adequate monitoring has become more important than ever. In addition to performing clinical evaluations and monitoring of peripheral blood counts, response assessment requires measurement of disease by the presence of the Philadelphia chromosome. Routine cytogenetics by G-banding are required at baseline to determine the presence of the Philadelphia chromosome as
146
S.J. Corey and J. Cortes
Fig.€2â•… Guidelines for managing a pediatric patient with CML. To prevent leukostasis, hydroxyurea (500 to 1,000€ mg/m2/d) may be given for several days. Imatinib dose should be started at 250€mg/m2/d. Dosage may be doubled if no response occurs. Therapy may be temporarily discontinued for neutropenia (<1â•›×â•›109/L) or thrombocytopenia (<50â•›×â•›109/L). For side effects, ondanestron may be given with imatinib. Upon attainment of cytogenetic response, qPCR of peripheral blood should be performed every 3 months. Bone marrow examination with cytogenetics should be performed yearly. Treatment options for imatinib/dasatinib resistance include cytarabine/interferon (IFN)a and allogeneic stem cell transplantations (SCT)
well as other chromosomal abnormalities that may have prognostic implications. During follow-up, cytogenetic studies of the bone marrow may be needed at least once per year to investigate the possible emergence of clonal evolution and the occurrence of chromosomal abnormalities in Ph-negative metaphases. This phenomenon is found in approximately 15% of patients and has rarely been associated with development of myelodysplastic syndromes and acute leukemias (Kovitz et€al. 2006). Patients must also be followed with real-time (quantitative) PCR to measure the transcript levels of Bcr-Abl. This is particularly important after a patient has achieved a complete cytogenetic remission since achievement of a major molecular remission correlates with an improved outcome. Unfortunately, the methodology used for PCR testing is not universal and there is considerable variability between the results obtained in different laboratories. Recently, a group of experts in the field proposed a harmonization that would convert the results obtained in different laboratories to an international standard (Hughes et€al. 2006). This process is still being developed.
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
147
Resistance to Kinase Inhibitors and New Therapeutic Strategies The optimal duration of imatinib therapy is not known (Goldman 2007). Despite anecdotal reports of patients maintaining remission off imatinib, most patients who discontinue therapy, despite having achieved undetectable disease by PCR, relapse. Thus, the current recommendation is that patients should remain on imatinib indefinitely. Resistance to imatinib therapy has been defined based on the response achieved at different times and can be either primary (refractory) or secondary. After 5 years of follow-up, it appears that the rate of resistance is approximately 3 to 4% per year and may decrease as patients remain on therapy longer (Druker et€al. 2006; Hughes et€al. 2003). Resistance occurs by several mechanisms (Michor et€al. 2005). The most common of these is the development of mutations in the Bcr-Abl kinase domain (Gorre et€ al. 2001; Shah et€ al. 2002). Mutations have been identified in 40 to 60% of patients affecting over 50 different residues. Mutations may confer different levels of resistance to imatinib, with some conferring only partial resistance (i.e., resistance can be overcome by higher concentrations of imatinib) and others conferring high Â�levels of resistance (i.e., T315I). Additional mechanisms for resistance include amplification or overexpression of Bcr-Abl. Non-Bcr-Abl related mechanisms of resistance have also been described. These include overexpression of multi-drug resistance phenotype that may expel imatinib out of the cell and decreased activity of the hOCT1 transporter that is required for importing imatinib into the cell (Thomas et€ al. 2004). The emergence of salvage pathways (in particular, the increased expression of the Src kinase Lyn) has been reported in some patients after imatinib failure, although the actual incidence of this mechanism among patients who fail imatinib therapy is not known (Dai et€ al. 2004; Donato et€ al. 2003). In addition, it has been suggested that the CML leukemic stem cell remains in a Â�dormant state and is insensitive to imatinib and could thus be responsible for the Â�persistence of disease (Elrick et€al. 2005). A second generation of tyrosine kinase inhibitors has been developed to Â�overcome resistance to imatinib. The Food and Drug Administration approved dasatinib (Sprycel™) in 2006 and nilotinib (Tisigna™) in 2008 for the treatment of patients with CML who are resistant or intolerant to imatinib. Whereas dasatinib targets both Src and Abl kinases, nilotinib does not inhibit Src kinases (Hazarika et€al. 2008). Other inhibitors still in clinical trials are: Bosutinib (SKI606) and INNO-406. All of these small molecule inhibitors are more potent than imatinib in blocking Â� Bcr-Abl tyrosine kinase activity and are active against a wide variety of Bcr-Abl kinase domain mutations associated with resistance to imatinib, except for the T315I mutation. The in€ vitro IC50 for imatinib against Bcr-Abl is 260€nM, whereas that for dasatinib it is 0.8€nM, for nilotinib 13€nM. In addition, dasatinib and Â�bosutinib inhibit several members of the Src family of kinases, and INNO-406 inhibits Lyn. Lyn has been involved in mediating Bcr-Abl salvage pathways or contributing to B-cell ALL. The role of Src kinases in CML is otherwise unknown.
148
S.J. Corey and J. Cortes
Dastinib, nilotinib, and bosutinib have shown significant clinical activity in patients with CML who have failed prior therapy with imatinib (Talpaz et€al. 2006; Kantarjian et€al. 2006; Cortes et€al. 2007). The results achieved with those agents already in phase II trials are summarized in Table€ 2. Complete cytogenetic responses have been achieved in 49% of patients with dasatinib, 40% with nilotinib, and 32% with bosutinib, although the follow-up time is different for each study, which may affect the overall response rate as responses continue improving over time. Still, these studies demonstrate that responses can be achieved in a significant number of patients after failure of response to imatinib therapy. Responses occur in patients with a wide range of different mutations or without mutations, with the one common exception being patients with the T315I mutation. Albeit with short follow-up, responses have been mostly durable and have translated into a survival benefit compared to historical controls treated with other modalities of therapy after imatinib failure. These agents have become the preferred therapy for patients who fail imatinib therapy. A COG phase I study of dasatinib was recently completed (Aplenc et al. 2007). To date, the drug has been well tolerated at 75€mg orally daily. Generally, dasatinib therapy has similar toxicity profile to that of imatinib. Management of drug interactions, myelosuppression, and edema follows the same guidelines (see above). Since the T315I mutation is still insensitive to all these new inhibitors, there are several agents that are being developed to counter this mutation (Quintas-Cardama et€ al. 2007a). Some of them are tyrosine kinase inhibitors whose binding to the kinase domain is not affected by the bulky residue of isoleucine in T315I. Among them, MK-0457 (formerly known as VX-680) has provided the most clinical data. This agent inhibits Aurora and Jak2 kinases in addition to Bcr-Abl. VX-680 is highly potent in€ vitro against Bcr-Abl with or without the T315I, inhibiting the proliferation of cell lines with either isoform of the kinase. Early results of an ongoing Table€2â•… Response to second generation tyrosine kinase inhibitors in patients with CML in chronic phase after imatinib failure Percentage Response Dasatinib Nilotinib Bosutinib Response â•… CHR 91 74 84 â•… MCyR 59 56 42 â•… CCyR 49 40 32 Toxicity (grade 3–4) â•… Thrombocytopenia 22–49 33 9 â•… Neutropenia 34–49 31 6 â•… Pleural effusion ~6 <1 <1 â•… Ý transaminases 2 1–4 <5 â•… RASh 0.5 2 6 â•… Ý lipase 11 15 <5 â•… Ý bilirubin 0 9 <5 Median FU (mo) 15 12 3
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
149
trial have shown clinical efficacy in patients with CML who failed imatinib and carry the T315I. Other similar multi-kinase inhibitors are being investigated both preclinically and in the clinic. Other agents, such as homoharringtonine (HHT), decrease protein synthesis with some selectivity for Bcr-Abl in CML cells. HHT is being investigated in patients who failed at least two prior tyrosine kinase inhibitors or who have the T315I mutation with some early suggestion of favorable clinical activity (Quintas-Cardama et€al. 2007b).
Indications for Transplantation Because of the lower incidence of treatment-related mortality in pediatric bone marrow transplantation, allogeneic stem cell transplants had been widely preferred for children with an HLA-matched sibling donor (Pulsipher 2004). According to the European BMT study on childhood CML, those who received sibling-matched transplants had a 76% 5 year survival for pediatric patients in CP and 46% for AP/BP (Cwynarski et€al. 2003). Much of the survival rested on prevention of graftversus-host disease. Those with grade 0 to 1 had 91% survival, but survival dropped to 61% for those with grades 2 to 4. Children who received a matched, unrelated donor transplant during CP had 65% 5 year survival. Transplant recipients have fared less well in other studies (Unal et€al. 2007; Creutzig et€al. 1996). Regardless of the availability of a sibling donor, there remains significant morbidity associated with severe immunosuppression, short- and long-term complications associated with graft-versus-host disease, and costs of frequent clinic visits, laboratory monitoring for transplant-associated complications, and hospitalization. Considering the 5-year probability of survival with imatinib is currently 90% (at least in adults), one should no longer recommend transplantation for pediatric patients with CML in CP as first line of therapy since the 5-year probability of survival is inferior. This is particularly true for patients receiving match unrelated donor transplants. Instead, evaluation for transplantation should be considered if (1) there is failure to reach response landmarks, (2) identification of the T315I mutation, and (3) emergence of a blast crisis. Patients who fail imatinib should be offered a trial with a second generation tyrosine kinase inhibitor (Guilhot et€al. 2007). Those pediatric patients with an HLA-matched sibling donor should be offered the option of transplant after salvage tyrosine kinase inhibitor fails. Adequate monitoring is critical to ensure the best outcome, as patients who are not responding optimally could be considered for alternative therapies. Landmark responses consist of cytogenetic response by 6 months, major cytogenetic responses by 12 months, and complete cytogenetic response by 18 months. Because there is no approved inhibitor active against T315I mutation, identification of this mutation would prompt the search for a stem cell transplant donor, either related or unrelated. Until a donor is identified, recruitment into clinical trials of investigational agents, for example, VX-680, that block T315I kinase activity should be encouraged. There is very limited experienced of the outcome of patients with T315I receiving an allogeneic transplant.
150
S.J. Corey and J. Cortes
Future Therapies The clinical success of imatinib and related tyrosine kinase inhibitors has led to the emergence of a number of critical biological and therapeutic questions. Must a CML stem cell be eradicated to achieve cure? How can one attack the quiescent CML stem cell? Because of their greater potency, are second generation Bcr-Abl kinase inhibitors better than imatinib for patients with untreated CML? How long should therapy be continued with either imatinib or second generation Bcr-Abl kinase inhibitor? Which clinical parameters dictate when to switch from one BcrAbl kinase inhibitor to another or undergo a stem cell transplant? What therapies can be developed to target the CML stem cell? How does one best manage relapse after transplantation? Continued research will give us the answer to these questions. If the recent past is predictive of the future, we might be able to get the answers to these important questions from our clinical specialists in adult leukemia.
References Melo JV, Barnes DJ. Chronic myeloid leukaemia as a model of disease evolution in human cancer. Nat Rev Cancer 2007;7:441–53. Sherbenou DW, Druker BJ. Applying the discovery of the Philadelphia chromosome. J Clin Invest 2007;117:2067–74. Nowell PC, Hungerford DA. A minute chromosome in human granulocytic leukemia. Science 1960;132:1497–501. Rowley JD. Letter: A new consistent chromosomal abnormality in chronic myelogenous leukaemia identified by quinacrine fluorescence and Giemsa staining. Nature 1973;243:290–3 Ben-Neriah Y, Daley GQ, Mes-Masson AM, Witte ON, Baltimore D. The chronic myelogenous leukemia-specific P210 protein is the product of the Bcr/Abl hybrid gene. Science 1986;233:212–4. Shtivelman E, Lifshitz B, Gale RP, Canaani E. Fused transcript of Abl and Bcr genes in chronic myelogenous leukaemia. Nature 1985;315:550–4. Chan LC, Karhi KK, Rayter SI, et€al. A novel Abl protein expressed in Philadelphia chromosome positive acute lymphoblastic leukaemia. Nature 1987;325:635–7. Diekmann D, Brill S, Garrett MD, et€ al. Bcr encodes a GTPase-activating protein for p21rac. Nature 1991;351:400–2. Maru Y, Witte ON. The Bcr gene encodes a novel serine/threonine kinase activity within a single exon. Cell 1991;67:459–68. Abelson HT, Rabstein LS. Lymphosarcoma: virus-induced thymic-independent disease in mice. Cancer Res 1970;30:2213–22. Wang JY, Ledley F, Goff S, Lee R, Groner Y, Baltimore D. The mouse c-Abl locus: molecular cloning and characterization. Cell 1984;36:349–56. Groffen J, Heisterkamp N, Reynolds FH, Jr., Stephenson JR. Homology between phosphotyrosine acceptor site of human c-Abl and viral oncogene products. Nature 1983;304:167–9. Daley GQ, Van Etten RA, Baltimore D. Induction of chronic myelogenous leukemia in mice by the P210Bcr/Abl gene of the Philadelphia chromosome. Science 1990;247:824–30. Lydon NB, Druker BJ. Lessons learned from the development of imatinib. Leuk Res 2004;28 Suppl 1:S29–38. Buchdunger E, Zimmermann J, Mett H, et€al. Inhibition of the Abl protein-tyrosine kinase in€vitro and in€vivo by a 2-phenylaminopyrimidine derivative. Cancer Res 1996;56:100–4.
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
151
Druker BJ, Tamura S, Buchdunger E, et€ al. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat Med 1996;2:561–6. le Coutre P, Mologni L, Cleris L, et€al. In vivo eradication of human Bcr/Abl-positive leukemia cells with an Abl kinase inhibitor. J Natl Cancer Inst 1999;91:163–8. Druker BJ, Talpaz M, Resta DJ, et€al. Efficacy and safety of a specific inhibitor of the Bcr-Abl tyrosine kinase in chronic myeloid leukemia. N Engl J Med 2001;344:1031–7. Shah NP, Skaggs BJ, Branford S, et€al. Sequential Abl kinase inhibitor therapy selects for compound drug-resistant Bcr-Abl mutations with altered oncogenic potency. J Clin Invest 2007;117:2562–9. Pendergast AM, Quilliam LA, Cripe LD, et€al. Bcr-Abl-induced oncogenesis is mediated by direct interaction with the SH2 domain of the GRB-2 adaptor protein. Cell 1993;75:175–85. Sattler M, Mohi MG, Pride YB, et€al. Critical role for Gab2 in transformation by Bcr/Abl. Cancer Cell 2002;1:479–92. Skorski T, Bellacosa A, Nieborowska-Skorska M, et€al. Transformation of hematopoietic cells by Bcr/Abl requires activation of a PI-3k/Akt-dependent pathway. EMBO J 1997;16:6151–61. Danhauser-Riedl S, Warmuth M, Druker BJ, Emmerich B, Hallek M. Activation of Src kinases p53/56lyn and p59hck by p210Bcr/Abl in myeloid cells. Cancer Res 1996;56:3589–96. Carlesso N, Frank DA, Griffin JD. Tyrosyl phosphorylation and DNA binding activity of signal transducers and activators of transcription (STAT) proteins in hematopoietic cell lines transformed by Bcr/Abl. J Exp Med 1996;183:811–20. Ilaria RL, Jr., Van Etten RA. P210 and P190(Bcr/Abl) induce the tyrosine phosphorylation and DNA binding activity of multiple specific STAT family members. J Biol Chem 1996;271:31704–10. Salgia R, Li JL, Ewaniuk DS, et€al. Bcr/Abl induces multiple abnormalities of cytoskeletal function. J Clin Invest 1997;100:46–57. Skorski T. Bcr/Abl regulates response to DNA damage: the role in resistance to genotoxic treatment and in genomic instability. Oncogene 2002;21:8591–604. Schuster C, Forster K, Dierks H, et€al. The effects of Bcr-Abl on C/EBP transcription-factor regulation and neutrophilic differentiation are reversed by the Abl kinase inhibitor imatinib mesylate. Blood 2003;101:655–63. Linet MS, Ries LA, Smith MA, Tarone RE, Devesa SS. Cancer surveillance series: recent trends in childhood cancer incidence and mortality in the United States. J Natl Cancer Inst 1999;91:1051–8. Millot F, Traore P, Guilhot J, et€al. Clinical and biological features at diagnosis in 40 children with chronic myeloid leukemia. Pediatrics 2005;116:140–3. Cortes JE, Talpaz M, O’Brien S, et€al. Staging of chronic myeloid leukemia in the imatinib era: an evaluation of the World Health Organization proposal. Cancer 2006;106:1306–15. Demetri GD, von Mehren M, Blanke CD, et€ al. Efficacy and safety of imatinib mesylate in advanced gastrointestinal stromal tumors. N Engl J Med 2002;347:472–80. O’Brien SG, Guilhot F, Larson RA, et€al. Imatinib compared with interferon and low-dose cytarabine for newly diagnosed chronic-phase chronic myeloid leukemia. N Engl J Med 2003;348:994–1004. Druker BJ, Guilhot F, O’Brien SG, et€al. Five-year follow-up of patients receiving imatinib for chronic myeloid leukemia. N Engl J Med 2006;355:2408–17. Baccarani M, Saglio G, Goldman J, et€ al. Evolving concepts in the management of chronic myeloid leukemia: recommendations from an expert panel on behalf of the European LeukemiaNet. Blood 2006;108:1809–20. Peng B, Lloyd P, Schran H. Clinical pharmacokinetics of imatinib. Clin Pharmacokinet 2005;44:879–94. Deininger MW, O’Brien SG, Ford JM, Druker BJ. Practical management of patients with chronic myeloid leukemia receiving imatinib. J Clin Oncol 2003;21:1637–47. Cohen MH, Moses ML, Pazdur R. Gleevec for the treatment of chronic myelogenous leukemia: U.S. Food and Drug Administration regulatory mechanisms, accelerated approval, and orphan drug status. Oncologist 2002;7:390–2.
152
S.J. Corey and J. Cortes
Esmaeli B, Prieto VG, Butler CE, et€al. Severe periorbital edema secondary to STI571 (Gleevec). Cancer 2002; 95:881–7. Kerkela R, Grazette L, Yacobi R, et€al. Cardiotoxicity of the cancer therapeutic agent imatinib mesylate. Nat Med 2006; 12:908–16. Verweij J, Casali PG, Kotasek D, et€al. Imatinib does not induce cardiac left ventricular failure in gastrointestinal stromal tumours patients: analysis of EORTC-ISG-AGITG study 62005. Eur J Cancer 2007;43:974–8. Atallah E, Kantarjian H, Cortes J. In reply to Cardiotoxicity of the cancer therapeutic agent imatinib mesylate. Nat Med 2007;13:14; author reply 5–6. Atallah E, Durand JB, Kantarjian H, Cortes J. Congestive heart failure is a rare event in patients receiving imatinib therapy. Blood 2007;110:1233–7. Bond M, Bernstein ML, Pappo A, et€al. A phase II study of imatinib mesylate in children with refractory or relapsed solid tumors: A Children’s Oncology Group study. Pediatr Blood Cancer 2007;50:254–258. Millot F, Guilhot J, Nelken B, et€ al. Results of a phase II trial testing interferon-alpha 2b and cytarabine in children and adolescents with chronic myelogenous leukemia. Pediatr Blood Cancer 2006;47:555–9. Millot F, Guilhot J, Nelken B, et€al. Imatinib mesylate is effective in children with chronic myelogenous leukemia in late chronic and advanced phase and in relapse after stem cell transplantation. Leukemia 2006;20:187–92. Champagne MA, Capdeville R, Krailo M, et€al. Imatinib mesylate (STI571) for treatment of children with Philadelphia chromosome-positive leukemia: results from a Children’s Oncology Group phase 1 study. Blood 2004;104:2655–60. Berman E, Nicolaides M, Maki RG, et€al. Altered bone and mineral metabolism in patients receiving imatinib mesylate. N Engl J Med 2006;354:2006–13. Peng B, Hayes M, Resta D, et€al. Pharmacokinetics and pharmacodynamics of imatinib in a phase I trial with chronic myeloid leukemia patients. J Clin Oncol 2004;22:935–42. Kovitz C, Kantarjian H, Garcia-Manero G, Abruzzo LV, Cortes J. Myelodysplastic syndromes and acute leukemia developing after imatinib mesylate therapy for chronic myeloid leukemia. Blood 2006;108:2811–3. Hughes T, Deininger M, Hochhaus A, et€ al. Monitoring CML patients responding to treatment with tyrosine kinase inhibitors: review and recommendations for harmonizing current methodology for detecting Bcr-Abl transcripts and kinase domain mutations and for expressing results. Blood 2006;108:28–37. Goldman JM. How I treat chronic myeloid leukemia in the imatinib era. Blood 2007;110:2242–2249. Hughes TP, Kaeda J, Branford S, et€al. Frequency of major molecular responses to imatinib or interferon alfa plus cytarabine in newly diagnosed chronic myeloid leukemia. N Engl J Med 2003;349:1423–32. Michor F, Hughes TP, Iwasa Y, et€ al. Dynamics of chronic myeloid leukaemia. Nature 2005;435:1267–70. Gorre ME, Mohammed M, Ellwood K, et€al. Clinical resistance to STI-571 cancer therapy caused by Bcr-Abl gene mutation or amplification. Science 2001;293:876–80. Shah NP, Nicoll JM, Nagar B, et€al. Multiple Bcr-Abl kinase domain mutations confer polyclonal resistance to the tyrosine kinase inhibitor imatinib (STI571) in chronic phase and blast crisis chronic myeloid leukemia. Cancer Cell 2002;2:117–25. Thomas J, Wang L, Clark RE, Pirmohamed M. Active transport of imatinib into and out of cells: implications for drug resistance. Blood 2004;104:3739–45. Dai Y, Rahmani M, Corey SJ, Dent P, Grant S. A Bcr/Abl-independent, Lyn-dependent form of imatinib mesylate (STI-571) resistance is associated with altered expression of Bcl-2. J Biol Chem 2004;279:34227–39. Donato NJ, Wu JY, Stapley J, et€al. Bcr-Abl independence and LYN kinase overexpression in chronic myelogenous leukemia cells selected for resistance to STI571. Blood 2003;101:690–8.
Chronic Myeloid Leukemia: Pathophysiology and Therapeutics
153
Elrick LJ, Jorgensen HG, Mountford JC, Holyoake TL. Punish the parent not the progeny. Blood 2005;105:1862-6. Hazarika M, Jiang X, Liu Q, et€al. Tasigna for chronic and accelerated phase Philadelphia chromosome–positive chronic myelogenous leukemia resistant to or intolerant of imatinib. Clin Cancer Res 2008;14:5325–31. Talpaz M, Shah NP, Kantarjian H, et€al. Dasatinib in imatinib-resistant Philadelphia chromosomepositive leukemias. N Engl J Med 2006;354:2531–41. Kantarjian H, Giles F, Wunderle L, et€ al. Nilotinib in imatinib-resistant CML and Philadelphia chromosome-positive ALL. N Engl J Med 2006;354:2542–51. Cortes J, Kantarjian H, Baccarani M, et€al. A phase I/II study of SKI-606, a dual inhibitor of Src and Abl kinases, in adult patients with Philadelphia chromosome positive (Ph+) chronic meylogenous leukemia (CML) or acute lymphoblastic leukemia relapsed, refractory or intolerant of imatinib. Blood 2007;108:54a, (abstract 168) Aplenc R, Strauss LC, Shusterman S, et al. Pediatric phase I trial and pharmacokinetic (PK) study of dasatinib: a report from the Children’s Oncology Group. J Clin Oncol 2007;25 (Suppl):14094. Quintas-Cardama A, Kantarjian H, Cortes J. Flying under the radar: the new wave of Bcr-Abl inhibitors. Nat Rev Drug Discov 2007;6:834–48. Quintas-Cardama A, Kantarjian H, Garcia-Manero G, et€ al. Phase I/II study of subcutaneous homoharringtonine in patients with chronic myeloid leukemia who have failed prior therapy. Cancer 2007;109:248–55 Pulsipher MA. Treatment of CML in pediatric patients: should imatinib mesylate (STI-571, Gleevec) or allogeneic hematopoietic cell transplant be front-line therapy? Pediatr Blood Cancer 2004;43:523–33. Cwynarski K, Roberts IA, Iacobelli S, et€al. Stem cell transplantation for chronic myeloid leukemia in children. Blood 2003;102:1224–31. Unal S, Fidan G, Tavil B, Cetin M, Cetinkaya DU. Allogeneic hematopoietic stem cell transplantation in pediatric chronic myelogenous leukemia cases: Hacettepe experience. Pediatr Transplant 2007;11:645–9. Creutzig U, Ritter J, Zimmermann M, Klingebiel T. [Prognosis of children with chronic myeloid leukemia: a retrospective analysis of 75 patients]. Klin Padiatr 1996;208:236–41. Guilhot F, Apperley J, Kim DW, et€al. Dasatinib induces significant hematologic and cytogenetic responses in patients with imatinib-resistant or -intolerant chronic myeloid leukemia in accelerated phase. Blood 2007;109:4143–50.
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes Lia Gore
Introduction New technologies have improved our understanding of hematopoietic and leukemia biology, and expanding knowledge of underlying molecular pathways and genomic aberrations is facilitating the development of new therapies for relapsed and refractory disease (Bhojwani et€al. 2006; Carroll et€al. 2006). Myelodysplasia and associated syndromes constitute a broad range of bone marrow dysfunction, characterized by variable clinical features; cytopenias due to bone marrow production and/or dysfunction and insufficiencies are typical. Clinical symptoms are frequently due to anemia, neutropenia, and/or thrombocytopenia, and transfusion dependence is common although not universal. The progression from myelodysplastic syndromes (MDS) to leukemia, particularly in children, is unequivocal, although the time course over which this occurs is highly variable. Risk factors for progression and lability of disease have been identified, and great effort is being made to identify the molecular triggers in MDS that are both causative and clinically relevant. This chapter presents a brief review of selected targets and new agents under investigation for MDS, the majority of which are in early phase clinical trials, and some of which remain in preclinical exploration. The novelty of such agents stems from their more selective nature in capitalizing on a characteristic that is unique or more predominant in cells with the aberrant MDS phenotype, compared to their normal counterparts. The current rate of drug development and the availability of new agents far outpace the realistic ability for any review to be comprehensive in nature. However, an attempt is made to focus on a number or promising molecularly and/or mechanistically targeted agents, with the hope of characterizing pathways that may be utilized more effectively to treat patients while minimizing short- and long-term toxicities. The therapy of choice for MDS in children has traditionally been
L. Gore () The University of Colorado Cancer Center and The Center for Cancer and Blood Disorders, The Children’s Hospital, University of Colorado Denver, Pediatrics Mail Stop 8302, P.O. Box€6511, Aurora, CO 80045, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_9, © Springer Science+Business Media, LLC 2010
155
156
L. Gore
hematopoietic stem cell transplantation as soon as feasible. As our understanding of the disease takes on a more molecular basis, analysis of the targets for therapy is increasingly likely to take on a more cross platform nature and the timing to transplantation may change.
Myelodysplastic Syndromes Incidence Reporting of MDS to the Surveillance, Epidemiology, and End Results (SEER) program in the US began in 2001; prior to this time, statistical information about the incidence and case rate were more difficult to extract from databases and relied on literature reviews and meta-analyses. MDS occurs in adults with a substantially higher frequency than in children, the incidence increases with each decade of life, and men have a higher incidence than women, particularly over the age of 60. Based on SEER data, there are estimated to be at least 10,000 cases of MDS per year in the US, with an incidence of up to 54.7 per 100,000 patients per year in the older age cohorts (http:// seer.cancer.gov/registries). However, there are recent suggestions that the incidence reported by SEER may be significantly underestimated (Rollison et€al. 2008). Based on an analysis of 1.7 million Medicare patients in the Medicare Standard Analytic file for 2003, an age-adjusted incidence of 181 per 100,000 patients, or approximately 76,000 patients in the Medicare system at the time of the analysis (Finn 2009) is postulated. SEER data for 2001 to 2003 in persons under the age of 20 indicate that there is an approximately equal distribution between males and females, and that the rate remains steady over the first two decades of life at 0 to 0.3 per 100,000 patients per year (Ma et€al. 2007). Overall, it is estimated that MDS represents less than 5% of hematologic malignancies in children. However, 15 to 20% of pediatric Acute Myelogenous Leukemias (AML) may have a preceding MDS (Locatelli et€al. 1995).
Classification Classification of MDS is sometimes separated from that of myelofibrosis or the myeloproliferative syndromes (Cervantes et€al. 2009) where the French–American– British (FAB) and World Health Organization (WHO) (Bennett et€al. 1982; Jaffe et€al. 2001) have developed stringent uniform criteria based on laboratory findings and clinical correlates. In addition, the International Prognosis Scoring System (IPSS) is frequently used to identify risk groups based on variables such as percent bone marrow blasts, karyotype, and number of cytopenias (Greenberg et€al. 1997; Kaushansky 1998). Patients are categorized as having low, intermediate-1, intermediate-2, or high risk disease based on IPSS score and prognostic information is available to
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
157
guide clinical practice and patient education on the basis of the scores. Classification of MDS in children has met with more challenges, largely because the FAB and WHO criteria do not account for some of the features unique to pediatric patients with these syndromes, as parallel diseases in adults do not exist. As a result, the pediatric WHO criteria were developed in 2003 to include myeloproliferative and myelodysplastic disease (Hasle et€al. 2003). The Category, Cytology, Cytogenetics (CCC) system uses descriptive terminology and etiology to group diseases, but does not include the myeloproliferative syndromes, including juvenile myelomonocytic leukemia (JMML) (Mandel et€al. 2002). Table€1 attempts to collate several of the diagnostic criteria categories from these different systems. One challenge has been to determine which scoring system is most appropriate to help determine both diagnostic and prognostic information for children with MDS. Lorand-Metze and colleagues evaluated the FAB and WHO classifications, together with IPSS score, and found that in 150 patients, ages 12 to 90, the younger patients had a relatively higher proportion of WHO “unclassified” disease, with cellular findings most similar to childhood MDS (Lorand-Metze et€ al. 2004). Similarly, Occhipinti and colleagues compared the CCC and pediatric WHO criteria in a retrospective chart review that outlines a number of advantages and disadvantages noted when applying each staging system to a series of pediatric cases. They concluded that the CCC and pediatric WHO criteria were superior in being able to classify and predict outcomes in pediatric patients with MDS, than the FAB and adult WHO criteria overall, however, felt that both the CCC and pediatric WHO had limitations in individual cases (Occhipinti et€ al. 2005). There remain discrepancies in several areas: (1) how to classify disorders with cytopenias other than refractory anemia; (2) whether 20 or 30% blasts should be the cut-off between MDS and AML, and the role of cytogenetics in diagnosis and prognostic systems; (3) the lack of prognostic information in the FAB classification system; (4) the broad use of the WHO “unclassifiable” category for unusual cases, and how to best accommodate those patients; and (5) that the pediatric IPSS system currently excludes patients with Down syndrome, other congenital marrow failure syndromes, or therapy-related disease, and provides no prognostic information for patients with JMML.
Etiology In children, MDS typically arises in one of three settings: Children who have genetic or metabolic syndromes, which predispose them to bone marrow dysfunction and failure, and eventual leukemic transformation; children with de€novo MDS who have no known predisposition or triggering exposure; and children who have had prior exposure to agents known to induce marrow dysplasia and/or aplasia. Regardless of the cause of the development of MDS, the overwhelming majority of children who develop myelodysplasia will eventually progress to leukemic transformation of their disease, which is typically resistant to therapeutic intervention.
Leukocytosis, thrombocytopenia frequent Leukocytosis, thrombocytopenia frequent
Refractory anemia with excess blasts (RAEB) Refractory anemia with excess blasts in transformation (RAEB-T)
BCR-ABL negative chronic myeloid leukemia (Ph-CML)
20–29% blasts
2–19% blasts
10–19% blasts ³20% basophils No BCR-ABL fusion or Ph chromosome Myelodysplastic syndromes (pediatric and adult WHO) Pediatric Refractory cytopenia (RC) – includes <2% blasts RARS, if meets adult criteria Anemia not always present
Myelodysplastic/myeloproliferative disease Monocytosis >1e9/L Juvenile myelomonocytic leukemia (JMML) <20% blasts No BCR-ABL fusion or Ph chromosome Monocytosis >1e9/L Chronic myelomonocytic leukemia <20% blasts (CMML) (*must be only in patients No BCR-ABL fusion or Ph chromosome with prior chemotherapy)
Down syndrome disease Transient abnormal myelopoiesis (TAM) Myeloid leukemia of down syndrome
<5% blasts Immature erythropoiesis Decreased megakaryocytes or micromegakaryocytes 5–19% blasts May or may not have Auer rods 20–29% blasts
<20% blasts No BCR-ABL fusion or Ph chromosome May or may not have monosomy 7 <20% blasts No BCR-ABL fusion or Ph chromosome Dyplasia in ³1 myeloid lineages or clonal cytogenetic abnormality detected 10–19% blasts No BCR-ABL fusion or Ph chromosome detected
³30% blasts Higher incidence of M7 morphology
Table€1â•… Combined adult and pediatric WHO classifications with IPSS prognostic scoring of myelodysplastic syndromes Clinical/peripheral blood Disease appearance Bone marrow appearance Myeloproliferative disease <10% blasts Transient myeloproliferative syndrome Usually normal maturation Cellularity usually increased and morphology ³1 myeloid cell lines increased Acute leukemia Dysplasia in ³1 myeloid lines ³30% blasts common ³20% blasts Cellularity usually increased, but may be decreased
IPSS 1.5–2
IPSS 0.5–1.5
IPSS 0-0.5
N/A
N/A
N/A
N/A N/A
N/A
N/A
IPSS prognostic scoring
Two or more cytopenias Few or no blasts No Auer rods <1e9/L monocytes Two or more cytopenias Few or no blasts No Auer rods <1e9/L monocytes Cytopenias <5% blasts No Auer rods <1e9/L monocytes Cytopenias <5% blasts Auer rods with or without <1e9/L monocytes Cytopenias No or rare blasts No Auer rods Anemia Normal or increased platelet count <5% blasts
Refractory cytopenia with multilineage dysplasia (RCMD)
N/A not applicable
MDS associated with isolated deletion of 5q
Myelodysplastic syndrome, unclassified (MDS-U)
Refractory anemia with excess blasts 2 (RAEB-2)
Refractory anemia with excess blasts (RAEB-1)
Refractory cytopenia with multilineage dysplasia and ringed sideroblasts (RCMD-RS)
Anemia No blasts
Anemia Few or no blasts
Refractory anemia with ringed sideroblasts (RARS)
Adult Refractory anemia (RA)
IPSS 0.5 One myeloid cell line dysplasia <5% blasts No Auer rods Normal or increased megakaryocytes with hypolobulated nuclei <5% blasts Isolated del(5q) cytogenetics No Auer rods
IPSS 0
IPSS 2.0
IPSS 1.0
IPSS 0.5
IPSS 0.5
IPSS 0
IPSS 0
Uni- or multilineage dysplasia 10–19% blasts Auer rods may or may not be present
<5% blasts <15% ringed sideroblasts Erythroid dysplasia only <5% blasts ³15% ringed sideroblasts Erythroid dysplasia only ³10% cells with dysplasia of two or more myeloid lines <5% blasts No Auer rods <15% ringed sideroblasts ³10% cells with dysplasia of two or more myeloid lines <5% blasts No Auer rods ³15% ringed sideroblasts Uni- or multilineage dysplasia 5–9% blasts No Auer rods
160
L. Gore
Predisposing genetic syndromes associated with the development of MDS include entities such as Fanconi anemia, Shwachman-Diamond syndrome, Diamond-Blackfan anemia, Bloom syndrome, Ataxia-telangiectasia, Kostmann syndrome, Nonan syndrome, neurofibromatosis type I, and Down syndrome (Alter 1996; Alter 2002; Homans et€ al. 1993; Creutzig et€ al. 1996; Rischewski et€ al. 2000). Other conditions associated with acquired predisposition to MDS include paroxysmal nocturnal hemoglobinuria, aplastic anemia, dysmegakaryocytopoiesis, selected congenital metabolic acidemias and acidurias, Griscelli syndrome, and Langerhans cell histiocytosis (Surico et€ al. 2000; Cetin et€ al. 1998; Ohara et€ al. 1997), among others. The development of MDS following chemotherapy or toxin exposure is typically thought to be less than 1 to 2% of leukemia cases overall. Inciting agents often implicated are the epipodophyllotoxins, especially etoposide (Winick et€al. 1993; Pui et€al. 1991), or alkylating agents such as cyclophosphamide, chlorambucil, melphalan, busulfan, and BCNU (carmustine) (Tew et€al. 2001; Smith et€al. 1996). Antimetabolite therapy may be implicated as well, as has been observed in patients who have altered thiopurine methyltransferase (TPMT) drug metabolism and elevated erythrocyte 6-thioguanine nucleotide levels (Bo et€al. 1999). In addition, exposure to toxins that are known to induce marrow aplasia, such as benzene and similar organic solvents, has been implicated. Patients with prior exposures as the purported etiology of MDS will often have rearrangements of the MLL gene or deletions of (or portions of) chromosomes 5 or 7 (Felix et€al. 1995; Privitera et€al. 1992; Yesilipek et€al. 1994; Maris et€al. 1997; Harrison et€al. 1995; Chantrain et€ al. 2000). Therapy-related MDS/AML (t-MDS/t-AML) is noted to have a particularly poor overall survival. Recently, Chakraborty and colleagues have reported that accelerated telomere loss precedes and is associated with t-MDS/AML following autologous stem cell transplantation for lymphomas (Chakraborty et€al. 2009). The authors suggest that this may reflect clonal expansion or altered telomere regulation in the preleukemic condition. As such, therapies directed toward normalization of telomere structure and function would be a potential intervention to reverse this process. Genetic alterations in MDS are distinctive, and involve a wide array of cytogenetic structural and gene dysregulation abnormalities (Maris et€al. 1997; Taketani et€al. 2003; Tartaglia et€al. 2003; Lo Coco et€al. 1997; Perkins et€al. 1997; Block et€al. 2002). As mentioned above, the most commonly identified abnormalities are of chromosomes 5 and 7, with monosomy and partial deletions being most frequent. Selected other genetic abnormalities, including mutations in critical target genes such as AML1/RUNX1, JAK2, PTPN11, and other members of the RAS/ Raf/Mek pathway are presented in Table€2. MDS is essentially ineffective hematopoiesis, typically with high proliferative rates and increased apoptosis. A variety of rare hematopoietic myelodysplastic and myeloproliferative disorders have been identified, including the chronic myelomonocytic leukemias, atypical chronic myeloid leukemia, and JMML; only JMML is found with any frequency among the already infrequent pediatric MDS disorders. The 4th edition of the World Health Organization classification of
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes 161
Table€2╅ Selected genetic abnormalities identified in pediatric myelodysplastic syndrome patients Cytogenetic abnormality noted Gene(s) associated Clinical subgroup correlates Reference 1q Morerio et€al. (2006), Mogul et€al. (1997) De novo and t-MDS t(1;3)(p36;q21) MEL1-PRDM16 t-MDS, t-AML Block et€al. (2002), Lahortiga et€al. (2004) t(1;7)(p11;p11) MDS with preceeding Horsman et€al. (1988) marrow hypoplasia t(1;17)(p36;q21) t-MDS, t-AML Block et€al. (2002) t(1;22) Perkins et€al. (1997) t(2;8)(p23;p11.2) MOZ t-MDS Imamura et€al. (2003) t(3;5)(q21;q31) t-MDS, t-AML Block et€al. (2002) 3q21q26 t-MDS, t-AML Stark et€al. (1994), Block et€al. (2002), Perkins et€al. (1997) t(3;21)(q26;q22) MDS1-EVI1 t-MDS, t-AML, CML Rubin et€al. (1990), Lo Coco et€al. (1997), Lahortiga et€al. (2004) del(4q)╛+╛t(9;22)(q34;q11) KIT Dalla Torre et€al. (2002), Galili et€al. (2007) t(5;12)(q31;p12-p13) KIT MDS with eosinophilia Galili et€al. (2007), Pellier et€al. (1996) Monosomy 5 or 5qPerkins et€al. (1997) t(6;9)(p23;q34) t-MDS, t-AML Block et€al. (2002), Perkins et€al. (1997) Monosomy 7 or 7qPerkins et€al. (1997) t(8;14)(q24;q32) t-MDS, t-AML Block et€al. (2002) t(8;16)(p11;p13) MOZ-CBP t-MDS, t-AML Block et€al. (2002), Imamura et€al. (2003) t(8;21)(q22;q22) AML1/RUNX1-ETO Lo Coco et€al. (1997), Perkins et€al. (1997) Trisomy 8 Perkins et€al. (1997) Pentasomy 8 Wong and Wu (2000) inv(8)(p11;q13) AML t(9;22)(q34;q11) t-MDS, t-AML Block et€al. (2002), Perkins et€al. (1997) Block et€al. (2002), Taki et€al. (1997), Perkins et€al. (1997) MLL╅ MLL-CBP Epipodophyllotoxin 11q23 abnormalities exposure ╅ t(9;11) (q13;q13) ╅ t(10;11)(p13;q13) ╅ t(11;16)(q23;p13) ╅ t(11;19)(q13;q13) (continued)
Gene(s) associated
Clinical subgroup correlates
Reference
Trisomy 11 Perkins et€al. (1997) 11p15 t-MDS, t-AML Stark et€al. (1994), Block et€al. (2002) 12p13 and 12q13 ETV6 t-MDS, t-AML Block et€al. (2002), Perkins et€al. (1997), Tosi et€al. (1998) 13q12 Adjacent to BRCA2 MDS Coignet et€al. (1999) Inversion 16 Perkins et€al. (1997) 20q- and 20q11.2-12 NNAT MDS, AML Perkins et€al. (1997), Kuerbitz et€al. (2002) 21q22 AML1 Lo Coco et€al. (1997), Perkins et€al. (1997), Taketani et€al. (2003) Trisomy 21 Perkins et€al. (1997) Selected gene mutations relevant in pediatric myelodysplastic syndromes Gene Fusion partners Clinical correlates Reference ETV6 Tosi et€al. (1998) FLT3 Co-duplication with MLL AML Jamal et€al. (2001) JAK2 AML, JMML Tartaglia et€al. (2003), Hellstrom-Lindberg and Cazzola (2008) Aberrant histone Imamura et€al. (2003) MOZ CBP acetylation regulation TIF2 p300 Side et€al. (1998) NF1 JMML not associated with clinical neurofibromatosis NNAT MDS, AML Kuerbitz et€al. (2002) NPM1 Zhang et€al. (2007) PPARgamma loss Silveira et€al. (2009) RAS May be associated with JMML Mahgoub et€al. (1998); Side et€al. (1998) MLL translocations P53, TP53 deletion Gafanovich et€al. (1999), Silveira et€al. (2009) PTPN11 JMML Kratz et€al. (2005), Bentires-Alj et€al. (2004), Loh et€al. (2004) WT1 Rosenfeld et€al. (2003) AML acute myeloid leukemia; JMML juvenile myelomonocytic leukemia; MDS myelodysplastic syndrome; t-AML therapy-related acute myeloid leukemia; t-MDS therapy-related myelodysplastic syndrome
Table€2╅ (continued) Cytogenetic abnormality noted
162 L. Gore
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
163
hematopoietic tumors, adopted in 2009, will replace the use of the term “myeloproliferative disorder” with “myeloproliferative neoplasm” to better describe the myriad molecular abnormalities associated with the malignant nature of these diseases. Regardless of the cause of the development of MDS, there is general agreement that the treatment in children should be the same. One particular population has separated itself somewhat: adults with 5q31 deletions appear to have a better outcome when treated with the immunomodulatory agent lenalidomide than patients with other cytogenetic findings. This will be discussed later in this chapter, and the significance in pediatrics is not yet known due to the rarity of 5q31 in pediatric patients, and the current lack of any clinical studies evaluating the use of lenalidomide in children with MDS.
The Relationship Between Acute Myeloid Leukemia and MDS The AML are addressed elsewhere in this text. However, AML relates significantly to this discussion. If not treated, MDS will invariably progress to a leukemic state, most often AML. AML comprises only 15 to 20% of childhood leukemia cases, but accounts for one-third of childhood leukemia deaths (Woods et€al. 2001). Current therapies are intensive and carry with them a significant risk for acute toxicity and late effects. It is doubtful that significant gains will be achieved by continued intensification of conventional chemotherapy regimens without paying a significant price of morbidity to patients. In addition, patients whose AML arises from a preceding MDS typically have more aggressive and refractory disease than those patients in whom AML arises de€novo. This makes the development of targeted therapies with greater efficacy and less toxicity of paramount importance. Fortunately, much has been learned about the biology and genetics of MDS and AML, leading to the identification of a number of candidate therapeutic targets. One of the great success stories of biologically-targeted therapy is the use of all-trans retinoic acid (ATRA) in the treatment of acute promyelocytic leukemia (APL). ATRA was initially integrated into therapeutic strategies based on the observations of clinical activity prior to the identification of a specific molecular mechanism of action. Based on clinical observations, investigators in China observed complete remission in patients with both newly diagnosed and relapsed acute promyelocytic leukemia (APML) (Huang et€ al. 1988). ATRA is now thought to act by several mechanisms, one of which is to drive leukemic blasts to terminal differentiation, which in effect, targets the class II mutation patterns. ATRA has become part of standard therapy for APML, and since its inclusion in treatment regimens, the cure rates for APML have risen substantially. The use of ATRA in MDS is now evolving to single agent and combination studies, most typically added to a demethylating agent or a histone deacetylase inhibitor, which will be discussed below.
164
L. Gore
Chromatin-Based Transcriptional Therapy Cancer cells contain complex genetic and epigenetic alterations leading to aberrant gene expression that result in the malignant phenotype. These changes characteristically involve global hypomethylation, promoter hypermethylation, and histone modification, leading to altered chromatin structure and function, as well as inappropriate gene silencing (Singal and Ginder 1999; Esteller et€ al. 2001; Das and Singal 2004; Vidal et€ al. 2007). As such, inhibition of DNA methyltransferase (DNMT) and histone deacetylation (HDAC) are the two most common therapeutic approaches to transcriptional therapy to date, and each will be discussed below more specifically. Other targets include the histone acetyltransferases, histone methyltransferases, and histone demethylases. Epigenetic and chromatin changes occur as a result of recruitment of multiprotein containing complexes to DNA (Cameron et€ al. 1999; Rountree et€ al. 2001; Rountree et€al. 2000), and the interaction between DNA methylation and histone acetylation are thought to play a major role in both normal and oncogenic development (Jones and Wolffe 1999). Zhou and colleagues have shown that promoter DNA methylation changes are important in MDS and are associated with changes in gene expression patterns relevant to cancer development and progression, including among others, alterations in TNF superfamily, HOX B4, EVI1, and CEPBB genes (Zhou et€al. 2008). In pediatric MDS, two genes, CALCA and CDKN2B, are reported to be frequently methylated (Vidal et€al. 2007). This suggests that pediatric MDS displays aberrant methylation, and that methylation targeted therapies may be similarly appropriate for the treatment, as they are currently used in the adult population. The neuronatin (NNAT) gene on chromosome 20q11.2-q12 is associated with the loss of heterozygosity in pediatric MDS and AML. Differential methylation patterns in CpG islands of exon 1 of NNAT have been demonstrated (Kuerbitz et€ al. 2002), and the treatment with 5-aza-2¢-deoxycytidine can reactivate NNAT expression and re-establish normal methylation patterns in€ vitro. This may be a promising approach to patients with MDS who harbor abnormalities of NNAT. It is also of note that several AML-associated fusion proteins containing the leukemia zinc finger gene MOZ, including MOZ-CBP, MOZ-TIF2, and MOZ-p300, are thought to act by aberrant regulation of histone acetylation, thus suggesting a possible role for histone modification in the therapy of these leukemias (Imamura et€al. 2003). Table€3 summarizes the more common agents being explored in MDS.
DNA Methyltransferase (DNMT) Inhibitors DNA methylation in promoter regions of genes is a control mechanism for gene transcription. When cytosine methylation occurs, promoters are suppressed, as is gene transcription. DNA hypermethylation in promoter regions and consequent inactivation of tumor suppressor genes, including p15INK4B and p21WAF1/CIP1, is thought to play a role in the pathogenesis of MDS and AML (Vidal et€al. 2007;
Immunomodulatory agents
Hydroxamic acid derivative Cyclic tetrapeptide Synthetic benzamide derivative Benzamide
Butyric acid pro-drug Polar hybrid hydroxamic acid derivatives
Short chain fatty acids
Histone deacetylase inhibitors Target
Thalidomide Lenalidomide (CC-5013, Revlimid®) Pomalidomide CC-4047)
Valproic acid Sodium phenylbutyrate AN-9 (Pivanex) Hexamethylene bisacetamide Vorinostat (SAHA, Zolinza®) Belinostat (PXD101) Panobinostat (LBH589) Trichostatin A Depsipeptide (FK228, Romidepsin™) Entinostat (MS275, SNDX 275) CI-994 MGCD0103
Agent
Table€3â•… Methylation and histone-targeted agents of interest in myelodysplastic syndrome Target Agent Methylation targeted agents Hypomethylated DNA 5-azacitidine (Vidaza®) Methylated DNA, Hypomethylated DNA 5-aza-2¢-deoxycitidine (Decitabine®) Binds 3¢ untranslated region DNA MG-98 oligonucleotide methyltransferase mRNA; growth inhibition Riboside pro-drug; cytidine deaminase inhibitor Zebularine
uM uM nM uM nM uM nM
I, II, IV I, II, IV I, II, IV I, II, IV I I I
I, IIa I, IIa
Range of reported in€vitro inhibition mM mM uM
HDAC inhibition specificity
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes 165
166
L. Gore
Uchida et€al. 1997; Jones and Laird 1999; Jiang et€al. 2009). DNA methylation inhibitors are incorporated as cytosine bases, irreversibly binding DNA methyltransferase I (DNMT1), and thus preventing DNA methylation, which ultimately results in altered gene transcription. Several DNA methylation inhibitors, including 5-azacytidine (Vidaza®), decitabine (5-aza-2-deoxycitadine, Dacogen®), and zebularine, are being studied in adults with MDS and acute leukemia. Combinations of the DNA methylase and histone deacetylase inhibitors (HDACi) are being extensively studied in clinical trials in adults, and results have been found to be superior to the prior approach of best supportive care methodologies, but not necessarily superior to single agent treatment (Garcia-Manero et€al. 2006; Scott et€al. 2007; Shaker et€al. 2003; Silverman et€al. 2006). Zebularine inhibits cytosine deaminase and DNA cytosine methyltransferase, but in€vivo did not demonstrate hypomethylating activity when compared to azacitidine or decitabine (Flotho et€al. 2009). Use of any of these agents has not advanced significantly in pediatric MDS to date.
Histone Deacetylases (HDACs) and HDAC Inhibitors (HDACi) Histone deacetylases catalyze the removal of acetyl groups from lysine residues on substrate proteins, thus regulating activity and/or function of a variety of cellular components. Core histones and non-histone proteins are substrates for HDACs. Deacetylation of the core histones is associated with regions of chromatin that are transcriptionally silenced or repressed. Acetylation of the non-histone substrate proteins regulates the activity of a variety of other effector proteins. HDACs show increased expression over normal progenitors are mutated or deleted in a variety of human malignancies. Proteins implicated in tumorigenesis, such as p53, Bcl6, Hsp90, Stat3, and E2F, as well as a variety of fusion proteins interact with HDACs to mediate inappropriate repression of target genes. HDAC inhibition likely leads to changes in transcription of genes important in cell cycle and apoptosis regulation, including p21WAF1/CIP1, p53, RB, bcl2, bcl6, bclxl, and mcl-1. Several HDAC inhibitors, including valproic acid, suberanilohydroxamic acid (SAHA, vorinostat, ZolinzaTM), entinostat (MS-275, now SNDX-275), and depsipeptide (FK228, RomidepsinTM), are under wide-spread investigation in MDS and the hematologic malignancies, both alone and in combination (Kuendgen et€al. 2006; Kuendgen et€al. 2005; Rosato et€al. 2003); their evaluation has been more limited in pediatric applications. A phase I study of depsipeptide in pediatric patients with refractory solid tumors established the maximum tolerated dose to be 13€ mg/m2 when administered as a weekly 4-h infusion for 3 weeks on a 28-day cycle (Fouladi et€al. 2006). To date, no studies in pediatric patients with hematological malignancies or MDS have been conducted. A phase I study of vorinostat in pediatric patients has been completed, but is not yet published. A combination study of vorinostat and etoposide in children with refractory cancers, including hematologic malignancies is under way.
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
167
Maturation-Directed Therapies Treatment with ATRA leads to dissociation of the HDAC complexes and eventual transcriptional activation of the repressed genes, leading to resumption of cellular differentiation. This has been a particularly effective treatment for APML, and early phase clinical trials have incorporated ATRA into combination therapy for adults with MDS (Murgo 2001). Pediatric use of ATRA is commonplace in APML; however, the knowledge about dosing and tolerability gained in the leukemia experience has generally not been incorporated into MDS treatment.
Immunomodulatory Drugs (IMiDs) and Anti-angiogenic Agents Inhibition of angiogenesis leading to growth arrest of MDS cells has been evaluated clinically with several agents. Thalidomide, while shown to induce major erythroid lineage response in adults with MDS, has also been associated with significant neurological toxicities as well as the known teratogenic effects that have limited its use. IMiDs, immunomodulatory agents that are structurally and functionally similar to thalidomide, are thought to act by anti-inflammatory, antiproliferative, and proapoptotic mechanisms, and are under intense investigation. Lenalidomide (Revlimid®, CC-5013) is a less neurotoxic second-generation immunomodulatory agent that is thought to induce direct cytotoxic effect on the del(5q31) clone, and significant clinical response in patients harboring 5q31 deletions have been noted (Sekeres and List 2005; Sekeres and List 2006; Sekeres et€al. 2008; Chang et€al. 2006; List et€al. 2006). Lenalidomide received FDA approval for use in transfusion-dependent adult patients with 5q-MDS after a pivotal study demonstrated significant improvement in red blood cell transfusion-independence (Raza et€ al. 2008), and current National Comprehensive Cancer Network Clinical Practice Guidelines recommend the use of lenalidomide for lower-risk MDS patients with del(5q) and transfusion-dependent anemia. Lenalidomide is also approved for use in combination with dexamethasone in adults with multiple myeloma who have received at least one prior therapy, with a significant improvement in complete and partial response rates, and in time for progression of disease when compared to dexamethasone alone. Other immunomodulatory affects of this class of agents includes anti-inflammatory effects and decrease in TNF-alpha, co-stimulatory affects of anti-CD3 stimulated T-cells (including enhanced T-cell proliferation), augmentation of NK cell cytotoxic activity against tumor cells, and inhibition of T-regulatory cell proliferation and suppressor function. Pomalidomide (CC-4047) is a second-generation IMiD in early development. While there are phase II data published on the use of thalidomide in a variety of pediatric diseases, including Crohn’s disease, systemic lupus erythematosis, Behcet’s disease, and a variety of pediatric malignancies (Gilheeney et€al. 2007; Kieran et€al. 2005; Lopez-Aguilar et€al. 2008), there are no published data on the pharmacokinetics or tolerability of lenalidomide or pomalidomide in pediatric patients to date.
168
L. Gore
Other Targeted Therapies A subset of adults with MDS, AML, CML, and ALL over-express the WT1 gene, and current immunotherapy-based approaches, including WT1-specific cytotoxic lymphocyte generation, WT1 anti-sense, and a WT1 vaccine, are being explored (Rosenfeld et€al. 2003). Alterations of WT1 in pediatric MDS have not been described. Small molecule tyrosine kinase inhibitors (TKIs) have revolutionized the therapy of CML, Philadelphia positive ALL, renal cell carcinoma, pancreatic, colon, and lung cancers among others. Boehrer and colleagues have demonstrated that the use of erlotinib, a small molecule inhibitor of the epidermal growth factor receptor (EGFR), can induce antineoplastic activity in MDS and AML cells via off-target inhibition of JAK2 leading to apoptosis, cell-cycle arrest, and differentiation (Boehrer et€al. 2008). This illustrates the possibility of broadened use of TKIs in MDS and other hematologic malignancies. Another TKI, imatinib mesylate, has shown activity in patients with MDS who have t(5;12) and deletion of 4q clones (Galili et€al. 2007), presumably due to its inhibitory activity against c-KIT and the known involvement of c-KIT in these patients.
Combination Therapies Significant numbers of clinical trials are focusing on combining one or more of the above classes of agents. The major efforts were initially to couple a methylationtargeted agent with a histone deacetylase inhibitor. There are now completed studies with a variety of combinations of demethylating agents (azacitidine, decitabine, and zebularine) with HDAC inhibitors, including valproic acid, vorinostat, entinostat, LBH589, PXD101, depsipeptide, CI-994, MG-98, and MGCD0103, in adults. Other combinations focus on a differentiating agent plus another agent, often a demethylating and/or HDAC inhibitor, such as decitabine and valproic acid, azacitidine and entinostat, azacitidine and ATRA, valproic acid and ATRA, ATRA and arsenic trioxide, or a triplet of ATRA, arsenic trioxide and valproic acid, or azacitidine, valproic acid and ATRA (Soriano et€al. 2007). A combination study of vorinostat and cis-retinoic acid was performed in pediatric patients with selected solid tumors. The dose limiting toxicities noted when vorinostat was dosed at 180€mg/m2 daily were neutropenia, thrombocytopenia, anorexia, and hypophosphatemia. Grade 3 and 4 transaminitis, leucopenia, and lymphopenia were also noted. There were no objective response, but three patients, with pineoblastoma, neuroblastoma, and medulloblastoma, had stable disease for 5, 6, and 7 cycles, respectively. The patient with medulloblastoma had to be removed from the treatment due to persistent and profound thrombocytopenia. The MTD for the combination of vorinostat and 13 cRA was vorinostat 180€mg/m2 per day 4â•›×â•›per week, with 13 cRA given at 80€mg/m2/dose bid days 1 to 14 (Fouladi et€al. 2008). One study of azacitidine in combination with valproic acid allowed children over the age of two to enroll, and there are anecdotal reports of the use of this combination in children with a variety of refractory cancers, although no formal studies have been published to date.
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
169
The epigenetic regulation of normal DNA structure and function affects scaffolding proteins during DNA assembly. The control of complex protein folding is known to be dysregulated in a variety of malignancies. The proteasome inhibitors, which affect ubiquitin-mediated protein degradation, and the heat shock protein (HSP) inhibitors, which alter chaperone protein function, have been evaluated in clinical trials. A phase I study of the first-in-class proteasome inhibitor bortezomib has been completed in pediatrics, and it is now being incorporated into combination therapy in the pediatric acute leukemias. HSP90, HSP70, and HSP27 show increased expression in ALL, AML, and MDS cells compared to controls. In addition, the combination of etoposide and 17-AAG, an HSP90 inhibitor, shows synergistic cell killing against FLT3+ leukemias in€vitro (Yao et€al. 2007). 17-AAG has been evaluated in two pediatric phase I studies, and demonstrated reasonable safety and tolerability but no substantial antitumor activity (Weigel et€al. 2007; Bagatell et€al. 2007). Three pediatric patients studied by Bagatell et€al. with refractory AML/MDS (2) and ALL (1) were treated with single agent 17-AAG in a separate study arm and demonstrated stability of disease for at least 8 weeks. These data suggest the possibility of combining an HSP90 inhibitor together with etoposide and a DNMT-targeted agent in pediatric patients with MDS to affect methylation and chaperone protein interactions in a complementary fashion. In pediatrics, there has been significant lag in developing combination studies for MDS outside of the phase III setting when MDS was formerly included in AML trials, in part due to the limited numbers of patients available for such enrollment, and with relatively little interest in such a small patient population on the part of pharmaceutical manufacturers. Contemporary phase III AML studies do not include MDS patients, and there are no current phase III trials evaluating de€novo MDS in children. Another problem in the process of analyzing these agents in pediatrics has been the lack of biologic correlative studies performed within the context of a clinical trial to show target modulation. This is critical to better evaluate the array of new and older agents, both alone and in combination.
Stem Cell Transplantation for Myelodysplastic Syndromes To date, the only truly curative therapy for MDS has been allogeneic stem cell transplantation. A variety of conditioning regimens have been evaluated in adults, including myeloablative, non-myeloablative, or reduced-intensity conditioning. This is in part due to the fact that most adults going to SCT have more co-morbidities than pediatric patients with the same disease. These concomitant medical conditions and the associated cumulative organ dysfunction compromise the chance of successful transplantation outcomes. In addition, pediatric patients who develop MDS have typically been taken to transplant relatively soon after their initial diagnosis without being treated with chemotherapy or MDS directed therapies, such as demethylating agents. In adults, treatment with demethylating agents is more common, since those with MDS are often elderly and not good candidates for stem cell transplantation. Many adult transplant programs now incorporate the IPSS and WHO classification schemes into their decision-making regarding timing of transplant, and most centers
170
L. Gore
restrict transplant for MDS to patients age 50 to 55 and under due to the excess morbidity and mortality noted in older patients. In contrast, most pediatric transplant centers assume that the patient with a diagnosis of MDS will undergo transplant as soon as feasible following diagnosis; the IPSS or other risk-based classification schemes do not usually play a major role in decision-making. The exception to this is the child who is highly transfusion-dependent or who has increasing evidence of clonality or leukemic progression on serial bone marrow evaluations, who is usually prepared for transplant more urgently. The principle limitation to transplant in pediatrics is most often limited to finding a suitable HLA-matched donor. With the increasing use of cord blood, and in some centers, HLA haploidentical individuals as a source of donor cells, the pool of available donors has expanded significantly. Overall and event free survival for pediatric MDS patients with SCT are reported to range from approximately 10 to 52%, with multiple variables affecting outcome (Stary et€al. 2005; Woodard et€al. 2006; Kindwall-Keller and Isola 2009). Similarly, the range of transplant-related morbidities, incidence of graft vs. host disease (GVHD), and transplant-related mortality range even more widely. Patients who have a higher bone marrow blast involvement and those with t-MDS/t-AML appear to have poorer outcomes, most often due to death from refractory or recurrent disease (Woodard et€ al. 2006). The poorer outcome for patients with higher blast involvement suggests that perhaps these patients should undergo pretransplant therapy aimed at reducing the clonal population. Several studies in the adult transplant literature have evaluated either a demethylating agent or lenalidomide (Platzbecker et€al. 2007) prior to allogeneic SCT, and shown this approach to be both safe and feasible. Efforts at post-transplant immunomodulation for relapse have also included epigenetic modifiers (Kroger 2008). Attempts to abrogate the severity of graft-vs.-host disease have been investigated using vorinostat (Reddy et€ al. 2004). Czibere and colleagues describe the use of azacitidine with donor lymphocyte infusion for the relapse of AML or MDS following allogeneic transplant (Czibere et€ al. 2006). One compelling approach for MDS would be to combine the epigenetic modifiers and/or immunomodulatory agents prior to and during stem cell transplantation, and then to use post-transplant maintenance therapy with epigenetic modulators as clinically and biologically indicated based on donor chimerism and molecular response. This could potentially provide definitive curative therapy with transplantation, even in the setting of reduced intensity conditioning regimens, while offering pre- and post-transplant maintenance medical therapy, to reduce disease burden prior to transplantation, and to sustain epigenetic and/or immunomodulatory effects following transplant.
Conclusions A wide variety of targeted agents are being studied for the treatment of MDS and the consequent leukemias in adult patients, and early results are promising, particularly in the realm of epigenetic modification of the cells of interest. While supportive
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
171
care efforts have improved and newer agents are available to enhance support and decrease side effects for patients with MDS, these are unlikely to alter the biological features that drive ultimate disintegration into a fatal disease. Further, molecular characterization of the group of syndromes collectively known as MDS will be important in the development of tailored therapy. Taken together, there are a large number of biologically compelling approaches to the treatment of MDS. A significant challenge is to determine the optimal combinations of one or several agents in conjunction with more traditional chemotherapy that will improve cure rates and decrease short- and long-term morbidity associated with the treatment for these diseases. Unfortunately, many of these combination trials have been slow to be tested in children, leaving the pediatric clinician with few options in an otherwise active area of investigation. Multi-institutional and international collaborations for the MDS, due to their rarity in children, will be of paramount importance to better gauge utility and effectiveness of such approaches in this population (Kroger 2008).
References Bhojwani D, Kang H, Moskowitz NP, et€al. Biologic pathways associated with relapse in childhood acute lymphoblastic leukemia: a Children’s Oncology Group study. Blood 2006:108(2):711–717. Carroll WL, Bhojwani D, Min DJ, et€al. Childhood acute lymphoblastic leukemia in the age of genomics. Pediatr Blood Cancer 2006:46(5):570–578. Rollison DE, Howlader N, Smith MT, et€ al. Epidemiology of myelodysplastic syndromes and chronic myeloproliferative disorders in the United States, 2001–2004, using data from the NAACCR and SEER programs. Blood 2008:112(1):45–52. Finn R. Medicare analysis suggests MDS is “severely underestimated.” Oncol Rep 2009(Spring): 21(1):37–38. Ma X, Does M, Raza A, et€al. Myelodysplastic syndromes: incidence and survival in the United States. Cancer 2007:109(8):1536–1542. Locatelli F, Zecca M, Pession A, et€al. Myelodysplastic syndromes: the pediatric point of view. Haematologica 1995:80(3):268–279. Cervantes F, Dupriez B, Pereira A, et€al. New prognostic scoring system for primary myelofibrosis based on a study of the International Working Group for Myelofibrosis Research and Treatment. Blood 2009:113(13):2895–2901. Bennett JM, Catovsky D, Daniel MT, et€al. Proposals for the classification of the myelodysplastic syndromes. Br J Haematol 1982:51(2):189–199. Jaffe ES, Harris NL, Stein H, et€ al. Tumours of the haematolpoietic and lyphoid tissues. In: Tumours WHOCo, editor. Pathology and Genetics of Tumours of the Haematopoietic and Lymphoid Tissues. Lyon: International Agency for Research on Cancer; 2001 Greenberg P, Cox C, LeBeau MM, et€al. International scoring system for evaluating prognosis in myelodysplastic syndromes. Blood 1997:89(6):2079–2088. Kaushansky. Erratum. Blood 1998:91(3):1100. Hasle H, Niemeyer CM, Chessells JM, et€al. A pediatric approach to the WHO classification of myelodysplastic and myeloproliferative diseases. Leukemia 2003:17(2):277–282. Mandel K, Dror Y, Poon A, et€al. A practical, comprehensive classification for pediatric myelodysplastic syndromes: the CCC system. J Pediatr Hematol Oncol 2002:24(7):596–605. Lorand-Metze I, Pinheiro MP, Ribeiro E, et€ al. Factors influencing survival in myelodysplastic syndromes in a Brazilian population: comparison of FAB and WHO classifications. Leuk Res 2004:28(6):587–594.
172
L. Gore
Occhipinti E, Correa H, Yu L, et€al. Comparison of two new classifications for pediatric myelodysplastic and myeloproliferative disorders. Pediatr Blood Cancer 2005:44(3):240–244. Alter BP. Fanconi’s anemia and malignancies. Am J Hematol 1996:53(2):99–110. Alter BP. Bone marrow failure syndromes in children. Pediatr Clin North Am 2002:49(5): 973–988. Homans AC, Verissimo AM, Vlacha V. Transient abnormal myelopoiesis of infancy associated with trisomy 21. Am J Pediatr Hematol Oncol 1993:15(4):392–399. Creutzig U, Ritter J, Vormoor J, et€al. Myelodysplasia and acute myelogenous leukemia in Down’s syndrome. A report of 40 children of the AML-BFM Study Group. Leukemia 1996:10(11): 1677–1686. Rischewski JR, Clausen H, Leber V, et€al. A heterozygous frameshift mutation in the Fanconi anemia C gene in familial T-ALL and secondary malignancy. Klin Padiatr 2000:212(4):174–176. Surico G, Muggeo P, Rigillo N, et€al. Concurrent Langerhans cell histiocytosis and myelodysplasia in children. Med Pediatr Oncol 2000:35(4):421–425. Cetin M, Hicsonmez G, Gogus S. Myelodysplastic syndrome associated with Griscelli syndrome. Leuk Res 1998:22(9):859–862. Ohara A, Kojima S, Hamajima N, et€al. Myelodysplastic syndrome and acute myelogenous leukemia as a late clonal complication in children with acquired aplastic anemia. Blood 1997:90(3):1009–1013. Winick NJ, McKenna RW, Shuster JJ, et€al. Secondary acute myeloid leukemia in children with acute lymphoblastic leukemia treated with etoposide. J Clin Oncol 1993:11:209–217. Pui CH, Ribeiro RC, Hancock ML, et€al. Acute myeloid leukemia in children treated with epipodophyllotoxins for acute lymphoblastic leukemia. N Engl J Med 1991:325(24):1682–1687. Tew KD, Colvin OM, Chabner BA. Alkylating agents. In: Chabner BA, Longo DL, editors. Cancer Chemotherapy and Biotherapy. Third ed. Phliadelphia: Lippincott Williams & Wilkins; 2001, pp 397–398. Smith MA, McCaffrey RP, Karp JE. The secondary leukemias: challenges and research directions. J Natl Cancer Inst 1996:88:407–418. Bo J, Schroder H, Kristinsson J, et€al. Possible carcinogenic effect of 6-mercaptopurine on bone marrow stem cells: relation to thiopurine metabolism. Cancer 1999:86(6):1080–1086. Felix CA, Hosler MR, Winick NJ, et€al. ALL-1 gene rearrangements in DNA topoisomerase II inhibitor-related leukemia in children. Blood 1995:85(11):3250–3256. Privitera E, Wang W, Raimondi SC. Monosomy 7 and unbalanced t(1;7) in an adolescent boy with myelodysplastic syndrome. Leukemia 1992:6(7):742–745. Yesilipek MA, Luleci G, Velipasaoglu S, et€al. Monosomy 7 myelodysplasia in childhood. Two case reports. Acta Haematol 1994:92(1):36–38. Maris JM, Wiersma SR, Mahgoub N, et€ al. Monosomy 7 myelodysplastic syndrome and other second malignant neoplasms in children with neurofibromatosis type 1. Cancer 1997:79(7): 1438–1446. Harrison KJ, Massing B, McKenna C, et€al. Molecular cytogenetic analysis of monosomy 7 in pediatric patients with myelodysplastic syndrome. Am J Hematol 1995:48(2):88–91. Chantrain C, Vermylen C, Michaux L, et€al. Clonal monosomy 7 and 5q – in a child with myelodysplastic syndrome. Pediatr Hematol Oncol 2000:17(6):505–509. Chakraborty S, Sun CL, Francisco L, et€al. Accelerated telomere shortening precedes development of therapy-related myelodysplasia or acute myelogenous leukemia after autologous transplantation for lymphoma. J Clin Oncol 2009:27(5):791–798. Taketani T, Taki T, Takita J, et€al. AML1/RUNX1 mutations are infrequent, but related to AMLM0, acquired trisomy 21, and leukemic transformation in pediatric hematologic malignancies. Genes Chromosomes Cancer 2003:38(1):1–7. Tartaglia M, Niemeyer CM, Fragale A, et€al. Somatic mutations in PTPN11 in juvenile myelomonocytic leukemia, myelodysplastic syndromes and acute myeloid leukemia. Nat Genet 2003:34(2):148–150. Lo Coco F, Pisegna S, Diverio D. The AML1 gene: a transcription factor involved in the pathogenesis of myeloid and lymphoid leukemias. Haematologica 1997:82(3):364–370.
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
173
Perkins D, Brennan S, Carstairs K, et€al. Regional cancer cytogenetics: a report on 1,143 diagnostic cases. Cancer Genet Cytogenet 1997:96(1):64–80. Block AW, Carroll AJ, Hagemeijer A, et€al. Rare recurring balanced chromosome abnormalities in therapy-related myelodysplastic syndromes and acute leukemia: report from an international workshop. Genes Chromosomes Cancer 2002:33(4):401–412. Woods WG, Neudorf S, Gold S, et€al. A comparison of allogeneic bone marrow transplantation, autologous bone marrow transplantation, and aggressive chemotherapy in children with acute myeloid leukemia in remission. Blood 2001:97(1):56–62. Huang ME, Yu-chen Y, Shu-rong C, et€al. Use of all-trans retinoic acid in the treatment of acute promyelocytic leukemia. Blood 1988:72:567–572. Singal R, Ginder GD. DNA methylation. Blood 1999:93(12):4059–4070. Esteller M, Corn PG, Baylin SB, et€al. A gene hypermethylation profile of human cancer. Cancer Res 2001:61(8):3225–3229. Das PM, Singal R. DNA methylation and cancer. J Clin Oncol 2004:22(22):4632–4642. Vidal DO, Paixao VA, Brait M, et€al. Aberrant methylation in pediatric myelodysplastic syndrome. Leuk Res 2007:31(2):175–181. Cameron EE, Bachman KE, Myohanen S, et€ al. Synergy of demethylation and histone deacetylase inhibition in the re-expression of genes silenced in cancer. Nat Genet 1999:21: 103–107. Rountree MR, Bachman KE, Herman JG, et€ al. DNA methylation, chromatin inheritance, and cancer. Oncogene 2001:20(24):3156–3165. Rountree MR, Bachman KE, Baylin SB. DNMT1 binds HDAC2 and a new co-repressor, DMAP1, to form a complex at replication foci. Nat Genet 2000:25(3):269–277. Jones PL, Wolffe AP. Relationships between chromatin organization and DNA methylation in determining gene expression. Semin Cancer Biol 1999:9(5):339–347. Zhou L, Opalinska JB, Sohal DP, et€al. Integrative genomic analysis reveals aberrant epigenetic marks in MDS that can be seen in peripheral blood leucocytes. In: Dunbar CE, editor; 2008; San Francisco, CO. American Society of Hematology. p 223. Kuerbitz SJ, Pahys J, Wilson A, et€ al. Hypermethylation of the imprinted NNAT locus occurs frequently in pediatric acute leukemia. Carcinogenesis 2002:23(4):559–564. Imamura T, Kakazu N, Hibi S, et€al. Rearrangement of the MOZ gene in pediatric therapy-related myelodysplastic syndrome with a novel chromosomal translocation t(2;8)(p23;p11). Genes Chromosomes Cancer 2003:36(4):413–419. Uchida T, Kinoshita T, Nagai H, et€al. Hypermethylation of the p15INK4B gene in myelodysplastic syndromes. Blood 1997:90(4):1403–1409. Jones PA, Laird PW. Cancer epigenetics comes of age. Nat Genet 1999:21(2):163–167. Jiang Y, Dunbar A, Gondek LP, et€ al. Aberrant DNA methylation is a dominant mechanism in MDS progression to AML. Blood 2009:113(6):1315–1325. Garcia-Manero G, Kantarjian HM, Sanchez-Gonzalez B, et€al. Phase 1/2 study of the combination of 5-aza-2¢-deoxycytidine with valproic acid in patients with leukemia. Blood 2006:108(10): 3271–3279. Scott SA, Lakshimikuttysamma A, Sheridan DP, et€al. Zebularine inhibits human acute myeloid leukemia cell growth in€vitro in association with p15INK4B demethylation and reexpression. Exp Hematol 2007:35(2):263–273. Shaker S, Bernstein M, Momparler LF, et€al. Preclinical evaluation of antineoplastic activity of inhibitors of DNA methylation (5-aza-2¢-deoxycytidine) and histone deacetylation (trichostatin A, depsipeptide) in combination against myeloid leukemic cells. Leuk Res 2003:27(5): 437–444. Silverman LR, McKenzie DR, Peterson BL, et€ al. Further analysis of trials with azacitidine in patients with myelodysplastic syndrome: studies 8421, 8921, and 9221 by the Cancer and Leukemia Group B. J Clin Oncol 2006:24(24):3895–3903. Flotho C, Claus R, Batz C, et€al. The DNA methyltransferase inhibitors azacitidine, decitabine and zebularine exert differential effects on cancer gene expression in acute myeloid leukemia cells. Leukemia 2009:23:1019–1028.
174
L. Gore
Kuendgen A, Schmid M, Schlenk R, et€al. The histone deacetylase (HDAC) inhibitor valproic acid as monotherapy or in combination with all-trans retinoic acid in patients with acute myeloid leukemia. Cancer 2006:106(1):112–119. Kuendgen A, Knipp S, Fox F, et€al. Results of a phase 2 study of valproic acid alone or in combination with all-trans retinoic acid in 75 patients with myelodysplastic syndrome and relapsed or refractory acute myeloid leukemia. Ann Hematol 2005:84 Suppl 1:61–66. Rosato RR, Almenara JA, Grant S. The histone deacetylase inhibitor MS-275 promotes differentiation or apoptosis in human leukemia cells through a process regulated by generation of reactive oxygen species and induction of p21CIP1/WAF1 1. Cancer Res 2003:63(13): 3637–3645. Fouladi M, Furman WL, Chin T, et€al. Phase I study of depsipeptide in pediatric patients with refractory solid tumors: a Children’s Oncology Group report. J Clin Oncol 2006:24(22): 3678–3685. Murgo AJ. Clinical trials of arsenic trioxide in hematologic and solid tumors: overview of the National Cancer Institute Cooperative Research and Development Studies. Oncologist 2001:6 Suppl 2:22–28. Sekeres MA, List A. Lenalidomide (Revlimid, CC-5013) in myelodysplastic syndromes: is it any good? Curr Hematol Rep 2005:4(3):182–185. Sekeres MA, List A. Immunomodulation in myelodysplastic syndromes. Best Pract Res Clin Haematol 2006:19(4):757–767. Sekeres MA, Maciejewski JP, Giagounidis AA, et€al. Relationship of treatment-related cytopenias and response to lenalidomide in patients with lower-risk myelodysplastic syndromes. J Clin Oncol 2008:26(36):5943–5949. Chang DH, Liu N, Klimek V, et€al. Enhancement of ligand-dependent activation of human natural killer T cells by lenalidomide: therapeutic implications. Blood 2006:108(2):618–621. List A, Dewald G, Bennett J, et€al. Lenalidomide in the myelodysplastic syndrome with chromosome 5q deletion. N Engl J Med 2006:355(14):1456–1465. Raza A, Reeves JA, Feldman EJ, et€al. Phase 2 study of lenalidomide in transfusion-dependent, low-risk, and intermediate-1 risk myelodysplastic syndromes with karyotypes other than deletion 5q. Blood 2008:111(1):86–93. Gilheeney SW, Lyden DC, Sgouros S, et€al. A phase II trial of thalidomide and cyclophosphamide in patients with recurrent or refractory pediatric malignancies. Pediatr Blood Cancer 2007:49(3): 261–265. Kieran MW, Turner CD, Rubin JB, et€al. A feasibility trial of antiangiogenic (metronomic) chemotherapy in pediatric patients with recurrent or progressive cancer. J Pediatr Hematol Oncol 2005:27(11):573–581. Lopez-Aguilar E, Sepulveda-Vildosola AC, Betanzos-Cabrera Y, et€al. Phase II study of metronomic chemotherapy with thalidomide, carboplatin-vincristine-fluvastatin in the treatment of brain stem tumors in children. Arch Med Res 2008:39(7):655–662. Rosenfeld C, Cheever MA, Gaiger A. WT1 in acute leukemia, chronic myelogenous leukemia and myelodysplastic syndrome: therapeutic potential of WT1 targeted therapies. Leukemia 2003: 17(7):1301–1312. Boehrer S, Ades L, Braun T, et€al. Erlotinib exhibits antineoplastic off-target effects in AML and MDS: a preclinical study. Blood 2008:111(4):2170–2180. Galili N, Cerny J, Raza A. Current treatment options: impact of cytogenetics on the course of myelodysplasia. Curr Treat Options Oncol 2007:8(2):117–128. Soriano AO, Yang H, Faderl S, et€al. Safety and clinical activity of the combination of 5-azacytidine, valproic acid, and all-trans retinoic acid in acute myeloid leukemia and myelodysplastic syndrome. Blood 2007:110(7):2302–2308. Fouladi M, Park JR, Sun J, et€al. A phase I trial and pharmacokinetic (PK) study of vorinostat (SAHA) in combination with 13 cis-retinoic acid (13cRA) in children with refractory neuroblastomas, medulloblastomas, primitive neuroectodermal tumors (PNETs), and atypical teratoid rhabdoid tumors. In: Grunberg SM, editor; 2008; Chicago, IL. American Society of Clinical Oncology. p 542s.
Molecularly Targeted Therapies in Pediatric Myelodysplastic Syndromes
175
Yao Q, Weigel B, Kersey J. Synergism between etoposide and 17-AAG in leukemia cells: critical roles for Hsp90, FLT3, topoisomerase II, Chk1, and Rad51. Clin Cancer Res 2007:13(5): 1591–1600. Weigel BJ, Blaney SM, Reid JM, et€al. A phase I study of 17-allylaminogeldanamycin in relapsed/ refractory pediatric patients with solid tumors: a Children’s Oncology Group study. Clin Cancer Res 2007:13(6):1789–1793. Bagatell R, Gore L, Egorin MJ, et€al. Phase I pharmacokinetic and pharmacodynamic study of 17-N-allylamino-17-demethoxygeldanamycin in pediatric patients with recurrent or refractory solid tumors: a pediatric oncology experimental therapeutics investigators consortium study. Clin Cancer Res 2007:13(6):1783–1788. Stary J, Locatelli F, Niemeyer CM. Stem cell transplantation for aplastic anemia and myelodysplastic syndrome. Bone Marrow Transplant 2005:35 Suppl 1:S13–S16. Woodard P, Barfield R, Hale G, et€ al. Outcome of hematopoietic stem cell transplantation for pediatric patients with therapy-related acute myeloid leukemia or myelodysplastic syndrome. Pediatr Blood Cancer 2006:47(7):931–935. Kindwall-Keller T, Isola LM. The evolution of hematopoietic SCT in myelodysplastic syndrome. Bone Marrow Transplant 2009:43:597–609. Platzbecker U, Mohr B, von Bonin M, et€al. Lenalidomide as induction therapy before allogeneic stem cell transplantation in a patient with proliferative CMML-2 and del(5q) not involving the EGR1 locus. Leukemia 2007:21(11):2384–2385. Kroger N. Epigenetic modulation and other options to improve outcome of stem cell transplantation in MDS. San Francisco, CA: American Society of Hematology; 2008. Reddy P, Maeda Y, Hotary K, et€al. Histone deacetylase inhibitor suberoylanilide hydroxamic acid reduces acute graft-versus-host disease and preserves graft-versus-leukemia effect. Proc Natl Acad Sci USA 2004:101(11):3921–3926. Czibere A, Graef T, Lind J, et€al. 5-Azacitidine in combination with donor lymphocyte infusion for treatment of patients with MDS or AML relapsing after allogeneic stem cell transplantation. Blood 2006:108:5341. Morerio C, Rapella A, Tassano E, et€al. Gain of 1q in pediatric myelodysplastic syndromes. Leuk Res 2006:30(11):1437–1441. Mogul MJ, Brady K, Brothman AR, et€ al. Myelodysplastic syndrome presenting with clonal rearrangement isolated to chromosomal region 1q. Cancer Genet Cytogenet 1997:95(2): 210–212. Lahortiga I, Agirre X, Belloni E, et€al. Molecular characterization of a t(1;3)(p36;q21) in a patient with MDS. MEL1 is widely expressed in normal tissues, including bone marrow, and it is not overexpressed in the t(1;3) cells. Oncogene 2004:23(1):311–316. Horsman DE, Massing BG, Chan KW, et€al. Unbalanced translocation (1;7) in childhood myelodysplasia. Am J Hematol 1988:27(3):174–178. Stark B, Jeison M, Shohat M, et€al. Involvement of 11p15 and 3q21q26 in therapy-related myeloid leukemia (t-ML) in children. Case reports and review of the literature. Cancer Genet Cytogenet 1994:75(1):11–22. Rubin CM, Larson RA, Anastasi J, et€al. t(3;21)(q26;q22): a recurring chromosomal abnormality in therapy-related myelodysplastic syndrome and acute myeloid leukemia. Blood 1990:76(12): 2594–2598. Dalla Torre CA, de Martino Lee ML, Yoshimoto M, et€al. Myelodysplastic syndrome in childhood: report of two cases with deletion of chromosome 4 and the Philadelphia chromosome. Leuk Res 2002:26(6):533–538 Pellier I, Le Moine PJ, Rialland X, et€al. Myelodysplastic syndrome with t(5;12)(q31;p12-p13) and eosinophilia: a pediatric case with review of literature. J Pediatr Hematol Oncol 1996:18(3): 285–288. Wong WY, Wu SQ. Pentasomy 8 in pediatric myelodysplastic syndrome. Cancer Genet Cytogenet 2000:121(2):220–222. Taki T, Sako M, Tsuchida M, et€al. The t(11;16)(q23;p13) translocation in myelodysplastic syndrome fuses the MLL gene to the CBP gene. Blood 1997:89(11):3945–3950.
176
L. Gore
Tosi S, Giudici G, Mosna G, et€al. Identification of new partner chromosomes involved in fusions with the ETV6 (TEL) gene in hematologic malignancies. Genes Chromosomes Cancer 1998:21(3): 223–229. Coignet LJ, Lima CS, Min T, et€al. Myeloid- and lymphoid-specific breakpoint cluster regions in chromosome band 13q14 in acute leukemia. Genes Chromosomes Cancer 1999:25(3): 222–229. Jamal R, Taketani T, Taki T, et€al. Coduplication of the MLL and FLT3 genes in patients with acute myeloid leukemia. Genes Chromosomes Cancer 2001:31(2):187–190. Hellstrom-Lindberg E, Cazzola M. The role of JAK2 mutations in RARS and other MDS. Hematology Am Soc Hematol Educ Program 2008:2008:52–59. Side LE, Emanuel PD, Taylor B, et€al. Mutations of the NF1 gene in children with juvenile myelomonocytic leukemia without clinical evidence of neurofibromatosis, type 1. Blood 1998:92(1): 267–272. Zhang Y, Zhang M, Yang L, et€ al. NPM1 mutations in myelodysplastic syndromes and acute myeloid leukemia with normal karyotype. Leuk Res 2007:31(1):109–111. Silveira CG, Oliveira FM, Valera ET, et€al. New recurrent deletions in the PPARgamma and TP53 genes are associated with childhood myelodysplastic syndrome. Leuk Res 2009:33(1):19–27. Mahgoub N, Parker RI, Hosler MR, et€al. RAS mutations in pediatric leukemias with MLL gene rearrangements. Genes Chromosomes Cancer 1998:21(3):270–275. Gafanovich A, Ramu N, Krichevsky S, et€al. Microsatellite instability and p53 mutations in pediatric secondary malignant neoplasms. Cancer 1999:85(2):504–510. Kratz CP, Niemeyer CM, Castleberry RP, et€al. The mutational spectrum of PTPN11 in juvenile myelomonocytic leukemia and Noonan syndrome/myeloproliferative disease. Blood 2005:106(6):2183–2185. Bentires-Alj M, Paez JG, David FS, et€al. Activating mutations of the noonan syndrome-associated SHP2/PTPN11 gene in human solid tumors and adult acute myelogenous leukemia. Cancer Res 2004:64(24):8816–8820. Loh ML, Vattikuti S, Schubbert S, et€al. Mutations in PTPN11 implicate the SHP-2 phosphatase in leukemogenesis. Blood 2004:103(6):2325–2331.
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma Megan S. Lim and Mitchell S. Cairo
Introduction Childhood non-Hodgkin lymphoma (NHL) accounts for approximately 6 to 7% of all childhood cancers diagnosed below the age of 15 in the United States and consists of four major histological subtypes of both intermediate and high-grade NHL including Burkitt lymphoma (BL), lymphoblastic lymphoma (LL), diffuse large B-cell lymphoma (DLBCL), and anaplastic large cell lymphoma (ALCL) (Cairo et€al. 2005). While there are a large number of other histological subtypes of NHL that may occur in children, they only account for <5% of all cases of NHL diagnosed in the United States and Western Europe. Posttransplant lymphoproliferative disease (PTLD) may also occur in children following either solid organ transplantation or mismatched or T-cell depleted allogeneic stem cell transplantation and will be the subject of a separate chapter in this monograph. We will focus our discussion on the four major histological subtypes that account for >95% of all NHL in children (Cairo et€al. 2005; Pinkerton 2005). BL accounts for approximately 40% of NHL diagnosed in children, and with the recent use of short and intensive multi-agent chemotherapy, the prognosis has improved dramatically over the past two decades (children with localized disease experience a ³95% 5-year event-free survival (EFS) and patients with more advanced disease a 60 to 90% EFS) (Table€1) (Cairo et€al. 2005; Pinkerton 2005). LL is the second most common histology of NHL diagnosed in children and accounts for approximately 30% of all cases and in patients with localized disease, depending on the intensity of therapy administered, 60 to 90% 5-year EFS and patients with more advanced disease, 80 to 90% 5-year EFS (Table€ 1) (Cairo et€ al. 2005; Pinkerton
M.S. Cairo (*) Division of Pediatric Blood and Marrow Transplantation, New York-Presbyterian Morgan Stanley Children’s Hospital, Columbia University, 3959 Broadway, CHN 10-03, New York, NY 10032, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_10, © Springer Science+Business Media, LLC 2010
177
178 Table€1╅ Childhood non-Hodgkin lymphoma Histology Frequency (%) Burkitt lymphoma (BL) 40 Lymphoblastic lymphoma (LL) 30 Diffuse large B-cell lymphoma 20 (DLBCL) Anaplastic large cell lymphoma 10 EFS event free survival
M.S. Lim and M.S. Cairo
Localized % EFS 95 60–90 95
Advanced % EFS 60–90 80–90 80–90
80–90
60–80
2005). DLBCL occurs in approximately 20% of all children diagnosed with NHL under the age of 15. Patients with localized disease have a >95% 5-year EFS and patients with more advanced disease, 60 to 80% 5-year EFS (Table€1). ALCL is the least common form of childhood NHL and accounts for approximately 10% of all cases of NHL diagnosed in children under 15 years of age and has a 5-year EFS for localized disease of 80 to 90% and in patients with more advanced disease a 5-year EFS of 60 to 80% (Table€1) (Cairo et€al. 2005; Pinkerton 2005). Despite these high cure rates with intensive multi-agent chemotherapy, there is still an emergent need to identify new targets and pathways that can be strategically exploited for more rational therapeutic approaches. Pinkerton (2005) in a more recent review, noted there are still significant challenges in the diagnosis and treatment of childhood NHL. In patients with an excellent prognosis following multi-agent and intensive chemotherapy, there is still a high risk of morbidity, some mortality, prolonged hospitalization and late effects. This group of patients may benefit more from a more targeted therapeutic approach that could potentially reduce the above complications. Secondly, there is a group of patients who still only have 60 to 80% chance of being cured with intensive multi-agent chemotherapy and there is considerable room to improve upon the prognosis in this subgroup of patients. Lastly, in patients who relapse after front line therapy for childhood NHL, the prognosis is generally considered dismal with overall survival rates ranging only between 10 and 30%, and therefore, new therapeutic modalities need to be developed to retrieve this poor risk group of relapsing or refractory children with NHL (Cairo et€ al. 2005; Pinkerton 2005). We recently summarized state of the science of what is known regarding the biology of childhood NHL. In the following discussion, we will focus on the potential molecular and pathway targets that could be potentially exploited for new rational drug design and therapeutics for childhood NHL.
Diffuse Large B-Cell Lymphoma DLBCL accounts for approximately 20% of all childhood and adolescent NHL. In the last international French-American-British (FAB) study in children and adolescents with newly diagnosed B-NHL, we reported a 80 to 90% 4-year EFS in
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
179
children with DLBCL, depending on stage, LDH, primary site location, and response to initial induction therapy (Cairo et€al. 2007; Patte et€al. 2007). Significant progress has occurred in the last 5 years in the understanding of molecular pathogenesis of adult DLBCL and only recently have we identified similar findings in childhood DLBCL. Rosenwald et€al. (2002) has identified three genetic subtypes of adult DLBCL including germinal center (GC-derived), activated B-cell-like (ABC) and type 3, all of which have a differential prognosis following standard CHOP therapy (Fig.€1a). Furthermore, this same group of investigators also demonstrated five distinct molecular signature patterns that identified four subquartile prognostic groups who had differential outcomes ranging from 10 to 70% following standard CHOP therapy (Fig.€1b) (Rosenwald et€al. 2002) We recently examined the genetic origin of children with DLBCL treated on the FAB/lymphoma malignancy B (LMB) 96 international B-NHL study and demonstrated that >80% of children and adolescents with DLBCL had the GC phenotype, which may in part account for the
Fig.€1â•… (a) Gene expression profiles of subgroups of adult DLBCL and expression of signature genes. Level of expression of variables in the outcome predictor and the scores in the three subgroups of lymphoma. (b) Gene expression profile signature score predicts for overall survival in adult DLBCL. Kaplan-Meier estimates of overall survival among all patients. Rosenwald et€al. 2002. Copyright © 2002 Massachusetts Medical Society. All rights reserved
180
M.S. Lim and M.S. Cairo
improved prognosis in children versus adults with DLBCL (Miles et€al. 2007). A fourth subtype of DLBCL includes the primary mediastinal B-cell lymphomas (PMBL). The prognosis of PMBL in children and adolescents is significantly inferior compared to other forms of childhood and adolescent B-NHL with only an approximately 65% 4-year EFS following therapy on the most recent international FAB/LMB 96 trial (Patte et€al. 2007). More recently, gene expression profiling of adult PMBL was compared to adult GCB and ABC DLBCL and was demonstrated to have a significantly different molecular signature pattern (Fig.€2a) (Rosenwald et€al. 2003). Furthermore, the molecular signature of adult PMBL in fact resembled more that of Hodgkin disease (HD) than the other forms of adult DLBCL (Fig.€2b) (Rosenwald et€al. 2003; Savage et€al. 2003). Shipp et€al. has recently proposed that PMBL, otherwise known as mediastinal large B-cell lymphoma (MLBCL), is part of a developmental continuum between classical HD (cHD) and classical DLBCL (Savage et€al. 2003; Abramson and Shipp 2005). One of the most commonly analyzed molecular targets in DLBCL has been the BCL-6 gene, particularly in the subtype of germinal center-derived DLBCL
Fig.€2╅ (a) Identification of a PMBL gene expression. Hierarchical clustering identified a set of 23 PMBL signature genes that were more highly expressed in most lymphoma with a clinical diagnosis of PMBL than in lymphomas assigned to the GCB or ABC DLBCL subgroups. Each row presents gene expression measurements from a single Lymphochip microarray feature representing the genes indicated. Each column represents a single lymphoma biopsy sample. Relative gene expression is depicted according to the color scale shown. (b) Relationship of PMBL to Hodgkin lymphoma. Relative gene expression is shown in primary PMBLs (average of all biopsy samples), the PMBL cell line K1106, three Hodgkin lymphoma (HL) cell lines, and six GCB DLBCL cell lines, according to the color scale shown in (a). Reproduced from Rosenwald et€al. 2003. Copyright 2003 The Rockfeller University Press
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
181
Fig.€2╅ (continued)
(GC DLBCL). BCL-6 gene is localized on chromosome 3q27 and is genetically rearranged in over 35% of adults with the GC DLBCL subtype (Ye et€al. 1993; Lo Coco et€al. 1994). The BCL-6 protooncogene is a nuclear phosphoprotein belonging to the POZ/Zinc finger (ZF) family of transcription factors that encodes for a transcriptional repressor required for development of germinal centers and has been implicated in the pathogenesis of GC-derived DLBCL (Chang et€al. 1996). Deregulation of BCL-6 expression recapitulates the pathogenesis of human DLBCL in mice (Cattoretti et€al. 2005). A variety of stimuli, including B-cell receptor (BCR) activation, mitogenic stimulation, and CD40 receptor engagement, appear to downregulate BCL-6 expression (Allman et€ al. 1996). Transcriptional regulation of the BCL-6 gene may be disrupted by somatic hypermutation in the five noncoding regions of
182
M.S. Lim and M.S. Cairo
the BCL-6 gene in GC-derived DLBCL (Migliazza et€al. 1995; Pasqualucci et€al. 1998) or may occur due to tumor-associated chromosomal translocations, which juxtapose heterologous promoters to the coding exons of BCL-6, causing its deregulated expression by a mechanism previously described as “promoter substitution” (Ye et€al. 1995). BCL-6 is thought to function in GC formation presumably by its ability to repress the transcription of specific target genes. BCL-6 is believed to repress STAT 6 mediated, IL4-induced expression of Ig germ line transcripts, therefore, presumably modulating the isotype switch toward IgE in€ vitro and in€vivo. Additionally, a number of other putative BCL-6 network genes involved in apoptosis, cell cycle proliferation, cell differentiation and activation, and inflammation have also been proposed as potential targets (Harris et€al. 1999; Baron et€al. 2002). More recently, BCL-6 has been identified to control the expression of CD80, a costimulatory molecule, in normal GC B-cells (Niu et€ al. 2003). We recently analyzed BCL-6 interacting proteins by tandem mass spectrometry by coimmunoprecipitation with an anti-BCL-6 antibody in a GC-derived DLBCL line (Miles et€al. 2005a). Importantly, BCL-6 was identified to bind to a number of proteins including several histone deacetylators, including histone deacetylase (HDAC) 1,2,4,5,7 and 9, and a number of other important proteins, including BCL-11A, IRF4, JunB, c-Jun, among 61 other proteins, from a variety of functional categories including transcriptional regulators, binding activity, signal transduction activities, etc. (Fig.€3) (Miles et€al. 2005a).
Fig.€3â•… Cellular location and molecular function of BCL-6-interacting proteins. The relative proportions of identified proteins are shown for functional classification. Miles et€al. 2005a. Copyright © 2005 American Society for Biochemistry and Molecular Biology
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
183
BCL-6 appears to be regulated by a similar set of enzymatic reactions that control chromatin remodeling. Histones similar to regulating chromatin remodeling also appear to influence the regulation of BCL-6. Histone acetyltransferases (HATs) regulate open chromatin structures allowing transcriptional activation while deacetylation by HDAC promotes chromatin condensation and transcriptional inactivation. Histone deacetylase inhibitors (HDACIs) inhibit the removal of the acetyl groups from the lysine tails of the histone and help maintain DNA and an open chromosome structure thereby promoting transcriptional activation (O’Connor 2005). Similarly, BCL-6 is subject to acetylation by a p300 acetylase (HAT), which completely abrogates its transcriptional repressor function, and thereby promotes a number of BCL-6 dependent genes to be transcribed and activated (Bereshchenko et€al. 2002). Alternatively, deacetylation of BCL-6 by HDACs promotes BCL-6 to serve as a transcriptional repressor and inhibit the activities on a number of BCL-6 dependent genes (O’Connor 2005; Bereshchenko et€al. 2002). Recent studies have demonstrated that HDACIs promote growth inhibition, apoptosis, and anti-tumor activity in BCL-6 genetic rearranged DLBCL (O’Connor 2005, Bereshchenko et€al. 2002). Therefore, HDACIs can potentially modulate gene expression in GC-derived DLBCLs by one of two mechanisms, including influencing the aceytlation state of BCL-6, and second, by modulating the balance of open and closed chromatin available for transcription (O’Connor 2005). There have been a number of Phase 1 and Phase 2 clinical trials in adults with advanced cancer including those with DLBCLs that have investigated the safety and efficacy of HDACIs (O’Connor 2005; Piekarz et€al. 2001; Kelly et€al. 2003a; Prince et€al. 2007 (abstract)). There are currently four major classes of HDACIs, including short-chain fatty acids, which include valproic acid, hydroxamic acids including suberoylanilide hydroxamic acid (SAHA) (vorinostat), cyclotetrapeptides including depsipeptide and benzamides (MS-275). Future studies will be required to test the safety, efficacy, and mechanism of anti-tumor activity of these HDACIs in GC-derived DLBCL. A number of these agents including depsipeptide and SAHA are already in Phase I investigation in children with a variety of solid tumors (Fouladi et€al. 2006, 2007). An alternative approach to inhibiting BCL-6 is the use of specific BCL-6 peptide interference peptides (Polo et€al. 2004). Polo et€al. developed a peptide that binds to BCL-6 and blocks corepressor recruitment and disrupts BCL-6 mediated repression and reactivates BCL-6 target genes. Peptides were generated against the BTB binding domain (BBD motif) that is critical in BCL-6 interaction with other key target genes involved in transcriptional repression. BBD peptides were demonstrated to induce apoptosis, cell cycle arrest, suppress normal germinal center formation, and inhibit growth of B-cell lymphomas in€vivo (Polo et€al. 2004). We recently identified that only a minority of children and adolescents with DLBCL treated on the FAB/LMB96 international B-NHL trial were derived from the ABC subtype of DLBCL (Miles et€al. 2005b (abstract)). In adults with ABCDLBCL, there is a significant difference in their gene expression signature compared to the more common GC-DLBCL subtype (Fig.€ 2a) (Rosenwald et€ al. 2003). Recently, Ngo in the Staudt laboratory utilized a doxycycline inducible retroviral vector for the expression of small hairpin RNAs (shRNAs) to construct a library
184
M.S. Lim and M.S. Cairo
targeting 2,500 human genes, and subsequently utilized this library to infect cell lines representing both ABC-DLBCL and GC-DLBCL subtypes (Ngo et€al. 2006). In this shRNA-transduced study, CARD11 shRNAs specifically targeted genes of the NF-kB pathway following engagement of MALT 1 and BCL10 (Ngo et€al. 2006). These studies identified that CARD11 was the key upstream signaling component likely responsible for the ongoing constitutive NF-kB activity previously demonstrated in ABC-DLBCL (Ngo et€al. 2006). CARD11 and other associated genes are considered to be attractive targets for ABC-DLBCL since inhibition of CARD11 signaling would likely only affect the lymphoid system (Table€2) (Ngo et€al. 2006; Thome 2004). Although there are no inhibitors of CARD11 available for clinical use today, these studies suggest that future development of CARD11 inhibitors could potentially be utilized in targeted therapy for ABC-derived DLBCL. Besides CARD11, there are a number of other targets in the NF-kB pathway that can be potentially inhibited to downregulate NF-kB gene expression and transcriptional activation of a variety of genes promoting cell proliferation and survival (Table€2) (Karin 2006; Van Waes 2007). There are a number of signaling pathways that induce NF-kB activation and translocation to the nucleus to induce gene activation (Fig.€4) (Van Waes 2007). In the classical pathway of NF-kB activation, NF-kB homodimers or heterodimers are retained in the cytoplasm by binding to specific inhibitors named the inhibitors of NF-kBs (IkBs). Briefly, the IkB complex is formed by a number of IkB kinases (IKKs) and a number of NF-kB dimers, including RelA, cRel, RelB, p50/NF-kB1, and p52/NF-kB2. These heterodimeric and Table€2╅ Molecular targeting of DLBCL and BL Subtype Gene structure Function GC-DLBCL
BCL-6 26-S proteasome
Transcriptional repressor BCL-6 interference peptide Ubiquitin-mediated protein degradation
ABC-DLBCL
BLIMP-1 CARD11 NF-kB
Tumor suppressor Transcriptional activator Transcriptional activator
PMBL-DLBCL BL
NF-kB C-MYC
Transcriptional activator Proto-oncogene, Transcriptional activator
Inhibitors/modifier HDACs[Vorinostat, Depsipeptide] BBD motif inhibitor Proteosome inhibitor â•… [Bortezomib] Unknown Unknown Rituximab Bortezomib Small IKK inhibitors Small IKK inhibitors Benozochazepine (Bz-423) GSI Anti-sense oligonucleotides SiRNAs, or ShRNAs
DLBCL diffuse large B-cell lymphoma, BL Burkitt lymphoma, GC-DLBCL germinal cell-Diffuse large B-cell lymphoma, ABC-DLBCL activated B-cell like-diffuse large B-cell lymphoma, PMBL-DLBCL primary mediastinal B-cell lymphoma-diffuse large B-cell lymphoma, BL Burkitt lymphoma, HDACs histone deacetylase inhibitors, BBD BTB binding domain, GSI Gamma secretase inhibitor
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
185
Fig.€4â•… NF-kB activation in pathogenesis and therapy of cancer. (a) NF-kB activation in cancer development is linked with chronic exposure to bacteria, certain viral products, chemical promoters, and carcinogens and reactive oxygen species, which cause repeated DNA damage. Induction also occurs in response to cytotoxic chemotherapy and ionizing radiation. (b) classic NF-kB pathway activation occurs in response to these bacterial, viral, chemical, or physical stimuli; by aberrant cytokine, integrin, and growth factor ligand/receptor activation (e.g., TNFR, IL-1R, integrin a6h4,EG FR,H ER-2/neu, and other serum factors); expression of activating fusion proteins following translocations (e.g., BCR-ABL and MALT1); or aberrant activation by intermediate kinases (e.g., P I3K, CK2, and AKT). Intermediate kinases convey signals to the IkB complex formed by IKKa, IKKb, and IKKg, and IKKb and CK2 phosphorylate IkB, marking it for ubiquitination by E3 ligase bTrCP (SCF) and proteasome degradation. P105/RELA or cREL is processed to NF-kB1 (p50)/RELA or cREL heterodimers, which translocate to the nucleus and bind promoters of genes regulating proliferation, apoptosis, migration inflammation, angiogenesis, and innate immunity. (c) Alternative pathway. The alternative pathway may be activated by other TNF family members via the NF-kB inducing kinase and involves IKKa/IKKa homodimers, which activate NF-kB2/p100 for processing into p52/RelB heterodimers. The RelB/p52 heterodimer then translocates into the nucleus to bind the promoter of genes whose products are important for the malignant phenotype in some cancers and B-cell development and adaptive immunity. (d) In certain leukemias, overexpression of BCL3 can activate NF-nB2. E, repression of gene activation can occur in the presence of intact p14ARF and p53, which favor replacement of CBP/p300 with histone deacetylases. Red highlighted inhibitors of NF-kB activation under clinical and preclinical investigation include proteasome, bTrCP and IKK antagonists, and inhibitors of receptor and intermediate kinases involved in activation. Abbreviations: NIK NF-kB inducing kinase, IL-1R interleukin 1 receptor, EGFR epidermal growth factor receptor, TNFR TNF receptor, TRAF TNF receptor-associated factor, TAK transforming growth factor-b-activated kinase, FAK focal adhesion kinase, CK2 casein kinase 2, PI3K phosphatidylinositol 3-kinase, PDK 3-phosphoinositidedependent protein kinase. Originally published in Van Waes 2007. Copyright © 2007 American Association for Cancer Research
186
M.S. Lim and M.S. Cairo
homodimeric combinations of NF-kB are initially expressed in an active form in the cytoplasm. However, upon IKK phosphorylation, the NF-kB bound IkB proteins are targeted for polyubiquitination and rapid degradation by the ubiquitin ligase complex allowing NF-kB dimers to translocate to the nucleus where they promote the transcriptional activation of several hundred target genes (Fig.€ 4) (Karin 2006; Van Waes 2007). A number of drugs are in development, particularly the IKK small molecule inhibitors, which in preclinical studies appear to promote tumor cell apoptosis. Some of the more recent IKK selective antagonists, PS1145 or ML120b derivatives, have already demonstrated significant antiproliferative, cytotoxic, and anti-tumor effects (Van Waes 2007; Lam et€al. 2005). There are a number of new drugs in development that are also capable of targeting NF-kB activation and may be useful in the future for treatment of ABC-derived DLBCL and potentially PMBL, as discussed below. An important component of NF-kB activation is the polyubiquitination of IkBs that occurs within the proteasome. A novel class of proteasome inhibitors for cancer therapy has been in development over the past decade and the prototype, bortezomib, was recently approved by the FDA for the treatment of multiple myeloma (Table€ 2) (Voorhees and Orlowski 2006). Bortezomib was demonstrated to have broad anti-tumor activity in a number of tumor cell lines and in several murine xenograft models (Adams et€al. 1999). Proteasome inhibitors block the degradation of IkB, thereby sequestering NF-kB in the cytoplasm and downregulating its transcriptional activity (Voorhees and Orlowski 2006; Traenckner et€al. 1994). Phase 2 studies of bortezomib in adult B-NHL have already been completed and bortezomib is currently under investigation in combination with CHOP therapy in adult patients with DLBCL (Goy et€ al. 2005; O’Connor et€ al. 2005; Mounier et€ al. 2007). It remains to be determined whether bortezomib will be selectively active in subtypes of DLBCL that have an overexpressed NF-kB signal transduction pathway including ABC-derived DLBCL and PMBL (Table€2). Another agent that may be useful in targeting subtypes of DLBCL that have an overexpression of the NF-kB signal transduction pathways is the anti-CD20 antibody, rituximab. The mechanism of action of rituximab is believed to be multiÂ�factorial and possibly includes complement-dependent cytotoxicity (CDC), antibody-dependent cellular cytotoxicity (ADCC), and direct induction of tumor cell apoptosis. The specific mechanism by which rituximab may affect the NF-kB signaling pathway has recently been elucidated (Jazirehi et€al. 2005). Rituximab appears to upregulate the Raf-1 kinase inhibitor protein (RKIP). RKIP physically associates with NF-kB inducing kinase (NIK), tumor growth factor-B activating kinase 1 (TAK1), and IKK, preventing them from signaling further downstream (Jazirehi et€al. 2005). Upregulation of RKIP expression interferes with NF-kB signaling, leading to the diminished expression of the Bcl-xL gene product, which normally has an anti-apoptotic role (Jazirehi et€al. 2005). Therefore, rituximab may have several mechanisms of action with respect to subtypes of DLBCL, including inducing CDC and ADCC as well as inhibiting NF-kB signaling (Jazirehi et€al. 2005). Recently, B lymphocyte-induced maturation protein 1 (BLIMP-1), which has previously been demonstrated to be a transcriptional repressor expressed in germinal
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
187
center B-cells and in plasma cells, is required for terminal B-cell differentiation into plasma cells (Turner et€ al. 1994; Angelin-Duclos et€ al. 2000). Pasqualucci et€ al. (2006) recently demonstrated that the BLIMP 1 gene is inactivated by structural alterations in approximately 25% of ABC-derived DLBCLs. These studies suggest that BLIMP 1 may function as a tumor suppressor gene whose inactivation may contribute to ABC-derived DLBCL lymphomagenesis by inhibiting post-GC differentiation of B-cells into plasma cells. Although at the present time there are no known mediators to induce BLIMP 1 expression or activation, possibly in the future there may be the potential for developing a specific drug that can upregulate BLIMP 1 in patients with ABC-derived DLBCL and turn on the tumor suppressor property of BLIMP 1 (Table€2) (Pasqualucci et€al. 2006). The last major subtype of DLBCL is the PMBL, which accounts for <5% of all subtypes of DLBCL in children and adolescents (Cairo et€ al. 2005; Pinkerton 2005). Recent genetic profiling studies in PMBL suggest that PMBL resembles HD more and resembles the other two subtypes of DLBCL less, including ABC and GC derivatives (Fig.€2a) (Rosenwald et€al. 2003). Differential expression of PMBL, GC and ABC-DLBCL demonstrates a significant upregulation of a number of NF-kB target genes in PMBL (Fig.€5a) (Feuerhake et€al. 2005). Subsequently, Lam et€ al. (2005), utilizing several small IKK inhibitors such as PS1145 and MLX105, demonstrated selective toxicity in the subgroup of DLBCL that had overexpression of genes within the NF-kB signal transduction pathway (Fig.€5b). These studies and the studies mentioned with ABC-derived DLBCL suggest that
Fig.€ 5â•… (a) Differential expression of NF-kB target genes in an independent series of primary MLBCLs and DLBCLs. The independent data set includes all primary MLBCLs (38 tumors) and DLBCLs (26 tumors: 13 GC-type and 13 ABC-like) that were made available at the NIH Lymphoma/ Leukemia Molecular Profiling Project website. MLBCLs versus DLBCLs. This independent series of primary MLBCLs had significantly higher expression of 79% (26/33) of the NFkB target genes that were identified in our series and represented on both platforms; and GC-type versus ABC-like DLBCLs. This research was originally published in Blood. Feuerhake et€al. 2005. (b) Gene expression profiling of ABC DLBCL and PMBL cells treated with IKK inhibitors. Genes that were downregulated in K1106 cells treated with 25€Amol/L MLX105 for the indicated times. Color intensity, ratio of gene expression in MLX105 treated versus untreated cells. Originally published in Lam et€al. 2005. Copyright © 2007 American Association for Cancer Research
188
M.S. Lim and M.S. Cairo
specific inhibitors of the NF-kB signal transduction pathway may potentially have a selective and targeted toxic effect in subtypes of DLBCL, including ABC and PMBL (Table€2).
Burkitt Lymphoma The prognosis for childhood and adolescents with BL has improved dramatically over the last 25 years (Cairo et€al. 2003). Results from the most recent FAB/LMB 96 international childhood and adolescent B-NHL study demonstrated a 95 to 99% 5-year EFS in patients with localized disease and a 60 to 90% 5-year EFS in patients with advanced disease (Cairo et€al. 2007; Patte et€al. 2007). However, there are subgroups of children and adolescents with BL that respond poorly to standard therapy, including those newly diagnosed patients who have a poor response to reduction COP therapy, those patients with initial central nervous system and bone marrow involvement, and those patients who relapse off standard therapy (Cairo et€al. 2007). Additionally, the morbidity of short and intense multi-agent chemotherapy is extremely high in this population, with 65 to 70% of patients developing Grade 3 and 4 infections and/or mucositis with prolonged hospitalizations (Cairo et€al. 2007). Therefore, new targeted therapeutic modalities for childhood and adolescent BL is needed not only for newly diagnosed patients with either a poor prognosis or for those patients that are experiencing high morbidities with current therapy and is definitely needed for patients whose disease reoccurs, who have a <15% chance of overall survival (Cairo et€al. 2007; Patte et€al. 2007). Although almost 100% of childhood and adolescents with BL have an associated 8q24 (C-MYC) gene rearrangement, there are a number of other cytogenetic aberrations that we recently identified in this patient population in the most recent FAB/ LMB 96 international B-NHL study (Poirel et€al. 2003 (abstract)). Abnormalities in 1q+ was the most common chromosomal aberration. However, in a multivariable analysis, controlling for stage, LDH, and group classification deletion of 13q and 7q+, each were associated with significantly poor outcome compared to the remainder of all patients treated with the same therapy (Poirel et€ al. 2003 (abstract)). Therefore, specific cytogenetic aberrations are associated with a poor prognosis in childhood BL and may help us identify specific molecular targets that can be utilized for therapeutic purposes. Recently, a large group of adults and a small group of children with BL were analyzed by oligonucleotide microarray analysis (Dave et€ al. 2006). Dave et€ al. (2006) demonstrated significant genomic differences between BL and all subtypes of DLBCL with a significant increase in the BL subgroup in C-MYC targeted genes and GC-derived targeted genes but with a concomitant significant decrease in MHC-1 and NF-kB target genes compared to DLBCL (Fig.€6). C-MYC, a well known protooncogene and transcription factor, is a major regulator of a large network of genes that promote cell proliferation and survival. In a recent highly sophisticated analysis using an algorithm for the reconstruction of
189
Fig.€6â•… Differential gene signatures of C-MYC (a), MHC class I (c), and NF-kB (d), target genes between BL and subtype of DLBCL. The average relative expression of genes that distinguish Burkitt lymphoma from each subgroup of diffuse large-B-cell lymphoma (activated B-cell–like, germinal-center B-cell–like, and primary mediastinal) are categorized into gene expression signatures: C-MYC and its target genes (a); MHC class I genes (c); and genes targeted by the NF-kB signaling pathway27 (d). Relative gene expression is depicted according to the color scale shown. Dave et€al. 2006. Copyright © 2006 Massachusetts Medical Society. All rights reserved
190
M.S. Lim and M.S. Cairo
accurate cellular networks (ARACNe), Basso et€ al. elucidated a hierarchical organized subnetwork of genes that interact with C-MYC (Basso et€al. 2005). Three hundred and thirty-six expression profiles from different subpopulations of B-cell phenotypes were utilized to deconvolute the C-MYC cellular network. In this analysis, Basso et€ al. (2005) identified that the C-MYC subnetwork included 56 genes that were directly connected to C-MYC and another 2,007 genes that were connected through an intermediate that they labeled secondary neighbors (Fig.€7a, b). As illustrated in Fig.€7a, there are over 2,000 genes that are connected through the C-MYC subnetwork and, in Fig.€7b, is a representation of the first 500 genes including the 56 first neighbors and 44 of the most statistically significant second neighbors (Fig.€ 7a, b) (Basso et€ al. 2005). These results suggest that it may be possible to develop a molecular targeted approach to specific genes downstream from C-MYC for tumors such as BL, which are highly regulated by overexpression of C-MYC. One of the most notable C-MYC subhubs and a recently known C-MYC target identified in the analysis was BYSL (Basso et€al. 2005). A preliminary analysis suggested that the BYSL network may regulate cell proliferation, nucleic acid, metabolism, and ribosomal biogenesis and could serve as a molecular target in future studies (Basso et€al. 2005). Basso et€al. (2005) suggest that the structure of the network indicates that the subhubs are crucial mediators of regulatory function and should be considered in the future as priority targets for molecular targeted cancer therapy. To date, there are no specific approved drugs that target C-MYC or any of the critically important genes in the subhub network of C-MYC. However, there are several developments both preclinical and in early clinical trials that may provide a targeted approach for C-MYC in BL. A novel benzodiazepine, Bz-423 was demonstrated to induce apoptosis and inhibit proliferation in both normal and transformed lymphoid cells (Blatt et€al. 2002; Boitano et€al. 2003). Initial investigations of this agent targeted lymphocytes in a mouse model of lupus and arthritis and demonstrated significant therapeutic benefits in autoimmune disease (Bednarski et€ al. 2003). More recently, Bz-423 has been demonstrated to rapidly and specifically deplete C-MYC protein in a post-translational mechanism of inducing anti-proliferation and apoptosis (Sundberg et€al. 2006). Sundberg et€al. demonstrated that Bz-423
Fig.€7╅ (continued)╅ genes connected through an intermediate (second neighbors). For representation purposes, only the first 500 genes are shown, including all 56 first neighbors and the 444 most statistically significant second neighbors. Red or pink nodes represent first neighbor target genes for which ChIP data is available or not available, respectively; yellow and light yellow nodes represent second neighbor target genes for which ChIP data is available or not available, respectively; MYC is shown in green; white nodes represent genes for which no MYC-related information is available. The complete list of genes, including gene symbol, Affymetrix ID and LocusLink ID, is given in the Supplementary Table€4 online (see original manuscript). (b) The first neighbors of the MYC subnetwork. The size of each circle is proportional to the number of gene interactions. For hubs with more than 100 interactions, the exact number of first neighbors is shown beside the gene symbol. Reprinted by permission from Macmillan Publishers Ltd: Basso et€al. 2005, copyright 2005
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
191
Fig.€7â•… (a, b) The MYC subnetwork. (a) A MYC-specific subnetwork was obtained by including all the genes that have pâ•›<â•›10−7 based on their pairwise mutual information with MYC. The faster bin-counting estimator was used with an error tolerance e = 0.15. The MYC subnetwork includes 56 genes directly connected to MYC (first neighbors; represented by larger circles) and 2,007
192
M.S. Lim and M.S. Cairo
inhibited f1f0-ATPase activity, which inhibited respiratory chain function in generating superoxide and thereby led to proteasomal degradation of C-MYC protein. The effects of Bz-423 not only decreased C-MYC protein, but also modulated expression of several cell cycle regulatory proteins dependent on C-MYC expression (Table€2) (Sundberg et€al. 2006). It is well known that NOTCH1 signaling is required for the regulation and maintenance of hematopoietic stem cells and is critically important to the development of T-cells (Allman et€al. 2002). NOTCH1 signaling appears to regulate a number of processes including proliferation, cell growth, differentiation, and apoptosis. Activating NOTCH mutations have recently been identified in over 50% of human T-acute lymphoblastic leukemia (ALL) (Weng et€al. 2004). Small molecule inhibitors of the gamma-secretase complex (GSI) inhibit NOTCH1 signaling and promote apoptosis and cell cycle arrest in T-ALL leukemia cells (O’Neil et€al. 2006). Palomero et€al. (2006) demonstrated that C-MYC is a critical mediator of NOTCH1 signaling and that NOTCH1 activation appears upstream of C-MYC and in part regulates cell proliferation and growth by directly regulating C-MYC activation. In vitro studies with NOTCH1 signaling inhibitors such as GSIs significantly downregulate C-MYC expression (Palomero et€al. 2006). However, the slow kinetics of C-MYC downregulated by GSI treatment suggest that NOTCH1 probably only regulates a fraction of C-MYC expression, and therefore, GSI-type inhibitors may only partially inhibit C-MYC induced cell proliferation and growth (Table€2) (Palomero et€al. 2006). A third potential approach to target C-MYC in BL is the use of antisense oligodeoxynucleotides specifically designed against specific regions within the C-MYC gene. In previous in€ vitro studies, phosphodiester and phosphorothioate anti-sense oligodeoxynucleotides induced significant inhibition in BL cell growth (Williams et€al. 1997). We further demonstrated that the addition of cationic lipids significantly reduces the time and the dose required of C-MYC antisense oligodeoxynucleotides to inhibit BL in€vitro growth (Williams et€al. 1996). There has been, however, little development of C-MYC antisense oligodeoxynucleotides for a variety of reasons, including different C-MYC point rearrangements in different subtypes of BL, and the difficulty of delivering antisense oligodeoxynucleotides clinically, and more importantly, the uptake of antisense deoxyolidonucleotides in BL cells in€vivo. However, further development is underway for exploring similar approaches using small hairpin or small interference RNAs that also target specific regions of the C-MYC gene (Table€2).
Anaplastic Large Cell Lymphoma ALCL is a distinct subtype of aggressive CD30-positive peripheral T-cell NHL that makes up about 10 to 15% of pediatric NHL (Drexler et€al. 2000; Perkins 2000). The World Health Organization (WHO) classification of lymphomas recognizes two subtypes of ALCL, a primary cutaneous ALCL that forms part of a clinical spectrum of CD30-positive cutaneous T-cell lymphoproliferative disorders and
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
193
systemic ALCL (Greenland et€ al. 2001; Kadin and Carpenter 2003). Systemic ALCLs are further subdivided into two clinically significant subtypes based on expression of anaplastic lymphoma kinase (ALK) (Drexler et€al. 2000; Kadin and Carpenter 2003; Falini 2001). In the pediatric population, primary cutaneous and ALK negative ALCL are very rare (Tomaszewski et€al. 1999). The majority of pediatric ALCL (90 to 95%) express ALK due to specific genetic events that lead to overexpression and constitutive activation of ALK (Kutok and Aster 2002). In about 80% of tumors, ALK overexpression arises due to a t(2;5)(p23;q35) translocation that juxtaposes the ALK gene on chromosome 2p23 to the nucleophosmin (NPM) gene on 5q35. This fusion gene encodes a chimeric, constitutively activated tyrosine kinase, NPMALK, consisting of the N-terminal portion of the NPM fused to the catalytic domain of ALK (Fig.€8). The fusion with NPM results in constitutive activation of the ALK kinase in a deregulated and ectopic manner (Drexler et€al. 2000; Pulford et€al. 2004).
Fig.€8╅ Chromosomal translocation t(2;5)(p23;q35) leads to NPM-ALK. The chromosomal translocation involving 2p35 (ALK) and 5q35 (NPM) results in the fusion gene product, NPM-ALK, which encodes a chimeric tyrosine kinase. The c-terminal portion of ALK tyrosine kinase fuses to the N-terminal portion of NPM, which leads to constitutive expression and activation of the NPMALK oncogenic protein. The fusion protein can be rapidly detected in routine tissue sections using the ALK-1 antibody, which demonstrates the fusion protein in the nucleus and the cytoplasm of tumor cells
194
M.S. Lim and M.S. Cairo
Over 13 alternative fusion partners for ALK have been identified and involve clathrin heavy polypeptide-like gene (CLTC) (Bridge et€al. 2001), nonmuscle tropomyosin 3 gene (TPM3) (Lamant et€al. 1999), nonmuscle tropomyosin 4 (TPM4) (Meech et€ al. 2001), moesin (MSN) (Tort et€ al. 2001), Trk fusion gene (TFG) (Hernandez et€ al. 1999), cysteinyl-tRNA synthetase (CARS) (Cools et€ al. 2002), nonmuscle myosin heavy chain (MYH9) (Lamant et€al. 2003), RAN binding protein (RANBP2) (Ma et€al. 2003), and 5-aminoimidazole-4-carboxamide ribonucleotide formyltransferase/IMP cyclohydrolase gene (ATIC) (Ma et€al. 2000). Most of the partner proteins contain an oligomerization domain that is required for constitutive activation of the ALK kinase (Drexler et€al. 2000; Lamant et€al. 1999; Meech et€ al. 2001). The downstream signaling pathways induced by the variant partner gene products are not known, although gene expression profiling studies comparing TPM3-ALK positive and NPM-ALK positive ALCL indicate overlapping pathways that may be shared (Bohling et€al. 2008). The overexpression of ALK secondary to the creation of a fusion protein is detectable by immunohistochemistry on tissue biopsies, providing a very rapid and effective means to evaluate the presence of ALK translocation in over 90% of cases. This provides a rapid and routine method of identifying tumors that may be amenable to molecular targeted therapies. ALK expression has been strongly correlated with clinical outcome. Patients with ALK-positive ALCLs have a favorable prognosis and favorable clinical features including low age at diagnosis, low stage, low risk IPI staging, and 80 to 90% 5-year survival, although there is an increased incidence of late (2 years or later after therapy) relapse. In contrast, ALK-negative ALCL usually occurs in older adults and is associated with high stage disease and an approximately 40% 5-year survival (Kadin and Carpenter 2003; Falini 2001). Constitutive activation of the oncogenic ALK tyrosine kinase leads to activation of a multiple, diverse downstream signaling cascade involved in cell proliferation, survival, transformation, and anti-apoptosis (Fig.€9). It is believed that these signaling cascades sustain survival and viability of NPM-ALK-positive tumor cells (Drexler et€al. 2000; Pulford et€al. 2001; Pulford et€al. 2004). This enables the application of two general strategies for development of molecularly targeted therapeutic agents in ALCLs. The first strategy is to target the ALK expression or activity. The other is to target any one of several downstream signaling pathways that are activated by ALK. Indeed, most studies to date have focused on targeting downstream signaling pathways, which will be discussed in the subsequent sections. A variety of approaches to ablate the expression or function of the ALK fusion protein have been attempted, including hammerhead ribozyme-mediated cleavage of the NPM-ALK fusion gene product (Hubinger et€al. 2003) and siRNAmediated ablation (Ritter et€al. 2003). Although both hammerhead ribozyme and siRNA effectively decreased the expression and activity of ALK and its mediators, such as AKT and ERK in cell lines (HeLa and 293) cotransfected with ALK, the results were not as significant in cell lines that expressed endogenous NPM-ALK with repeated transfections over 8 days leading to significant reduction of NPMALK protein but without induction of apoptosis. These results are attributed to the long half life of the NPM-ALK protein (48€hours). More recently, Piva et€al. (2006)
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
195
Fig.€9╅ Activation of multiple signaling pathways by NPM-ALK. The NPM-ALK fusion tyrosine kinase recruits a number of adaptor proteins via interactions with phosphorylated tyrosine residues. A complex network of signaling pathways including RAS/MAPK, JAK/STAT and PI3-K/ AKT are activated as a result of the activation of NPM-ALK. These signaling pathways play a critical role in enhancing the survival and proliferation of ALCLs
demonstrated a more effective approach to ablation of ALK by using lentiviral transduction of shRNA, targeting a sequence coding for the catalytic domain of ALK. Lentiviral transduction of human ALCL cells with shRNA against ALK led to G1 cell cycle arrest and apoptosis in€ vitro and tumor growth inhibition and regression in€vivo. These results indicate that sustained ablation of ALK expression is required for efficacious therapy of ALCLs and clinical application of this approach may have limited value at this time as a single modality therapeutic option. Indeed, the downregulation of NPM-ALK by siRNA has been shown to augment the chemosensitivity to chemotherapeutic agents such as doxorubicin (Hsu et€al. 2007). Another strategy involves the use of small molecule inhibitors that may be delivered orally with increased bioavailability and better pharmacokinetics. Clearly, the benefit of this type of agent has been demonstrated by the immense success of imatinib for the treatment of chronic myeloid leukemias with BCR-ABL translocations. Indeed, there is clinical and therapeutic rationale to identify potent, selective, and efficacious inhibitors of all of the 75 to 80 known tyrosine kinases in the human genome. Furthermore, because the ALK tyrosine kinase is not normally expressed in many cell types, inhibition of the cancer-specific tyrosine kinase by small molecule
196
M.S. Lim and M.S. Cairo
inhibitors would be an attractive option for tumor-specific therapy. The development of small molecule kinase inhibitors for ALK has been somewhat slow in gaining enthusiasm by pharmaceutical companies. This is attributed to the low incidence of the disease and thus, low economic potential. However, several groups have recently reported the development of small molecule inhibitors of ALK (Wan et€al. 2006; Galkin et€ al. 2007) that have varying degrees of potency, selectivity and efficacy. One group of quinazoline-type compounds, designated WHI-P131 (4-(4¢hydroxyphenyl)amino-6,7-dimethoxyquinazoline) and WHI-P154 (4-[(3¢-bromo4¢-hydroxyphenyl)amino]-6,7-dimethoxyquinazoline) inhibited proliferation and induced apoptotic cell death of the ALK-positive T-cell lymphoma cells (Marzec et€ al. 2005). These compounds abrogated STAT3 tyrosine phosphorylation in a Jak3-independent manner and directly inhibited the kinase activity of ALK with IC50 ranging from 5 to 10€mM. Another group of compounds developed by Cephalon represent cell-permeable fused pyrrolocarbazole (FP)-derived small molecule inhibitors of ALK. CEP-14083 and CEP-14513 exhibited potent ALK inhibitory activity in both in€ vitro enzymatic assays (IC50â•›<â•›5€ nM) and cell-based assays of ALK tyrosine phosphorylation (IC50â•›<â•›30€nM) (Wan et€al. 2006). Treatment of CEP14513 or CEP-14083 led to cell-cycle G1 arrest, followed by accumulation of subG1 population in ALK-positive ALCL cells, including Sudhl-1 and Karpas-299 cells. The G1 arrest that occurred as a result of ALK inhibition correlated with the inhibition of PI3K/AKT and STAT3 activation. These compounds, however, have unfavorable physical properties, and thus, are not suitable for in€vivo use. Another promising group of compounds, identified through a cellular screen used to search for compounds that are cytotoxic to NPM-ALK transfected cells, is NVP-TAE684 (Galkin et€al. 2007). This compound blocked the growth of ALCL-derived cell lines with IC50 of 2 to 10€nM. Furthermore, it was effective in rapid and sustained inhibition of ALK phosphorylation and its downstream effectors and resulted in cell cycle arrest and cellular apoptosis. More importantly, the compound also inhibited the development of tumors in xenograft models of ALCL. The favorable pharmacokinetic properties of NVP-TAE684 in mice, including high bioavailability, half-life, and tissue distribution, support future efforts in evaluating these small molecule kinase inhibitors for treatment of ALCLs. Activated ALK induces multiple downstream signaling pathways that represent potential targets for therapy. The most well-studied of these are the PI3K/AKT pathway and the JAK3/STAT3 pathway. ALK exerts anti-apoptotic activity via activation of PI3K/AKT pathway, leading to AKT-mediated phosphorylation of Bad proteins and inhibition of Bad pro-apoptotic activity (Bai et€al. 2000). ALK also results in activation of STAT 3 by either direct phosphorylation by ALK or via JAK 3 activation that leads to upregulation of survivin and Bcl-XL, which results in promotion of cell survival (Zamo et€al. 2002; Amin et€al. 2003; Khoury et€al. 2003). More recent evidence indicates that other pathways including FRAP/mTOR (Vega et€ al. 2006; Marzec et€ al. 2007a), MEK/ERK (Marzec et€ al. 2007b), and HSP90 (Bonvini et€al. 2002) are important mediators of ALK signaling and contribute to tumor cell survival in ALCLs. Because of the observed activation of both proliferation and anti-apoptotic functions by ALK, it has been hypothesized that therapeutic
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
197
approaches that block ALK effects on signaling pathways may be effective. This has been attempted in experimental systems by a variety of approaches, including selective inhibition of STAT3 by adenoviral transfection of a mutated STAT3 molecule (Amin et€ al. 2004), small molecular inhibition of JAK3/STAT3 using cucurbitacin I (JSI-124) (Shi et€ al. 2006), and more recently, ablation of STAT3 expression by antisense oligonucleotides (Chiarle et€al. 2005), which impaired the growth of human and mouse NPM-ALK tumors in€vivo. The inhibition of the PI3K by wortmannin and the inhibition of AKT by small molecule inhibitors (Marzec et€al. 2007a) resulted in decreased proliferation and increased apoptosis of ALCLs. In addition, selective inhibition of the mTOR pathway, which is downstream of AKT by rapamycin and mTOR-specific siRNAs resulted in cell cycle arrest and cellular apoptosis (Vega et€al. 2006). A group of proteasome inhibitors, particularly the benzoquinone ansamycin geldanamycin (GA), have also been identified as agents that interfere with signal transduction of HSP90 client proteins that require HSP90 for the correct confirmation and stability. NPM-ALK is a client protein of HSP90. A recent study demonstrated that an analog of GA, 17-AAG, disturbed NPM-ALK/HSP90 complex formation, thereby destabilizing and degrading the NPM-ALK fusion protein (Bonvini et€al. 2004). Our recent study evaluating the proteomic consequences of inhibition of HSP90 chaperone function in ALCL (Schumacher et€al. 2007) indeed demonstrated the efficacy of GA in decreasing the expression of NPM-ALK and inducing cell cycle arrest in a dose and time-dependent manner. GA is currently being evaluated in phase I clinical trials at the National Cancer Institute for advanced epithelial cancer, malignant lymphoma, and sarcoma. Another group of potential therapeutic targets are represented by the adaptor proteins that are recruited to the autoactivated NPM-ALK chimeric fusion protein. p130Cas (Ambrogio et€ al. 2005) and pp60c-src (Cussac et€ al. 2004) represent two adaptor proteins that bind to phosphorylated and active NPM-ALK and mediate cellular transformation. Inhibition of Src-kinase by small molecules (PPI) or siRNA-mediated ablation of pp60c-src resulted in decreased cell viability of ALCLs (Cussac et€al. 2004). Cytokine and growth factor receptors, which are upregulated in ALCLs, represent another group of therapeutic targets. ALCLs express CD30, a cytokine receptor that is a member of the tumor necrosis factor receptor (TNFR) family (Gruss and Dower 1995; Schneider and Hubinger 2002). Expression of CD30 has been associated with increased proliferation and decreased apoptosis via activation of the nuclear factor kB (NFkB) pathway in Hodgkin lymphoma, but to a much lesser degree in ALCL (Horie et€ al. 1999; Mir et€ al. 2000). A soluble form of CD30 (sCD30) has been detected in the serum of patients with ALCL (Gause et€al. 1992; Nadali et€ al. 1995), and is thought to arise from proteolytic cleavage of the cell surface CD30 by a zinc metalloproteinase (Hansen et€al. 1995). Elevated levels of sCD30 at the time of diagnosis in patients with ALCL are associated with high tumor burden, a lower relapse-free survival, and overall survival, suggesting that this may be a specific indicator for a higher risk of treatment failure (Zinzani et€al. 1998). Furthermore, the expression of CD30 is regulated by ALK.
198
M.S. Lim and M.S. Cairo
Because of its effects on cell proliferation and apoptosis, immunomodulatory therapy using antibodies directed at the ligand binding site for CD30 have been suggested as a means to transduce the apoptotic effects of CD30 binding. Two murine monoclonal antibodies directed against the ligand binding region of CD30 have shown anti-tumor effects in animal models (Tian et€al. 1995). A newly developed human monoclonal antibody has shown anti-tumor activity and is currently entering clinical trials (Borchmann et€al. 2003). SGN-30, a chimeric monoclonal antibody that binds to CD30 and inhibits growth of ALCLs, is currently being evaluated in a Phase I/II pilot study in children with CD30-positive recurrent ALCL (Children’s Oncology Group Trial ANHL06P1). CD52, Campath-1 protein, is another target that has been proposed as a possible target for antibody-mediated therapy. Initial studies on small numbers of pediatric ALK-positive ALCL suggest that the majority (70%) of tumors do express CD52 (Perkins et€ al. 2004 (abstract)) and may respond to use of anti-CD52 immunotherapy. Our recent cDNA microarray analysis of ALCL-derived cell lines and tissue biopsies of pediatric ALCLs indicated the overexpression of IL-2 receptor (CD25) relative to reactive T-cells (Fillmore et€al. 2002; Lim and Elenitoba-Johnson 2006). Additional studies demonstrate that CD25 expression is highly correlated with ALK expression in ALCLs (Lim and Elenitoba-Johnson 2006) and is often accompanied by high levels of serum IL-2 receptors (Janik et€al. 2004). This suggests that use of immune-based therapies such as denileukin diftitox or humanized monoclonal anti-CD25 (daclizumab or basiliximab) would provide an alternative therapeutic approach. There are limited reports of this approach in ALK-positive ALCLs. In€ vitro studies from our laboratory indicate that the aberrant overexpression of IL-2 receptor in ALCLs provides a rational target for therapy using denileukin diftitox. Treatment of ALCL-derived cell lines with denileukin diftitox resulted in a dose- and time-dependent decrease in cell viability, which was associated with increased cellular apoptosis (Lim et al. 2006). The authors are currently evaluating the clinical therapeutic utility of denileukin diftitox in a Phase II pilot study in children, adolescents, and young adults with relapsed or refractory ALCL.
Lymphoblastic Lymphoma LLs represent genetically and clinically heterogeneous neoplasms of immature or precursor lymphoid cells that comprise approximately 30% of the NHLs that occur in children and young adults. The majority of LLs arise from immature T-cells (Crist et€ al. 1988; Uckun et€ al. 1997), while fewer than 10% of LLs are of B-precursor cell origin. The predominant sites of disease in T-LL are the anterior mediastinum and supradiaphragmatic lymph nodes, whereas B-precursor LL is more commonly localized to peripheral lymph nodes and extranodal sites such as skin, soft tissues, and bone, with a predilection for the head and neck regions (Soslow et€al. 1999; Lin et€al. 2000; Neth et€al. 2000; Maitra et€al. 2001). A higher
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
199
incidence of bone marrow and central nervous system involvement is seen in T-LL compared to B-precursor LL. Precursor T-LLs are characterized by a cortical thymocyte immunophenotype. Whereas most T-cell leukemias arise from earlier stages of intrathymic T-cell differentiation (stage I prothymocytes), the majority of T-cell LLs arise from intermediate (stage II) thymocytes (Crist et€al. 1988). Precursor B-LLs most often display an early pre-B or pre B-phenotype (CD19, CD10, and TdT with variable CD20, CD22, HLA-DR and cytoplasmic immunoglobulin expression) (Perkins 2000; Maitra et€al. 2001). ALL and LL are thought to represent a spectrum of a single disease entity, which is termed T (or B) lymphoblastic leukemia/lymphoma in the REAL (Harris et€al. 1994) and WHO (Jaffe 2001) classification schemes. By convention, bone marrow involvement by the tumor must be <25% in LL. When the extent of marrow involvement is >25%, the disease is classified as ALL. Clinically, both B-precursor and T-cell LLs are effectively treated using ALLbased therapies, which include multidrug systemic chemotherapy and central nervous system prophylaxis (Reiter et€al. 2000; Hoelzer et€al. 2002; Goldberg et€al. 2003). Long-term survival rates of 85 to 90% for localized disease and 65 to 85% for advanced stage disease have been achieved, paralleling those seen in ALL (Goldberg et€al. 2003; Link et€al. 1997; Patte et€al. 2001). Although not observed in all studies, evidence suggests that early consolidation with high-dose methotrexate may also be beneficial in both adult and pediatric patients with T-LL (Goldberg et€al. 2003; Asselin 2001; Thomas et€al. 2004). LL and ALL are characterized by many of the same underlying cytogenetic abnormalities, which occur with a frequency of 50 to 90% (Thomas and Kantarjian 2001). As a part of developing antigen recognition capabilities, T and B-lymphocytes rearrange a series of V (variable), D (diversity), J (joining), and C (constant) genetic regions to produce unique antigen receptors. The antigen binding domains of T-cell receptors (TCR) and immunoglobulins (Ig) are unique and serve as clonal markers. Chromosomal translocations are the most frequent cytogenetic abnormalities that occur in T-LL (Table€3). These chromosomal translocations frequently juxtapose promoter and enhancer elements from TCR genes next to transcription factors genes such as HOX11, TAL1, and LYL1, which are normally transcriptionally silent in normal T-lymphocytes (Sandlund et€ al. 1996; Goldsby and Carroll 1998; Heerema et€ al. 1998; Ferrando et€ al. 2002).The most common translocations in T-LL involve TCRad (14q11), TCRb (7q32-36), and TCRg (7p15), with the more immature cellular phenotypes favoring the TCRg or TCRd loci (Table€3). While the prognostic significance of these cytogenetic features is unknown in LL, many of these same cytogenetic findings have not been shown to have prognostic significance in pediatric T-ALL (Heerema et€al. 1998). Recent developments in molecular cytogenetics including fluorescent in situ hybridization and mutation analysis revealed cytogenetically cryptic aberrations in most cases of T-ALL. Gene expression profiling studies using DNA microarrays have revealed that T-ALLs recapitulate specific stages of thymocyte maturation and have also identified potential targets for molecular therapies. Furthermore, there is support for the accumulation of multiple genetic abnormalities in the pathogenesis
200
M.S. Lim and M.S. Cairo
Table€3╅ Cytogenetic and molecular changes in lymphoblastic lymphomas Involved Function of fusion gene or gene(s) Protein(s) expressed oncogene Translocations involving TCR genes t(7;10)(q34;q24) and TLX1 t(10;14)(q24;q11) (HOX11) t(1;14)(p32;q11) and t(1;7)(p32;q34) t(7;9)(q34;q32) t(7;19)(q34;p13) t(11;14)(p15;q11) t(11;14)(p13;q11) and t(7;11)(q35;p13) t(1;7)(p34;q34) Formation of fusion genes t(10;11)(p13;q14) (often cryptic)
Transcription factor
TAL1
Class II homeodomaincontaining HLH type II
Transcription factor
TAL2 LYL1 LMO1 LMO2
HLH type II HLH type II LIM-only domain LIM-only domain
Transcription factor Transcription factor Protein–protein interaction Protein–protein interaction
LCK
SRC family of tyrosine kinase
Signal transduction
CALM/
ENTH motif containing Zinc fingers/ leucine zippercontaining Homeodomain protein HLH type II
AF10
t(1;9)(q23;p13)
PBX1
21q addition
E2A Unknown
Transcription factor
Transcription factor Transcription factor
TCR T-cell receptors, ENTH epsin N-terminal homology, HLH helix-loop-helix
of T-ALL/LL. Gene expression studies performed to identify global expression signatures associated with T-ALL (Ferrando et€al. 2002) have identified major oncogenic pathways associated with HOX11, TAL1, and LYL1 activation and help define molecular subtypes within the morphologically homogeneous disease (Fig.€ 10) (Ferrando and Look 2003). These seminal observations indicated the overexpression of key oncogenes in the majority of T-ALL and occurred frequently in the absence of translocations suggesting alternate mechanism of oncogene activation. Distinct molecular subtypes with varying treatment outcomes were identified based on the microarray analyses. Those with HOX11 expression were associated with a 92â•›±â•›8% 5-year survival, whereas those with TAL1 and LYL1 exhibited a much lower survival rate at 43â•›±â•›19% and 33â•›±â•›19%, respectively. Furthermore, subtypes with specific oncogene activation were associated with specific states of thymocyte differentiation. Those with expression of HOX11 tended to arise from earlier cortical thymocytes, whereas those associated with TAL1 expression possessed mature, late cortical thymocyte differentiation (Fig.€11) (Ferrando and Look 2003). To elucidate the genetic basis for T-LL, Raetz and coworkers (2006) compared the global gene expression profiles of T-ALL and T-LL using an Affymetrix U133A
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma Table€4╅ Molecular targets for ALCL and LL Gene product or Subtype pathway Cellular function ALCL ALK Tyrosine kinase
201
Inhibitors/modifier
NVP-TAE684 CEP-11988 WHI-131, WHI-154 Wortmannin, LY PI3K/AKT Protein kinase ShRNA STAT3 Transcription factor AG490, Curcumin JAK3 Tyrosine kinase 17-AAG and Geldanamycin HSP90 Chaperone PPI pp60SRC Protein kinase Rapamycin, CCI-779 Protein kinase mTOR U01206 Protein kinase MAPK/ERK Denileukin diftitox, Growth cytokine receptor IL-2R Daclizumab SGN-30 Cytokine receptor CD30 Compound E Protooncogene LL NOTCH Benozochazepine (Bz-423) Protooncogene C-MYC Rapamycin, CCI-779 Protein kinase mTOR ALCL anaplastic large cell lymphoma, LL lymphoblastic lymphoma, ALK anaplastic lymphoma kinase, PI3K phosphatidylinositol-3-kinase, STAT3 signal transducer and activator of transcription 3, JAK3 janus kinase 3, HSP90 heat shock protein 90, mTOR mammalian target of rapamycin, MAPK/ERK mitogen-activated protein kinase/extracellular signal-regulated kinase, IL-2R interleukin-2 receptor, ShRNA small hairpin RNA, 17-AAG 17-N-allylamino-17-demethoxygeldanamycin, PPI protein–protein interaction, CCI-779 cell cycle inhibitor-779, SGN-30 Seattle Genetics’ anti-CD30 monoclonal antibody
GeneChip composed of over 22,283 probe sets. Unsupervised hierarchical clustering of ten T-ALL bone marrow samples and nine T-LL tissue samples led to the complete segregation of T-ALL and T-LL into distinct groups. Furthermore, they identified over 200 genes, which were able to differentiate between the diseases using significance analysis of microarrays. The results indicate the clear differences in the gene expression profiles between T-ALL and T-LL, suggesting the underlying differences in the biology of the two entities. Future studies in the identification of downstream target genes of the commonly overexpressed oncogenes will lead to identification of potential targets for therapy. The main categories of functional molecular abnormalities identified in LL and those that represent potential targets for therapy are discussed below. Defects in cell cycle control due to genetic loss of INK4 locus (p16/p14) and aberration of Rb1 and p53 are common in T-LL (Okuda et€al. 1995; Cayuela et€al. 1996). In addition, gene rearrangements can result in deregulated expression of transcription factors that lead to aberrant cell cycle control. For example, CALM-AF10 T-ALLs overexpress BMI1 that leads to suppression of p16. Another frequent underlying genetic event in T-LL is dysregulation of TAL1 (1p32), which is estimated to occur with a frequency of 30% (Baer 1993; Bash et€al. 1995). Tumor-specific alterations in TAL1 occur by distinctive mechanisms. In a small percentage of cases, TAL1 is juxtaposed next to
202
M.S. Lim and M.S. Cairo
Fig.€10╅ Hierarchical cluster analysis of T-ALL gene expression data. The tree (top) shows the relatedness of gene expression among samples and is color-coded according to the expression of three major T-ALL transcription factor oncogenes, TAL1, HOX11, and LYL1. Each column represents a T-ALL mRNA sample and each row one of the genes in the microarray, selected based on a permutation test for genes best distinguishing among HOX11+, TAL1+, LYL1+, and unclassified samples (Other). Gene expression values are color-coded, as indicated by the scale beneath the graph. HOX11+, TAL1+, and LYL1+ cases are grouped in major branches of the dendrogram, while the unclassified cases are split into two groups, one related to HOX11+ cases (HOX11-like, branch H2) and the other related to TAL1+ cases (TAL1-like, branch T2). Three cases in the H2 cluster express high levels of HOX11L2, a HOX11-related gene, demonstrating the ability of the hierarchical analysis to cluster together cases with related mechanisms of transformation. Three cases expressing the MLL-ENL fusion oncogene were clustered together based on the gene expression signature as a major independent group (branch M). Numerous cytogenetic and molecular abnormalities (discrete symbols at the bottom of the tree) are found in cases that were clustered by the analysis, supporting the hypothesis that the gene expression profiles define major subgroups of T-ALL that arise from different multistep molecular pathogenetic pathways. From Ferrando and Look 2003. Copyright Elsevier, used with permission
TCR promoter/enhancer sequences as the result of t(1;14) or t(1;7) chromosomal translocations. More commonly, deletions of a regulatory region of TAL1 are observed that position TAL1 next to the promoter of the active SIL gene (Brown et€al. 1990; Janssen et€al. 1993). In both cases, deregulated expression of the full-length TAL1 gene occurs. TAL1 is a basic helix-loop-helix DNA binding protein that heterodimerizes with E2A proteins (E47 and E12), regulating transcription of target
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
MLL-ENL LYL1+
γδ CD34, BCL2 CD7, IL7R
HOX11+/ HOX11L2+
TAL1+
203 CD8+ CD4-
TCRD TCRG
αβ
CD4- CD8+
Double Negative Thymocytes
CD10, CD1A CD1B, CD1C CD45
CD6, LCK CD3D, CD3E TCRB, TCRA
Late Cortical
Early Cortical CD4+ CD8+
CD8CD4+
Mature Single Positive T-cells
Double Positive Thymocytes Fig.€ 11╅ Correspondence of gene expression signatures of LYL1+, MLL-ENL+, HOX11+ and TAL1+ T-ALL samples with recognized stages of thymocyte differentiation. LYL1+ cases have a gene expression signature corresponding to that of the most immature normal T-cell precursors (CD4/CD8 double-negative cells), which express CD34 but not CD4, CD8, or CD3. Normal thymocytes can become committed to either the gamma-delta or the alpha-beta lineages of T-cell maturation. MLL-ENL+ T-ALL appears to correspond to thymocytes with aberrant differentiation along the gamma-delta lineage. As thymocytes within the alpha-beta lineage mature, they lose CD34 expression, while gaining CD4 and then CD8, eventually becoming double-positive thymocytes. HOX11+ cases have gene expression signatures corresponding to early double-positive cells that express CD1 and CD10 (early cortical thymocytes). TAL1+ corresponds to thymocytes that express both T-cell receptor RNAs and CD3 (late cortical thymocytes). Normal thymocytes that have undergone both positive and negative selection can then proceed through a final step of differentiation, in which they downregulate the expression of either CD4 or CD8 to become mature single-positive T cells, ready for export from the thymus. From Ferrando and Look 2003. Copyright Elsevier, used with permission
genes involved in cell growth and differentiation. TAL1 is normally expressed in megakaryocyte, mastocyte, and eyrthroid lineages but not in normal T-cells, and has been shown to have critical functional roles in early hematopoiesis (Delabesse et€al. 1997; Hall et€ al. 2003). Recent studies have shown that TAL1, similar to PMLRARa and AML-1/ETO, contributes to leukemogenesis by repressing gene expression and inducing a differentiation arrest in developing thymocytes (O’Neil et€ al. 2004). TAL1 overexpression negatively regulates the p16 gene through interference with the E boxes sequences of the p16 promoter (Hansson et€al. 2003). E2A, which is deregulated as a result of t(1;19)(q23;p13) in B-LL, positively regulates several cyclin-dependent kinase inhibitors’ promoters and negatively affects cell growth. Furthermore, HOX11 can act on the cell cycle by its interactions of protein serine/ threonine phosphatases such as PP2A and PP1 (Kawabe et€al. 1997). Although the use of inhibitors of transcription factors has not been well developed, cell cycle regulators can provide targets for therapy (Fahraeus et€al. 1996). Mutations in several key oncogenes have been identified in T-ALL. Gain of function NOTCH1 mutations have been recently identified as one of the most
204
M.S. Lim and M.S. Cairo
Â� common genetic abnormalities in T-ALL, occurring in more than 50% of cases (Weng et€ al. 2004). NOTCH1 encodes a transmembrane receptor that regulates normal T-cell development. Activated NOTCH1 has been shown to potently induce T-ALL in murine models, and in a recent analysis of 96 diagnostic bone marrow samples from children with T-ALL, at least one activating mutation in the extracellular heterodimerization domain and/or the C-terminal PEST domain of NOTCH1 was present in 54 cases (56.2%). NOTCH1 mutations were observed in all molecularly defined subtypes of T-ALL, characterized by HOX11, TAL1, and LYL1 activation. Furthermore, recent studies using expression-profiling screens have led to the identification of C-MYC as a direct target of NOTCH in NOTCH-dependent T-ALL cell lines. Inhibitors of C-MYC interfered with the growth promoting effects of activated NOTCH1. C-MYC appears to be an important developmentally regulated mediator of I signaling in certain stages of thymocyte development. The discovery of aberrant NOTCH signaling highlights the potential use of gamma secretase inhibitors that interfere with NOTCH signaling as a potential therapy for T-LL. Furthermore, immunomodulatory agents such as benzodiazepine Bx-423, which target C-MYC protein for rapid and specific degradation (Sundberg et€ al. 2006), may have a role in therapy of T-LL. A number of tyrosine kinase genes are activated as a result of chromosomal translocations or by gene mutation. Although rare, activating mutations (internal tandem duplications or point mutations) of FLT3, a receptor tyrosine kinase important in the development of hematopoietic stem cells, occur in T-ALL, with very immature phenotype expressing LYL1, LMO2, and the KIT receptor (Paietta et€al. 2004). Similarly activating mutations of LCK and ABL1 interfere with preTCR and TCR signaling and provide a survival and proliferative advantage to T-ALL. ABL1 is a ubiquitously expressed cytoplasmic tyrosine kinase which plays a role in TCR signaling and is involved in the NUP214-ABl1 fusion found in approximately 6% of T-ALL. This fusion gene is expressed in amplified episomes, which are not cytogenetically visible. The resultant constitutive activation of ABL1 leads to activation of survival and proliferation pathways, which are sensitive to imatinib (Griesinger et€ al. 2002). Although they are rare and occur in only 10 to 20% of T-ALL, they represent a very viable mode of targeted therapy using selective small molecule kinase inhibitors that are effective and improve bioavailability and pharmacokinetic properties. Future studies will determine whether T-LLs harbor activating mutations of FLT3, ABL1, LCK, and NOTCH1. Another downstream target of NOTCH activation identified by an innovative approach, utilizing protein microarrays composed of 108 epitopes representing 82 phosphorylated signaling proteins (Chan et€ al. 2007), is the mTOR pathway. Simultaneous blockade of the mTOR pathway by rapamycin and the NOTCH pathway using GSI resulted in synergistic suppression of T-ALL growth. Similarly, the activation of phosphatidylinositol 3-kinase (PI3-K) (Barata et€al. 2004), Akt (Sade et€ al. 2004), extracellular signal regulated kinase ½ (Talora et€ al. 2003) STAT5 (Kelly et€ al. 2003b), and nuclear factor kB (Bellavia et€ al. 2000) in T-ALL, all represent targets for combinatorial therapeutic targeting as a novel therapeutic approach for T-LL.
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
205
Cytogenetic and molecular abnormalities in B-LL are less well-characterized. Classical chromosomal translocations which occur in B-precursor ALL, such as hyperdiploidy, t(12;21), t(1;19) and t(9;22), appear to occur less frequently in B-precursor LL (Maitra et€ al. 2001; Head and Behm 1995). Reported cytogenetic abnormalities in B-precursor LL include additional material from the 21q locus (Maitra et€al. 2001). As expected, precursor B-LLs will usually show monoclonal Ig gene rearrangements and lack evidence of somatic hypermutation (Hojo et€al. 2001).
Summary and Future Directions The prognosis for childhood and adolescent NHL is excellent to outstanding with multi-agent chemotherapy without concomitant surgery and radiotherapy. The major future challenges are mainly to discover the molecular and cellular pathogenesis of malignant transformation (lymphomagenesis) to develop strategies of prevention, early identification, and more selective targeted therapies. A significant increase in new knowledge about the genetic basis for malignant transformation and prognosis has been identified in the last 10 years in adult DLBCL. It remains to be determined whether childhood and adolescent DLBCL has similar or disparate genetic and cellular mechanisms as adult DLBCL. While we have commonly treated BL and DLBCL under similar childhood chemotherapy regimens, biologically the two histological subtypes are dramatically different and further investigations will be required to identify the exact genetic differences that may lead to different therapeutic approaches. Primary mediastinal large B-cell lymphoma is likely a distinct genetic entity from other forms of DLBCL and may genetically resemble HD more than DLBCL. These childhood and adolescent primary mediastinal B-large cell lymphomas require more intense genetic and cellular investigations to help determine the correct therapeutic approach. Significant progress has been made in the identification of molecular targets for therapy in ALCLs. The challenges for the future of molecular targeted therapeutics in ALCLs lie in deciphering the complexity and diversity of the signaling pathways that are activated as a result of the NPM-ALK overexpression. Determining which of the multiple overlapping and redundant pathways are critical in promoting cell survival, proliferation, adhesion, and invasive phenotype is not clear and remains technically challenging with the model systems available. It would seem logical that inhibition of multiple pathways with either nonselective small molecular inhibitors or multiple inhibitors that target multiple proteins would represent reasonable approaches. Enhanced chemosensitivity that is observed with many signal transduction inhibitors in combination with routine chemotherapeutic agents would provide another alternative approach. On a socio-economic level, obtaining financial support to carry out clinical trials on a disease group that is considered an “orphan” or “rare” disease is another significant challenge and one that would benefit from international multi-institutional collaborative efforts and support from the Children’s Oncology Group.
206
M.S. Lim and M.S. Cairo
There is still a paucity of information on the genetic basis of LL-associated lymphomagenesis. Do the same genetic and cellular events that regulate T-ALL occur in T-LL? Are the transcription factors identified in the leukemogenesis process of T-ALL similar or disparate in T-LL? Few targets have been identified in T-LL compared to B-NHL and ALCL. New target identification in T-LL will facilitate the development of selective targeted therapy and hopefully reduce the morbidity currently associated with 2 years of prolonged therapy in T-LL. The integration of new agents into treatment approaches is needed to improve outcome for patients with a poor prognosis. Moreover, it is anticipated that future development of agents specifically targeted to underlying transforming pathways may replace more toxic conventional chemotherapeutic drugs. Acknowledgmentsâ•… The authors would like to thank Erin Morris for her assistance in the development of this manuscript and Sean Park for his preparation of the ALCL figures. Supported in part by grants from the Pediatric Cancer Research Foundation, Andrew Gargiso Foundation, Sonia Scaramella Fund and National Institutes of Health.
References Cairo, M. S., Raetz, E. and Perkins, S. L. 2005. Non-Hodgkin’s lymphoma in children. In Cancer Medicine, ed. D. W. Kufe, R. C. Bast, W. N. Hait, W. K. Hong, R. Pollock, R. R. Weichselbaum, T. Gansler, J. F. Holland & E. Frei, pp. 1962–1976. Ontario: BC Decker Pinkerton, R. 2005. Continuing challenges in childhood non-Hodgkin’s lymphoma. Br J Haematol 130:480–488. Cairo, M. S., Gerrard, M., Sposto, R., et€al. 2007. Results of a randomized international study of high-risk central nervous system B non-Hodgkin lymphoma and B acute lymphoblastic leukemia in children and adolescents. Blood 109:2736–2743. Patte, C., Auperin, A., Gerrard, M., et€ al. 2007. Results of the randomized international FAB/ LMB96 trial for intermediate risk B-cell non-Hodgkin lymphoma in children and adolescents: it is possible to reduce treatment for the early responding patients. Blood 109:2773–2780. Rosenwald, A., Wright, G., Chan, W. C., et€al. 2002. The use of molecular profiling to predict survival after chemotherapy for diffuse large-B-cell lymphoma. N Engl J Med 346:1937–1947. Miles, R. R., Cairo, M. S., Satwani, P., Zwick, D. L., et€al. 2007. Immunophenotypic identification of possible therapeutic targets in pediatric non-Hodgkin lymphomas: A Children’s Oncology Group report. Br J Haematol 138:506–512 Rosenwald, A., Wright, G., Leroy, K., et€al. 2003. Molecular diagnosis of primary mediastinal B cell lymphoma identifies a clinically favorable subgroup of diffuse large B cell lymphoma related to Hodgkin lymphoma. J Exp Med 198:851–862. Savage, K. J., Monti, S., Kutok, J. L., et€al. 2003. The molecular signature of mediastinal large B-cell lymphoma differs from that of other diffuse large B-cell lymphomas and shares features with classical Hodgkin lymphoma. Blood 102:3871–3879. Abramson, J. S. and Shipp, M. A. 2005. Advances in the biology and therapy of diffuse large B-cell lymphoma: moving toward a molecularly targeted approach. Blood 106:1164–1174. Ye, B. H., Rao, P. H., Chaganti, R. S., et€al. 1993. Cloning of bcl-6, the locus involved in chromosome translocations affecting band 3q27 in B-cell lymphoma. Cancer Res 53:2732–2735. Lo Coco, F., Ye, B. H., Lista, F., et€al. 1994. Rearrangements of the BCL6 gene in diffuse large cell non-Hodgkin’s lymphoma. Blood 83:1757–1759.
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
207
Chang, C. C., Ye, B. H., Chaganti, R. S., et€ al. 1996. BCL-6, a POZ/zinc-finger protein, is a sequence-specific transcriptional repressor. Proc Natl Acad Sci U S A 93:6947–6952. Cattoretti, G., Pasqualucci, L., Ballon, G., et€al. 2005. Deregulated BCL6 expression recapitulates the pathogenesis of human diffuse large B cell lymphomas in mice. Cancer Cell 7:445–455. Allman, D., Jain, A., Dent, A., et€ al. 1996. BCL-6 expression during B-cell activation. Blood 87:5257–5268. Migliazza, A., Martinotti, S., Chen, W., et€ al. 1995. Frequent somatic hypermutation of the 5¢ noncoding region of the BCL6 gene in B-cell lymphoma. Proc Natl Acad Sci U S A 92:12520–12524. Pasqualucci, L., Migliazza, A., Fracchiolla, N., et€al. 1998. BCL-6 mutations in normal germinal center B cells: evidence of somatic hypermutation acting outside Ig loci. Proc Natl Acad Sci U S A 95:11816–11821. Ye, B. H., Chaganti, S., Chang, C. C., et€al. 1995. Chromosomal translocations cause deregulated BCL6 expression by promoter substitution in B cell lymphoma. Embo J 14:6209–6217. Harris, M. B., Chang, C. C., Berton, M. T., et€ al. 1999. Transcriptional repression of Stat6dependent interleukin-4-induced genes by BCL-6: specific regulation of iepsilon transcription and immunoglobulin E switching. Mol Cell Biol 19:7264–7275. Baron, B. W., Anastasi, J., Thirman, M. J., et€ al. 2002. The human programmed cell death-2 (PDCD2) gene is a target of BCL6 repression: implications for a role of BCL6 in the downregulation of apoptosis. Proc Natl Acad Sci U S A 99:2860–2865. Niu, H., Cattoretti, G. and Dalla-Favera, R. 2003. BCL6 controls the expression of the B7-1/CD80 costimulatory receptor in germinal center B cells. J Exp Med 198:211–221. Miles, R. R., Crockett, D. K., Lim, M. S., et€al. 2005. Analysis of BCL6-interacting proteins by tandem mass spectrometry. Mol Cell Proteomics 4:1898–1909. O’Connor, O. A. 2005. Targeting histones and proteasomes: new strategies for the treatment of lymphoma. J Clin Oncol 23:6429–6436. Bereshchenko, O. R., Gu, W. and Dalla-Favera, R. 2002. Acetylation inactivates the transcriptional repressor BCL6. Nat Genet 32:606–613. Piekarz, R. L., Robey, R., Sandor, V., et€al. 2001. Inhibitor of histone deacetylation, depsipeptide (FR901228), in the treatment of peripheral and cutaneous T-cell lymphoma: a case report. Blood 98:2865–2868. Kelly, W. K., Richon, V. M., O’Connor, O., et€al. 2003a. Phase I clinical trial of histone deacetylase inhibitor: suberoylanilide hydroxamic acid administered intravenously. Clin Cancer Res 9:3578–3588. Prince, H. M., George, D., Patnaik, A., et€al. 2007. Phase 1 study of oral LBH589, a novel deacetylase (DAC) inhibitor in advanced solid tumors and non-Hodgkin’s lymphoma (abstract). J Clin Oncol 25:3500. Fouladi, M., Furman, W. L., Chin, T., et€ al. 2006. Phase I study of depsipeptide in pediatric patients with refractory solid tumors: a Children’s Oncology Group report. J Clin Oncol 24:3678–3685. Fouladi, M., Park, J., Sun, J., et€al. 2007. A Phase 1 trial of vorinostat in children with refractory solid tumors: A children’s oncology group study (abstract). J Clin Oncol 25:9569. Polo, J. M., Dell’Oso, T., Ranuncolo, S. M., et€ al. 2004. Specific peptide interference reveals BCL6 transcriptional and oncogenic mechanisms in B-cell lymphoma cells. Nat Med 10:1329–1335. Miles, R., Raphael, M., McCarthy, K., et€al. 2005. Diffuse large B-cell lymphomas in pediatric patients demonstrate a marked predominance of germinal center cell phenotype (abstract). Ann Oncol 16:v61. Ngo, V. N., Davis, R. E., Lamy, L., et€al. 2006. A loss-of-function RNA interference screen for molecular targets in cancer. Nature 441:106–110. Thome, M. 2004. CARMA1, BCL-10 and MALT1 in lymphocyte development and activation. Nat Rev Immunol 4:348–359. Karin, M. 2006. Nuclear factor-kappaB in cancer development and progression. Nature 441:431–436.
208
M.S. Lim and M.S. Cairo
Van Waes, C. 2007. Nuclear factor-kappaB in development, prevention, and therapy of cancer. Clin Cancer Res 13:1076–1082. Lam, L. T., Davis, R. E., Pierce, J., et€al. 2005. Small molecule inhibitors of IkappaB kinase are selectively toxic for subgroups of diffuse large B-cell lymphoma defined by gene expression profiling. Clin Cancer Res 11:28–40. Voorhees, P. M. and Orlowski, R. Z. 2006. The proteasome and proteasome inhibitors in cancer therapy. Annu Rev Pharmacol Toxicol 46:189–213. Adams, J., Palombella, V. J., Sausville, E. A., et€al. 1999. Proteasome inhibitors: a novel class of potent and effective antitumor agents. Cancer Res 59:2615–2622. Traenckner, E. B., Wilk, S. and Baeuerle, P. A. 1994. A proteasome inhibitor prevents activation of NF-kappa B and stabilizes a newly phosphorylated form of I kappa B-alpha that is still bound to NF-kappa B. Embo J 13:5433–5441. Goy, A., Younes, A., McLaughlin, P., et€al. 2005. Phase II study of proteasome inhibitor bortezomib in relapsed or refractory B-cell non-Hodgkin’s lymphoma. J Clin Oncol 23:667–675. O’Connor, O. A., Wright, J., Moskowitz, C., et€ al. 2005. Phase II clinical experience with the novel proteasome inhibitor bortezomib in patients with indolent non-Hodgkin’s lymphoma and mantle cell lymphoma. J Clin Oncol 23:676–684. Mounier, N., Ribrag, V., Haioun, C., et€ al. 2007. Efficacy and toxicity of two schedules of R-CHOP plus bortezomib in front-line B lymphoma patients: A randomized phase II trial from the Groupe d’Etude des Lymphomes de l’Adulte (GELA) (abstract). J Clin Oncol 25:8010. Jazirehi, A. R., Huerta-Yepez, S., Cheng, G., et€al. 2005. Rituximab (chimeric anti-CD20 monoclonal antibody) inhibits the constitutive nuclear factor-{kappa}B signaling pathway in nonHodgkin’s lymphoma B-cell lines: role in sensitization to chemotherapeutic drug-induced apoptosis. Cancer Res 65:264–276. Turner, C. A., Jr., Mack, D. H. and Davis, M. M. 1994. Blimp-1, a novel zinc finger-containing protein that can drive the maturation of B lymphocytes into immunoglobulin-secreting cells. Cell 77:297–306. Angelin-Duclos, C., Cattoretti, G., Lin, K. I., et€ al. 2000. Commitment of B lymphocytes to a plasma cell fate is associated with Blimp-1 expression in€vivo. J Immunol 165:5462–5471. Pasqualucci, L., Compagno, M., Houldsworth, J., et€al. 2006. Inactivation of the PRDM1/BLIMP1 gene in diffuse large B cell lymphoma. J Exp Med 203:311–317. Feuerhake, F., Kutok, J. L., Monti, S., et€al. 2005. NFkappaB activity, function, and target-gene signatures in primary mediastinal large B-cell lymphoma and diffuse large B-cell lymphoma subtypes. Blood 106:1392–1399. Cairo, M. S., Sposto, R., Perkins, S. L., et€al. 2003. Burkitt’s and Burkitt-like lymphoma in children and adolescents: a review of the Children’s Cancer Group experience. Br J Haematol 120:660–670. Poirel, H., Heerema, N., Swansbury, J., et€al. 2003. Prognostic value of recurrent chromosomal alterations in pediatric B-cell non-Hodgkin lymphoma (NHL): A report of 238 cases from the international FAB/LMB96 study (abstract). Blood 102:1420. Dave, S. S., Fu, K., Wright, G. W., et€al. 2006. Molecular diagnosis of Burkitt’s lymphoma. N Engl J Med 354:2431–2442. Basso, K., Margolin, A. A., Stolovitzky, G., et€al. 2005. Reverse engineering of regulatory networks in human B cells. Nat Genet 37:382–390. Blatt, N. B., Bednarski, J. J., Warner, R. E., et€al. 2002. Benzodiazepine-induced superoxide signals B cell apoptosis: mechanistic insight and potential therapeutic utility. J Clin Invest 110:1123–1132. Boitano, A., Ellman, J. A., Glick, G. D., et€ al. 2003. The proapoptotic benzodiazepine Bz-423 affects the growth and survival of malignant B cells. Cancer Res 63:6870–6876. Bednarski, J. J., Warner, R. E., Rao, T., et€al. 2003. Attenuation of autoimmune disease in Fasdeficient mice by treatment with a cytotoxic benzodiazepine. Arthritis Rheum 48:757–766. Sundberg, T. B., Ney, G. M., Subramanian, C., et€al. 2006. The immunomodulatory benzodiazepine Bz-423 inhibits B-cell proliferation by targeting C-MYC protein for rapid and specific degradation. Cancer Res 66:1775–1782. Allman, D., Aster, J. C. and Pear, W. S. 2002. Notch signaling in hematopoiesis and early lymphocyte development. Immunol Rev 187:75–86.
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
209
Weng, A. P., Ferrando, A. A., Lee, W., et€al. 2004. Activating mutations of NOTCH1 in human T cell acute lymphoblastic leukemia. Science 306:269–271. O’Neil, J., Calvo, J., McKenna, K., et€al. 2006. Activating Notch1 mutations in mouse models of T-ALL. Blood 107:781–785. Palomero, T., Lim, W. K., Odom, D. T., et€ al. 2006. NOTCH1 directly regulates C-MYC and activates a feed-forward-loop transcriptional network promoting leukemic cell growth. Proc Natl Acad Sci U S A 103:18261–18266. Williams, S. A., Gillan, E. R., Knoppel, E., et€al. 1997. Effects of phosphodiester and phosphorothioate antisense oligodeoxynucleotides on cell lines which overexpress C-MYC: implications for the treatment of Burkitt’s lymphoma. Ann Oncol 8 Suppl 1:25–30. Williams, S. A., Chang, L., Buzby, J. S., et€ al. 1996. Cationic lipids reduce time and dose of C-MYC antisense oligodeoxynucleotides required to specifically inhibit Burkitt’s lymphoma cell growth. Leukemia 10:1980–1989. Drexler, H. G., Gignac, S. M., von Wasielewski, R., et€al. 2000. Pathobiology of NPM-ALK and variant fusion genes in anaplastic large cell lymphoma and other lymphomas. Leukemia 14:1533–1559. Perkins, S. L. 2000. Work-up and diagnosis of pediatric non-Hodgkin’s lymphomas. Pediatr Dev Pathol 3:374–390. Greenland, C., Dastugue, N., Touriol, C., et€al. 2001. Anaplastic large cell lymphoma with the t(2;5) (p23;q35) NPM/ALK chromosomal translocation and duplication of the short arm of the nontranslocated chromosome 2 involving the full length of the ALK gene. J Clin Pathol 54:152–154. Kadin, M. E. and Carpenter, C. 2003. Systemic and primary cutaneous anaplastic large cell lymphomas. Semin Hematol 40:244–256. Falini, B. 2001. Anaplastic large cell lymphoma: pathological, molecular and clinical features. Br J Haematol 114:741–760. Tomaszewski, M. M., Moad, J. C. and Lupton, G. P. 1999. Primary cutaneous Ki-1(CD30) positive anaplastic large cell lymphoma in childhood. J Am Acad Dermatol 40:857–861. Kutok, J. L. and Aster, J. C. 2002. Molecular biology of anaplastic lymphoma kinase-positive anaplastic large-cell lymphoma. J Clin Oncol 20:3691–3702. Pulford, K., Morris, S. W. and Turturro, F. 2004. Anaplastic lymphoma kinase proteins in growth control and cancer. J Cell Physiol 199:330–358. Bridge, J. A., Kanamori, M., Ma, Z., et€al. 2001. Fusion of the ALK gene to the clathrin heavy chain gene, CLTC, in inflammatory myofibroblastic tumor. Am J Pathol 159:411–415. Lamant, L., Dastugue, N., Pulford, K., et€al. 1999. A new fusion gene TPM3-ALK in anaplastic large cell lymphoma created by a (1;2)(q25;p23) translocation. Blood 93:3088–3095. Meech, S. J., McGavran, L., Odom, L. F., et€al. 2001. Unusual childhood extramedullary hematologic malignancy with natural killer cell properties that contains tropomyosin 4 – anaplastic lymphoma kinase gene fusion. Blood 98:1209–1216. Tort, F., Pinyol, M., Pulford, K., et€al. 2001. Molecular characterization of a new ALK translocation involving moesin (MSN-ALK) in anaplastic large cell lymphoma. Lab Invest 81:419–426. Hernandez, L., Pinyol, M., Hernandez, S., et€al. 1999. TRK-fused gene (TFG) is a new partner of ALK in anaplastic large cell lymphoma producing two structurally different TFG-ALK translocations. Blood 94:3265–3268. Cools, J., Wlodarska, I., Somers, R., et€al. 2002. Identification of novel fusion partners of ALK, the anaplastic lymphoma kinase, in anaplastic large-cell lymphoma and inflammatory myofibroblastic tumor. Genes Chromosomes Cancer 34:354–362. Lamant, L., Gascoyne, R. D., Duplantier, M. M., et€ al. 2003. Non-muscle myosin heavy chain (MYH9): a new partner fused to ALK in anaplastic large cell lymphoma. Genes Chromosomes Cancer 37:427–432. Ma, Z., Hill, D. A., Collins, M. H., et€al. 2003. Fusion of ALK to the Ran-binding protein 2 (RANBP2) gene in inflammatory myofibroblastic tumor. Genes Chromosomes Cancer 37:98–105. Ma, Z., Cools, J., Marynen, P., et€ al. 2000. Inv(2)(p23q35) in anaplastic large-cell lymphoma induces constitutive anaplastic lymphoma kinase (ALK) tyrosine kinase activation by fusion to ATIC, an enzyme involved in purine nucleotide biosynthesis. Blood 95:2144–2149.
210
M.S. Lim and M.S. Cairo
Bohling, S. D., Jenson, S., Crockett, D. K., et€al. 2008. Analysis of gene expression profiling of TPM3-ALK positive anaplastic large cell lymphoma reveals overlapping and unique patterns with that of NPM-ALK positive anaplastic large cell lymphoma. Leuk Res 32:383–393 Pulford, K., Morris, S. W. and Mason, D. Y. 2001. Anaplastic lymphoma kinase proteins and malignancy. Curr Opin Hematol 8:231–236. Hubinger, G., Wehnes, E., Xue, L., et€al. 2003. Hammerhead ribozyme-mediated cleavage of the fusion transcript NPM-ALK associated with anaplastic large-cell lymphoma. Exp Hematol 31:226–233. Ritter, U., Damm-Welk, C., Fuchs, U., et€al. 2003. Design and evaluation of chemically synthesized siRNA targeting the NPM-ALK fusion site in anaplastic large cell lymphoma (ALCL). Oligonucleotides 13:365–373. Piva, R., Chiarle, R., Manazza, A. D., et€al. 2006. Ablation of oncogenic ALK is a viable therapeutic approach for anaplastic large-cell lymphomas. Blood 107:689–697. Hsu, F. Y., Zhao, Y., Anderson, W. F., et€al. 2007. Downregulation of NPM-ALK by siRNA causes anaplastic large cell lymphoma cell growth inhibition and augments the anti cancer effects of chemotherapy in€vitro. Cancer Invest 25:240–248. Wan, W., Albom, M. S., Lu, L., et€al. 2006. Anaplastic lymphoma kinase activity is essential for the proliferation and survival of anaplastic large-cell lymphoma cells. Blood 107:1617–1623. Galkin, A. V., Melnick, J. S., Kim, S., et€al. 2007. Identification of NVP-TAE684, a potent, selective, and efficacious inhibitor of NPM-ALK. Proc Natl Acad Sci U S A 104:270–275. Marzec, M., Kasprzycka, M., Ptasznik, A., et€al. 2005. Inhibition of ALK enzymatic activity in T-cell lymphoma cells induces apoptosis and suppresses proliferation and STAT3 phosphorylation independently of Jak3. Lab Invest 85:1544–1554. Bai, R. Y., Ouyang, T., Miething, C., et€ al. 2000. Nucleophosmin-anaplastic lymphoma kinase associated with anaplastic large-cell lymphoma activates the phosphatidylinositol 3-kinase/ Akt antiapoptotic signaling pathway. Blood 96:4319–4327. Zamo, A., Chiarle, R., Piva, R., et€al. 2002. Anaplastic lymphoma kinase (ALK) activates Stat3 and protects hematopoietic cells from cell death. Oncogene 21:1038–1047. Amin, H. M., Medeiros, L. J., Ma, Y., et€al. 2003. Inhibition of JAK3 induces apoptosis and decreases anaplastic lymphoma kinase activity in anaplastic large cell lymphoma. Oncogene 22:5399–5407. Khoury, J. D., Medeiros, L. J., Rassidakis, G. Z., et€al. 2003. Differential expression and clinical significance of tyrosine-phosphorylated STAT3 in ALK+ and ALK- anaplastic large cell lymphoma. Clin Cancer Res 9:3692–3699. Vega, F., Medeiros, L. J., Leventaki, V., et€al. 2006. Activation of mammalian target of rapamycin signaling pathway contributes to tumor cell survival in anaplastic lymphoma kinase-positive anaplastic large cell lymphoma. Cancer Res 66:6589–6597. Marzec, M., Kasprzycka, M., Liu, X., et€al. 2007a. Oncogenic tyrosine kinase NPM/ALK induces activation of the rapamycin-sensitive mTOR signaling pathway. Oncogene 26:5606–5614. Marzec, M., Kasprzycka, M., Liu, X., et€al. 2007b. Oncogenic tyrosine kinase NPM/ALK induces activation of the MEK/ERK signaling pathway independently of c-Raf. Oncogene 26:813–821. Bonvini, P., Gastaldi, T., Falini, B., et€ al. 2002. Nucleophosmin-anaplastic lymphoma kinase (NPM-ALK), a novel Hsp90-client tyrosine kinase: down-regulation of NPM-ALK expression and tyrosine phosphorylation in ALK(+) CD30(+) lymphoma cells by the Hsp90 antagonist 17-allylamino,17-demethoxygeldanamycin. Cancer Res 62:1559–1566. Amin, H. M., McDonnell, T. J., Ma, Y., et€al. 2004. Selective inhibition of STAT3 induces apoptosis and G(1) cell cycle arrest in ALK-positive anaplastic large cell lymphoma. Oncogene 23:5426–5434. Shi, X., Franko, B., Frantz, C., et€al. 2006. JSI-124 (cucurbitacin I) inhibits Janus kinase-3/signal transducer and activator of transcription-3 signalling, downregulates nucleophosmin-anaplastic lymphoma kinase (ALK), and induces apoptosis in ALK-positive anaplastic large cell lymphoma cells. Br J Haematol 135:26–32. Chiarle, R., Simmons, W. J., Cai, H., et€al. 2005. Stat3 is required for ALK-mediated lymphomagenesis and provides a possible therapeutic target. Nat Med 11:623–629. Bonvini, P., Dalla Rosa, H., Vignes, N., et€al. 2004. Ubiquitination and proteasomal degradation of nucleophosmin-anaplastic lymphoma kinase induced by 17-allylamino-demethoxygeldana-
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
211
mycin: role of the co-chaperone carboxyl heat shock protein 70-interacting protein. Cancer Res 64:3256–3264. Schumacher, J. A., Crockett, D. K., Elenitoba-Johnson, K. S., et€al. 2007. Proteome-wide changes induced by the Hsp90 inhibitor, geldanamycin in anaplastic large cell lymphoma cells. Proteomics 7:2603–2616. Ambrogio, C., Voena, C., Manazza, A. D., et€al. 2005. p130Cas mediates the transforming properties of the anaplastic lymphoma kinase. Blood 106:3907–3916. Cussac, D., Greenland, C., Roche, S., et€al. 2004. Nucleophosmin-anaplastic lymphoma kinase of anaplastic large-cell lymphoma recruits, activates, and uses pp60c-src to mediate its mitogenicity. Blood 103:1464–1471. Gruss, H. J. and Dower, S. K. 1995. Tumor necrosis factor ligand superfamily: involvement in the pathology of malignant lymphomas. Blood 85:3378–3404. Schneider, C. andHubinger, G. 2002. Pleiotropic signal transduction mediated by human CD30: a member of the tumor necrosis factor receptor (TNFR) family. Leuk Lymphoma 43:1355–1366. Horie, R., Gattei, V., Ito, K., et€al. 1999. Frequent expression of the variant CD30 in human malignant myeloid and lymphoid neoplasms. Am J Pathol 155:2029–2041. Mir, S. S., Richter, B. W. and Duckett, C. S. 2000. Differential effects of CD30 activation in anaplastic large cell lymphoma and Hodgkin disease cells. Blood 96:4307–4312. Gause, A., Jung, W., Schmits, R., et€al. 1992. Soluble CD8, CD25 and CD30 antigens as prognostic markers in patients with untreated Hodgkin’s lymphoma. Ann Oncol 3 Suppl 4:49–52. Nadali, G., Vinante, F., Stein, H., et€al. 1995. Serum levels of the soluble form of CD30 molecule as a tumor marker in CD30+ anaplastic large-cell lymphoma. J Clin Oncol 13:1355–1360. Hansen, H. P., Kisseleva, T., Kobarg, J., et€al. 1995. A zinc metalloproteinase is responsible for the release of CD30 on human tumor cell lines. Int J Cancer 63:750–756. Zinzani, P. L., Pileri, S., Bendandi, M., et€al. 1998. Clinical implications of serum levels of soluble CD30 in 70 adult anaplastic large-cell lymphoma patients. J Clin Oncol 16:1532–1537. Tian, Z. G., Longo, D. L., Funakoshi, S., et€al. 1995. In vivo antitumor effects of unconjugated CD30 monoclonal antibodies on human anaplastic large-cell lymphoma xenografts. Cancer Res 55:5335–5341. Borchmann, P., Treml, J. F., Hansen, H., et€al. 2003. The human anti-CD30 antibody 5F11 shows in€vitro and in€vivo activity against malignant lymphoma. Blood 102:3737–3742. Perkins, S. L., Lones, M. A., Davenport, V., et€al. 2004. CD52 is highly expressed and may provide an excellent target for monoclonal antibody (alemtuzumab) therapy in childhood nonHodgkin’s lymphoma: a report from the Children’s Oncology Group (abstract). Pediatr Blood Cancer 43:366. Fillmore, G. C., Lin, Z., Bohling, S. D., et€al. 2002. Gene expression profiling of cell lines derived from T-cell malignancies. FEBS Lett 522:183–188. Lim, M. S., Elenitoba-Johnson, K. S. 2006. Mass spectrometry-based proteomic studies of human anaplastic large cell lymphoma. Mol Cell Proteomics 5:1787–1798. Lim, M. S., Tygeson, J., Seiler, C., Crockett, D. K., Satwani, P., Perkins, S. L., Cairo, M. S., Elenitoba-Johnson, K. S. J. 2006. Aberrant expression of IL-2R in pediatric anaplastic large cell lymphoma: cellular and proteomic analysis of denileukin difititox (ONTAK) as a potential therapeutic agent. American Society of Hematology Annual Meeting, 2006. Blood 108 (11):2051. Janik, J. E., Morris, J. C., Pittaluga, S., et€al. 2004. Elevated serum-soluble interleukin-2 receptor levels in patients with anaplastic large cell lymphoma. Blood 104:3355–3357. Crist, W. M., Shuster, J. J., Falletta, J., et€al. 1988. Clinical features and outcome in childhood T-cell leukemia-lymphoma according to stage of thymocyte differentiation: a Pediatric Oncology Group Study. Blood 72:1891–1897. Uckun, F. M., Gaynon, P. S., Sensel, M. G., et€al. 1997. Clinical features and treatment outcome of childhood T-lineage acute lymphoblastic leukemia according to the apparent maturational stage of T-lineage leukemic blasts: a Children’s Cancer Group study. J Clin Oncol 15: 2214–2221. Soslow, R. A., Baergen, R. N. and Warnke, R. A. 1999. B-lineage lymphoblastic lymphoma is a clinicopathologic entity distinct from other histologically similar aggressive lymphomas with blastic morphology. Cancer 85:2648–2654.
212
M.S. Lim and M.S. Cairo
Lin, P., Jones, D., Dorfman, D. M., et€al. 2000. Precursor B-cell lymphoblastic lymphoma: a predominantly extranodal tumor with low propensity for leukemic involvement. Am J Surg Pathol 24:1480–1490. Neth, O., Seidemann, K., Jansen, P., et€ al. 2000. Precursor B-cell lymphoblastic lymphoma in childhood and adolescence: clinical features, treatment, and results in trials NHL-BFM 86 and 90. Med Pediatr Oncol 35:20–27. Maitra, A., McKenna, R. W., Weinberg, A. G., et€al. 2001. Precursor B-cell lymphoblastic lymphoma. A study of nine cases lacking blood and bone marrow involvement and review of the literature. Am J Clin Pathol 115:868–875. Harris, N. L., Jaffe, E. S., Stein, H., et€al. 1994. A revised European-American classification of lymphoid neoplasms: a proposal from the International Lymphoma Study Group. Blood 84:1361–1392. Jaffe, E. 2001. Pathology and genetics of tumours of haematopoietic and lymphoid tissues. In World Health Organization Classification of Tumors, ed. E. Jaffe, N. L. Harris, H. Stein & J. W. Vardiman, pp. 1–351. Washington, DC: IARC Press. Reiter, A., Schrappe, M., Ludwig, W. D., et€al. 2000. Intensive ALL-type therapy without local radiotherapy provides a 90% event-free survival for children with T-cell lymphoblastic lymphoma: a BFM group report. Blood 95:416–421. Hoelzer, D., Gokbuget, N., Digel, W., et€al. 2002. Outcome of adult patients with T-lymphoblastic lymphoma treated according to protocols for acute lymphoblastic leukemia. Blood 99:4379–4385. Goldberg, J. M., Silverman, L. B., Levy, D. E., et€al. 2003. Childhood T-cell acute lymphoblastic leukemia: the Dana-Farber Cancer Institute acute lymphoblastic leukemia consortium experience. J Clin Oncol 21:3616–3622. Link, M. P., Shuster, J. J., Donaldson, S. S., et€al. 1997. Treatment of children and young adults with early-stage non-Hodgkin’s lymphoma. N Engl J Med 337:1259–1266. Patte, C., Auperin, A., Michon, J., et€ al. 2001. The Societe Francaise d’Oncologie Pediatrique LMB89 protocol: highly effective multiagent chemotherapy tailored to the tumor burden and initial response in 561 unselected children with B-cell lymphomas and L3 leukemia. Blood 97:3370–3379. Asselin, B. 2001. Improved event-free survival (EFS) with high dose methotrexate (HDM) in T-cell lymphoblastic leukemia (T-ALL) and advanced lymphoblastic lymphoma (T-NHL): a Pediatric Oncology Group (POG) study (abstract). Proc ASCO 1464. Thomas, D. A., O’Brien, S., Cortes, J., et€al. 2004. Outcome with the hyper-CVAD regimens in lymphoblastic lymphoma. Blood 104:1624–1630. Thomas, D. A. and Kantarjian, H. M. 2001. Lymphoblastic lymphoma. Hematol Oncol Clin North Am 15:51–95, vi. Sandlund, J. T., Downing, J. R. and Crist, W. M. 1996. Non-Hodgkin’s lymphoma in childhood. N Engl J Med 334:1238–1248. Goldsby, R. E. and Carroll, W. L. 1998. The molecular biology of pediatric lymphomas. J Pediatr Hematol Oncol 20:282–296. Heerema, N. A., Sather, H. N., Sensel, M. G., et€al. 1998. Frequency and clinical significance of cytogenetic abnormalities in pediatric T-lineage acute lymphoblastic leukemia: a report from the Children’s Cancer Group. J Clin Oncol 16:1270–1278. Ferrando, A. A., Neuberg, D. S., Staunton, J., et€al. 2002. Gene expression signatures define novel oncogenic pathways in T cell acute lymphoblastic leukemia. Cancer Cell 1:75–87. Ferrando, A. A. and Look, A. T. 2003. Gene expression profiling in T-cell acute lymphoblastic leukemia. Semin Hematol 40:274–280. Raetz, E. A., Perkins, S. L., Bhojwani, D., et€al. 2006. Gene expression profiling reveals intrinsic differences between T-cell acute lymphoblastic leukemia and T-cell lymphoblastic lymphoma. Pediatr Blood Cancer 47:130–140. Okuda, T., Shurtleff, S. A., Valentine, M. B., et€al. 1995. Frequent deletion of p16INK4a/MTS1 and p15INK4b/MTS2 in pediatric acute lymphoblastic leukemia. Blood 85:2321–2330.
New Therapeutic Frontiers for Childhood Non-Hodgkin Lymphoma
213
Cayuela, J. M., Madani, A., Sanhes, L., et€al. 1996. Multiple tumor-suppressor gene 1 inactivation is the most frequent genetic alteration in T-cell acute lymphoblastic leukemia. Blood 87:2180–2186. Baer, R. 1993. TAL1, TAL2 and LYL1: a family of basic helix-loop-helix proteins implicated in T cell acute leukaemia. Semin Cancer Biol 4:341–347. Bash, R. O., Hall, S., Timmons, C. F., et€al. 1995. Does activation of the TAL1 gene occur in a majority of patients with T-cell acute lymphoblastic leukemia? A pediatric oncology group study. Blood 86:666–676. Brown, L., Cheng, J. T., Chen, Q., et€al. 1990. Site-specific recombination of the tal-1 gene is a common occurrence in human T cell leukemia. Embo J 9:3343–3351. Janssen, J. W., Ludwig, W. D., Sterry, W., et€al. 1993. SIL-TAL1 deletion in T-cell acute lymphoblastic leukemia. Leukemia 7:1204–1210. Delabesse, E., Bernard, M., Landman-Parker, J., et€ al. 1997. Simultaneous SIL-TAL1 RT-PCR detection of all tal(d) deletions and identification of novel tal(d) variants. Br J Haematol 99:901–907. Hall, M. A., Curtis, D. J., Metcalf, D., et€al. 2003. The critical regulator of embryonic hematopoiesis, SCL, is vital in the adult for megakaryopoiesis, erythropoiesis, and lineage choice in CFU-S12. Proc Natl Acad Sci U S A 100:992–997. O’Neil, J., Shank, J., Cusson, N., et€al. 2004. TAL1/SCL induces leukemia by inhibiting the transcriptional activity of E47/HEB. Cancer Cell 5:587–596. Hansson, A., Manetopoulos, C., Jonsson, J. I., et€al. 2003. The basic helix-loop-helix transcription factor TAL1/SCL inhibits the expression of the p16INK4A and pTalpha genes. Biochem Biophys Res Commun 312:1073–1081. Kawabe, T., Muslin, A. J. and Korsmeyer, S. J. 1997. HOX11 interacts with protein phosphatases PP2A and PP1 and disrupts a G2/M cell-cycle checkpoint. Nature 385:454–458. Fahraeus, R., Paramio, J. M., Ball, K. L., et€ al. 1996. Inhibition of pRb phosphorylation and cell-cycle progression by a 20-residue peptide derived from p16CDKN2/INK4A. Curr Biol 6:84–91. Paietta, E., Ferrando, A. A., Neuberg, D., et€ al. 2004. Activating FLT3 mutations in CD117/ KIT(+) T-cell acute lymphoblastic leukemias. Blood 104:558–560. Griesinger, F., Janke, A., Podleschny, M., et€ al. 2002. Identification of an ETV6-ABL2 fusion transcript in combination with an ETV6 point mutation in a T-cell acute lymphoblastic leukaemia cell line. Br J Haematol 119:454–458. Chan, S. M., Weng, A. P., Tibshirani, R., et€al. 2007. Notch signals positively regulate activity of the mTOR pathway in T-cell acute lymphoblastic leukemia. Blood 110:278–286. Barata, J. T., Silva, A., Brandao, J. G., et€al. 2004. Activation of PI3K is indispensable for interleukin 7-mediated viability, proliferation, glucose use, and growth of T cell acute lymphoblastic leukemia cells. J Exp Med 200:659–669. Sade, H., Krishna, S. and Sarin, A. 2004. The anti-apoptotic effect of Notch-1 requires p56lckdependent, Akt/PKB-mediated signaling in T cells. J Biol Chem 279:2937–2944. Talora, C., Campese, A. F., Bellavia, D., et€al. 2003. Pre-TCR-triggered ERK signalling-dependent downregulation of E2A activity in Notch3-induced T-cell lymphoma. EMBO Rep 4:1067–1072. Kelly, J. A., Spolski, R., Kovanen, P. E., et€al. 2003b. Stat5 synergizes with T cell receptor/antigen stimulation in the development of lymphoblastic lymphoma. J Exp Med 198:79–89. Bellavia, D., Campese, A. F., Alesse, E., et€al. 2000. Constitutive activation of NF-kappaB and T-cell leukemia/lymphoma in Notch3 transgenic mice. Embo J 19:3337–3348. Head, D. R. and Behm, F. G. 1995. Acute lymphoblastic leukemia and the lymphoblastic lymphomas of childhood. Semin Diagn Pathol 12:325–334. Hojo, H., Sasaki, Y., Nakamura, N., et€al. 2001. Absence of somatic hypermutation of immunoglobulin heavy chain variable region genes in precursor B-lymphoblastic lymphoma: a study of four cases in childhood and adolescence. Am J Clin Pathol 116:673–682. Edwards H, Xie C, LaFiura KM, Dombkowski A, Buck S, Boerner J, Taub JW, Matherly LH, Ge Y. 2009. RUNX1 regulates phosphoinositide 3-kinase/AKT pathway: role in chemotherapy sensitivity in acute megakaryocytic leukemia. Blood 114:2744–2752.
Molecular Targeting of Post-transplant Lymphoproliferative Disorders Michael Wang and Thomas G. Gross
Introduction Post-transplant lymphoproliferative disorders (PTLD) represent a heterogeneous group of life-threatening lymphoproliferative disorders that can be observed in a transplant recipient. PTLD can occur in patients after solid organ transplantation (SOT) because of immunosuppression to prevent graft rejection (Penn et€al. 1969), and continues to be a major cause of morbidity and mortality seen in about 10% of pediatric SOT recipients. There is a higher incidence in children following SOT than in adults (Ho et€al. 1988; Swerdlow et€al. 2000), with highest incidence of 20% following heart-lung transplant. PTLD occurs in hematopoietic stem cell transplantation (HSCT) recipients secondary to the immunosuppression of pre-HSCT preparative regimens, and the post-HSCT immunosuppression to prevent graft vs host disease (GVHD). PTLD in HSCT occurs at a lower rate than following SOT (approximately 1%), with the vast majority occurring within 6€months following HSCT (Bhatia et€al. 1996; Curtis et€al. 1999). Accordingly, few cases of PTLD have been reported after autologous HSCT (Lones et€al. 2000; Nash et€al. 2003). PTLD is associated with Epstein–Barr virus (EBV) and inadequate EBV immunity in the majority of cases. PTLD following HSCT is essentially all EBV-associated. EBVnegative PTLD occurs following SOT in as many as 30% of cases. (Leblond et€al. 2001). The pathogenesis, treatment strategies and outcome differ from EBVpositive PTLD, as EBV-negative disease tends to require more aggressive therapy and portends a worse prognosis. This chapter will focus on EBV positive PTLD and molecularly targeted therapies in its prevention and treatment. The World Health Organization classification of PTLD contains three categories (Harris et€al. 1999). Early lesions, which are usually felt to not represent “true PTLD,” are characterized by reactive plasmacytic hyperplasia with normal tissue architecture retained. When normal tissue architecture is disrupted by the lymphoproliferative T.G. Grossâ•›(*) Division of Hematology/Oncology/BMT, The Ohio State University College of Medicine, 700 Children’s Drive, Columbus, OH 43205, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_11, © Springer Science+Business Media, LLC 2010
215
216
M. Wang and T.G. Gross
process, then the diagnosis of PTLD is appropriate. Polymorphic PTLD comprises different cells types, B cells, T cells, and histiocytes, and can be either monoclonal or polyclonal. Monomorphic PTLD is diagnosed when the lesions contain a more homogeneous population of transformed lymphocytes and resemble frank lymphoma such as B cell lymphoma (Burkitt, Burkitt-like lymphoma, diffuse large B cell lymphoma, or plasma cell myeloma), T cell lymphoma (peripheral T cell lymphoma, gd T cell lymphoma, hepatosplenic T cell lymphoma, T/NK lymphoma), and other types, such as Hodgkin disease-like and plasmacytoma-like lymphomas. Of interest, lymphoblastic and anaplastic large cell histologies are not associated with PTLD or immunodeficiency. Following SOT, PTLD are usually recipient in origin; however, following HSCT, PTLD is usually donor origin. Risk factors for developing PLTD after SOT include: type and intensity of immunosuppressive therapy, transplant from an EBV-seropositive donor into a seronegative recipient, and small bowel or heart-lung transplant (Opelz and Dohler 2004). HSCT PTLD risk factors are: T cell depletion, ATG as prophylaxis or therapy for GVHD, HLAmismatched transplant, and immunodeficiency as the primary diagnosis (Gross et€al. 1999; Baker et€al. 2003). The signs and symptoms of PTLD reported by patients are diverse and a high index of suspicion is necessary to make the diagnosis (Collins et€al. 2001). Symptoms may include those suggestive of acute mononucleosis, but PTLD can involve almost any organ system, including the central nervous system. Fulminant disease can mimic sepsis, GVHD, or a hemophagocytic syndrome. Most PTLD presents as lymphoproliferation with extranodal involvement being the rule, not the exception. Recently, several investigators have demonstrated that EBV-PTLD is associated with increase in EBV DNA load in the peripheral blood of patients following SOT and HSCT (Stevens et€al. 2001; Holmes et€al. 2002; Wagner et€al. 2004; Greenfield et€al. 2006). Polymerase chain reaction (PCR) used to measure EBV DNA has been suggested as a method to make a preemptive diagnosis of PTLD. However, using quantitative real-time PCR assays relies on the assumption that growth of EBVinfected B cells correlates with increases in peripheral blood EBV DNA. This assumption has been demonstrated to be true with CMV monitoring in HSCT patients (Boeckh et€al. 2003). The main difference is that CMV monitoring reflects active viral replication, whereas EBV-PTLD is associated with latent virus. Thus, detection of elevated levels of EBV DNA may point toward a diagnosis of PTLD, but signs and symptoms along with correlative imaging and tissue biopsy, if at all possible, must be analyzed before therapy is started. New approaches to EBV monitoring include assessing EBV-specific cytotoxic T cells (CTLs). One method measures interferon-g using either Elispot or flow cytometry assaying intracellular cytokine. The use of Elispot in combination with EBV DNA viral load has been reported to have a very high positive predictive value in predicting PTLD development (Smets et€al. 2002). MHC-class I peptide tetramers can quantify CTLs-specific EBV antigens, both lytic and latent viral antigens, and have been used to measure the recovery of EBV-specific CTL after HSCT. The predictive value of developing PTLD appears much higher with low numbers of EBV-specific CTL rather than with high levels of EBV DNA (Meij et€al. 2003).
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
217
Current strategies to treat EBV-PTLD have largely involved the experience of the clinician or retrospective case series of patients. There is no clear consensus therapy as the disease is heterogeneous and patient co-morbidities may limit therapeutic options. Most first-line treatment strategies involve reducing or changing the type of immunosuppressive medications. This has been most effective in patients who have polymorphic PTLD or early stage disease; however, it may be effective even in monoclonal or disseminated disease. Surgery and radiation have also shown good efficacy for localized disease, though this is feasible only in a minority of patients. For patients who do not respond to reduction of immunosuppression, especially those who have monomorphic PTLD or lymphoma, chemotherapy has been used. Standard lymphoma regimens are associated with increased toxicities in the transplant population. In children with PTLD following SOT, it appears that a low-dose regimen, that is, cyclophosphamide and prednisone, may be as effective as standard lymphoma regimens without as much toxicity (Gross et€al. 2005). A number of targeted therapies are currently being evaluated and will be the focus of discussion of this chapter.
Epstein–Barr Virus Infection of Humans To discuss potential preventative and treatment strategies of PTLD, a review of human EBV infection is necessary (Dolcetti and Masucci 2003; Young and Rickinson 2004). EBV is a human gamma-herpes virus that establishes latency in about 90% of the adult population. EBV infects naïve B cells by entry through the gp350-CD21 interaction causing a polyclonal expansion of EBV-transformed lymphoblasts in oropharyngeal lymphoid tissue. This leads to a lytic infection usually in tonsillar B cells and serves as the site of production of infectious EBV particles for spread of the virus to subsequent hosts. In most B cells that disseminate throughout the body, viral latency is determined by one of three latent gene transcription programs. The first is the “growth” program, type III latency, where all of the latent genes are expressed: Epstein–Barr nuclear antigens (EBNAs -1, -2, -3A, -3B, -3C, -LP), latent membrane protein (LMP1, LMP2A and LMP2B), polyadenylated viral RNAs (EBER1 and EBER2), and BamH1 RNAs (BARTs). Latency type III promotes proliferation and survival of infected naïve B cells without differentiation. This is accomplished in part by the expression of EBNA-2, which mimics Notch signaling (Hofelmayr et€al. 2001). The control of this EBV-driven B cell expansion is by T cells, with both CD4(+) and CD8(+) cells playing important roles. This interplay and ultimate establishment of latency represents a careful co-evolution of the human immune system and EBV (Thorley-Lawson 2001). After exiting the cell cycle, EBNA-2, the main latent gene transactivator of the latency type III program and inhibitor of B cell differentiation, is turned off. These B cells now enter type II latency where B lymphoblasts migrate to lymphoid follicles and differentiate into a germinal center phenotype. Differentiation of B cells is now driven by LMP1 and LPM2A independent of antigen signals from follicular dendritic cells or T cell help. LMP2A suppresses B cell receptor signaling but
218
M. Wang and T.G. Gross
enhances signaling through Syk and Src – family kinases for B cell survival (Dykstra et€al. 2001). The survival signals normally provided by T cell help come from LMP1, which is a functional mimic of CD40. Both LMP2A and LMP1 are transmembrane proteins, ligand-independent and constitutively active, organizing themselves in lipid rafts (Kilger et€al. 1998). B cells that survive type II latency join the pool of long-lived memory B cells where EBV persists for the lifetime of the host. Type I latency is characterized by the absence of EBV gene expression except for the periodic expression of EBNA-1 that allows the viral episome to be passaged extrachromosomally during cell division. A latency 0 stage has been suggested where EBER 1 and 2 and possibly BARTs are expressed, but the remainder of the EBV genome is silenced. With the lack of EBV gene expression in latency stages 0 and 1, host immune memory established after the primary infection with EBV would be unable to detect EBV infected cells, and EBV can persist by evading immune detection, and hence, establishing a lifelong infection (Table€1).
Host Immune Response Against EBV Much of the understanding and monitoring of primary EBV infection is derived from early serological work that investigated IgM and IgG titers to EBNAs expressed in latently infected cells. Most of these data today are derived from patients with infectious mononucleosis (IM) (Pedneault et€al. 1998). When patients present with IM, they usually have high titer IgM to viral capsid antigen (VCA) and rising IgG titers to VCA and early antigen (EA). The IgG anti-EA response falls more rapidly and to lower levels than the IgG anti-VCA. In addition, IgG responses to EBNA2 occur in the acute phase, while IgG anti-EBNA-1 responses occur in late in convalescence. Antibody responses to LMP1 or LMP2 are not seen in IM. Healthy EBV carriers are IgG anti-VCA, anti-gp350 neutralizing antibody and, IgG anti-EBNA-1 positive, but can also carry stable antibody reactivity against other EBV proteins such as EA (Rickinson and Kieff 2007). Serologic studies can be misleading in transplant patients due to the immunosuppression and/or passive antibody transfer from blood products or supplemental gammaglobulin. Cellular immune responses are vital for controlling primary EBV infection and subsequent EBV reactivation. The importance of the innate and adaptive cellular immune system is best exemplified through understanding genetic defects in patients susceptible to EBV disease, such as X-linked lymphoproliferative disorder (XLP), or signaling lymphocyte activation molecule (SLAM)-associated protein deficiency (Veillette et€ al. 2007), and X-linked inhibitor-of-apoptosis (XIAP) (Rigaud et€al. 2006). The defect in innate cellular immunity caused by mutations in these genes has been shown to alter normal development of NKT cells, likely due to abnormal peripheral homeostasis. Following EBV infection, patients with these diseases also demonstrate abnormalities in adaptive immune responses. SAP mutations have been shown to influence the ability of CD4(+) T cells to secrete certain cytokines
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
219
Table€1╅ Potential targeted treatment strategies for EBV associated PTLD Target Desired result Potential agents Potential obstacles EBV latent genes BHRF1 (Bcl-2 homologue) Enhance None currently Delivery & tumor apoptosis specificity LMP1 (upregulates NF-Kb Enhance None currently Delivery & tumor gene expression) apoptosis specificity LMP2A (increases LMP1 Enhance None currently Delivery & tumor half-life via RAS/PIK/akt) apoptosis specificity EBNA1 (can inhibit p53) Enhance None currently Delivery & tumor apoptosis specificity EBV lytic genes BCRF1 (IL-10 homologue) Enhance cellular None currently Delivery & tumor specificity immune response BARF1 Enhance cytokine None currently Delivery & tumor response specificity None currently Delivery & tumor Enhance viral gp350, ZEBRA VCA, specificity & potential replication & thymidine kinase viral of increase infected cell lysis DNA polymerase B-cell pool and future PTLD risk Cellular proteins Death receptors (TRAIL, Enhance Monoclonal Tumor specificity etc) apoptosis antibody Tumor specificity & Monoclonal Tumor cell surface antigens Enhance tumor activity in immune kill antibody compromised host (apoptosis, ADCC, etc) Growth signaling pathways Growth inhibition Sirolimus, Efficacy and toxicities everolimus remain (mTOR, etc) to be determined Enhanced immune response to EBV Humoral response vaccine Inhibit EBV Anti-gp350 Does not prevent infection infection & activity in immune compromised host None currently Activity in immune Cellular response vaccine Enhance control compromised host of EBV B-cell proliferation Limited to few Activity in immune Adoptive EBV-specific Enhance control centers compromised host cellular therapy of EBV B-cell proliferation
and mediate B cell help (B cell antibody production). In addition, SAP mutations decrease the proliferation, cytokine production, and cytotoxicity of CD8(+) T cells. The main defect in XIAP deficiency patients is increased activation-induced cell death of B cells, CD4(+) and CD(8+) T cells. Cellular responses to EBV infection in immunocompetent individuals have been well characterized (Callan et€ al. 1996; Rickinson et€ al. 2000; Hislop et€ al. 2007).
220
M. Wang and T.G. Gross
CD8(+) T cells undergo a multifold expansion during IM. In some individuals, this can be seen as large expansions of Vb-restricted CD8(+) T cells to epitopes from both latent and lytic cycle proteins. Lytic epitopes have been shown by tetramer staining to represent up to 50% of the CD8(+) T cell pool, whereas latent epitopes accounted for less than 5% (Callan et€al. 1998). The lytic response is directed against immediate early antigens (BZLF1 and BRLF1) and early antigens (BMLF1, BMRF1, BALF2, and BALF5). Most responses directed against latent epitopes include EBNA3A, -3B, -3C, LMP2 and EBNA1. Rarely is a response seen against EBNA-2, EBNA-LP, or LMP1. The pool of antigen expanded T cells undergo contraction, and after 1 to 2€years, CD8(+) T cells specific for lytic antigens represent 2 to 5% of the T cell pool and have an effector memory T cell phenotype (Callan et€al. 1998; Hislop et€al. 2002). In contrast, CD8(+) T cells specific directed against latent antigens only represent 0.5 to 2% of peripheral T cells but have a central memory phenotype (Catalina et€ al. 2001). CD4(+) T cell immunity has been less characterized than CD8(+) T cells (Maini et€al. 2000; Williams et€al. 2005). Clones that recognize lytic-antigens BZLF1, BHRF1, gp350 and gp110 have been characterized. CD4(+) T cell memory to latentantigens has been more extensively studied, and the hierarchy of immunodominance is different from that of CD8(+) T cells (Woodberry et€ al. 2005). Most memory responses are focused on EBNA-1, EBNA-2, and EBNA-3C proteins. These clones secrete IFN-g and display a TH1 phenotype. Whether they have T cell effector function in€vivo is unknown. NK cells also play a role in EBV immunity by linking adaptive and cellular immunity. It is likely that increased NK cell numbers function to initially control primary EBV infected B cells and protect humans during early infection. Additionally, they secrete cytokines that enhance T cell immunity, which ultimately resolves the primary EBV infection (Williams et€al. 2005). EBV evades the immune system by cytokine modulation and the production of a viral homologue of IL-10 (vIL-10, the product of EBV gene BCRF1) during the lytic infection. vIL-10 binds to the IL-10 receptor with reduced affinity and is associated with decreased secretion of IL-12 and interferon-g (Moore et€ al. 2001). This may result in suppression of Th1 T cell responses as infected cells transition from lytic to latent infection (Bejarano and Masucci 1998). Viral IL-10 also downregulates the expression of transporters associated with antigen presentation (TAP) proteins and interferes with CTL recognition of EBV-infected B cells (Zeidler et€al. 1997). Finally, during lytic infection, a soluble receptor for colony stimulating factor 1, BARF1, is expressed. BARF1 may inhibit cellular immune function by blocking antiviral cytokine (interferon-a) release from monocytes (Cohen and Lekstrom 1999).
EBV Targets Apoptotic Resistance Apoptosis is potential target of EBV. Early studies revealed that expression of EBV latency genes could prevent death from serum withdrawal (Gregory et€ al. 1991). BHRF1 is a Bcl-2 homologue that inhibits apoptosis induced by Fas ligand and TRAIL (Kawanishi et€ al. 2002). LMP1 upregulates multiple anti-apoptotic genes (bfl-1, mcl-1, A20, clAP2) that are downstream targets of NF-kb (Young et€ al.
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
221
1999), and Bcl-2 (Henderson et€ al. 1991). LMP1 via phosphoinositide-3 kinase (PI3K) (Lambert and Martinez 2007) induces IL-10 production that alters T cell immunity. LMP2A signaling can constitutively activate the RAS/PI3K/Akt pathway and prolongs the half-life of LMP1. It is assumed that the major function EBNA1 is maintenance of the EBV genome in latency 1 by binding ubiquitin-specific protease 7 (USP7) to prevent p53 stabilization by deubiquitination (Saridakis et€ al. 2005). EBNA2 can block apoptosis by binding Nur77 and inhibiting cytochrome C release from mitochondria after it translocates out of the nucleus into the cytoplasm (Lee et€al. 2002). Finally, death receptors may play a role as EBV lymphoblasts derived from PTLD patients exhibit resistance to TRAIL-induced apoptosis (Snow et€ al. 2006). To date, there are no clinical data on the use of anti-apoptosis agents to treat EBV disease of EBV-PTLD, though there are several interesting potential targets.
Molecular Targets for EBV-PTLD Vaccines Targeting EBV antigens with vaccines is an attractive therapeutic option for all malignancies associated with EBV, including nasopharyngeal carcinoma and endemic Burkitt lymphoma. Recently, an EBV recombinant gp350 vaccine completed a phase II, randomized, double-blind, placebo controlled trial for IM (Sokal et€al. 2007). The rationale for choosing gp350 as an immunologic target was that EBV predominantly infects B cells through binding of gp350 to CD21 (Nemerow et€al. 1987). Therefore, a neutralizing anti-gp350 antibody response should theorectically prevent EBV infection of B cells. This study showed that the recombinant gp350 vaccine decreased the number of symptomatic infections from 10% in the control group to 2% in the vaccine treated group. However, it did not reduce the incidence of asymptomatic EBV infections. The utility of vaccines that induce a humoral response to EBV for PTLD development remains to be determined. It is unclear if this vaccination approach would decrease the risk of PTLD for HSCT or SOT recipient who are EBV naïve prior to transplant. It is possible that gp350 vaccine might decrease EBV viremia and hence reduce the risk of EBV related posttransplant disease even in EBV positive patients. Vaccines to produce effective cellular immune responses should have greater utility in prevention of EBV related disease; however, progress in the development of a vaccine that produces an effective cellular response to latent virus infection has been slow in coming.
Anti-Viral Medications The role of anti-viral medications has not been substantiated even as prophylaxis of patients at risk for developing EBV-PTLD as thymidine kinase inhibitors do not suppress the proliferation of EBV-infected B cells during latency. These medications
222
M. Wang and T.G. Gross
cannot eradicate latently infected lymphocytes as EBV-thymidine kinase and BGLF4 gene product, which are necessary to phosphorylate gancyclovir and convert it to its active cytotoxic form, are expressed only during the lytic infection. One potential novel therapy to enhance the efficacy of gancyclovir would be to intentionally induce EBV lytic infection. This approach takes advantage of the expression of virus encoded kinases (thymidine kinase) during lytic infection. Arginine butyrate can induce viral thymidine kinase and has been shown to treat patients with refractory PTLD when used with gancyclovir (Mentzer et€ al. 2001). A recent phase I/II trial supports this hypothesis (Perrine et€al. 2007). Additionally, in€vitro, gemcitabine and doxorubicin have been shown to induce EBV lytic infection, and when used with gancyclovir inhibit EBV-lymphoproliferative disease in SCID mice (Feng et€ al. 2004). Similar synergy between rituximab, dexamethasone, and gancyclovir in€vitro has been shown to induce EBV lytic infection (Daibata et€al. 2005). The histone deactylase inhibitor valproic acid when given with chemotherapy can induce EBV lytic infection and increased chemotherapy tumor, killing in€vitro and in€vivo; however, it has not been used with an anti-viral such as gancyclovir (Feng and Kenney 2006).
Immunosuppression The more T cell immunosuppressive a regimen, including such agents as calcineurin inhibitors, OKT3 and anti-thymocyte globulin, the higher the risk of PTLD. Withdrawing immunosuppression has been an effective treatment and the “gold standard” for PTLD; however, this potentially jeopardizes the allograft. Recently, the role of proliferation signal inhibitors has been studied in order to better understand their role in malignancies post-transplantation (Pascual 2007). The mTOR pathway is nearly ubiquitously activated in PTLD (El-Salem et€ al. 2007). Everolimus and sirolimus are macrolide antibiotics that are immunosuppressive. They are proliferation signal inhibitors that have been shown to inhibit the growth of PTLD-like cell lines in€vitro and in€vivo (Majewski et€al. 2000), and may have the potential to prevent PTLD in high risk populations (Majewski et€al. 2003). A potential mechanism for the efficacy of proliferation signal inhibitors is that they inhibit the IL-10 transduction pathway (Nepomuceno et€al. 2003). In patients with PTLD who experienced graft rejection, sirolimus has been able to rescue the graft and decrease EBV viremia (Sindhi et€al. 2001). Proliferation signal inhibitors offer potential as not only immunosuppression but, additionally, prophylaxis for patients at increased risk to develop PTLD.
Monoclonal Antibodies Monoclonal antibodies that recognize antigens (CD20, CD21, CD24) on B cells have been used in patients who fail a trial of reduction of their immunosuppression. Now that rituximab, a monoclonal antibody that recognizes CD20, has become
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
223
commercially available and licensed to treat lymphoproliferative disease, the approach to treat PTLD has changed dramatically. Rituximab is an IgG1-kappa chimeric monoclonal antibody that is made from the variable regions from heavy and light chains of a murine anti-CD20 antibody and human IgG1 and kappa constant regions (Reff et€al. 1994). CD20 may play a role in the development and differentiation of B-cells into plasma cells. It is expressed on over 80% of EBV-PTLD (Swerdlow 1992). In vitro, rituximab induces both complement and antibody dependent cytotoxicity, apoptosis, and enhances the cytotoxic effects of chemotherapy (Demidem et€al. 1997). It is an attractive therapy to choose because it is easy to administer, has limited co-morbid side effects and does not limit EBV-CTL function; however, the cost of rituximab is significant and long-term suppression of B-cells has been observed with it use (Gross 2007). Rituximab has been successfully used to treat PTLD in children and adults, and response rates have ranged from 44 to 65%, even as a single agent (Lee et€ al. 2007). The largest prospective study using rituximab as a single agent included 43 SOT patients with PTLD (Choquet et€al. 2006). Three-fourths of the patients had advanced stage disease at the time of treatment. The overall response rate at 80€days was 44% and the overall survival at 1 year was 67%. The median survival was 15€ months. A smaller phase II trial found similar results; however, EBVpositivity and a shorter time from transplantation to diagnosis correlated with response to rituximab. In this study, none of the EBV-negative PTLD patients responded to rituximab (Oertel et€al. 2005). Data regarding the use of rituximab with chemotherapy are limited. A phase II study using the low-dose regimen of cyclophoshamide and prednisone plus rituximab in children, adolescents, and young adults with PTLD after SOT is currently being conducted by the Children’s Oncology Group.
T Cell Adoptive Immunotherapy EBV-CTL can be used to contain the virus or potentially kill virus-infected tumor cells; therefore, EBV-specific viral antigens recognized by host T cells in EBVPTLD are an attractive target for cellular immunotherapy (Bollard et€ al. 2003; Fujita et€al. 2008). Autologous and allogeneic EBV-CTLs have been evaluated by numerous groups (Savoldo et€ al. 2006; Haque et€ al. 2007). The advantages of autologous EBV-CTL are that they do not increase graft rejection and they can increase EBV-specific cellular immunity. A major limitation of T cell adoptive immunotherapy is our understanding of how to maintain long-term memory pools of infused T cells in the SOT recipients. Additionally, it takes 4 to 6€weeks to generate EBV-CTL, which may be problematic for patients with EBV-PTLD, and it is possible, though difficult, to generate EBV-CTL from EBV-seronegative recipients pretransplant, the highest risk population. This technology requires significant commitment of resources and expertise and is highly regulated to administer, making it unavailable but to a few centers.
224
M. Wang and T.G. Gross
Conclusion PTLD is a heterogeneous group of diseases that has multiple potential molecular targets. Determining the right combination of therapies for any given patient with EBV-PTLD remains a major clinical problem. Simply reducing immunosuppression may no longer be sufficient and the use of proliferation signal inhibitors may improve both graft function, spare calcineurin inhibitor use, and treat PTLD. Many patients who develop EBV-PTLD cannot tolerate multi-agent chemotherapy or low-dose chemotherapy approaches may not establish long-term remissions for all patients. The use of antiviral medications will not likely be as effective as agents that can initiate EBV lytic infection. Early studies suggest this to be a promising therapeutic approach. EBV-specific vaccines may one day show efficacy in EBVnegative patients, but vaccines that stimulate an effective cellular response have been difficult to produce. New and more specific monoclonal antibodies look promising since they do not interfere with graft function, but long-term immunological consequences need to be evaluated. T cell therapy has shown the most promise, but remains non-feasible for the majority of patients. Other strategies of enhancing EBV-specific T cell immunity that would provide viral clearance and/or protective immunity against EBV-PTLD would be of great benefit. Though there are multiple anti-apoptotic targets, both cellular and viral genes, to date no agents have been identified as being feasible for clinical trials. To improve outcome of EBV-PTLD, as in most cancer, combination modality therapy with noncross reacting toxicity and complementary mechanism of tumor cell clearance will be required.
References Baker, K. S., T. E. DeFor, et€al. (2003). “New malignancies after blood or marrow stem-cell transplantation in children and adults: incidence and risk factors.” J Clin Oncol 21(7): 1352–8. Bejarano, M. T. and M. G. Masucci (1998). “Interleukin-10 abrogates the inhibition of Epstein-Barr virus-induced B-cell transformation by memory T-cell responses.” Blood 92(11): 4256–62. Bhatia, S., N. K. Ramsay, et€al. (1996). “Malignant neoplasms following bone marrow transplantation.” Blood 87(9): 3633–9. Boeckh, M., W. G. Nichols, et€al. (2003). “Cytomegalovirus in hematopoietic stem cell transplant recipients: current status, known challenges, and future strategies.” Biol Blood Marrow Transplant 9(9): 543–58. Bollard, C. M., B. Savoldo, et€ al. (2003). “Adoptive T-cell therapy for EBV-associated posttransplant lymphoproliferative disease.” Acta Haematol 110(2–3): 139–48. Callan, M. F., N. Steven, et€al. (1996). “Large clonal expansions of CD8+ T cells in acute infectious mononucleosis.” Nat Med 2(8): 906–11. Callan, M. F., L. Tan, et€al. (1998). “Direct visualization of antigen-specific CD8+ T cells during the primary immune response to Epstein-Barr virus in€vivo.” J Exp Med 187(9): 1395–402. Catalina, M. D., J. L. Sullivan, et€al. (2001). “Differential evolution and stability of epitope-specific CD8(+) T cell responses in EBV infection.” J Immunol 167(8): 4450–7. Choquet, S., V. Leblond, et€al. (2006). “Efficacy and safety of rituximab in B-cell post-transplantation lymphoproliferative disorders: results of a prospective multicenter phase 2 study.” Blood 107(8): 3053–7.
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
225
Cohen, J. I. and K. Lekstrom (1999). “Epstein-Barr virus BARF1 protein is dispensable for B-cell transformation and inhibits alpha interferon secretion from mononuclear cells.” J Virol 73(9): 7627–32. Collins, M. H., K. T. Montone, et€al. (2001). “Autopsy pathology of pediatric posttransplant lymphoproliferative disorder.” Pediatrics 107(6): E89. Curtis, R. E., L. B. Travis, et€al. (1999). “Risk of lymphoproliferative disorders after bone marrow transplantation: a multi-institutional study.” Blood 94(7): 2208–16. Daibata, M., K. Bandobashi, et€al. (2005). “Induction of lytic Epstein-Barr virus (EBV) infection by synergistic action of rituximab and dexamethasone renders EBV-positive lymphoma cells more susceptible to ganciclovir cytotoxicity in€vitro and in€vivo.” J Virol 79(9): 5875–9. Demidem, A., T. Lam, et€ al. (1997). “Chimeric anti-CD20 (IDEC-C2B8) monoclonal antibody sensitizes a B cell lymphoma cell line to cell killing by cytotoxic drugs.” Cancer Biother Radiopharm 12(3): 177–86. Dolcetti, R. and M. G. Masucci (2003). “Epstein-Barr virus: induction and control of cell transformation.” J Cell Physiol 196(2): 207–18. Dykstra, M. L., R. Longnecker, et€al. (2001). “Epstein-Barr virus coopts lipid rafts to block the signaling and antigen transport functions of the BCR.” Immunity 14(1): 57–67. El-Salem, M., P. N. Raghunath, et€al. (2007). “Constitutive activation of mTOR signaling pathway in post-transplant lymphoproliferative disorders.” Lab Invest 87(1): 29–39. Feng, W. H., G. Hong, et€ al. (2004). “Lytic induction therapy for Epstein-Barr virus-positive B-cell lymphomas.” J Virol 78(4): 1893–902. Feng, W. H. and S. C. Kenney (2006). “Valproic acid enhances the efficacy of chemotherapy in EBV-positive tumors by increasing lytic viral gene expression.” Cancer Res 66(17): 8762–9. Fujita, Y., C. M. Rooney, et€al. (2008). “Adoptive cellular immunotherapy for viral diseases.” Bone Marrow Transplant 41(2): 193–8. Greenfield, H. M., M. I. Gharib, et€al. (2006). “The impact of monitoring Epstein-Barr virus PCR in paediatric bone marrow transplant patients: can it successfully predict outcome and guide intervention?” Pediatr Blood Cancer 47(2): 200–5. Gregory, C. D., C. Dive, et€ al. (1991). “Activation of Epstein-Barr virus latent genes protects human B cells from death by apoptosis.” Nature 349(6310): 612–4. Gross, T. G. (2007). “Post-transplant lymphoproliferative disease in children following solid organ transplant and rituximab – the final answer?” Pediatr Transplant 11(6): 575–7. Gross, T. G., J. C. Bucuvalas, et€ al. (2005). “Low-dose chemotherapy for Epstein-Barr viruspositive post-transplantation lymphoproliferative disease in children after solid organ transplantation.” J Clin Oncol 23(27): 6481–8. Gross, T. G., M. Steinbuch, et€ al. (1999). “B cell lymphoproliferative disorders following hematopoietic stem cell transplantation: risk factors, treatment and outcome.” Bone Marrow Transplant 23(3): 251–8. Haque, T., G. M. Wilkie, et€al. (2007). “Allogeneic cytotoxic T-cell therapy for EBV-positive posttransplantation lymphoproliferative disease: results of a phase 2 multicenter clinical trial.” Blood 110(4): 1123–31. Harris, N. L., E. S. Jaffe, et€al. (1999). “World Health Organization classification of neoplastic diseases of the hematopoietic and lymphoid tissues: report of the Clinical Advisory Committee meeting-Airlie House, Virginia, November 1997.” J Clin Oncol 17(12): 3835–49. Henderson, S., M. Rowe, et€ al. (1991). “Induction of bcl-2 expression by Epstein-Barr virus latent membrane protein 1 protects infected B cells from programmed cell death.” Cell 65(7): 1107–15. Hislop, A. D., N. E. Annels, et€al. (2002). “Epitope-specific evolution of human CD8(+) T cell responses from primary to persistent phases of Epstein-Barr virus infection.” J Exp Med 195(7): 893–905. Hislop, A. D., G. S. Taylor, et€al. (2007). “Cellular responses to viral infection in humans: lessons from Epstein-Barr virus.” Annu Rev Immunol 25: 587–617. Ho, M., R. Jaffe, et€ al. (1988). “The frequency of Epstein-Barr virus infection and associated lymphoproliferative syndrome after transplantation and its manifestations in children.” Transplantation 45(4): 719–27.
226
M. Wang and T.G. Gross
Hofelmayr, H., L. J. Strobl, et€al. (2001). “Activated Notch1 can transiently substitute for EBNA2 in the maintenance of proliferation of LMP1-expressing immortalized B cells.” J Virol 75(5): 2033–40. Holmes, R. D., K. Orban-Eller, et€al. (2002). “Response of elevated Epstein-Barr virus DNA levels to therapeutic changes in pediatric liver transplant patients: 56-month follow up and outcome.” Transplantation 74(3): 367–72. Kawanishi, M., S. Tada-Oikawa, et€al. (2002). “Epstein-Barr virus BHRF1 functions downstream of Bid cleavage and upstream of mitochondrial dysfunction to inhibit TRAIL-induced apoptosis in BJAB cells.” Biochem Biophys Res Commun 297(3): 682–7. Kilger, E., A. Kieser, et€al. (1998). “Epstein-Barr virus-mediated B-cell proliferation is dependent upon latent membrane protein 1, which simulates an activated CD40 receptor.” EMBO J 17(6): 1700–9. Lambert, S. L. and O. M. Martinez (2007). “Latent membrane protein 1 of EBV activates phosphatidylinositol 3-kinase to induce production of IL-10.” J Immunol 179(12): 8225–34. Leblond, V., N. Dhedin, et€ al. (2001). “Identification of prognostic factors in 61 patients with posttransplantation lymphoproliferative disorders.” J Clin Oncol 19(3): 772–8. Lee, J. J., M. S. Lam, et€al. (2007). “Role of chemotherapy and rituximab for treatment of posttransplant lymphoproliferative disorder in solid organ transplantation.” Ann Pharmacother 41(10): 1648–59. Lee, J. M., K. H. Lee, et€al. (2002). “Epstein-Barr virus EBNA2 blocks Nur77- mediated apoptosis.” Proc Natl Acad Sci USA 99(18): 11878–83. Lones, M. A., I. Kirov, et€al. (2000). “Post-transplant lymphoproliferative disorder after autologous peripheral stem cell transplantation in a pediatric patient.” Bone Marrow Transplant 26(9): 1021–4. Maini, M. K., N. Gudgeon, et€al. (2000). “Clonal expansions in acute EBV infection are detectable in the CD8 and not the CD4 subset and persist with a variable CD45 phenotype.” J Immunol 165(10): 5729–37. Majewski, M., M. Korecka, et€al. (2003). “Immunosuppressive TOR kinase inhibitor everolimus (RAD) suppresses growth of cells derived from posttransplant lymphoproliferative disorder at allograft-protecting doses.” Transplantation 75(10): 1710–7. Majewski, M., M. Korecka, et€ al. (2000). “The immunosuppressive macrolide RAD inhibits growth of human Epstein-Barr virus-transformed B lymphocytes in€vitro and in€vivo: a potential approach to prevention and treatment of posttransplant lymphoproliferative disorders.” Proc Natl Acad Sci USA 97(8): 4285–90. Meij, P., J. W. van Esser, et€al. (2003). “Impaired recovery of Epstein-Barr virus (EBV) – specific CD8+ T lymphocytes after partially T-depleted allogeneic stem cell transplantation may identify patients at very high risk for progressive EBV reactivation and lymphoproliferative disease.” Blood 101(11): 4290–7. Mentzer, S. J., S. P. Perrine, et€al. (2001). “Epstein-Barr virus post-transplant lymphoproliferative disease and virus-specific therapy: pharmacological re-activation of viral target genes with arginine butyrate.” Transpl Infect Dis 3(3): 177–85. Moore, K. W., R. de Waal Malefyt, et€al. (2001). “Interleukin-10 and the interleukin-10 receptor.” Annu Rev Immunol 19: 683–765. Nash, R. A., R. Dansey, et€al. (2003). “Epstein-Barr virus-associated posttransplantation lymphoproliferative disorder after high-dose immunosuppressive therapy and autologous CD34selected hematopoietic stem cell transplantation for severe autoimmune diseases.” Biol Blood Marrow Transplant 9(9): 583–91. Nemerow, G. R., C. Mold, et€al. (1987). “Identification of gp350 as the viral glycoprotein mediating attachment of Epstein-Barr virus (EBV) to the EBV/C3d receptor of B cells: sequence homology of gp350 and C3 complement fragment C3d.” J Virol 61(5): 1416–20. Nepomuceno, R. R., C. E. Balatoni, et€al. (2003). “Rapamycin inhibits the interleukin 10 signal transduction pathway and the growth of Epstein Barr virus B-cell lymphomas.” Cancer Res 63(15): 4472–80.
Molecular Targeting of Post-transplant Lymphoproliferative Disorders
227
Oertel, S. H., E. Verschuuren, et€al. (2005). “Effect of anti-CD 20 antibody rituximab in patients with post-transplant lymphoproliferative disorder (PTLD).” Am J Transplant 5(12): 2901–6. Opelz, G. and B. Dohler (2004). “Lymphomas after solid organ transplantation: a collaborative transplant study report.” Am J Transplant 4(2): 222–30. Pascual, J. (2007). “Post-transplant lymphoproliferative disorder – the potential of proliferation signal inhibitors.” Nephrol Dial Transplant 22(Suppl 1): i27–35. Pedneault, L., N. Lapointe, et€ al. (1998). “Natural history of Epstein-Barr virus infection in a prospective pediatric cohort born to human immunodeficiency virus-infected mothers.” J Infect Dis 177(4): 1087–90. Penn, I., W. Hammond, et€ al. (1969). “Malignant lymphomas in transplantation patients.” Transplant Proc 1(1): 106–12. Perrine, S. P., O. Hermine, et€al. (2007). “A phase 1/2 trial of arginine butyrate and ganciclovir in patients with Epstein-Barr virus-associated lymphoid malignancies.” Blood 109(6): 2571–8. Reff, M. E., K. Carner, et€al. (1994). “Depletion of B cells in€vivo by a chimeric mouse human monoclonal antibody to CD20.” Blood 83(2): 435–45. Rickinson, A. B., M. F. Callan, et€ al. (2000). “T-cell memory: lessons from Epstein-Barr virus infection in man.” Philos Trans R Soc Lond B Biol Sci 355(1395): 391–400. Rickinson, A. B. and E., Kieff (2007). Epstein-Barr Virus. Fields Virology. In: D. M. Knipe, P.M., Howley (eds). Philadelphia, Lippencott Williams & Wilkens. 2: 2655–700. Rigaud, S., M. C. Fondaneche, et€ al. (2006). “XIAP deficiency in humans causes an X-linked lymphoproliferative syndrome.” Nature 444(7115): 110–4. Saridakis, V., Y. Sheng, et€ al. (2005). “Structure of the p53 binding domain of HAUSP/USP7 bound to Epstein-Barr nuclear antigen 1 implications for EBV-mediated immortalization.” Mol Cell 18(1): 25–36. Savoldo, B., J. A. Goss, et€al. (2006). “Treatment of solid organ transplant recipients with autologous Epstein Barr virus-specific cytotoxic T lymphocytes (CTLs).” Blood 108(9): 2942–9. Sindhi, R., S. Webber, et€ al. (2001). “Sirolimus for rescue and primary immunosuppression in transplanted children receiving tacrolimus.” Transplantation 72(5): 851–5. Smets, F., D. Latinne, et€al. (2002). “Ratio between Epstein-Barr viral load and anti-Epstein-Barr virus specific T-cell response as a predictive marker of posttransplant lymphoproliferative disease.” Transplantation 73(10): 1603–10. Snow, A. L., S. L. Lambert, et€al. (2006). “EBV can protect latently infected B cell lymphomas from death receptor-induced apoptosis.” J Immunol 177(5): 3283–93. Sokal, E. M., K. Hoppenbrouwers, et€ al. (2007). “Recombinant gp350 vaccine for infectious mononucleosis: a phase 2, randomized, double-blind, placebo-controlled trial to evaluate the safety, immunogenicity, and efficacy of an Epstein-Barr virus vaccine in healthy young adults.” J Infect Dis 196(12): 1749–53. Stevens, S. J., E. A. Verschuuren, et€al. (2001). “Frequent monitoring of Epstein-Barr virus DNA load in unfractionated whole blood is essential for early detection of posttransplant lymphoproliferative disease in high-risk patients.” Blood 97(5): 1165–71. Swerdlow, A. J., C. D. Higgins, et€al. (2000). “Risk of lymphoid neoplasia after cardiothoracic transplantation. a cohort study of the relation to Epstein-Barr virus.” Transplantation 69(5): 897–904. Swerdlow, S. H. (1992). “Post-transplant lymphoproliferative disorders: a morphologic, phenotypic and genotypic spectrum of disease.” Histopathology 20(5): 373–85. Thorley-Lawson, D. A. (2001). “Epstein-Barr virus: exploiting the immune system.” Nat Rev Immunol 1(1): 75–82. Veillette, A., Z. Dong, et€ al. (2007). “Consequence of the SLAM-SAP signaling pathway in innate-like and conventional lymphocytes.” Immunity 27(5): 698–710. Wagner, H. J., Y. C. Cheng, et€al. (2004). “Prompt versus preemptive intervention for EBV lymphoproliferative disease.” Blood 103(10): 3979–81. Williams, H., K. McAulay, et€al. (2005). “The immune response to primary EBV infection: a role for natural killer cells.” Br J Haematol 129(2): 266–74.
228
M. Wang and T.G. Gross
Woodberry, T., T. J. Suscovich, et€al. (2005). “Differential targeting and shifts in the immunodominance of Epstein-Barr virus-specific CD8 and CD4 T cell responses during acute and persistent infection.” J Infect Dis 192(9): 1513–24. Young, L. S., C. W. Dawson, et€al. (1999). “Epstein-Barr virus and apoptosis: viral mimicry of cellular pathways.” Biochem Soc Trans 27(6): 807–12. Young, L. S. and A. B. Rickinson (2004). “Epstein-Barr virus: 40€ years on.” Nat Rev Cancer 4(10): 757–68. Zeidler, R., G. Eissner, et€al. (1997). “Downregulation of TAP1 in B lymphocytes by cellular and Epstein-Barr virus-encoded interleukin-10.” Blood 90(6): 2390–7.
Part II
Solid Tumors
Molecularly Targeted Therapies for Astrocytomas Ian F. Pollack
Introduction Astrocytomas constitute the largest group of central nervous system (CNS) neoplasms during childhood, and incorporate tumors with diverse histological features and biological characteristics (Pollack 1994). Whereas patients with superficial low-grade gliomas are often cured with surgery alone, patients with deep-seated or malignant gliomas continue to have a suboptimal prognosis, despite recent improvements in surgery and adjuvant therapy. The prognosis remains particularly grim for patients with malignant gliomas, such as glioblastoma multiforme (GBM) and anaplastic astrocytoma, which generally lead to death within several years after diagnosis (Finlay et€ al. 1995). The poor response of these tumors to conventional therapies reflects a resistance of malignant glioma cells to undergo apoptosis in response to DNA damage, which may result from mutations of tumor suppressor and cell cycle control genes and aberrant activation of growth and survival signaling pathways. Although molecular pathways leading to tumorigenesis have been clearly established for adult malignant gliomas (Louis 1997), the involvement of such pathways in pediatric glial neoplasia remains inferential. For example, because in adult gliomas, platelet-derived growth factor receptor (PDGFR) and epidermal growth factor receptor (EGFR) have been observed to play important roles in tumor proliferation, and have therefore constituted logical targets for molecularly targeted therapies, such approaches have concurrently been explored in pediatric gliomas. This chapter reviews the molecular features of childhood gliomas associated with disease progression and prognosis, and discusses recent molecularly targeted therapeutic strategies.
I.F. Pollack (*) Department of Neurosurgery, Children’s Hospital of Pittsburgh, 3705 Fifth Avenue, Pittsburgh, PA, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_12, © Springer Science+Business Media, LLC 2010
231
232
I.F. Pollack
Molecular Features of Tumorigenesis Low-Grade Gliomas Gliomas are broadly classified as low-grade (grades 1 and 2) and high-grade (grades 3 and 4) (Kleihues et€al. 1993). Juvenile pilocytic astrocytomas are the most common low-grade gliomas, and frequently occur in the cerebellar and cerebral cortices, where they are generally amenable to complete surgical resection. However, these lesions can arise from midline structures, such as the hypothalamus, where they are not amenable to gross total resection without unacceptable morbidity. Studies within the last few years have demonstrated that pilocytic astrocytomas commonly exhibit molecular changes leading to constitutive activation of the BRAF gene, which may constitute a promising therapeutic target (Jones et al. 2008). Although other consistent molecular abnormalities have not been observed in these tumors, a subset of pilocytic astrocytomas exhibits deletions of the chromosome 17q (Von Deimling et€al. 1993a), the site of the neurofibromin gene that is commonly altered in patients with NF1. Patients with NF1 have a significantly increased frequency of pilocytic astrocytomas (Guttmann et€al. 2000), and NF1-deleted transgenic mice have been used as an animal model for low-grade gliomas (Dasgupta et€al. 2005). However, an association between NF1 loss and tumor development in sporadic pilocytic astrocytomas has not been confirmed. Recently, Wong et€al. (2005) using microarray-based screening identified in a small institutional series of low-grade gliomas, identified a series of genetic alterations in tumors that had progressed after initial therapy. Non-pilocytic low-grade astrocytomas include a diverse group of lesions, the most common being grade II fibrillary astrocytomas, which resemble the infiltrative low-grade gliomas commonly seen in adults. Whereas adult low-grade gliomas often exhibit deletions of chromosome 17p, the site of the TP53 gene, as well as mutations of the TP53 gene (James et€al. 1989), this pattern has not been frequently observed in childhood lesions (Lang et€al. 1994; Griffin et€al. 1988; Cheng et€al. 2000). These findings suggest that lesions in these two age groups may arise from a different series of molecular events. This distinction also fits with the dramatically different prognoses observed in childhood versus adult low-grade gliomas. Whereas the latter frequently progress to grade III and IV lesions over time, this pattern of disease progression is less common in childhood gliomas.
High-Grade Gliomas Pediatric high-grade astrocytomas include anaplastic astrocytoma and oligoastrocytoma (grade III) and GBM (grade IV). These lesions have historically had a poor prognosis with conventional therapies, including surgery, irradiation, and cytotoxic chemotherapy (Finlay et€al. 1995). In contrast to the extensive efforts that have been directed at characterizing the patterns of genomic abnormalities in adult
Molecularly Targeted Therapies for Astrocytomas
233
high-grade gliomas (Louis 1997; Ichimura et€al. 2000), there is comparatively little information regarding the molecular features of pediatric malignant gliomas. One clinical pathway that has been observed in adults involves so-called “de€novo” or primary development of a glioblastoma, which is frequently associated with amplification and/or rearrangement of the EGFR gene, but without mutation of the TP53 gene. These tumors also characteristically show loss of chromosome 10q involving the region of the PTEN gene (von Deimling et€al. 1993b; Louis 1997; Collins 1999). Studies from our group and others indicate that EGFR amplification and PTEN deletions are observed in only about 10% of pediatric malignant gliomas, suggesting that this pathway is less commonly implicated than in adults (Cheng et€al. 1999; Raffel et€al. 1999; Sung et€al. 2000; Pollack et€al. 2006a). In a recent analysis of the cohort from the Children Cancer Group 945 (CCG-945) study, mutations of PTEN were observed in only 1 of 62 tumors, and amplification of EGFR was almost equally uncommon (Pollack et€al. 2006a). However, expression of high levels of the EGFR protein product was commonly observed (Bredel et€al. 1999b), suggesting that this receptor may play a role in tumor growth, despite the lack of amplification. Similarly, loss of heterozygosity involving chromosome 10 in the vicinity of the PTEN locus was observed in a subset of these tumors (Pollack et€al. 2006a), suggesting that PTEN may be inactivated by a mechanism other than mutation. More recently, studies in adult primary malignant gliomas indicate that these tumors may include several molecularly defined subsets of lesions, although the applicability of these subsets to the childhood age group remains uncertain (Verhaak et al. 2010). A second clinical pathway of tumorigenesis involves the development of a malignant astrocytoma from a pre-existing low-grade glioma, which evolves into an anaplastic astrocytoma, and ultimately becomes a GBM (von Deimling et€ al. 1993b; Louis 1997; Collins 1999). Such tumors characteristically exhibit mutations of the TP53 gene, overexpression of PDGFRa, and a low frequency of EGFR amplification. It is of interest that recent analyses of childhood malignant gliomas have identified TP53 mutations in 40% of tumors and an even higher incidence in grade 4 lesions (Sung et€al. 2000; Pollack et€al. 2001b), comparable to the frequency in secondary adult malignant gliomas. However, such lesions appear to arise de€novo in children, rather than progressing from a pre-existing low-grade lesion, which calls into question whether pediatric malignant gliomas are truly analogous to the adult secondary gliomas from a molecular and biological standpoint. In that regard, our observation in both an institutional cohort (Pollack et€al. 1997) and the CCG-945 study group (Pollack et€al. 2002) of a worse prognosis in children whose malignant gliomas had TP53 mutations than those with non-mutated tumors contrasts with the lack of such an association in adult lesions. In addition, secondary adult malignant gliomas have been noted to frequently have mutations in the IDH1 gene, changes that are relatively uncommon in pediatric malignant gliomas (Parsons et al., 2008; Hartmann et al., 2009), further suggesting that these groups of tumors are biologically distinct. A third group of adult gliomas arises by a distinct pathway, with characteristic deletions of chromosomes 1p and 19q (Ino et€ al. 2000). Such tumors often have oligodendroglial features and carry a better prognosis than other groups of malignant
234
I.F. Pollack
gliomas, although the genetic loci that confer the more favorable prognosis remain undefined. Although 1p and 19q alterations are observed in about 20% of childhood malignant gliomas, a favorable association between 1p/19q loss and outcome was not observed in the CCG-945 cohort (Pollack et€al. 2003). One consequence of the genetic alterations seen in malignant gliomas is that these tumors exhibit abnormalities in signal transduction that drive invasiveness, proliferation, and cell survival. A significant percentage of tumors exhibit autocrine stimulation via receptors for EGF, PDGF, (Nistér et€al. 1991), Â�insulin-like growth factor (Glick et€al. 1989), and fibroblast growth factor (FGF) (Gross et€al. 1990) by producing both the growth factor and its receptor. These tumors also promote their growth by paracrine interactions, secreting proteins, such as vascular endothelial growth factor (Plate et€al. 1992), FGF, and PDGF, which stimulate adjacent tumor cells and support proliferation of endothelial cells, contributing to an angiogenic phenotype. Cell survival is further promoted by dysregulation of PI3K/Akt signaling, stimulated both by upstream receptor signaling and loss of PTEN expression, disinhibiting Akt activation. As with the above molecular data, it is important to recognize that these observations have largely been made in adult high-grade gliomas, and work is ongoing to determine whether similar changes occur in pediatric gliomas. In parallel with their autocrine and paracrine stimulation of cell proliferation, malignant gliomas in adults characteristically exhibit dysregulation of cell cycle control gene products as a result of homozygous deletion, mutation, or epigenetic silencing of p16 or p15, which encode inhibitors of cyclin-dependent kinases; amplification of cyclin-dependent kinase 4; or deletion or mutation of the retinoblastoma susceptibility gene RB1 (Collins 1999). Such molecular alterations have also been observed in pediatric malignant gliomas (Cheng et€al. 1999). Another molecular feature associated with malignant gliomas in both children and adults is the acquisition of a drug-resistant phenotype, which in part accounts for the poor response of these tumors to many conventional chemotherapeutic agents, particularly alkylating agents. Several studies have noted that both adult and pediatric malignant gliomas commonly express high levels of alkylguanine DNA alkyltransferase (AGT), the proximal mechanism of resistance to alkylators, such as the nitrosoureas and temozolomide (Friedman et€al. 1998b; Pollack et€al. 2006b). Recent studies have noted that the response of these tumors to such agents correlates with levels of AGT expression and activity (Jaeckle et€al 1998; Esteller et€al. 2000; Hegi et€al. 2005; Pollack et€al. 2006b). The association between tumor AGT expression status and outcome was demonstrated in children with malignant gliomas in the CCG-945 cohort (Pollack et€ al. 2006b), in which those with high levels of AGT expression had a substantially worse progression-free survival after nitrosourea-based therapy than those with low levels. These results were independently validated in the ACNS0126 study of the Children’s Oncology Group (COG) among children treated with temozolomide during and after irradiation. In addition to AGT, a number of other factors have been associated with resistance to temozolomide, including deficiency of mismatch repair (MMR) protein expression (Friedman et€al. 1998b) and alterations of base-excision repair function.
Molecularly Targeted Therapies for Astrocytomas Table€1╅ Molecularly targeted therapies Inhibitors of growth factor receptors PDGF inhibition ╅ Imatinib ╅ Sunitinib ╅ AZD2171 EGFR inhibition ╅ Gefitinib ╅ Erlotinib ╅ Lapatanib ╅ Cetuximab Inhibitors of downstream signaling Farnesyl transferase inhibition ╅ Tipifarnib ╅ Lonafarnib PKC inhibition ╅ Enzastaurin MAPK cascade/Raf inhibition ╅ Sorafenib PI3K/Akt pathway (mTOR) inhibition ╅ Sirolimus ╅ Temsirolimus ╅ Everolimus Angiogenesis inhibition VEGFR/multitargeted kinase inhibition ╅ SU5416 ╅ Valatinib ╅ ZD6474 ╅ Sunitinib ╅ CEP-7055 ╅ Sorafenib ╅ AZD2171 VEGF inhibition ╅ Bevacizumab Integrin inhibition ╅ Cilengitide Other putative antiangiogenics ╅ Thalidomide ╅ Lenalidomide ╅ Cox-2 inhibitors ╅ Rofecoxib ╅ Endostatin ╅ Angiostatin
235
Counteracting drug resistance phenotype AGT inhibition â•… O6-Benzylguanine PARP Inhibition â•… INO-1001 ABT888 HDAC Inhibition â•… SAHA â•… Depsipeptide â•… Valproic acid HSP Inhibition â•… Geldanamycin â•… 17-AAG Proteosome inhibition â•… Bortezomib Cell cycle modulation â•… UCN-01 â•… CYC202 â•… Flavopiridol Immunologic or ligand-based therapies Vaccine-based targets Antibody therapy â•… Cetuximab â•… Bevacizumab Radiolabeled intracavitary antibody therapy â•… Anti-EGFR â•… Anti-Tenascin Immunotoxins â•… Tf-CRM107 â•… IL13-PE38QQR â•… TP-38 â•… IL4-PE
Molecularly Based Therapeutic Strategies The availability of novel agents for therapy of gliomas has dramatically increased during the last decade. The portfolio of available compounds can be broadly categorized into the following major groups: (1) inhibitors of growth factor receptors expressed on glioma cells; (2) inhibitors of signal transduction pathway intermediates;
236
I.F. Pollack
(3) inhibitors of angiogenesis; (4) agents that counteract drug resistance; and (5) immunological or ligand-based agents targeted against molecules overexpressed in gliomas (Table€ 1). Illustrative examples of such agents are discussed subsequently, and results (in most cases preliminary) are presented, where available.
Inhibitors of Growth Factor Receptors Because growth factor receptors, particularly EGFR and PDGFR, were recognized to be commonly overexpressed or amplified in malignant gliomas, these have constituted a major target for drug development. A variety of strategies have been used in preclinical studies for growth factor receptor blockade, including neutralizing antibodies, antisense techniques, and dominant negative mutants. For example, Nitta and Sato (1994) observed that c-sis antisense oligonucleotides inhibited growth in€vitro of A172 glioma cells, which express large quantities of PDGF-BB and the b receptor. Similarly, growth of U87 and U343 glioma cells transfected with a dominant negative construct for PDGF-A, was suppressed in€vitro, and these cells exhibited diminished tumorigenicity in a nude mouse model (Shamah et€al. 1993). Our previous studies have also demonstrated significant antiproliferative activity in€ vitro using a PDGF neutralizing antibody (Pollack et€ al. 1991). Other groups have noted significant activity of antibodies designed to neutralize EGFR or to target the most common, constitutively active EGFR mutant (EGFR vIII) (Mishima et€al. 2001). Similarly, in non-CNS tumors, monoclonal antibodies directed against EGFR (e.g., C225, Cetuximab, ImClone Systems) have demonstrated efficacy in preclinical models as well as clinical studies (Baselga et€al. 2000). Although the above results confirm that blocking critical growth factor pathways can interfere with cell proliferation in selected tumor cell lines, the applicability of these results for glioma therapeutics is limited by the challenges imposed by the blood–brain barrier, which impairs drug access to the tumor after systemic administration. An alternative approach involves the use of “small molecule” inhibitors of receptor signaling. During the last several years, molecular modeling approaches have allowed the development of “designer inhibitors,” which selectively target one or more tyrosine kinase receptors by competitive inhibition of ATP-binding sites involved in receptor activation. A brief summary of several promising agents is provided below.
Platelet-Derived Growth Factor Receptor as a Target PDGF was originally identified as a potent mitogen for fibroblasts, glial cells, and smooth muscle. The ligand is a disulfide-linked dimer composed of A, B, C, and D polypeptide chains, the B chain exhibiting homology to the v-sis oncogene isolated from simian sarcoma virus-transformed cells (Lokker et€ al. 2002). The various PDGF isoforms (AA, AB, BB, CC, DD) bind with differential affinity to two cellsurface PDGF receptors (Nistér et€al. 1991; Lokker et€al. 2002). Concurrent expression
Molecularly Targeted Therapies for Astrocytomas
237
of one or more of these ligands and their receptors has been observed in a high percentage of malignant gliomas and the acquisition of PDGF receptor overexpression represents a potential step in the transition from grade II to grade III gliomas in adults (Hermanson et€al. 1996), which is supported by studies in transgenic mouse glioma models (Dai et€al. 2001) showing that overexpression of PDGF alone leads to low-grade gliomas, whereas combined PDGF overexpression with other genetic changes leads to a more malignant phenotype. From a mechanistic perspective, PDGF and its receptors have been implicated in autocrine and paracrine growth stimulation in malignant gliomas (Nistér et€al. 1991; Lokker et€al. 2002). Ligand binding to the receptor leads to receptor dimerization, autophosphorylation of tyrosine residues, and phosphorylation of a series of signaling intermediates, such as RAS, phospholipase C-gamma (PLCg), and phosphatidylinositol 3-kinase (PI3K). Previous studies have indicated that PDGF is a potent mitogen for malignant glioma cells (Pollack et€ al. 1991), and that inhibition of PDGFR activation inhibits glioma growth in preclinical models (Nitta and Sato 1994). In addition, because of the role of PDGF in supporting glioma-induced angiogenesis (Wang et€ al. 1999), inhibition of PDGFR may provide a means for simultaneously blocking tumor growth and angiogenic activation. A number of PDGFR inhibitors are in various phases of clinical development. One agent that has been extensively tested is STI571 (CGP57148B, also known as Gleevec or Imatinib, Novartis Pharmaceutical Corp). This compound initially became a focus of clinical interest because of its potent inhibition of BCR-ABL, associated with Philadelphia chromosome-positive leukemias. Because of the dramatic hematologic and cytologic responses observed in preliminary studies (Druker et€al. 2001), phase III trials were conducted, which demonstrated superiority of this agent in terms of complete cytogenetic responses and freedom from disease progression in comparison to conventional therapy (O’Brien et€al. 2003). A second target for STI571 is c-kit, constitutively activated in a substantial percentage of gastrointestinal stromal tumors. Initial studies demonstrated striking efficacy in patients with advanced disease (Joensuu et€al. 2001). Because imatinib was also extremely potent in blocking PDGFR signaling (Buchdunger et€al. 2000), attention was focused on the application of this agent in tumors with PDGFR-driven proliferation. Preliminary studies in glioma cell lines demonstrated inhibition of proliferation in€vitro and delay of tumor growth in€vivo (Kilic et€al. 2000), although the results were less striking than those noted earlier for c-kit and BCR-ABL-dependent tumors. Based on these results, dose escalation studies in children and adults with malignant gliomas were initiated by several cooperative groups, including the North American Brain Tumor Consortium (NABTC) and Pediatric Brain Tumor Consortium (PBTC). An unexpected finding from several of these studies was that a subset of patients experienced intratumoral hemorrhage during treatment (Wen et€al. 2006). The etiology of this effect remains to be determined, but one potential mechanism relates to known effects on perivascular cells (Pietras et€al 2001). This observation, coupled with the fact that a paucity of responses or long-term survivors were noted in the pediatric phase I trial (Pollack et€al. 2007) or the NABTC study, limited enthusiasm for a pediatric phase II study.
238
I.F. Pollack
In contrast to these observations, a study from Europe detected intratumoral hemorrhage in only 1 of 51 patients, and four patients had partial radiological responses (Raymond et€al. 2004). In a second study, prolonged disease control was achieved in 3 of 6 high-grade glioma patients younger than 45€years of age versus only one of nine older than 45€years (Katz et€al. 2004). Combinations of STI571 and conventional chemotherapeutic agents have also been examined. Although the optimal agent to employ in this regard remains uncertain, promising results have been reported with hydroxyurea (1,000€mg/day) combined with 400€mg/day of STI571, with objective disease regression in 5 of 26 patients (Dresemann 2005). In a separate phase II study of this combination conducted by Reardon et€ al. (2005a), 9% of GBM patients achieved radiographic responses, while 35% had stable disease. The progression-free survival rate at 6€months was 26.3%, which compared favorably with results using single-agent cytotoxic chemotherapy. Combinatorial application for pediatric gliomas has, to date, not been systematically evaluated.
Epidermal Growth Factor Receptor as a Target EGFR, also known as erbB1, the protein product of the cellular homolog of the v-erbB oncogene, is one of a family of receptors that also includes erbB2 (HER2/neu), erbB3 (HER3), and erbB4 (HER4). The ligands for these receptors include EGF and TGFa, among others (Schlessinger 2000). EGFR is normally activated by ligand binding to its extracellular receptor domain (Ullrich and Schlessinger 1990). However, in some tumors with EGFR amplification, the gene is also rearranged, leading to constitutively active mutants (Libermann et€al. 1985; Wong et€al. 1992). The most common mutant, the EGFRvIII variant, is caused by deletions of exons 2 to 7. Because activated EGFR induces tyrosine phosphorylation of substrates that contribute to cell proliferation, excessive activation of this protein, either by ligand binding or mutation-induced constitutive signaling, may provide cells with a growth advantage under certain conditions. Although pediatric malignant gliomas do not share the high incidence of EGFR amplification noted in primary adult malignant gliomas, they commonly exhibit high levels of EGFR expression (Bredel et€al. 1999b), which suggests that excessive EGFR signaling may contribute to cell proliferation in these tumors as well. The important role of EGFR in brain tumor development is supported by the fact that targeted inhibition of this receptor, using dominant negative constructs, interfered with glioma proliferation (O’Rourke et€al. 1997). Similarly, antibody-based therapies targeted at wild-type or mutant EGFR family members have been shown to have efficacy in preclinical glioma models (Mishima et€al. 2001), which provided a rationale for small molecule-based strategies. One such agent to be tested in pediatric malignant gliomas was ZD 1839 (Iressa, Gefitinib, AstraZeneca), which is effective in blocking EGFR autophosphorylation and EGFR-dependent cell signaling in€vitro in cell lines that rely heavily on EGFR activation for proliferative stimulation, and tumor growth in€vivo in EGFR-dependent xenograft model systems (Wakeling et€al. 2002). Initial clinical data indicated that
Molecularly Targeted Therapies for Astrocytomas
239
gefitinib was well tolerated at effective doses (Albanell et€al. 2002), with the most common toxicity being an acneiform skin rash, and diarrhea being a common doselimiting toxicity. Results from several clinical studies for non-central nervous system solid tumors demonstrated activity of gefitinib as a single agent (Baselga et€al. 2002; Fukuoka et€al. 2003), although phase III studies in patients with advanced small cell lung cancer failed to demonstrate a convincing benefit of adding this agent to regimens, including either gemcitabine/cisplatin or paclitaxel/carboplatin (Herbst et€al. 2004). Although these results were initially a source of disappointment, recent reports have demonstrated that response to gefitinib is strongly influenced by tumor EGFR status, with a high percentage of objective responses among patients with gain-of-function mutations of the EGFR gene (Lynch et€al. 2004). These observations emphasize the importance of studies of tumor genotype and phenotype in considerations of study design and response analysis. With these issues in mind, Phase I/II studies of gefitinib were initiated within the NABTC (for adults with malignant glioma) and the PBTC (for children with recurrent malignant glioma and newly diagnosed brainstem malignant glioma). These studies included analyses of receptor expression and mutational status, in an effort to correlate molecular features with treatment responses. In the phase II component of the NABTC study, 7 of 55 patients were noted to have partial tumor regression, although the median times to progression were not superior to historical controls (Lieberman et€al. 2004). In the pediatric study (PBTC007), gefitinib was administered with irradiation to patients with newly diagnosed malignant gliomas with separate strata for children with brainstem gliomas and non-brainstem high-grade gliomas. After completing the phase I component of the study, which established the maximal tolerated dose, a phase II component was initiated. The results are currently pending. Clinical trials for gliomas have also been completed using other EGFR inhibitors. As with gefitinib, OSI-774 (Erlotinib, Tarceva, Genentech) reversibly inhibits EGFR by competition with the ATP binding site (Pollack et€al. 1999). In phase I studies, toxicities were comparable to ZD1839, with rash and diarrhea as common events (Hidalgo et€ al. 2001). Phase II studies in patients with non-CNS tumors demonstrated activity in terms of tumor response (Perez-Soler 2004), as did a trial in adults with malignant glioma, conducted by the NABTC (Prados et€al. 2006). However, as with gefitinib, median time to progression for the latter cohort was no better than historical control data. A phase I trial of erlotinib with RT in patients with GBM determined the toxicity and maximum-tolerated dose (Krishnan et€al. 2005). The median time to progression was 161€days, and the median survival time was 386€days. The combination of erlotinib with temozolomide and concurrent RT was also well tolerated, although it is unclear if this improved outcome (Peereboom et€al. 2006). A phase I study of this agent alone and in conjunction with temozolomide for children with recurrent disease of a variety of histologies was conducted in the Children’s Oncology Group (ADVL0214), although activity was insufficient to support a phase II study of this combination in gliomas. Because responses were observed to both gefitinib and erlotinib in studies for adults with malignant gliomas, efforts were made to determine whether a molecular
240
I.F. Pollack
signaling profile could distinguish responders from non-responders. Although one analysis failed to note a clear association between EGFR status and response (Lassman et€al. 2005), a subsequent study noted that responses were typically seen in tumors that had amplification of EGFR and expression of the EGFRvIII variant in conjunction with preservation of PTEN (Mellinghoff et€al. 2005), thereby suppressing constitutive Akt activation. An attempt to correlate response with EGFR amplification and expression is also planned in the pediatric glioma study. One other selective, reversible EGFR inhibitor that is in early phase trials is GW572016 (Lapatinib, GlaxoSmithKline, Research Triangle Park, NC) (Burris 2004), which has demonstrated activity against erbB2-expressing tumors (Rusnak et€ al. 2001), and has been tested in a study of the PBTC (PBTC-016), including patients with recurrent malignant gliomas, ependymomas, and medulloblastomas. After completing the phase I portion of the trial to define the maximal tolerated dose (MTD), a phase II study was initiated, which incorporated a molecular biology component for patients in whom a resection was planned. In such patients, lapatinib was administered for several days prior to surgery and then a portion of the resected tumor tissue was analyzed for suppression of the receptor target and other downstream molecular features. The results of this study will soon be available.
Inhibitors of Downstream Signaling Inhibition of intermediate and downstream components of growth factor signaling pathways, such as RAS, PI3K, PKC, and mTOR, is a promising strategy for interfering with the proliferation of malignant gliomas and other brain tumors (Pollack et€al. 1998). Because the number of inhibitors and targets that are being examined preclinically and in early clinical studies is vast, we focus below on the approaches that have been most intensively considered to date.
Inhibition of RAS Processing The response of cells to growth factors, such as EGF and PDGF, is mediated by cell-surface receptors that contain an extracellular domain that interacts with a ligand, a transmembrane domain that anchors the receptor to the cell membrane, and an intracytoplasmic domain that interacts with downstream signaling components. Activation induces conformational changes within the receptor and/or receptor cross-linking, and phosphorylation of tyrosine residues within their cytoplasmic domains (Ullrich and Schlessinger 1990). Phosphorylation exposes Src homology 2 (SH2) domains, binding sites for intracellular adapter molecules, such as Grb2 (McCormick 1993). The adapter proteins in turn associate with guanine nucleotide exchange factors (Pawson 1995), which facilitate the activation of RAS family guanine nucleotide triphosphatases by exchanging GDP for GTP. Excessive activation of
Molecularly Targeted Therapies for Astrocytomas
241
these pathways has been implicated in the development of a variety of human tumors. In certain tumor types, mutation of one of the RAS genes to a constitutively active protein product has been associated with tumorigenesis (Peddanna et€ al. 1996). Although such mutations are rare in gliomas, RAS activity is markedly elevated as a result of deregulation of upstream signaling elements, such as growth factor receptors (Prigent et€al. 1996; Guha et€al. 1997). These observations suggest that targeted inhibition of RAS-dependent signaling may constitute a therapeutically useful strategy for malignant gliomas. Because RAS is synthesized as a pro-peptide and undergoes post-translational modifications to associate with the inner surface of the plasma membrane, one strategy for blocking RAS activity has involved interference with those modifications. Inhibition of farnesylation has been one of the principal strategies used to date (Bredel et€al. 1998; Rowinsky et€al. 1999). Although farnesyl transferase inhibitors (FTIs) inhibit RAS farnesylation, their antiproliferative effects are not exclusively due to the effects on RAS as they target other G-proteins, including Rho-B (Lebowitz and Prendergast 1998; Jiang et€ al. 2000). Feldkamp et€ al (1999) and Bredel et€al. (1998) demonstrated that astrocytomas are amenable to growth inhibition by FTI’s, through a combination of antiproliferative and pro-apoptotic effects. Based on encouraging preclinical data, clinical trials have been initiated with several FTIs. Tipifarnib (R115777, Zarnestra; Johnson and Johnson) is a non-peptidomimetic FTI with in€ vivo activity against many human cancer cell lines and xenograft models (End et€ al. 2001). Although some activity was observed in patients with hematological malignancies (Alsina et€al. 2004), the results of Phase II and III trials in adults with solid tumors have been less promising (Rao et€ al. 2004; Van Cutsem et€al. 2004). In a recent phase II trial in adult malignant gliomas, tipifarnib showed a low level of activity, with 5 of 89 patients with GBM achieving partial responses (PRs) (Cloughesy et€al. 2006). Similar results were observed in a pediatric trial of the COG (ACNS0226) for children with progressive tumors. Among 31 evaluable patients with progressive high-grade glioma, 1 patient had a confirmed PR and 3 patients continued on therapy for at least 4 cycles. Among the 35 patients with brainstem glioma, 1 patient had a PR, and 4 patients continued on therapy with stable disease for at least 4 cycles (11%). The 6-month progressionfree survival was 14â•›±â•›6% for high-grade glioma and 3â•›±â•›3% for brainstem glioma. A trial of this agent with radiotherapy was recently completed for children with malignant brainstem gliomas (PBTC-014). Results are pending. A second FTI that was examined in the PBTC (PBTC003) was Lonafarnib (SCH66336, SARASARTM, Schering-Plough) (Kieran et€al. 2007), which has previously been evaluated in a variety of tumor types alone and in conjunction with conventional chemotherapeutic agents (Sharma et€al. 2002; Kim et€al. 2005). Fifty-three children with diverse tumor tumors were treated in the PBTC study: high-grade glioma (19), brain stem glioma (7), medulloblastoma (3), PNET (4), ependymoma (9), refractory low-grade glioma (5), meningioma (3), sarcoma (1), and gliomatosis cerebri (1). Forty-eight patients were evaluable for response: one with anaplastic astrocytoma demonstrated a PR lasting 13 courses and 12 demonstrated stable disease for 2 to 18 courses. Seven patients remained on therapy for at least 1€year.
242
I.F. Pollack
Inhibition of Protein Kinase C Another important signaling pathway in gliomas involves the activation of protein kinase C (PKC), which comprises a family of serine/threonine kinases that constitute an element of the signaling cascade of several growth factors that stimulate glioma cell proliferation, such as EGF and PDGF (Fields et€al. 1990; Nishizuka 1992). Activation of PKC leads to diverse biological effects, depending upon which isoforms are involved. These include phosphorylation of other effectors, such as RAF and MAPK, and influencing the activation status of RAS ((Kolch et€ al. 1993; Marais et€ al. 1998), which together contribute to the transduction of a proliferative signal to the nucleus. Levels of PKC activity have been noted to correlate with proliferative status in neoplastic astrocytes (Couldwell et€al. 1991). In addition, astrocytoma cells express levels of PKCa and e that are up to tenfold higher than in normal astrocytes. Glioma mitogens, such as PDGF and EGF, produce elevations in PKC activity that parallel increases in DNA synthesis (Couldwell et€al. 1992). In addition, pharmacological and antisense agents that target PKC isoforms overexpressed in astrocytomas diminish glioma proliferation in€ vitro (Pollack et€al. 1996; Yazaki et€al. 1996) and induce apoptosis (Bredel et€al. 1999a). LY317615 (enzastaurin, Eli Lilly and Company), is a bisindolylmaleimide-derived PKC inhibitor that has demonstrated antiangiogenic activity (Keyes et€al 2004) as well as antitumor activity in U87 glioma xenografts and a variety of other solid tumors (Graff et€al. 2005). Enzastaurin and its active metabolites prevent substrate phosphorylation by competing with the enzyme’s ATP binding site. The in€ vitro IC50s for PKCb inhibition is approximately 5€nM, whereas inhibition is 3- to 8-fold less potent for PKCa and 10- to 20-fold less potent for PKCg and PKCe. Additionally, this compound inhibits signaling through the PI3 kinase/AKT signaling pathway. Accordingly, inhibition of the PKC and AKT signaling pathways by enzastaurin suppresses phosphorylation of glycogen synthase kinase 3b, induces apoptosis and inhibits proliferation in cultured cell lines from a variety of human cancers, including gliomas, and also has antiangiogenic activity. In single-agent phase I studies, enzastaurin was well tolerated (Carducci et€ al. 2006). Fine and colleagues recently reported preliminary results of a phase II trial of enzastaurin in adult patients with recurrent high-grade gliomas (Fine et€al. 2005). Of the 87 patients evaluable for response, 22% were reported to have objective radiographic responses, and 5% had stable disease. The drug was well tolerated, and a maximum tolerated dose was not determined. Intra-tumoral hemorrhages were noted in seven patients; however, six of the seven had progressive disease at the time of the bleed. Based on these encouraging results, a phase III study was initiated in adult patients with GBM and lymphoma. A phase I study of this agent in children with recurrent brain tumors is also in progress (PBTC023).
MAPK Cascade Inhibitors Following the activation of the membrane-associated components of the various signaling pathways, downstream signals reach the nucleus by a variety of mechanisms.
Molecularly Targeted Therapies for Astrocytomas
243
One of the most critical pathways involves the mitogen-activated protein kinase (MAPK) cascade. This cascade involves multiple separate protein kinases. The most proximal kinase in the cascade is the RAF (MAPKKK) family, which includes at least three members. RAF, a serine/threonine kinase, is recruited to the cell membrane, stabilized by interaction with other proteins, and phosphorylated to an active form (Freed et€ al. 1994). Activated RAF phosphorylates and activates MAP/ERK kinase (MEK, also known as MAPKK), which subsequently activates MAPK (also known as ERK (extracellular signal-regulated kinase)) (Howe et€al. 1992). MAPKs are serine/threonine kinases that activate a number of additional downstream mediators that regulate transcription, protein translation, and cytoskeletal rearrangement. In an effort to block this signaling pathway, a number of agents that function as RAF kinase inhibitors have been developed. These include sorafenib (BAY 43-9006; Bayer Pharmaceuticals) (Wilhelm et€al. 2004; Jane et€al. 2006), which has been a focus of combinatorial studies for malignant gliomas within the NABTC. Results are pending. This agent has also been examined in a phase I study for recurrent pediatric malignancies in the Children’s Oncology Group, and a phase II study for brain tumors is under consideration. Because this agent also inhibits VEGF receptor activation, it may have applicability as an antiangiogenic agent (discussed below), independent of any direct cytotoxic activity against glioma cells.
Inhibitors of PI3K/Akt Pathways Another major pathway implicated in growth factor receptor-mediated signaling involves activation of AKT via PI3K. PI3K is a phospholipid kinase that contains a regulatory subunit, p85, and a catalytic subunit, p110. Upon cell-surface activation, PI3K phosphorylates phosphatidylinositol 4,5-biphosphate (PI4,5P2, PIP2) to form PI3,4P2, and PIP3. PIP3 leads to translocation of PDK1 and AKT. Akt is a serine/ threonine kinase that is phosphorylated at threonine 308 by PDK1 and at serine 473 by a second kinase. Activated AKT phosphorylates several proteins involved in cell survival signaling, such as BAD, forkhead transcription factor, glycogen synthase kinase, and the mammalian target of rapamycin (mTOR) (Cross et€ al. 1995; Cardone et€al. 1998). Under normal conditions, Akt activation is inhibited by PTEN, a phosphatase that converts PIP3 to PIP2 (Li and Sun 1998; Cantley and Neel 1999). However, PTEN is mutated in at least 40% of grade IV gliomas (Parsons et al. 2009), particularly in primary GBMs. The importance of this pathway in glioma development is highlighted by the fact that transfer of a wild-type PTEN gene to the PTEN-deleted U87 glioma cell line suppressed tumor growth, leading to cell cycle arrest (Li and Sun 1998). Conversely, greater PI3K/AKT activity has been associated with resistance to irradiation (Nakamura et€al. 2005). Based on these observations, inhibitors of the PI3K/AKT pathway appear to be logical therapeutic agents for GBM. To date, inhibitors of PI3K, such LY294002 (Nakamura et€al. 2005), have demonstrated promising activity in preclinical models,
244
I.F. Pollack
but the profile of toxicity has precluded their clinical use. Accordingly, interest has focused on blocking downstream targets of the Akt pathway. In particular, several inhibitors of the mTOR are being investigated in clinical trials. These include sirolimus (rapamycin, Rapamune®; Wyeth), temsirolimus (CCI-779; Wyeth), and everolimus (RAD-001, everolimus, Certican®; Novartis), all of which inhibit glioblastoma proliferation in culture and intracerebral xenografts. Temsirolimus demonstrated modest activity in recurrent gliomas in recent phase I/II studies (Chang et€al. 2005). Everolimus has also undergone phase I testing for children with recurrent malignant brain tumors and a phase II study is being developed. In a recent phase I study, administration of temsirolimus to patients with progressive GBM was well tolerated, with some objective responses (Galanis et€al. 2005). High levels of phosphorylated p70s6 kinase, as determined by immunohistochemistry in baseline tumor samples, seemed to correlate with response to treatment.
Inhibition of Angiogenesis VEGFR Inhibition In addition to direct effects of growth factor receptor inhibition on tumor proliferation, there is compelling evidence that these agents may have indirect effects on angiogenesis, which may be relevant for brain tumor therapeutics. Recent studies have confirmed that the secretion by tumor cells of vascular endothelial growth factor (VEGF), the most potent endothelial cell mitogen, depends heavily on EGFRmediated signaling (Maity et€ al. 2000) and that a significant component of the therapeutic efficacy of EGFR-targeting agents reflects this secondary effect on tumor angiogenesis (Ciardiello et€al. 2001; Huang et€al. 2002). Similarly, PDGF has been demonstrated to stimulate tumor angiogenesis (Yancopoulos et€ al. 2000) in addition to supporting the growth and survival of vascular pericytes and promoting glioma VEGF secretion (Wang et€al. 1999). In view of these paracrine interactions, an important element in evaluating the therapeutic utility of EGFR- and PDGFRtargeted inhibitors must focus on understanding the effects of these agents on not only the tumor cells themselves, but also on the surrounding vasculature (Ciardiello et€al. 2001; Pietras et€al. 2001). Because VEGF represents a major stimulatory factor for the initiation of angiogenesis, inhibition of ligand/receptor interactions has been a focus of recent attention for malignant gliomas. SU5416 (Sugen, Pfizer) was the first antiangiogenic agent studied in the PBTC, although development plans were ultimately halted after a negative phase III study for patients with colorectal carcinoma. Subsequently, a number of other agents have been examined in adult patients with GBM, including PTK787/ZK 222584 (valatinib, Novartis Pharmaceuticals and Schering AG), ZD6474 (Zactima™; AstraZeneca Pharmaceuticals), and CEP-7055 (Sanofi-Aventis), based on encouraging preclinical data in glioma xenografts in nude mice (Goldbrunner et€al. 2004; Reardon and Wen 2006). Clinical trials of other multitargeted kinase inhibitors that
Molecularly Targeted Therapies for Astrocytomas
245
block VEGF receptors among other kinase targets, such as sorafenib, AZD2171 (AstraZeneca), sunitinib (SU11248, Pfizer), have also been conducted in adults with malignant gliomas. A phase I study of sorafenib has been conducted in the Children’s Oncology Group, although there was not a specific focus on gliomas, and a phase II study is being considered. A phase I study of AZD2171 is currently ongoing in the Pediatric Brain Tumor Consortium (PBTC021).
VEGF Blockade A particularly promising agent is bevacizumab (Avastin®; Genentech, Inc.), a recombinant, humanized monoclonal antibody targeting VEGF (Ferrera et€ al 2005). This agent has been shown to cause decreased vascular permeability and increased apoptosis in intracranial xenografts of human glioblastoma and synergism has been observed with several chemotherapeutic agents (Ignoffo 2004). In a phase III placebo-controlled trial of 815 patients with metastatic colorectal cancer who were randomized to receive either bevacizumab plus irinotecan/5-FU/leucovorin or irinotecan/5-FU/leucovorin plus placebo, bevacizumab produced a significantly better rate and duration of response compared to placebo (Hurwitz et€ al. 2004). Bevacizumab also prolonged the time to progression compared with placebo for patients with metastatic renal cell cancer (Yang et€al. 2003). Stark-Vance (2005) has also treated adult patients with recurrent malignant glioma with bevacizumab plus irinotecan and reported 1 patient with a complete response (CR), 8 patients with PR, 11 patients with stable disease, among 21 patients treated. On the basis of the preclinical and clinical efficacy demonstrated in adult studies of recurrent solid tumors, a phase II study was initiated for patients with recurrent malignant glioma. In a preliminary communication, 20 of the first 32 patients (63%) reportedly demonstrated an objective response (19 PR and 1 CR) to treatment (Vredenburgh et€al. 2006) with a 6-month progression-free survival rate of 39%, which provided a basis for a host of additional studies in adults. These results also provided a basis for a phase II study of bevacizumab plus irinotecan initiated by the PBTC (PBTC022) for patients with recurrent malignant gliomas and those with brainstem malignant gliomas.
Other Angiogenic Agents In addition to inhibition of promoters of VEGF secretion, such as EGFR and PDGFR, and direct inhibition of VEGF receptors and VEGF itself, evaluation of a number of other agents that have affects on angiogenesis has been conducted in adult and pediatric malignant gliomas. Monotherapy with thalidomide (Thalomid®; Celgene Corporation) has been investigated for the treatment of GBM and malignant brainstem glioma because of its antiangiogenic effects. Although this agent has had
246
I.F. Pollack
only modest activity in this setting, with 4 of 36 objective radiologic responses in one study (Fine et€ al. 2000), more recent studies have examined lenalidomide (CC-5013, (Revlimid®; Celgene Corporation)), a more potent analog of thalidomide. A study of this agent for patients with recurrent brain tumors is being conducted by the PBTC (PBTC-018). Long-term stable disease has been observed in several patients with gliomas, and a phase II study is being considered. In addition, it has been observed that the combination of thalidomide and chemotherapy appears to be more active in patients with recurrent gliomas than either approach alone (Fine et€al. 2003), and combinatorial studies with this agent as well as CC-5013 are being considered. Other agents that are being examined as potential angiogenesis inhibitors include the avb3 integrin inhibitor cilengitide (EMD 121974; EMD Pharmaceuticals) and the previously mentioned PKC inhibitor enzastaurin (LY317615), which among its other effects has been noted to decrease VEGF levels in a mouse tumor model (Taga et€al. 2002; Keyes et€al. 2004). A phase I trial of cilengitide was conducted in the PBTC (PBTC-012), and the MTD was carried forward in a Phase II trial being conducted within the Children’s Oncology Group (ACNS0621). As noted above, a phase I trial of enzastaurin is in progress within the PBTC (PBTC023). Other angiogenic inhibitors of interest include cyclooxygenase 2 (COX-2) inhibitors, based on the association between COX-2 expression and angiogenesis (Reardon and Wen 2006). Both celecoxib (Celebrex, Pfizer Pharmaceuticals) and rofecoxib (Vioxx, Merck) have been combined with conventional chemotherapeutic agents in studies for both children and adults with recurrent brain tumors, including gliomas (Reardon et€al. 2005b; Tuettenberg et€al. 2005). Other agents include endostatin and angiostatin, which are natural product inhibitors of angiogenesis, and Atrasentan, a selective inhibitor of the endothelin A receptor, which has been proposed to be involved in regulating glioma-induced angiogenesis (Barnett et€ al. 2004; Phuphanich et€ al. 2005). In a phase I study of this agent, outcome results were disappointing with a median progression-free survival time of only 1.5€months.
Interference with Growth Factor Receptor and Survival Signaling May Potentiate Other Therapies Although signal transduction inhibition may have activity in decreasing proliferation and inducing apoptosis in a subset of tumors, the efficacy of these approaches may be enhanced by combining them with other therapeutic strategies. Preclinical data suggest potentiation of the activity of cytotoxic drugs by growth factor receptor inhibitors. EGFR-targeted monoclonal antibodies have been noted to enhance the effects of cisplatin (Baselga et€al. 2000), topotecan, gemcitabine (Bruns et€al. 2000), and taxol. In addition, marked radiosensitization has been achieved by EGFRspecific monoclonal antibodies (Huang et€al. 1999; Milas et€al. 2000) and dominant negative transfection (O’Rourke et€al. 1998). Recent studies also suggest that small molecule inhibitors of EGFR kinase activity, such as gefitinib (Ciardiello et€al. 2000;
Molecularly Targeted Therapies for Astrocytomas
247
Huang et€ al. 2002), and PDGFR kinase activity, such as imatinib (Topaly et€ al. 2001; Pietras et€al. 2002), may achieve similar potentiation of conventional therapies. The application of such combinatorial approaches in brain tumor therapeutics is already under way. Both adult and pediatric brain tumor trials have examined the combination of erlotinib and temozolomide, and pediatric studies in brainstem glioma have evaluated the combination of gefitinib and concurrent irradiation (PBTC007), and imatinib immediately following irradiation in newly diagnosed patients (PBTC006), with outcome comparisons versus historical control groups treated with irradiation alone (Pollack et€al. 2007). In addition to the applicability of combining growth factor receptor-targeted inhibitory strategies with conventional chemotherapy, recent studies indicate that combinations of growth factor receptor inhibitors with other molecularly targeted approaches may potentiate efficacy by circumventing resistance mechanisms (Bianco et€ al. 2003). For example, Bianco et€ al. (2003) observed that treatment resistance to EGFR inhibitors in tumors with PTEN deletions could be counteracted by independent inhibition of Akt signaling, suggesting a role for agents, such as rapamycin, that block downstream Akt signaling, in conjunction with EGFRtargeted therapies. In that context, preliminary data from a phase I trial of gefitinib in combination with rapamycin for the treatment of patients with recurrent malignant glioma have been reported (Reardon et€al. 2006b). The optimal agent(s) to combine with individual growth factor receptor inhibitors is a topic of intense research interest. Similarly, combinations of agents may be needed to effectively block angiogenic signaling. The combination of endostatin and SU5416 has been reported to achieve superior tumor growth inhibition in preclinical models, compared to treatment with either agent alone (Abdollahi et€al. 2003). Combinations of antiangiogenic agents with conventional chemotherapeutic agents have also shown promise. For example, patients with recurrent GBM who received thalidomide and carmustine had a response rate of 24%, which compared favorably with carmustine alone (Fine et€al. 2003). Moreover, the combination of thalidomide and temozolomide in patients with GBM was more effective than thalidomide alone with respect to survival and response (Baumann et€al. 2004). Recent studies for patients with recurrent brain tumors, including children, have incorporated multiple antiangiogenic agents in a “cocktail,” in an effort to simultaneously block several targets potentially contributing to tumor-induced neovascularization.
Counteracting the Drug Resistance Phenotype Alkylguanine DNA Alkyltransferase Inhibition Drug resistance, either intrinsic or acquired, is frequently observed in tumors of the CNS, and many different mechanisms have been described. One particularly important mechanism involves the reversal of DNA damage induced by alkylating agents,
248
I.F. Pollack
such as BCNU and temozolomide. These agents cause alkylation at the O6-position of guanine, which can be repaired by alkylguanine-DNA alkyltransferase (AGT), also known as methylguanine DNA methyltransferase (MGMT), a DNA-binding protein with an alkyl group acceptor site. This protein has some characteristics of an enzyme; however, unlike a true enzyme, the receptor site on AGT becomes irreversibly alkylated and the AGT is permanently inactivated and rapidly degraded (Dolan et€al. 1993; Hotta et€al. 1995). Once the enzyme is depleted, further repair by this mechanism requires de€novo synthesis of the protein. Therefore, the ability to repair O6-alkylation of DNA is related to the number of AGT molecules and the rate of protein resynthesis (Belanich et€al. 1996). Although endogenous AGT levels vary by organ and cell type, many tumors express AGT levels much higher than their corresponding normal tissue, suggesting that this protein provides an acquired mechanism for drug resistance. Whereas brain tissue expresses low or undetectable levels of AGT, significant levels of expression have been observed in most brain tumor specimens from both children and adults, in many cases several orders of magnitude higher than normal brain (Silber et€al. 1993). Several groups have noted that AGT levels correlated inversely with response to alkylator-based therapy in preclinical models (Schold et€al. 1989; Hotta et€al. 1995). Moreover, several investigators have confirmed an adverse association between elevated AGT levels and clinical response to nitrosoureas and temozolomide (Belanich et€al. 1996; Esteller et€al. 2000; Jaeckle et€al. 1998; Hegi et€al. 2005; Pollack et€al. 2006b). Although the preferred substrate for AGT is O6-methylguanine on double-stranded DNA, the protein will target other O6-alkylated substrates, which if present in excess, can effectively deplete AGT activity. In particular, O6-benzylguanine (O6-BG) has been shown to reverse AGT-mediated resistance to alkylating agents in a variety of tumor cell lines in€vitro by saturating AGT (Dolan et€al. 1991; Gerson et€al. 1993), and in€vivo studies have confirmed that O6-BG potentiates the therapeutic effect of BCNU and other alkylating agents. In a clinical trial in adult patients with gliomas who received O6-BG as a bolus dose of 100€ mg/m2, AGT levels within the tumor 18€hours after O6-BG administration were below the level of detection in 11/11 patients (Friedman et€al. 1998a). No serious side effects were observed either with O6-BG alone or in combination with low doses of BCNU (Friedman et€al. 1998a). In a recent pediatric study, P9871, a fixed dose of 06-BG was administered with escalating doses of BCNU (dose levels: 25 to 78€ mg/m2). Although there were a number of objective responses, it was unclear whether this exceeded the rate that might have been observed with BCNU alone. In addition, a recently completed phase I study of the NABTT, involving 41 patients undergoing Gliadel wafer implantation, demonstrated that administration of a continuous infusion of 30€mg/m2/day O6-BG following the initial bolus maintained an extended interval of AGT suppression. Following confirmation of AGT suppression to unmeasurable levels in 11 of 13 patients who received a 2-day infusion prior to tumor resection and Gliadel placement, the postsurgical infusion duration was increased to 2€weeks (Weingart et€al. 2007). Although this was tolerated without excessive toxicity, it remains to be determined whether combining this approach with wafer implantation improves tumor response rate or progression-free survival
Molecularly Targeted Therapies for Astrocytomas
249
compared to Gliadel delivery alone. Such an approach was also initiated in the PBTC (PBTC-009), but in view of the high frequency with which tumor resections necessitated large openings in the ventricular wall, accrual of eligible patients was challenging, and the study ultimately closed. More recently, attention has been directed at combining O6-BG with temozolomide. 6 O -BG has been noted to increase tumor cell sensitivity to temozolomide by 2.5- to 10-fold in€vitro and to potentiate sensitivity to this agent in malignant glioma xenograft models (Wedge and Newlands 1996). This enhanced efficacy is to some extent counterbalanced by increased toxicity against normal tissues, such as hematopoietic progenitor cells (Fairbairn et€al. 1995), necessitating dose reductions of temozolomide. A Phase I trial of O6-BG and temozolomide on a 5-day schedule in pediatric patients has recently been completed. Patients received O6-BG IV over 60€ min, followed 30€ min later by oral temozolomide, dailyâ•›×â•›5€ days every 4€ weeks. This study had a two-part dose-escalation scheme. Initially, the dose of O6-BG was escalated from 60 to 90 to 120€mg/m2/dayâ•›×â•›5, while the temozolomide dose was held at 28€mg/m2/day. In subsequent cohorts, the dose of temozolomide was escalated while the dose of O6-BG was held constant at 120€mg/m2/day. Forty-one patients were entered and 32 were evaluable. The MTD was defined as an O6-BG dose of 120€mg/m2/day and a temozolomide dose of 75€mg/m2/day. Dose-limiting toxicity was myelosuppression. Several patients had objective responses or prolonged stable disease. Based on these data, the PBTC opened two phase II studies with temozolomide and O6-BG. One study (PBTC-015) used the above MTD daily for 5 days of each 28-day cycle, for patients with recurrent malignant gliomas and malignant brainstem gliomas. A second study (PBTC-005), used an O6-BG bolus with escalating doses of temozolomide, followed by a continuous O6-BG infusion for 48€hours.
Base-Excision Repair Inhibition Although AGT is a proximal mechanism of resistance to alkylating agents, other mediators are also involved. Friedman et€al. (1998b) noted that defects in mismatch repair can contribute to temozolomide resistance. For example, O6-methylguanine pairs with thymine rather than guanine, leading to repetitive cycles of futile mismatch repair, which results in growth arrest and/or apoptosis (D’Atri et€al. 1998). A deficiency of DNA mismatch repair bypasses this step and thereby renders cells resistant to the apoptotic effects of temozolomide in€ vitro (Liu et€ al. 1996) and in€ vivo (Friedman et€ al. 1997). Additional mechanisms of resistance are also involved, including alterations in the DNA base excision repair (Liu et€al. 1999) and apoptotic signaling pathways (Bocangel et€al. 2002). The importance of recognizing these alternative resistance mechanisms is that DNA mismatch repair deficiency can confer resistance to temozolomide, even after O6-BG mediated AGT inhibition (Friedman et€ al. 1997; Bocangel et€ al. 2002). Likewise, a series of studies have shown that two temozolomide-initiated adducts,
250
I.F. Pollack
N7-methylguanine and N3-methyladenine, are not susceptible to AGT demethylation and thus produce cytotoxicity independently of AGT or DNA mismatch repair activity (Tentori et€al. 2003). These lesions are repaired by the base excision repair pathway, involving the actions of N-methylpurine-DNA glycosylase, AP endonuclease, poly (ADP-ribose) polymerase (PARP), DNA polymerase b, X-ray repair cross complementing protein 1 (XRCC1), and ligase III. Tumor cells resistant to temozolomide due to DNA mismatch repair deficiency have been rendered susceptible by inhibition of base excision repair. One strategy to accomplish this has involved inhibition of PARP, one of the most abundant proteins in the nucleus, which binds to damaged DNA through a zinc finger domain, catalyzing the cleavage of NAD+ into nicotinamide and ADP-ribose and using the latter to synthesize branched nucleic acid polymers, which are covalently attached to nuclear acceptor proteins (Liu et€al. 1999; Tentori et€al 2002, 2003). Tentori et€al. (2002, 2003) have demonstrated that PARP inhibition increases susceptibility to apoptosis induced by temozolomide in€vitro and increases the antitumor activity of temozolomide treatment in mice with orthotopic xenografts of human GBM. INO-1001 (Inotek Pharmaceuticals Corporation) has recently been identified as a potent PARP inhibitor, and preclinical studies of this agent for malignant gliomas were conducted at the Duke University Medical Center. The combination of this agent with temozolomide was evaluated in athymic mice carrying AGT-negative, MMR-deficient or MMR proficient D-245 malignant glioma xenografts. Groups of ten mice carrying these xenografts were treated with varying doses of INO-1001 and temozolomide. Higher doses of temozolomide and INO-1001 resulted in significant tumor regression in MMR-deficient xenografts compared to temozolomide alone, suggesting that INO-1001 was able to reverse MMR deficiency-related resistance. A phase I clinical trial of this combination was recently initiated in adults with malignant gliomas and consideration has been given to a separate pediatric study. Studies with an alternate PARP inhibitor, ABT888, are also under development.
Other Contributors to Drug Resistance Recent studies have indicated that alterations in the patterns of gene expression and protein processing may influence the growth promoting and apoptosis-resisting phenotype of malignant glioma cells, which has suggested the use of several strategies. There is evidence that histone processing is altered in malignant gliomas. The acetylation and deacetylation of histone proteins of the nucleosomes in chromatin play an important role in the regulation of gene expression (Gray and Ekstrom 2001). Acetylation of histone lysine residues is associated with a relaxation of the DNA wrapped around the core histones, enhancing access by the transcriptional machinery; conversely, deacetylation condenses the nucleosome structure, restricting access to the DNA (Marks et€al. 2000). Accordingly, treatment of mammalian cells with inhibitors of histone deacetylase (HDAC) activity results in increased expression of a variety of genes. In recent years, HDAC inhibitors have been observed to have antitumor effects,
Molecularly Targeted Therapies for Astrocytomas
251
inhibiting cell growth, inducing terminal differentiation and preventing the formation of tumors in murine models (Yoshida et€al. 1995; Marks et€al. 2000). One of the first HDAC inhibitors to be evaluated clinically was phenylbutyrate. In a phase I/II study in adults with recurrent solid tumors, Gilbert et€ al. (2001) observed that 25% of patients had stable disease for >6€months. In addition, Baker et€al. (2002) reported a complete response in a recurrent multicentric malignant glioma treated with phenylbutyrate. Other HDAC inhibitors have been observed to independently inhibit tumor growth in€vitro and potentiate the activity of conventional cytotoxic agents (Marks et€al. 2001; Sawa et€al. 2004). As with phenylbutyrate, treated cells typically undergo cell cycle arrest with induction of p21 and, in some cases, apoptosis. Clinical trials of several histone deacetylase inhibitors, including valproic acid, depsipeptide, and suberoylanilide hydroxamic acid (SAHA), are currently under way or planned (Plumb et€al. 2003). A second strategy for modulating the availability of proteins that may influence apoptosis resistance involves inhibition of members of the heat shock protein (HSP) family, which are involved in the conformational maturation of a variety of key growth stimulating and apoptosis-inhibiting proteins. Inhibitors of HSP90, such as geldanamycin and 17-AAG, have been observed to potentiate the efficacy of conventional cytotoxic chemotherapeutic agents against gliomas (Graner and Bigner 2005) as well as other signaling modulators (Premkumar et€al. 2006), although their clinical applicability for gliomas has been limited by their extremely poor penetration of the blood–brain barrier. Other potential targets for reversing therapy resistance in malignant gliomas include inhibition of NF-kB, which is involved in counteracting downstream mediators of apoptotic signaling. In the absence of a clinically applicable direct pharmacological NF-kB inhibitor, this has been accomplished indirectly by inhibition of the ubiquitin/proteasome system, which regulates the post-transcriptional degradation mechanism of NF-kB, as well as proteins involved in cell cycle regulation, DNA transcription and repair, apoptosis, angiogenesis, and cell growth. Accordingly, there has been interest in drugs that target this system. One such agent, bortezomib (Velcade®; Millennium Pharmaceuticals, Inc.), has been noted to induce apoptosis in human GBM cell lines and primary GBM explants and is currently in preliminary testing in adult patients with recurrent or progressive gliomas (Yin et€al. 2005), with the eventual goal of combining this agent with a conventional cytotoxic chemotherapeutic agent. More developmental strategies for potentiating apoptosis include stimulation of signaling from TNF-related apoptosis-inducing ligands, such as TRAIL/Apo2L. Although this strategy has shown significant promise in glioma preclinical models in several laboratories (Pollack et€ al. 2001a; Walczak et€ al 1999), clinical applicability of this approach remains to be confirmed. Finally, recent studies have explored the use of inhibitors of cell cycle regulatory proteins, in order to either (1) promote passage of injured cells through the cell cycle, thereby bypassing damage repair steps or (2) block cell cycle progression at phases of the cell cycle that confer vulnerability to various classes of chemotherapeutic agents. In the former category are agents, such as 7-hydroxystaurosporine (UCN-01), which interfere with Chk1 kinase function. By blocking cdc25 phosphorylation,
252
I.F. Pollack
UCN-01 secondarily disinhibits cell cycle progression through the G2/M checkpoint, thereby abrogating the potential for cell cycle arrest and DNA repair after treatment with agents such as temozolomide (Hirose et€al. 2001). Conversely, other inhibitors, such as cyc202 (Cyclacel), flavopiridol, and numerous recently characterized small molecular inhibitors, block various members of the cyclin-dependent kinase family, thereby promoting cell cycle arrest, which potentially causes cells to remain in portions of the cell cycle that are selectively vulnerable to adjuvant modalities, such as irradiation and conventional chemotherapy.
Immunological or Ligand-Based Therapies Vaccine-based (active) immunotherapy against protein targets expressed in gliomas is a potentially promising immunological approach that has recently transitioned from preclinical to pilot clinical studies. As a cellularly mediated strategy, it is beyond the scope of the current chapter, but is nonetheless relevant from the standpoint of molecularly targeted approaches. An array of pharmacologically oriented immunotherapy approaches has also been developed in recent years. Although systemic application of antibody-based inhibitor strategies to neutralize or block receptor-mediated signaling in malignant gliomas is challenged by the presence of a blood–brain barrier, one strategy for circumventing this limitation has involved the use of antiangiogenic antibodies targeted to the tumor vasculature, such as bevacizumab, which obviates concerns about blood vessel penetration. A second strategy has involved intracavitary delivery of antibodies directed against proteins overexpressed in the tumor cells, such as tenascin or EGFR, linked to radioisotopes or toxins (Lorimer et€al. 1996; Reardon et€al. 2006a). Recent studies have suggested a potential survival benefit in appropriately selected patients using such intratumoral delivery approaches (Quang and Brady 2004). A third approach for achieving delivery of a high-molecular weight agent into the tumor involves convectional-enhanced delivery, which employs positive-pressure infusion of microliter volumes over a period of several days through stereotactically placed catheters (Bobo et€ al. 1994) to achieve bulk flow directly into the brain tumor or peritumoral brain. In preclinical studies, this approach allowed for the distribution of an active agent over wide areas of the brain in rodent and primate models (Laske et€al. 1997a). Subsequent clinical studies confirmed the feasibility of drug distribution over a wide area around a tumor, along the white matter tracts that provide principal pathways for tumor expansion. A common strategy for applying these techniques has been to use ligands or antibodies targeted against receptors overexpressed on brain tumor cells relative to normal brain, and coupled to a mutated toxin, such as pseudomonas exotoxin or diphtheria toxin. To date, a variety of receptors have been targeted, including EGFR, the interleukin 4 (IL4), and 13 (IL13) receptors, and the transferrin receptor (TfR). By eliminating or inactivating the receptor binding moiety of the native toxin and conjugating the active component of the toxin to a ligand or antibody for the tumor-associated receptor, the immunotoxin is selectively transferred into receptor-expressing tumor cells, but excluded from
Molecularly Targeted Therapies for Astrocytomas
253
normal cells, thereby facilitating tumor-specific cell killing (Pastan and FitzGerald 1991; Greenfield et€ al. 1987). Adverse effects have consistently related to edema or necrosis within peritumoral brain, rather than systemic toxicity, which calls attention to the importance of defining volumetric and concentration-related delivery parameters. A discussion of several of the convection-enhanced delivery approaches that have been examined in malignant gliomas is provided below.
Tf-CRM107 The first immunotoxin to undergo clinical testing for patients with brain tumors was Tf-CRM107, a conjugate of a mutated diphtheria toxin linked to human transferrin (Laske et€al. 1997b). Diphtheria toxin has A and B subunits, with the A subunit catalyzing the transfer of ADP ribose to elongation factor-2, thereby blocking the transfer of peptidyl-tRNA on ribosomes, which inhibits protein synthesis and kills the cell. The B subunit is responsible for toxin binding to the cell surface. In CRM107, the B chain is mutated, which decreases binding activity by four orders of magnitude, but leaves the toxin functions intact. Because transferrin receptors are highly expressed on rapidly dividing cells, such as malignant glioma cells (Recht et€al. 1990), in comparison to low levels of expression on normal glial cells, selective tumor targeting is feasible with local drug delivery. Preclinical studies demonstrated the efficacy of this toxin in inhibiting protein synthesis (Weaver and Laske 2003) with IC50 values in the low picomolar range. Activity of the conjugate was demonstrated against subcutaneous and intracranial glioma xenografts (Laske et€al. 1994). These results provided a foundation for a phase I clinical study of this agent in patients with recurrent malignant brain tumors to determine the maximal tolerated drug concentration and infusate volume (Laske et€ al. 1997b). All patients underwent stereotactic biopsy to confirm the presence of tumor, followed by implantation of 1 to 3 silastic infusion catheters into the tumor. Nine of 15 evaluable patients demonstrated more than a 50% decrease in tumor volume and two had complete responses. Although transient worsening in neurological status occurred during 3 of 44 infusions, this was controllable with steroids and hyperosmolar therapy. A subsequent phase II study used the maximal tolerated dose established in the above study followed by a second infusion cycle 4 to 10€weeks later. Forty-four patients were enrolled, and of the 34 evaluable for efficacy, five had complete responses and seven had partial responses. These encouraging results provided a rationale for a larger phase III trial in adults with recurrent malignant gliomas as well as studies for newly diagnosed patients. A pediatric trial has been considered but not undertaken.
IL13-PE38QQR IL13-PE38QQR (Neopharm) is a tumor-targeted cytotoxin that combines human IL13 and an enzymatically active portion of Pseudomonas exotoxin (PE). This exotoxin contains three domains: domain Ia binds to the PE receptors, domain II
254
I.F. Pollack
catalyzes the translocation of the toxin into the cytosol, and domain III contains the toxin moiety (Pastan and FitzGerald 1991). The immunotoxin is formed by expression from a chimeric gene that fused the human gene for IL13 with a mutated PE gene harboring amino acid substitutions in the receptor binding domain (Debinski et€ al. 1995). The rationale for developing this combination was the observation that malignant glioma cells express the IL13 receptor (IL13R) at high density, while normal brain tissue expresses little if any IL13R (Debinski et€al. 1999; Joshi et€al. 2000). In vitro studies have demonstrated that most malignant glioma cell lines are sensitive to IL13-PE38QQR at concentrations in the picomolar range, with inhibition of protein synthesis and loss of clonogenicity. In vivo glioma model studies in nude mice demonstrated significant activity with intratumoral delivery in an IL13R-dependent fashion. Following toxicology studies in rodent and primate models (Husain and Puri 2003), a series of phase I and II clinical trials were initiated in patients with recurrent malignant gliomas by the NABTT, the NABTC, and a multinational consortium. The NABTT study (Weingart et€al. 2003) involved the administration of the drug over 96€hours via two intratumoral catheters in two infusions for patients with unresectable malignant gliomas, with escalation of the infusion concentration. The NABTC study assessed the histological effects and safety of pre- and postoperative infusions of IL13-PE38QQR in patients with recurrent resectable malignant glioma. In the first stage of this study (Prados et€ al. 2002), after biopsy and catheter placement into the tumor on day 1, the drug was delivered via one intratumoral catheter for 48€hours on days 2 to 4. Resection was conducted on day 8, with a goal of en bloc removal of the tumor and catheter to determine the histologically effective dose to the primary tumor site. Following resection, two or three catheters were placed in the peritumoral brain, and the immunotoxin was administered for 96€hours, at a range of doses to assess toxicity in peritumoral brain. Not unexpectedly, the intratumoral preresection dose was effective in inducing tumor necrosis immediately around the infusion catheter, and given the lack of penetration into the surrounding brain and the fact that the treated area was resected, it had relatively low toxicity. Subsequently, in the second stage of the study, the preresection dose was eliminated and the postresection dose was escalated. Objective responses and long-term survival have been noted in patients with appropriately placed peritumoral catheters (Kunwar et€al. 2007). A third multinational study used a single infusion, with escalation of the infusion duration at a constant intratumoral infusion rate and concentration, followed by resection at week 3. After establishment of the maximally tolerated infusion duration, IL13-PE38QQR concentration was escalated to define the maximal tolerated concentration. Taken together, these studies provided a basis for a phase III randomized study comparing IL13-PE38QQR with GLIADELTM; unfortunately, preliminary results were reportedly disappointing. A pediatric trial (PBTC011C) had been developed, although its opening was placed on hold based upon preliminary analysis of the aforementioned adult trial.
Molecularly Targeted Therapies for Astrocytomas
255
TP-38 Human malignant gliomas commonly overexpress EGFR relative to normal astrocytes. As with the IL13-PE38QQR, TP-38 (IVAX Research, Inc) was designed to couple a receptor ligand, in this case TGFa, an endogenous ligand for EGFR, with a mutated form of the pseudomonas exotoxin. The interactions between TGFa and EGFR provide the basis for internalization of the toxin. Following animal toxicology and efficacy results, pilot studies of this agent in adults were conducted at Duke University and the University of California at San Francisco (Sampson et€al. 2003). Catheters were placed stereotactically with radiographic confirmation of catheter position prior to infusion. TP-38 was infused over 50€h at a fixed flow rate and volume, with escalation of the concentration and dose in successive cohorts. Patients were stratified based on the presence or absence of residual disease at the time of infusion. Toxicity was solely neurologic. In an initial report of this series, 2 of 15 patients with residual disease had objective responses and four remained progression-free for 1€year. Based on these promising results, further studies were initiated and a pediatric trial was briefly opened (PBTC-013), although closed secondary to funding constraints prior to achieving the study objectives. In addition, studies of another cytokine receptor targeted immunotoxin approach, using a circularly permutated IL4 conjugated to a mutated pseudomonas exotoxin have also been initiated in adult patients, based on encouraging preclinical data (Puri et€al. 1996) and evidence of activity in an initial pilot study (Weber et€ al. 2003). Because of these promising findings, a multicenter dose escalation study was planned followed by a phase II efficacy study.
Future Directions and Challenges Because of the poor responsiveness of malignant gliomas to conventional therapies, there is a pressing need to implement new approaches to improve the outcome of patients with these tumors. Studies in many laboratories have demonstrated that the proliferation and survival of malignant glioma cells is strongly influenced by a number of molecular pathways that stimulate tumor growth, promote angiogenesis, and inhibit apoptosis induction. These pathways have therefore emerged as promising targets for therapy, which have been effectively addressed by small molecule inhibitors as well receptor-directed cytotoxins. The major ongoing challenges are the identification of tumor genotypic and phenotypic features that predict response to individual agents and the discovery of rational combinations of agents to potentiate therapeutic efficacy. Given the molecular diversity and heterogeneity of malignant gliomas, it is unlikely that any single agent will have efficacy in more than a subset of tumors. Moreover, responses to individual drugs, if any, are likely to be transient at best. Accordingly, the appropriate use of these agents will likely require combinations of molecularly targeted compounds and conventional therapies or of several molecularly targeted agents
256
I.F. Pollack
administered as a “cocktail,” incorporating a tumor-tailored therapy paradigm based on the molecular features of the lesion to guide agent selection. Acknowledgmentâ•… This work was supported in part by NIH grant NSP0140923.
References Abdollahi, A., Lipson, K.E., Sckell, A., et€al. 2003. Combined therapy with direct and indirect angiogenesis inhibition results in enhanced antiangiogenic and antitumor effects. Cancer Res 63:8890–8898. Albanell, J., Royo, F., Auerbach, S., et€al. 2002. Pharmacokinetic studies of the epidermal growth factor receptor inhibitor ZD1839 in skin from cancer patients: histological and molecular consequences of receptor inhibition. J Clin Oncol 20:110–124. Alsina, M., Fonseca, R., Wilson, E.F., et€al. 2004. Farnesyltransferase inhibitor tipifarnib is well tolerated, induces stabilization of disease, and inhibits farnesylation and oncogenic/tumor survival pathways in patients with advanced multiple myeloma. Blood 103:3271–3277. Baker, M.J., Brem, S., et€ al. 2002. Complete response of a recurrent, multicentric malignant glioma patient treated with phenylbutyrte. J Neurooncol 59:239–242. Barnett, F.H., Scharer-Schuksz, M., Wood, M., et€al. 2004. Intra-arterial delivery of endostatin gene to in tumors prolongs survival and alters tumor vessel ultrastructure. Gene Ther .11:1283–1289. Baselga, J., Rischin, D., Ranson, M., et€al. 2002. Phase I safety, pharmacokinetic, and pharmacodynamic trial of ZD1839, a selective oral epidermal growth factor receptor tyrosine kinase inhibitor, in patients with five selected solid tumor types. J Clin Oncol 20:4292–4302. Baselga, J., Pfister, D., Cooper, M.R., et€al. 2000. Phase I studies of anti-epidermal growth factor receptor chimeric antibody C225 alone and in combination with cisplatin. J Clin Oncol 18:904–914. Baumann, F., Bjeljac, M., Kollias, S.S., et€ al. 2004. Combined thalidomide and temozolomide treatment in patients with glioblastoma multiforme. J Neurooncol 67:191–200. Belanich, M., Pastor, M., Randall, T., et€al. 1996. Retrospective study of correlation between the DNA repair protein alkyl-transferase and survival of brain tumor patients treated with carmustine. Cancer Res 56:783–788. Bianco, R., Shin, I., Ritter, C.A., Yakes, F.M., Basso, A., Rosen, N., Tsurutani, J., Dennis, P.A., Mills, G.B., Arteaga, C.L. 2003. Loss of PTEN/MMAC1/TEP in EGF receptor-expressing tumor cells counteracts the antitumor action of EGFR tyrosine kinase inhibitors. Oncogene 22: 2812–2822. Bobo, R.H., Laske, D., Akbasak, A., et€al. 1994. Convection-enhanced delivery of macromolecules in the brain. Proc Natl Acad Sci U S A 91:2076–2080. Bocangel, D.B., Finkelstein, S., Schold, S.C., et€al. 2002. Multifaceted resistance of gliomas to temozolomide. Clin Cancer Res 8:2725–2734. Bredel, M., Pollack, I.F., Freund, J.M., Hamilton, A.D., Sebti, S.M. 1998. Inhibition of Ras and related G-proteins as a therapeutic strategy for blocking malignant glioma growth. Neurosurgery 43:124–131. Bredel, M., Pollack, I.F., Freund, J.M., Rusnak, J., Lazo, J.S. 1999a. Protein kinase C inhibition by UCN-01 induces apoptosis in human glioma cells in a time-dependent fashion. J Neurooncol 41:9–20. Bredel, M., Pollack, I.F., Hamilton, R.L., James, C.D. 1999b. Epidermal growth factor receptor (EGFR) expression in high-grade non-brainstem gliomas of childhood. Clin Cancer Res 5:1786–1792. Bruns, C.J., Harbison, M.T., Davis, D.W., et€al. 2000. Epidermal growth factor receptor blockade with C225 plus gemcitabine results in regression of human pancreatic carcinoma growing orthotopically in nude mice by antiangiogenic mechanisms. Clin Cancer Res 6:1936–1948.
Molecularly Targeted Therapies for Astrocytomas
257
Buchdunger, E., Cioffi, C.L., Law, N., et€al. 2000. Abl protein-tyrosine kinase inhibitor STI571 inhibits in€vitro signal transduction mediated by c-kit and platelet-derived growth factor receptors. J Pharmacol Exp Ther 295:139–145. Burris, H.A., III. 2004. Dual kinase inhibition in the treatment of breast cancer: initial experience with the EGFR/ErbB-2 inhibitor lapitinib. Oncologist 9 (Suppl 3):10–15. Cantley, L.C., Neel, B.G. 1999. New insights into tumor suppression: PTEN suppresses tumor formation by restraining the phosphoinositide 3-kinase/AKT pathway. Proc Natl Acad Sci U S A 96:4240–4245. Cardone, M.H., Roy, N., Stennicke, H.R., et€al. 1998. Regulation of cell death protease caspase-9 by phosphorylation. Science 282:1318–1321. Carducci, M.A., Musib, L., Kies, M.S., et€al. 2006. Phase I dose escalation and pharmacokinetic study of enzastaurin, an oral protein kinase C beta inhibitor, in patients with advanced cancer. J Clin Oncol 24:4092–4099. Chang, S.M., Wen, P., Cloughesy, T., et€al. 2005. Phase II study of CCI-779 in patients with recurrent glioblastoma multiforme. Invest New Drugs 23:357–361. Cheng, Y., Ng, H.-K., Zhang, S.-F., et€al. 1999. Genetic alterations in pediatric high-grade astrocytomas. Hum Pathol 30:1284–1290. Cheng, Y., Pang, J.C., Ng, H.K., et€ al. 2000. Pilocytic astrocytomas do not show most of the genetic changes commonly seen in diffuse astrocytomas. Histopathology 37:437–444. Ciardiello, F., Caputo, R., Bianco, R., et€ al. 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. Ciardiello, F., Caputo, R., Bianco, R., et€al. 2000. Antitumor effect and potentiation of cytotoxic drugs activity in human cancer cells by ZD-1839 (Iressa), an epidermal growth factor receptorselective tyrosine kinase inhibitor. Clin Cancer Res 6:2053–2063. Cloughesy, T.F., Wen, P.Y., Robins, H.I., et€al. 2006.Phase II trial of tipifarnib in patients with recurrent malignant glioma either receiving or not receiving enzyme-inducing antiepileptic drugs: a North American Brain Tumor Consortium Study. J Clin Oncol 24:3651–3656 Collins, V.P. 1999. Progression as exemplified by human astrocytic tumors. Cancer Biol. 9:267–276. Couldwell, W.T., Uhm, J.H., Antel, J.P., Yong, V.W. 1991. Enhanced protein kinase C activity correlates with the growth rate of malignant gliomas in€vitro. Neurosurgery 29:880–887. Couldwell, W.T., Antel, J.P., Yong, V.W. 1992. Protein kinase C activity correlates with the growth rate of malignant gliomas: part II. Effects of glioma mitogens and modulators of protein kinase C. Neurosurgery 31:717–724. Cross, D.A., Alessi, D.R., Cohen, P., Andjelkovich, M., Hemmings, B.A. 1995. Inhibition of glycogen synthase kinase-3 by insul mediated by protein kinase B. Nature 378:785–789. Dai, C., Celestino, J.C., Okada, Y., Louis, D.N., Fuller, G.N., Holland, E.C. 2001. PDGF autocrine stimulation dedifferentiates cultured astrocytes and induces oligodendrogliomas and oligoastrocytomas from neural progenitors and astrocytes in€ vivo. Genes Dev 15:1913–1925. Dasgupta, B., Yi, Y., Chen, D.Y., Weber, J.D., Gutmann, D.H. 2005. Proteomic analysis reveals hyperactivation of the mammalian target of rapamycin pathway in neurofibromatosis 1-associated human and mouse brain tumors. Cancer Res 65:2755–2760. D’Atri, S., Tentori, L., Lacal, P.M., et€ al. 1998. Involvement of the mismatch repair system in temozolomide-induced apoptosis. Mol Pharmacol 54:334–341. Debinski, W., Gibo, D.M., Hulet, S.W., et€al. 1999. Receptor for interleukin 13 is a marker and therapeutic target for human high grade gliomas. Clin Cancer Res 5:985–990. Debinski, W., Obiri, N.I., Pastan, I., et€al. 1995. A novel chimeric protein composed of interleukin 13 and Pseudomonas exotoxin is highly cytotoxic to human carcinoma cells expressing receptors for interleukin 13 and interleukin 4. J Biol Chem 270:16775–16780. Dolan, M., Mitchell, R., Mummert, C., et€ al. 1991. Effect of O6-Benzylguanine analogs on sensitivity of tumor cells to the cytotoxic effects of alkylating agents. Cancer Res 51:3367–3372.
258
I.F. Pollack
Dolan, M., Pegg, A., Moschel, R., et€al. 1993. Effect ot O6-benzylguanine on the sensitivity of human colon tumor xenografts to 1,3-bis(2-chloroethyl)-1-nitrosourea. Biochem Pharmacol 46:285–290. Dresemann, G. 2005. Imatinib and hydroxyurea in pretreated progressive glioblastoma multiforme: a patient series. Ann Oncol 16:1702–1708. Druker, B.J., Talpaz, M., Resta, D.J., et€al. 2001. Efficacy and safety of a specific inhibitor of the BCR-ABL tyrosine kinase in chronic myeloid leukemia. N Engl J Med 344:1031–1037. End, D.W., Smets, G., Todd, A.V., et€ al. 2001. Characterization of the antitumor effects of the selective farnesyl protein transferase inhibitor R115777 in€ vivo and in€ vitro. Cancer Res 61:131–137. Esteller M., Garcia-Foncillas, J., Andion, E., et€al. 2000. Inactivation of the DNA-repair gene MGMT and the clinical response of gliomas to alkylating agents. N Engl J Med 343:1350–1354. Fairbairn, L., Watson, A., Rafferty, J., et€ al. 1995. O6-Benzylguanine increases sensitivity of human primary bone marrow cells to the cytotoxici effects of temozolomide. Exp Hematol 23:112–116. Feldkamp, M.M., Lau, N., Guha, A. 1999. Growth inhibition of astrocytoma cells by farnesyl transferase inhibitors is mediated by a combination of anti-proliferative, pro- apoptotic and anti-angiogenic effects. Oncogene 18:7514–7526. Ferrera, N., Hillan, K.J., Novotny, W. 2005. Bevacizumab (Avastin), a humanized anti-VEGF monoclonal antibody for cancer therapy. Biochem Biophys Res Commun 333:328–335. Fields, A.P., Tyler, G., Kraft, A.S., May, W.S. 1990. Role of nuclear protein kinase C in the mitogenic response to platelet-derived growth factor. J Cell Sci 96:107–114. Fine, H.A., Figg, W.D., Jaeckle, K., et€al. 2000. Phase II trial of the antiangiogenic agent thalidomide in patients with recurrent high-grade gliomas. J Clin Oncol 18:708–715. Fine, H.A., Kim, L., Royce, C., et€al. 2005. Results from phase II trial of enzastaurin (LY317615) in patients with recurrent high grade gliomas. J Clin Oncol 23:1504a. Fine, H.A., Wen, P.Y., Maher, E.A., et€al. 2003. Phase II trial of thalidomide and carmustine for patients with recurrent high-grade gliomas. J Clin Oncol 21:2299–2304. Finlay, J.L., Boyett, J.M., Yates, A.J., et€al. 1995. Randomized phase III trial in childhood highgrade astrocytoma comparing vincristine, lomustine, and prednisone with the eight-drugs-in1-day regimen. J Clin Oncol 13:112–123. Freed, E., Symons, M., MacDonald, S.G., McCormick, F., Ruggieri, R. 1994. Binding of 14-3-3 proteins to the protein kinase Raf and effects on its activation. Science 265:1713–1716. Friedman, H.S., Johnson, S.P., Dong, Q., et€al. 1997. Methylator resistance mediated by mismatch repair deficiency in a glioblastoma multiforme xenograft. Cancer Res 57:2933–2936. Friedman, H., Kokkinakis, D., Pluda, J., et€ al. 1998a. Phase I Trial of O6-Benzylguanine for patients undergoing surgery for malignant glioma. J Clin Oncol 16:3570–3575. Friedman, H.S., McLendon, R.E., Kerby, T., et€ al. 1998b. DNA mismatch repair and O6-alkylguanine-DNA alkyltransferase analysis and response to temodal in newly diagnosed malignant glioma. J Clin Oncol 16:3851–3857. Fukuoka, M., Yano, S., Giaccone, G., et€al. 2003. Multi-institutional randomized phase II trial of gefitinib for previously treated patients with advanced non-small-cell lung cancer. J Clin Oncol 21:2237–2246. Galanis, E., Buckner, J.C., Maurer, M.J., et€al. 2005. Phase II trial of temsirolimus (CCI-779) in recurrent glioblastoma multiforme: a North Central Cancer Treatment Group study. J Clin Oncol 23:5294–5304. Gerson, S., Zborowska, E., Norton, K., et€al. 1993. Synergistic efficacy of O6-benzylguanine and 1,3-bis(2-chloroethyl)-1-nitrosourea in a human colon cancer xenograft completely resistant to BCNU alone. Biochem Pharmacol 45:483–491. Gilbert, J., Baker, S.D., et€ al. 2001. A phase I dose escalation and bioavailability study of oral sodium phenylbutyrte in patients with refractory solid tumor malignancies. Clin Cancer Res 7:2292–2300. Glick, R.P., Gettleman, R., Patel, K., et€al. 1989. Insulin and insulin-like growth factor I in brain tumors: binding and in€vitro effects. Neurosurgery 24:791–797.
Molecularly Targeted Therapies for Astrocytomas
259
Goldbrunner, R.H., Bendszus, M., Wood, J., et€al. 2004. PTK787/ZK222584, an inhibitor of vascular endothelial growth factor receptor tyrosine kinases, decreases glioma growth and vascularization. Neurosurgery 55:426–432. Graff , J.R., McNulty, A.M., Hanna, K.R., et€al. 2005. The protein kinase Cb-selective inhibitor, enzastaurin (LY317615.HCL), suppresses signaling through the AKT pathway, induces apoptosis, and suppresses growth of human colon cancer and glioblastoma xenografts. Cancer Res 65:7462–7469. Graner, M.W. and Bigner, D.D. 2005. Chaperone proteins and brain tumors: potential targets and possible therapeutics. Neurooncology 7:260–277. Gray, S.G., Ekstrom, T.J. 2001. The human histone deacetylase family. Exp Cell Res 262:75–83. Greenfield, L., Johnson, V.G., Youle, R.J. 1987. Mutations in diphtheria toxin separate binding from entry and amplify immunotoxin selectivity. Science 238:536–539. Griffin, C.A., Hawkins, A.L., Packer, R.J., et€ al. 1988. Chromosome abnormalities in pediatric brain tumors. Cancer Res 48:175–180. Gross, J.L., Morrison, R.S., Eidsvoog, K., et€al. 1990. Basic fibroblast growth factor: a potential autocrine regulator of human glioma growth. J Neurosci Res 27:689–696. Guha, A., Feldkamp, M.M., Lau, N., Boss, G., Pawson, A. 1997. Proliferation of human malignant astrocytomas is dependent on Ras activation. Oncogene 15:2755–2765. Guttmann, D. H., Donahoe, J., Brown, T., James, C.D., Perry, A. 2000. Loss of neurofibromatosis 1(NF1) gene expression in NF1-associated pilocytic astrocytomas. Neuropathol Appl Neurobiol 26:361–367. Hartmann, C., Meyer, J., Balss, J., et al. 2009. Type and frequency of IDH1 and IDH2 mutations are related to astrocytic and oligodendroglial differentiation and age: a study of 1,010 diffuse gliomas. Acta Neuropathol 118:469–474. Hegi, M.E., Diserens, A.C., Gorlia, T., et€al. 2005. MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med 352:997–1003. Herbst, R.S., Giaccone, G., Schiller, J.H., et€al. 2004. Gefitinib in combination with paclitaxel and carboplatin in advanced non-small-cell lung cancer: a phase III trial – Intact 2. J Clin Oncol 22:785–794. Hermanson, M., Funa, K., Koopman, J., et€ al. 1996. Association of loss of heterozygosity on chromosome 17p with high platelet-derived growth factor a receptor expression in human malignant gliomas. Cancer Res 56:164–171. Hidalgo, M., Siu, L.L., Nemunaitis, J., et€al. 2001. Phase I and pharmacologic study of OSI-774, an epidermal growth factor receptor tyrosine kinase inhibitor, in patients with advanced solid malignancies. J Clin Oncol 19:3267–3279. Hirose, Y., Berger, M.S., Pieper, R.O. 2001. Abrogation of the Chk1-mediated G(2) checkpoint pathway potentiates temozolomide-induced toxicity in a p53-independent manner in human glioblastoma cells. Cancer Res 61:5843–5849. Hotta, T., Saito, Y., Fujita, H., et€ al. 1995. 06 alkylguanine-DNA Alkyltransferase activity in human malignant glioma and its clinical implications. J Neurooncol 21:135–140. Howe, L.R., Leevers, S.J., Gomez, N., et€al. 1992. Activation of the MAP kinase pathway by the protein kinase raf. Cell 71:335–342. Huang, S.M., Bock, J.M., Harari, P.M. 1999. Epidermal growth factor receptor blockade with C225 modulates proliferation, apoptosis, and radiosensitivity in squamous cell carcinomas of the head and neck. Cancer Res 59:1935–1940. Huang, S-M., Li, E., Armstrong, E.A., Harari, P.M. 2002. Modulation of radiation response and tumor-induced angiogenesis after epidermal growth factor receptor inhibition by ZD1839 (Iressa). Cancer Res 62:4300–4306. Hurwitz, H., Fehrenbache, L., Novotny, W., et€al. 2004. Bevacizumab plus irinotecan, fluorouracil, and leucovorin for metastatic colorectal cancer. N Engl J Med 350:2335–2342. Husain, S.R., Puri, R.K. 2003. Interleukin-13 receptor-directed cytotoxin for malignant glioma therapy: from bench to bedside. J Neurooncol 65:37–48. Ichimura, K., Bolin, M.B., Goike, H.M., et€al. 2000. Deregulation of the p14ARF/MDM2/p53 pathway is a prerequisite for human astrocytic gliomas with G1-S transition control gene abnormalities. Cancer Res 60:417–424.
260
I.F. Pollack
Ignoffo, R.J. 2004. Overview of bevacizumab: a new cancer therapeutic strategy targeting vascular endothelial growth factor. Am J Health Syst Pharm 61:S21–S26. Ino, Y., Zlatescu, M.C., Sasaki, H., et€ al. 2000. Long survival and therapeutic responses in patients with histologically disparate high-grade gliomas demonstrating chromosome 1p loss. J Neurosurg 92:983–990. Jaeckle, K.A., Eyre, H.J., Townsend, J.J., et€al: 1998. Correlation of tumor O6 methylguanine-DNA methyltransferase levels with survival of malignant astrocytoma patients treated with bis- chloroethylnitrosourea: a Southwest Oncology Group study. J Clin Oncol 16:3310–3315. James, C.D., Carlbom, E., Nordenskjoldk M., et€al. 1989. Mitotic recombination of chromosome 17 in astrocytomas. Proc Natl Acad Sci U S A 86:2858–2862. Jane, E.P., Premkumar, D.R., Pollack, I.F. 2006. Coadministration of sorafenib with rottlerin potently inhibits cell proliferation and migration in human malignant glioma cells. J Pharmacol Exp Ther 319:1070–1080. Jiang, K., Coppola, D., Crespo, N.C., et€al. 2000. The phosphoinositide 3-OH kinase/AKT2 pathway as a critical target for farnesyltransferase inhibitor-induced apoptosis. Mol Cell Biol 20:139–148. Joensuu, H., Roberts, P.J., Sarlomo-Rikala, M., et€al. 2001. Effect of the tyrosine kinase inhibitor STTI571 in a patient with a metastatic gastrointestinal stromal tumor. N Engl J Med 344:1052–1056. Jones, D. T., Kocialkowski, S., Liu, L., Pearson, D.M., Backlund, L. M., Ichimura K., Collins, V. P. 2008. Tandem duplication producing a novel oncogenic BRAF fusion gene defines the majority of pilocytic astrocytomas. Cancer Res. 68: 8673–8677. Joshi, B.H., Plautz, G.E., Puri, R.K. 2000. Interleukin-13 Receptor a chain: a novel tumor-associated transmembrane protein in primary explants of human malignant gliomas. Cancer Res 60:1168–1172. Katz, A., Barrios, C. H., Abramoff, R., et€al. 2004. Imatinib (STI 571) is active in patients with high-grade gliomas progressing on standard therapy. Proceedings ASCO Annual Meeting 2004. J Clin Oncol 23:117 (abstract) Keyes, K.A., Mann, L., Sherman, M., et€al. 2004. LY317615 decreases plasma VEGF levels in human tumor xenograft-bearing mice. Cancer Chemother Pharmacol 53:133–140. Kieran, M.W., Packer, R., Onar, A., et€ al. 2007. Phase I and pharmacokinetic study of the oral farnesyl-transferase inhibitor Sarasar™ (Lonafarnib–SCH66336) given twice daily to pediatric patients with advanced central nervous system tumors: A Pediatric Brain Tumor Consortium (PBTC) study. J Clin Oncol 25(21) 3137–3143. Kilic, T., Alberta, J.A., Zdunek, P.R., et€ al. 2000. Intracranial inhibition of platelet-derived growth factor-mediated glioblastoma cell growth by an orally active kinase inhibitor of the 2-phenylaminopyrimidine class. Cancer Res 60:143–150. Kim, E.S., Kies, M.S., Fossella, F.V., et€al. 2005. Phase II study of the farnesyltransferase inhibitor lonafarnib with paclitaxel in patients with taxane-refractory/resistant nonsmall cell lung carcinoma. Cancer 104:561–569. Kleihues, P., Burger, P.C., Scheithauer, B.W. 1993. Histological typing of tumours of the central nervous system. In: Sobin, L.H. (ed.) International histological classification of tumours, Vol. 21. Geneva: Springer, pp. 11–16. Kolch, W., Heldecker, G., Kochs, G., et€ al. 1993. Protein kinase Ca activates Raf-1 by direct phosphorylation. Nature 364:249–252. Krishnan, S., Brown, P., Ballman, K., et€al. 2005. Phase I trial of erlotinib with radiation therapy (RT) in patients with glioblastoma multiforme (GBM). J Clin Oncol 23:1513a. Kunwar, S., Prados, M.D., Chang, S.M., et€al. 2007. Cintredekin Besudotox Intraparenchymal Study Group. Direct intracerebral delivery of cintredekin besudotox (IL13-PE38QQR) in recurrent malignant glioma: a report by the Cintredekin Besudotox Intraparenchymal Study Group. J Clin Oncol 25:837–844. Lang, F.F., Miller, D.C., Pisharody, S., et€al. 1994. High frequency of p53 protein accumulation without p53 gene mutation in human juvenile pilocytic, low grade and anaplastic astrocytomas. Oncogene 9:949–954.
Molecularly Targeted Therapies for Astrocytomas
261
Laske, D.W., Ilercil, O., Akbasak, A., Youle, R.J., Oldfield, E.H. 1994. Efficacy of direct intratumoral therapy with targeted protein toxins for solid human gliomas in nude mice. J Neurosurg 80:520–526. Laske, D.W., Morrison, P.F., Lieberman, D.M., et€al.1997a. Chronic interstitial infusion of protein to primate brain: determination of drug distribution and clearance with single-photon emission computerized tomography imaging. J Neurosurg 87:586–594. Laske, D.W., Youle, R.J., Oldfield, E.H. 1997b. Tumor regression with regional distribution of targetted toxin TF-CRM107 in patients with malignant brain tumors. Nat Med 3:1362–1368. Lassman, A.B., Rossi, M.R., Razier, J.R., et€ al. 2005. Molecular study of malignant gliomas treated with epidermal growth factor receptor inhibitors: tissue analysis from North American Brain Tumor Consortium Trial 01-03 and 00-01. Clin Cancer Res 11:7841–7850. Lebowitz, P.F., Prendergast, G.C. 1998. Non-Ras targets of farnesyltransferase inhibitors: focus on Rho. Oncogene 17:1439–1445. Li, D.M., Sun, H. 1998. PTEN/MMAC1/TEP1 suppressed the tumorigenicity and induces G1 cell cycle arrest in human glioblastoma cells. Proc Natl Acad Sci U S A 95:15406–15411. Libermann, T.A., Nusbaum, H.R., Razon, N., et€al. 1985. Amplification, enhanced expression and possible rearrangement of EGF receptor gene in primary human brain tumors of glial origin. Nature 313:44–147. Lieberman, F.S., Cloughesy, T., Fine, H., et€ al. 2004. NABTC phase I/II trial of ZD-1839 for recurrent malignant gliomas and unresectable meningiomas. J Clin Oncol 22:1510. Liu, L., Markowitz, S., Gerson, S.L. 1996. Mismatch repair mutations override alkyltransferase in conferring resistance to temozolomide but not to 1,3-bis(2-chloroethyl)nitrosourea. Cancer Res 56:5375–5379. Liu, L., Taverna, P., Whitacre, C.M., et€al. 1999. Pharmacologic disruption of base excision repair sensitizes mismatch repair-deficient and -proficient colon cancer cells to methylating agents. Clin Cancer Res 5:2908–2917. Lokker, N.A., Sullivan, C.M., Hollenbach, S.J.,et€al. 2002. Platelet-derived growth factor (PDGF) autocrine signaling regulates survival and mitogenic pathways in glioblastoma cells: evidence that the novel PDGF-C and PDGF-D ligands may play a role in the development of brain tumors. Cancer Res 62:3729–3735. Lorimer, I.A., Keppler-Hafkemeyer, A., Beers, R.A., et€ al. 1996. Recombinant immunotoxins specific for a mutant epidermal growth factor receptor: targeting with a single chain antibody variable domain isolated by phage display. Proc Natl Acad Sci U S A 93:14815–14820. Louis, D.N. 1997. A molecular genetic model of astrocytoma histopathology. Brain Pathol 7:755–764. Lynch, T.J., Bell, D.W., Sordella, R., et€al. 2004. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non-small-cell lung cancer to gefitinib. N Engl J Med 350:2129–2139. Maity, A., Pore, N., Lee, J., Solomon, D., O’Rourke, D.M. 2000. Epidermal growth factor transcriptionally up-regulates vascular endothelial growth factor expression in human glioblastoma cells via a pathway involving phosphatidylinositol 3’-kinase and distinct from that induced by hypoxia. Cancer Res 60:5879–5886. Marais, R., Light, Y., Mason, C., Paterson, H., Olson, M.F., Marshall, C.J. 1998. Requirement of Ras-GTP-Raf complexes for activation of Raf-1 by protein kinase C. Science 280:109–112. Marks, P.A., Richon, V.M., Rifkind, R.A. 2000. Histone deacetylase inhibitors: inducers of differentiation or apoptosis of transformed cells. J Natl Cancer Inst 92:1210–1216. Marks, P., Rifkind, R.A., Richon, V.M., et€al. 2001. Histone deacetylases and cancer: causes and therapies. Nat Rev Cancer 1:194–202. McCormick, F. 1993. How receptors turn Ras on. Nature 363:15–16. Mellinghoff, I.K., Wang, M.Y., Vivanco, I., Haas-Kogan, D.A., Zhu, S., Dia, E.Q., et€ al. 2005. Molecular determinants of the response of glioblastomas to EGFR kinase inhibitors. N Engl J Med 353:2012–2024. Milas, L., Mason, K., Hunter, N., et€ al. 2000. In vivo enhancement of tumor radioresponse by C225 antiepidermal growth factor receptor antibody. Clin Cancer Res 6:323–325.
262
I.F. Pollack
Mishima, K., Johns, T.G., Luwor, R.B., et€ al. 2001. Growth suppression of intracranial xenografted glioblastomas overexpressing mutant epidermal growth factor receptors by systemic administration of monoclonal antibody (mAb) 806, a novel monoclonal antibody directed to the receptor. Cancer Res 61:5349–5354. Nakamura, J.L., Karlsson, A., Arvold, N.D., et€al. 2005. PKB/Akt mediates radiosensitization by the signaling inhibitor LY294002 in human malignant gliomas. J Neurooncol 71:215–222. Nishizuka, Y. 1992. Intracellular signaling by hydrolysis of phospholipids and activation of protein kinase C. Science 258:607–613. Nistér, M., Claesson-Welch, L., Eriksson, A., Heldin, C.-H., Westermark, B. 1991. Differential expression of platelet-derived growth factor receptors in human malignant glioma cell lines. J Biol Chem 266:16755–16763. Nitta, T. and Sato, K. 1994. Specific inhibition of c-sis protein synthesis and cell proliferation with antisense oligodeoxynucleotides in human glioma cells. Neurosurgery 34:309–314. O’Brien, S.G., Guilhot, F., Larson, R.A., et€al. 2003. Imatinib compared with interferon and lowdose cytarabine for newly diagnosed chronic-phase chronic myeloid leukemia. N Engl J Med 348:994–1004. O’Rourke, D.M., Kao, G.D., Singh, N., Park, B.W., et€ al. 1998. Conversion of a radioresistant phenotype to a more sensitive one by disabling erbB receptor signaling in human cancer cells. Proc Natl Acad Sci U S A 95:10842–10847. O’Rourke, D.M., Qian, X., Zhang, H.T., et€al. 1997. Trans receptor inhibition of human glioblastoma cells by erbB family ectodomains. Proc Natl Acad Sci U S A 94:3250–3255. Parsons, D.W., Jones, S., Zhang, X., et al. 2008. An integrated genomic analysis of human glioblastoma multiforme. Science 321:1807–1812. Pastan, I., FitzGerald, D. 1991. Recombinant toxins for cancer treatment. Science 254:1173. Pawson, T. 1995. Protein modules and signalling networks. Nature 373:573–580. Peddanna, N., Mendis, R., Holt, S., Verma, R.S. 1996. Genetics of colorectal cancer. Int J Oncol 9:327–335. Peereboom, D.M., Brewer, C.J., Suh, J.H., et€al. 2006. Phase II trial of erlotinib with temozolomide and concurrent radiation therapy in patients with newly diagnosed glioblastoma multiforme: final results. Neurooncology 8:448. Perez-Soler, R. 2004. The role of erlotinib (Tarceva, OSI 774) in the treatment of non-small cell lung cancer. Clin Cancer Res 10:4238s–4240s. Phuphanich S, Carson K, Grossman S et€al. 2005. A phase I evaluation of the safety of escalating doses of atrasentan in adults with recurrent malignant glioma (MG). J Clin Oncol 23:1526a. Pietras, K., Ostman, A., Sjoquist, M., et€ al. 2001. Inhibition of platelet-derived growth factor receptors reduces interstitial hypertension and increases transcapillary transport in tumors. Cancer Res 61:2929–2934. Pietras, K., Rubin, K, Sjoblom, T., et€al. 2002. Inhibition of PDGF receptor signaling in tumor stroma enhances antitumor effect of chemotherapy. Cancer Res 62:5476–5484. Plate, K.H., Breier, G., Weich, H.A., et€al. 1992. Vascular endothelial growth factor is a potential tumour angiogenesis factor in human gliomas in€vivo. Nature 359:845–848. Plumb, J.A., Finn, P.W., Williams, R.J., et€al. 2003. Pharmacodynamic response and inhibition of growth of human tumor xenografts by the novel histone deacetylase inhibitor PXD 101. Mol Cancer Ther 2:721–728. Pollack, I.F. 1994. Current concepts: brain tumors in children. N Engl J Med 331:1500–1507. Pollack, I.F., Bredel, M., Erff, M. 1998. The application of signal transduction inhibition as a therapeutic strategy for central nervous system tumors. Pediatr Neurosurg 29:228–244. Pollack, I.F., Erff, M., Ashkenazi, A. 2001a. Direct stimulation of apoptotic signaling by Apo2L/ tumor necrosis factor-related apoptosis-inducing ligand leads to selective killing of glioma cells. Clin Cancer Res 7:1362–1369 Pollack, I.F., Finkelstein, S.D., Burnham, J., et€al. 2003. The association between chromosome 1p loss and outcome in pediatric malignant gliomas: results from the CCG-945 cohort. Pediatr Neurosurg 39:114–121. Pollack, I.F., Finkelstein, S.D., Burnham, J., et€al. 2001b. Age and TP53 mutation frequency in childhood gliomas. Results in a multi-institutional cohort. Cancer Res 61:7404–7407.
Molecularly Targeted Therapies for Astrocytomas
263
Pollack, I.F., Finkelstein, S.D., Woods, J., et€al. 2002. Expression of p53 and prognosis in malignant gliomas in children. N Engl J Med 346:420–427. Pollack, I.F., Hamilton, R.L., James, C.D., et€ al. 2006a. Rarity of PTEN deletions and EGFR amplification in malignant gliomas of childhood: results from the CCG-945 cohort. J Neurosurg 105:418–424. Pollack, I.F., Hamilton, R.L., Finkelstein, S.D., et€al. 1997. The relationship between TP53 mutations and overexpression of p53 and prognosis in malignant gliomas of childhood. Cancer Res 57:304–309. Pollack, I.F., Hamilton, R.L., Sobol, R.W., et€ al. 2006b. O6-methylguanine-DNA methyltransferase expression strongly correlates with outcome in childhood malignant gliomas: results from the CCG-945 cohort. J Clin Oncol 24:3431–3437. Pollack, I.F., Jakacki, R.I., Blaney, S.M., et€al. 2007. Phase I trial of imatinib in children with newly diagnosed brainstem and recurrent malignant gliomas: a PBTC report. Neurooncology 9:145–160. Pollack, I.F., Kawecki, S., Lazo, J.S. 1996. 7-hydroxystaurosporine (UCN-01), a selective protein kinase C inhibitor, exhibits cytotoxicity against cultured glioma cells and potentiates the antiproliferative effects of BCNU and cisplatin. J Neurosurg 84:1024–1032. Pollack, I.F., Randall, M.S., Kristofik, M.P., et€al. 1991. Response of low-passage human malignant gliomas in€vitro to stimulation and selective inhibition of growth factor-mediated pathways. J Neurosurg 75:284–293. Pollack, V.A., Savage, D., Baker, D., et€ al. 1999. Inhibition of epidermal growth factor receptorassociated tyrosine phosphorylation in human carcinomas with CP-358,774. Dynamics of receptor inhibition in situ and antitumor effects in athymic mice. J Pharmacol Exp Ther 291:739–748. Prados, M., Lamborn, K.R., Chang, S., et€al. 2006. Phase I study of erlotinib HCl alone and combined with temozolomide in patients with recurrent malignant glioma. Neurooncology 8:67–78. Prados, M.D., Lang, F.F., Sherman, J.W., et€al. 2002. Convection-enhanced delivery (CED) by positive pressure infusion for intra-tumoral and peri-tumoral administration of IL13-PEQQR, a recombinant tumor-targeted cytotoxin in recurrent malignant glioma. Neurooncology 4:S78. Premkumar, D.R., Arnold, B., Pollack, I.F. 2006. Cooperative inhibitory effect of ZD1839 (Iressa) in combination with 17-AAG on glioma cell growth. Mol Carcinog 45:288–301. Prigent, S.A., Nagane, M., Lin, H., et€al. 1996. Enhanced tumorigenic behavior of glioblastoma cells expressing a truncated epidermal growth factor receptor is mediated through the RasSHC-GRB2 pathway. J Biol Chem 271:25639–25645. Puri, R.K., Hoon, D.S., Leland, P., et€al. 1996. Preclinical development of a recombinant toxin containing circularly permuted interleukin-4 and truncated Pseudomonas exotoxin for therapy of malignant astrocytoma. Cancer Res 56:5631–5637. Quang, T.S. and Brady, L.W. 2004. Radioimmunotherapy as a novel treatment regimen: 125I-labeled monoclonal antibody 425 in the treatment of high-grade brain gliomas. Int J Radiat Oncol Biol Phys 58:972–975. Raffel, C., Frederick, L., O’Fallon, J.R., et€al. 1999. Analysis of oncogene and tumor suppressor gene alterations in pediatric malignant astrocytomas reveals reduced survival for patients with PTEN mutations. Clin Cancer Res 5:4085–4090. Rao, S., Cunningham, D., de Gramont. A., et€al. 2004. Phase III double-blind placebo-controlled study of farnesyl transferase inhibitor R115777 in patients with refractory advanced colorectal cancer. J Clin Oncol 22:3950–3957. Raymond, E., Brandes, A., Van Oosterom, A., et€al. 2004. Multicentre phase II study of imatinib mesylate in patients with recurrent glioblastoma: an EORTC:NDDG/BTG intergroup study. Proceedings ASCO Annual Meeting 23:107 (abstract). Reardon, D.A., Akabani, G., Coleman, R.E., et€al. 2006a. Salvage radioimmunotherapy with murine iodine-131-labeled antitenascin monoclonal antibody 81C6 for patients with recurrent primary and metastatic malignant brain tumors: phase II study results. J Clin Oncol 24:115–122. Reardon, D.A., Egorin, M.J., Quinn, J.A., et€al. 2005a. Phase II study of imatinib mesylate plus hydroxyurea in adults with recurrent glioblastoma multiforme. J Clin Oncol 23:9359–9368. Reardon, D.A., Quinn, J.A., Vredenburgh, J., et€al. 2005b. Phase II trial of irinotecan plus celecoxib in adults with recurrent malignant glioma. Cancer 103:329–338.
264
I.F. Pollack
Reardon, D. A., Quinn, J. A., Vredenburgh, J.J., et€al. 2006b. Phase 1 trial of gefitinib plus sirolimus in adults with recurrent malignant glioma. Clin Cancer Res 12:860–868. Reardon, D. and Wen, P.Y. 2006. Therapeutic advances in the treatment of glioblastoma: rationale and potential role of targeted agents. Oncologist 11:152–164. Recht, L., Torres, C.O., Smith, T.W., Raso, V., Griffin, T.W. 1990. Transferrin receptor in normal and neoplastic brain tissue. Implications for brain tumor immunotherapy. J Neurosurg 72:941–945. Rowinsky, E K., Windle, J. J., Von Hoff, D.D. 1999. Ras protein farnesyltransferase: a strategic target for anticancer therapeutic development. J Clin Oncol 17:3631–3652. Rusnak, D.W., Affleck, K., Cockerill, S.G., et€al. 2001. The characterization of novel, dual ErbB-2/ EGFR, tyrosine kinase inhibitors: potential therapy for cancer. Cancer Res 61:7196–7203. Sampson, J.H., Akabani, G., Archer, G.E., et€al. 2003. Progress report of a phase I study of the intracerebral microinfusion of a recombinant chimeric protein composed of transforming growth factor (TGF)-a and a mutated form of the Pseudomonas exotoxin termed PE-38 (TP-38) for the treatment of malignant brain tumors. J Neurooncol 65:27–35. Sawa, H., Murakami, H., et€al. 2004. Histone deacetylase inhibitor, FK228, induces apoptosis and suppresses cell proliferation of human glioblastoma cells in€vitro and in€vivo. Acta Neuropathol 107:523–531. Schlessinger, J. 2000. Cell signaling by receptor tyrosine kinases. Cell 103:211–225. Schold, S.C., Brent, T.P., Von Hofe, E., et€al. 1989. 06 alkylguanine-DNA Alkyltransferase and sensitivity to procarbazine in human brain-tumor xenografts. J Neurosurg 70:573–577. Shamah, S.M., Stiles, C.D., Guha, A. 1993. Dominant-negative mutants of platelet-derived growth factor revert the transformed phenotype of human astrocytoma cells. Mol Cell Biol 13:7203–7212. Sharma, S., Kemeny, N., Kelsen, D.P., et€al. 2002. A phase II trial of farnesyl protein transferase inhibitor SCH 66336, given by twice-daily oral administration, in patients with metastatic colorectal cancer refractory to 5-fluorouracil and irinotecan. Ann Oncol 13:1067–1071. Silber, J.R., Mueller, B.A., Ewers, T.G., Berger, M.S. 1993. Comparison of O6-methylguanineDNA methyltransferase activity in brain tumors and adjacent normal brain. Cancer Res 53:3416–3420. Stark-Vance, V. 2005. Bevacizumab and CPT-11 in the treatment of relapsed malignant glioma. Neurooncology 7:369. Sung, T., Miller, D.C., Hayes, R.L., et€al. 2000. Preferential inactivation of the p53 tumor suppressor pathway and lack of EGFR amplification distinguish de€ novo high grade pediatric astrocytomas from de€novo adult astrocytomas. Brain Pathol 10:249–259. Taga, T., Suzuki, A., Gonzalez-Gomez, I., et€al. 2002. Alpha v-Integrin antagonist EMD 121974 induces apoptosis in brain tumor cells growing on vitronectin and tenascin. Int J Cancer 98:690–697. Tentori, L., Leonetti, C., Scarsella, M., et€al. 2003. Systemic administration of gpi 15427, a novel poly(adp-ribose) polymerase-1 inhibitor, increases the antitumor activity of temozolomide against intracranial melanoma, glioma, lymphoma. Clin Cancer Res 9:5370–5379. Tentori, L., Portarena, I., Torino, F., et€al. 2002. Poly(adp-ribose) polymerase inhibitor increases growth inhibition and reduces g(2)/m cell accumulation induced by temozolomide in malignant glioma cells. Glia 40:44–54. Topaly, J., Zeller, W.J., Fruehauf, S. 2001. Synergistic activity of the new ABL-specific tyrosine kinase inhibitor STI571 and chemotherapeutic drugs on the BCR-ABL-positive chronic myelogenous leukemia cells. Leukemia 15:342–347. Tuettenberg, J., Grobholz, R., Korn, T., et€ al. 2005. Continuous low-dose chemotherapy plus inhibition of cyclooxygenase-2 as an antiangiogenic therapy of glioblastoma multiforme. J Cancer Res Clin Oncol 131:31–40. Ullrich, A. and Schlessinger, J. 1990. Signal transduction by receptor with tyrosine kinase activity. Cell 61:203–212. Van Cutsem, E., van de, V., Karasek, P., Oettle, H., Vervenne, W.L., Szawlowski, A., et€al. 2004. Phase III trial of gemcitabine plus tipifarnib compared with gemcitabine plus placebo in advanced pancreatic cancer. J Clin Oncol 22:1430–1438.
Molecularly Targeted Therapies for Astrocytomas
265
von Deimling, A., Louis, D.N., Menon, A.G., et€al. 1993a. Deletions on the long arm of chromosome 17 in pilocytic astrocytoma. Acta Neuropathol 86:81–85. von Deimling, A., von Ammon, K., Schoenfeld, D., et€al. 1993b. Subsets of glioblastoma multiforme defined by molecular genetic analysis. Brain Pathol 3:19–26. Verhaak, R.G.W., Hoadley, K.A., Purdom, E., et al. 2010. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFR, IDH1, EGFR, and NF1. Cancer Cell 17:98–110. Vredenburgh, J., Desjardins, A., Herndon, J.E.II, et€al. 2006. Bevacizumab and irinotecan is an effective treatment for malignant gliomas. Neurooncology 8:454. Wakeling, A.E., Guy, S.P., Woodburn, J.R., et€al. 2002. ZD1839 (Iressa): An orally active inhibitor of epidermal growth factor signaling with potential for cancer therapy. Cancer Res 62:5749–5754. Walczak, H., Miller, R.E., Ariail, K., et€al. 1999. Tumoricidal activity of tumor necrosis factor-related apoptosis-inducing ligand in€vivo. Nat Med 5:157–163. Wang, D., Huang, H.J., Kazlauskas, A., Cavenee, W.K. 1999. Induction of vascular endothelial growth factor expression in endothelial cells by platelet-derived growth factor through the activation of phosphatidylinositol 3-kinase. Cancer Res 59:1464–1472. Weaver, M. and Laske, D.W. 2003. Transferrin ligand-targeted toxin conjugate (Tf-CRM107) for therapy of malignant gliomas. J Neurooncol 65:3–13. Weber, F.W., Floeth, F., Asher, A., et€ al. 2003. Local convection-enhanced delivery of IL4Pseudomonas exotoxin (NBI-3001) for treatment of patients with recurrent malignant glioma. Acta Neurochir 88(Suppl):93–103. Wedge, S. and Newlands, E. 1996. O6-Benzylguanine enhances the sensitivity of a glioma xenograft with low o6-alkylguanine-DNA alkyltransferase activity to temozolomide and BCNU. Br J Cancer 73:1049–1052. Weingart, J., Grossman, S.A., Carson, K.A., et€ al. 2007. Phase I trial of polifeprosan 20 with carmustine implant plus continuous infusion of intravenous O6-benzylguanine in adults with recurrent malignant glioma: new approaches to brain tumor therapy CNS consortium trial. J Clin Oncol 25:399–404. Weingart J, Tatter S, Rosenfeld S, et€al. 2003. Intratumoral convection-enhanced delivery of IL13PEQQR cytotoxin for recurrent malignant glioma without planned resection: a phase I/II study. Neurooncology 5:357. Wen, P.Y., Yung, W.K., Lamborn, K.R., et€al. 2006. Phase I/II study of imatinib mesylate for recurrent malignant gliomas: North American Brain Tumor Consortium Study 99-08. Clin Cancer Res 12:4899–4907. Wilhelm, S.M., Carter, C., Tang, L., et€al. 2004. BAY 43-9006 exhibits broad spectrum oral antitumor activity and targets the RAF/MEK/ERK pathway and receptor tyrosine kinases involved in tumor progression and angiogenesis. Cancer Res 64:7099–7109. Wong, A.J., Ruppert, J.M., Bigner, S.H., et€al. 1992. Structural alterations of the epidermal growth factor receptor gene in human gliomas. Proc Natl Acad Sci U S A 89:2965–2969. Wong, K.K., Chang, Y.M., Tsang, Y.T.M., et€al. 2005. Expression analysis of juvenile pilocytic astrocytomas by oligonucleotide microarray reveals two potential subgroups. Cancer Res 65:76–84. Yancopoulos, G.D., Davis, S., Gale, N.W., et€al. 2000. Vascular-specific growth factors and blood vessel formation. Nature 407:242–248. Yang, J.C., Haworth, L., Sherry, R.M., et€al. 2003. A randomized trial of bevacizumab, an anti-vascular endothelial growth factor antibody, for metastatic renal cancer. N Engl J Med 349:427–434. Yazaki, T., Ahmad, S., Chahlavi, A., et€ al. 1996. Treatment of glioblastoma U-87 by systemic administration of an antisense protein kinase C-a phosphorothioate oligodeoxynucleotide. Mol Pharmacol 50:236–242. Yin, D., Zhou, H., Kumagai, T., et€al. 2005. Proteasome inhibitor PS-341 causes cell growth arrest and apoptosis in human glioblastoma multiforme (GBM). Oncogene 24:344–354. Yoshida, M., Horinouchi, S., Beppu, T. 1995.Trichostatin A and trapoxin: novel chemical probes for the role of histone acetylation in chromatin structure and function. Bioessays 17:423–430.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy for Childhood Cancer Yoon-Jae Cho and Scott L. Pomeroy
Introduction Medulloblastomas are primitive embryonal tumors of the central nervous system arising exclusively in the cerebellum. They have a tendency to disseminate via CSF spaces throughout the brain and spine and are the most common malignant brain tumors in children (CBTRUS 2005). There has been considerable progress in our understanding of the mechanisms driving these tumors through the study of normal cerebellar development, familial cancer syndromes, genome-based analyses of human primary tumor samples (Table 1) and genetic mouse models. These studies have provided a framework for the development of several targeted therapies. This chapter focuses on our current molecular understanding of medulloblastoma and the targeted therapies in development and clinical trials.
Histopathological and Clinical Considerations Medulloblastomas are a heterogeneous group of small blue cell tumors. The 2007 World Health Organization (WHO) classification differentiates subtypes of medulloblastoma by specific, sometimes subtle, and histological features. The major recognized subtypes are classic, desmoplastic/nodular, and medulloblastoma with extensive nodularity, large cell, and anaplastic variants (Louis et€ al. 2007). Rarer subtypes include melanotic and medullomyoblastoma. Classic medulloblastomas are composed of sheets of small round blue cells with high nuclear to cytoplasmic ratios and, typically, uniform nuclei set in a fibrillary background. There may be signs of neuronal differentiation in the form
S. L. Pomeroy (*) Department of Neurology, Children’s Hospital Boston, 300 Longwood Avenue, Enders 270, Boston, MA 02115, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_13, © Springer Science+Business Media, LLC 2010
267
Nijmegen breakage syndrome
NBS1
8q21
Medulloblasotma, lymphoma, leukemia, rhabdomyosarcoma
Table€1â•… Cancer predisposition syndromes associated with medulloblastoma Syndrome Gene Locus Tumor spectrum Gorlin syndrome PTCH1 9q22.3 Medulloblastoma, basal cell carcinoma Turcot syndrome APC 5q21 Medulloblastoma, colorectal adenoma Li Fraumeni syndrome TP53 17p13.1 Medulloblastoma, leukemia, sarcomas, carcinoma Fanconi anemia, type D BRCA2 13q12.3 Medulloblasotma, leukemia, breast cancer, Wilms’ tumor
References Hahn et€al. (1996), Johnson et€al. (1996) Hamilton et€al. (1995) Pearson et€al. (1982), Barel et€al. (1998), Guran et€al. (1999) de Chadarevian et€al. (1985), Ruud and Wesenberg (2001), Offit et€al. (2003), Hirsch et€al. (2004), Tischkowitz et€al. (2004) Bakhshi et€al. (2003), Distel et€al. (2003)
268 Y.-J. Cho and S.L. Pomeroy
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
269
of Homer-Wright rosettes or occasional groups of ganglion cells. To a small degree, differentiation along glial lineages as well as nuclear pleomorphism can be seen. However, when nuclear pleomorphism is abundant, with cell molding, numerous mitotic figures and apoptotic bodies, the tumor is considered an anaplastic variant. Large cell medulloblastomas contain groups of cells with round nuclei, open chromatin, and a single nucleolus and often contains regions similar to the anaplastic phenotype (Louis et€al. 2007). Due to the considerable overlap of large cell and anaplastic medulloblastomas, these two subtypes are often thought of as a continuum of the same tumor. Medulloblastomas may also show varying degrees of a nodular architecture which largely defines the desmoplastic/nodular and medulloblastoma with extensive nodularity subgroups. These tumors are characterized by scattered nodules of neurocytic cells among desmoplastic regions containing larger and more pleomorphic nuclei. In the desmoplastic subtype, the “pale islands” of tumor cells are surrounded by a dense fibrous, reticulin rich stroma (McManamy et€ al. 2007). Medulloblastomas with extensive nodularity in comparison have less of the dense reticulin but instead a markedly expanded lobular architecture due to the fact that the reticulin-free zones become unusually large and rich in neuropil-like tissue. This confers a “grape-like” appearance on neuroimaging and gross pathology.
Risk Stratification Several studies have associated histological subtype with outcome, in particular, desmoplastic tumors with good prognosis and large cell/anaplastic variants with poor prognosis (Pomeroy et€al. 2002; Gajjar et€al. 2006). However, intratumoral heterogeneity is not uncommon and a single tumor may have regions suggestive of one subtype and other regions suggestive of a different subtype. This often makes definitive subtyping via traditional histopathology techniques difficult. Various molecular and/or cytogenetic features have been assigned to help define histological subtypes and although they are increasingly incorporated into diagnostic schemas, they are neither uniformly used nor standardized. Thus, current risk stratification schemas are based solely on clinical criteria. This includes the age at diagnosis, the extent of surgical resection, and the presence or absence of tumor dissemination at diagnosis. Specific multimodal treatment regimens based on this clinical stratification include surgical resection, age appropriate craniospinal irradiation, and conventional chemotherapy. This schema has afforded 5-year survival rates of up to 85% for standard risk patients and 60 to 80% for high risk groups (Crawford et€ al. 2007). Survivorship, however, almost universally comes at some cost with long-term neurological and neurocognitive deficits resulting from aggressive therapies on the developing nervous system. Thus, emphasis is
270
Y.-J. Cho and S.L. Pomeroy
placed on the identification of molecular targets to which less toxic targeted therapies can be developed and applied.
Cellular Origins of Medulloblastoma As medulloblastomas arise exclusively in the cerebellum, studies of cerebellar development have provided significant insight into its origin. Fate-mapping experiments in the mouse and chick reveal most, if not all, cell types in the mature Â�cerebellum are derived from two germinal zones: the ventricular zone and the rhombic lip (Hatten et€ al. 1997; Lin and Cepko 1999; Machold and Fishell 2005; Wang et€al. 2005a). The ventricular zone generates Purkinje neurons and various deep cerebellar nuclei, including the dentate nucleus. In addition, this region contributes Bergmann glia and other astroglial derivatives (Sottile et€al. 2006). Precursors from the rhombic lip, however, appear to be more restricted in their differentiation potential and are fated to become granule cell neurons, the most abundant neuronal population in the mature central nervous system (Alder et€al. 1996; Goldowitz and Hamre 1998). During development, granule neuron precursor cells stream from the rostral portion of the rhombic lip around the outermost boundary of the cerebellum forming the external granular layer (EGL). Here, they undergo tremendous proliferation and upon exiting the cell cycle migrate to the internal granular layer, where they become mature granule cell neurons. Mouse studies have shown Sonic Hedgehog (SHH) expressed from Purkinje neurons that regulate the proliferation of granule cell progenitors (Wallace 1999). This proliferation is completed within the first 3€weeks in the postnatal mouse, and the EGL is largely gone by 1€month of age in mouse and 1 to 2€years of age in humans (Wechsler-Reya and Scott 1999). Their abundance and prolonged proliferative phase during cerebellar development make granule neuron precursors an attractive candidate as the cell of origin in medulloblastoma. This is especially the case for desmoplastic medulloblastomas whose gene-expression patterns indicate dysregulated SHH signaling, a key regulator of granule cell proliferation and specification (Pomeroy et€al. 2002). The cell of origin for the majority of classic and large cell/anaplastic medulloblastomas is less clear. Conceivably, these subgroups might arise from granule cell precursors through mutations involving other molecular pathways. Alternatively, they might arise from different cells of origin (Katsetos et€ al. 1995). Indeed, the Wingless (WNT) signaling pathway is important in the support and maintenance of ventricular zone precursors and is pathologically activated in a subset of classic medulloblastomas (Ellison et€al. 2005; Clifford et€al. 2006; Salsano et€al. 2007). Recently, there has been great interest in characterizing subpopulations of tumor cells which have stem cell-like properties. These “cancer stem cells” have an inher-
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
271
ent capacity for self-renewal and multipotentiality and are proposed to serve as the root of cancers. They, therefore, represent a very attractive therapeutic target. The evidence for the existence of such cancer stem cells in medulloblastoma comes from the identification of a defined population of medulloblastoma cells expressing the neural stem cell marker, CD133 (Singh et€al. 2003, 2004; Peterson et€al. 2005). These CD133+ cells have an increased ability to propagate or “self-renew” in primary and secondary neurosphere cultures and xenograft assays when compared to CD133- medulloblastoma cells. In addition, the stem cell factor BMI1 is highly expressed in some medulloblastomas (Singh et€al. 2003, 2004). Mice lacking Bmi1 have decreased pools of progenitor cells in the subventricular and proliferative regions of the cerebellum with a distinct lack of robust neurosphere growth ex vivo (Leung et€al. 2004; Bruggeman et€al. 2005; Wiederschain et€al. 2007). The real-world implications for targeting cancer stem cells and the mechanisms that maintain their “stemness” still need to be tested in proof of principle studies. To our knowledge, there are no currently available therapies that directly target cancer stem cells in medulloblastoma. However, several pathways implicated in medulloblastoma pathogenesis, such as SHH signaling, WNT signaling, and NOTCH signaling, have also been implicated in the maintenance of normal stem cell compartments (Ruiz i Altaba et€al. 2002; Radtke and Clevers 2005; Reya and Clevers 2005; Androutsellis-Theotokis et€ al. 2006; Ward et€ al. 2006; Dreesen and Brivanlou 2007).
Molecular Features of Medulloblastoma and Targeted Strategies Several molecular pathways when dysregulated have been identified as pathogenic in medulloblastoma, either causally or in a supportive role. Current targeted strategies employ agents directed toward a single or a number of gene products. Alternatively, compounds that affect broader cellular functions have been developed (Fig.€1). The most prolific work involves targeting of the SHH pathway in medulloblastoma or pathways that cooperate with SHH activation in promoting tumorigenesis. This is largely due to the availability of robust preclinical mouse models to study, design, and test targeted therapies, and the discovery of cyclopamine, a plant derived SHH pathway inhibitor. Thus, current targeted therapies are somewhat biased toward the treatment of SHH-activated or desmoplastic/nodular subtypes of medulloblastoma. Some of these therapies have been proposed to work across all medulloblastoma subtypes; however, this has not been definitively proven. Small molecule screens have also generated numerous drugs directed at specific tyrosine kinase inhibition. Several of these compounds are in clinical trials and even clinical use for various types of cancer. Given many of these receptor tyrosine kinases are activated in medulloblastoma, these small molecule kinase inhibitors are currently being developed for trials in medulloblastoma as well.
272
Y.-J. Cho and S.L. Pomeroy Cyclopamine HhAntag
R1507 Ab
CXCL12
NVP-AEW541
TrkC EGFR/ERBB IGF1R PDGFR CXCR4
AMD3100 AMD3465
Lapatinib Erlotinib
PI3K
WNT
SHH
FZD
PTCH1
GLI Fused SUFU
GSK3B
+ VPA / Vorinostat
BMP2
differentiation apoptosis
Rapamycin Everolimus
AKT
NOTCH
γ-secretase
DSH
PtdIns
13-cis-retionic acid
SMO
APC β-catenin
DAPT
AXIN
mTOR
β-catenin
c-MYC TCF/LEF
N-MYC GLI
PROLIFERATION, SURVIVAL
Fig.€1╅ Molecular pathways and targeted therapies in medulloblastoma. Current pathways implicated in medulloblastoma tumorigenesis and reagents shown to inhibit proliferation and/or tumor cell survival in€vitro and in mouse models are outlined
Sonic Hedgehog The SHH pathway was first implicated in medulloblastoma tumorigenesis through the association of medulloblastoma with Gorlin syndrome, an autosomal dominant disorder caused by germline mutation of the SHH receptor PTCH1 (Hahn et€ al. 1996; Johnson et€al. 1996) Table 1. Approximately 2 to 5% of patients with Gorlin syndrome develop medulloblastoma, specifically the desmoplastic variant (Evans et€al. 1991a, b). Overall, activation of SHH signaling is seen in 20 to 30% of medulloblastomas with the majority of these tumors harboring inactivating mutations in the PTCH1 tumor suppressor gene (Pietsch et€al. 1997; Raffel et€al. 1997). PTCH1 is a multipass membrane protein which binds SHH. In the absence of SHH, PTCH1 sequesters and prevents the activation of Smoothened (SMO) at the cell membrane. When bound by SHH, PTCH1 releases SMO to transactivate the Glioma-associated Oncogene Homolog (GLI) family of transcription factors via Suppressor of Fused (SUFU), resulting in upregulation of transcriptional targets such as MYCN and Cyclin D1 (Pomeroy et€al. 2002; Kenney et€al. 2003; Oliver et€al. 2003) Table€2. Somatic activating mutations in SMO have been found in medulloblastomas, although with very low frequency, whereas upregulation of GLI family members is uniformly seen in desmoplastic medulloblastomas (Lam et€al. 1999). Conversely, truncating mutations in SUFU, which normally acts to sequester GLI in a cytoplasmic
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy Table€2╅ Mutations demonstrated in sporadic medulloblastoma % Mutation # of cases in study SHH signaling pathway PTCH1 10 7 of 68 20 5 of 24 10 5 of 46 SMO 4 1 of 21 6 1 of 15 SUFU 9 4 of 46 3 2 of 77 0 132 GLI3 0 12 WNT signaling pathway b-catenin 5 3 of 80 9 4 of 46 4 3 of 67 10 5 of 51 10 7 of 77 25 8 of 32 25a 27 of 109 APC 2 2 of 46 4 3 of 77 AXIN1 4 1of 23 0 0 of 46 1 1 of 86 5 2 of 39 AXIN2 1 1 of 116 Other TRP53 0 0 of 12 11 1 of 9 23 5 of 22 11 2 of 19 c-MYC genomic 16 4 of 24 amplification 6 4 of 69 5 4 of 77 1 1 of 71 6 4 of 71 N-MYC genomic amplification 21 4 of 19 13 2 of 15 6 4 of 77 a Nuclear localization of b-catenin
273
References Pietsch et€al. (1997) Raffel et€al. (1997) Thompson et€al. (2006) Lam et€al. (1999) Reifenberger et€al. (1998) Taylor et€al. (2002) Thompson et€al. (2006) Koch et€al. (2004) Erez et€al. (2002) Koch et€al. (2001) Huang et€al. (2000) Zurawel et€al. (1998) Eberhart et€al. (2000) Thompson et€al. (2006) Clifford et€al. (2006) Ellison et€al. (2005) Huang et€al. (2000) Thompson et€al. (2006) Yokota et€al. (2002) Thompson et€al. (2006) Dahmen et€al. (2001) Baeza et€al. (2003) Koch et€al. (2007) Saylors et€al. (1991) Adesina et€al. (1994) Badiali et€al. (1993) Ohgaki et€al. (1991) Rossi et€al. (2006) Ebinger et€al. (2006) Aldosari et€al. (2002) De Bortoli et€al. (2006) De Bortoli et€al. (2006) Bayani et€al. (2000) Michiels et€al. (2002) Aldosari et€al. (2002)
ubiquitination/degradation complex, have been reported in as many as 9% of sporadic medulloblastomas (Taylor et€al. 2002). Preclinical mouse models of SHH-activated medulloblastomas which faithfully recapitulate the human condition in many regards have been developed Table€ 3. Mice heterozygous for mutation of the murine homolog of the SHH receptor PTCH1 develop medulloblastomas at a low frequency (10 to 20% depending on
274 Table€3â•… Mouse models of medulloblastoma Mouse model Type Ptch1+/− Germline knockout Germline knockout Ptch1+/−;trp53−/− DNAlig4−/−;trp53−/− Germline knockout Parp1−/−;trp53−/− Germline knockout Brca2−/−;trp53−/− Conditional Xrcc4−/−;p53−/− Conditional Rb−/−;trp53−/− Conditional Ink4c−/−;trp53−/− Conditional Ku80−/−;Rag1−/−;trp53−/− Germline knockout Sufu+/−;trp53−/− Germline knockout ND2-SmoA1 transgenic Transgene (NeuroD2 promoter) SHH with AKT and/or IGF2 Retroviral infection at E13.5
Y.-J. Cho and S.L. Pomeroy
References Goodrich et€al. (1997) Wetmore et€al. (2001) Lee and McKinnon (2002) Tong et€al. (2003) Frappart et€al. (2007) Yan et€al. (2006) Marino et€al. (2000) Uziel et€al. (2006) Holcomb et€al. (2006) Lee et€al. (2007) Hallahan et€al. (2004) Rao et€al. (2004)
background strain). When crossed into a Trp53−/− background, latency of tumor formation is dramatically reduced and penetrance increases to 100% (Goodrich et€al. 1997). Another mouse model overexpressing an activated form of Smo under the control of a NeuroD2 promoter also develops medulloblastoma in 48% of mice regardless of Trp53 status (Hallahan et€ al. 2004). In utero injection of a SHH expressing retrovirus in E13.5d embryos results in medulloblastoma in as many as 76% of infected mice (Weiner et€al. 2002). There are two well-published inhibitors of the SHH pathway that work at the level of SMO: cyclopamine and HhAntag (Cooper et€al. 1998; Romer et€al. 2004). Inhibitors that target the downstream SHH signaling mediator GLI have also been developed but have yet to be studied in medulloblastomas specifically (Lauth et€al. 2007). Cyclopamine (11-deoxojervine) is a naturally occurring chemical that belongs to the group of steroidal jervatum alkaloids. It is derived from the corn lily Veratum californicum and was discovered in 1957 as a causative teratogen in endemic cases of cyclopic/holoprosencephalic lambs born to pregnant sheep grazing on the corn lily plant in Idaho pastures (Keeler 1970, 1978). HhAntag, on the other hand, was discovered by a high-throughput cell-based screening assay designed to identify small-molecule inhibitors of the SHH pathway (Williams et€al. 2003). HhAntag, a benzimidazole derivative, has a higher affinity for Smo than cyclopamine and blocks SHH signaling at greater than tenfold lower concentrations (Romer et€al. 2004). Both cyclopamine and HhAntag were very efficient at inhibiting medulloblastoma formation in Ptch1+/−;Trp53−/− mice. In addition, the treatment of mice which had spontaneous tumors or subcutaneously allografted tumors, inhibited tumor growth and promoted regression via induction of differentiation or apoptosis. Although lower doses neither completely nor permanently eliminated medulloblastomas in the Ptch1+/−;Trp53−/− mice, higher doses and longer treatments with HhAntag often resulted in complete tumor eradication and overall event free survival was markedly prolonged (Romer and Curran 2004). Importantly, both cyclopamine and HhAntag appeared to be very well tolerated and no noticeable adverse reactions were reported in mice, even at the highest doses
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
275
of HhAntag (Berman et€al. 2002; Romer and Curran 2004). However, further studies demonstrated that transient inhibition of the Hedgehog pathway by HhAntag caused permanent defects in bone growth (Kimura et€al. 2008). At the time of this writing, several SMO inhibitors have been recently developed and are in early phase clinical trials for adult and pediatric malignancies, including medulloblastoma.
WNT/b-Catenin Activation of the Wnt/beta-catenin pathway in medulloblastoma was first established by the association of medulloblastoma with Turcot syndrome, also referred to as Brain Tumor Polyposis syndrome. Turcot syndrome results from mutations in the tumor suppressor gene, adenomatous polyposis coli (APC) (Hamilton et€ al. 1995). APC is part of a multimeric complex which functions to sequester b-catenin in the cytosol, preventing its nuclear translocation where it normally interacts with the T-Cell Factor (TCF)/LEF family of transcription factors to transcriptionally regulate WNT target genes such as c-MYC, MYCN, and Cyclin D1(Gordon and Nusse 2006). Mutations found in APC prevent this sequestration of b-catenin allowing for its nuclear translocation and transcriptional activity. Activation of the WNT/b-catenin pathway as inferred from the nuclear localization of b-catenin occurs in 15 to 20% of medulloblastoma (Eberhart et€al. 2000; Salaroli et€al. 2007). However, mutations in the APC gene in sporadic medulloblastoma are rare, occurring in only 1 to 2% of tumors (Huang et€ al. 2000). Mutations in the b-catenin gene contribute to 5 to 10% of medulloblastomas (Clifford et€al. 2006). Mutations in other genes involved in the positive or negative regulation of the WNT pathway have been reported with rare frequency (Baeza et€ al. 2003; Koch et€al. 2007). Mutations found in the b-catenin gene confer stability to this oncogene by preventing its ubiquitination and thus degradation/downregulation. This results in stabilization of b-catenin, constitutive nuclear localization and thus transcriptional activation (Eberhart et€ al. 2000; Koch et€ al. 2001; Clifford et€ al. 2006). Interestingly, aCGH and SNP microarray studies have identified a subset of medulloblastomas in which the only cytogenetic feature is the numerical loss of one chromosome 6 copy. All of these tumors are of the classic hitological subtype and on immunohistological staining exhibit nuclear localization of b-catenin (Clifford et€al. 2006; Thompson et€al. 2006). Two reports have suggested better the overall survival of patients with this subtype of tumor, but larger scale analyses are necessary to substantiate these findings (Clifford et€al. 2006; Thompson et€al. 2006). Nonetheless, these findings clearly identify a subset of medulloblastomas with WNT/b-catenin pathway activation which would be amenable to targeted therapy of this pathway. Although there are currently no mouse models of medulloblastoma derived from aberrant WNT pathway activation, targeted therapies designed against the WNT pathway are being developed for use in other types of cancers. Preclinical studies have shown some promise of ZTM000990 and PKF118-310, and lead compounds
276
Y.-J. Cho and S.L. Pomeroy
targeting the canonical WNT signaling cascade. In addition, anti-WNT1 and anti-WNT2 monoclonal antibodies show in€ vitro effects in various tumor types (Barker and Clevers 2006).
NOTCH NOTCH family members have important roles in a variety of developmental processes via regulation of cell–cell interactions. The NOTCH proteins function as receptors for membrane bound ligands (Lutolf et€ al. 2002). On interaction with these ligands, the NOTCH protein is cleaved in two places. The first, just outside the membrane, releases the extracellular portion of Notch, which remains bound to the ligand and is endocytosed by the ligand-expressing cell. The second cleavage occurs just inside the inner leaflet of the cell membrane and is mediated by g-secretase (Shih Ie and Wang 2007). This releases the intracellular portion of the NOTCH protein, which then moves to the nucleus upregulating various genes, including HES1 and HES5. The NOTCH family member, NOTCH2, and downstream signaling partners are upregulated in some primary medulloblastomas suggesting their role in medulloblastoma pathogenesis. The upregulation of these genes is particularly evident in medulloblastomas derived from SHH pathway activation (Hallahan et€al. 2004; Dakubo et€al. 2006). Inhibitors of g-secretase have been developed to specifically target the NOTCH endoproteolytic cleavage step which is necessary for downstream NOTCH signaling (Dovey et€ al. 2001; Sastre et€ al. 2001; Shih Ie and Wang 2007). Treatment of medulloblastoma cell lines with N-S-phenyl-glycine-t-butyl ester (DAPT) in particular resulted in 40 to 60% decrease in total viable cell number after 48€hours of treatment (Hallahan et€ al. 2004). Primary mouse tumors derived from NeuroD1Smo transgenic mice and human primary tumor cell line xenografts had equal responses to DAPT or cyclopamine. Moreover, combined SHH and NOTCH pathway inhibition had an additive effect in these models suggesting the importance of NOTCH signaling in the setting of SHH activation. However, antitumoral effects afforded by DAPT diminished after 2€weeks, presumably due to an increase in drug metabolism (Hallahan et€ al. 2004). Regardless, these studies establish NOTCH signaling as an important regulator of medulloblastoma tumorigenesis and identify it as a target for therapies. Currently, there are no NOTCH pathway inhibitors in clinical trials for medulloblastoma, however several inhibitors are currently in development or in early phase trials (MK0752) for other tumor types.
IGFR The insulin-like growth factor-1 receptor (IGF1R) is a membrane-associated tyrosine kinase which together with its corresponding ligands, IGF1 and IGF2, is expressed during embryonic and early postnatal cerebellar development (Rotwein
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
277
et€al. 1988; Baserga et€al. 1994; Lee et€al. 1995). This expression is restricted to granule neuron precursor cells and decreases significantly during postnatal development (LeRoith et€al. 1993). IGF1R, like other receptor tyrosine kinases, signals through several cascades, including phosphatidylinositol-3¢ kinase (PI3K) and its second messengers. High-throughput gene expression profiling of primary medulloblastomas has shown increased expression of IGF2, mainly in desmoplastic/nodular tumors (Glick et€al. 1993; LeRoith et€al. 1993; Pomeroy et€al. 2002). Moreover, homozygous deletion of the IGF2 gene completely eliminates tumor formation in Ptch1+/− mice, suggesting IGF2 and its downstream signaling partners are essential for SHH mediated medulloblastoma tumorigenesis (Hahn et€al. 2000). In addition, retroviral transduction with combinations of SHH, IGF2, and activated AKT into Nestin-expressing neural progenitors induced tumor formation in mice, with the combination of SHH/AKT (48%) having the strongest oncogenic transformation effect compared to SHH/IGF2 (39%) and SHH alone (15%) (Rao et€al. 2004). Importantly, neither IGF2 nor AKT caused tumors when expressed independently emphasizing SHH signaling as the dominant pathway in medulloblastoma pathogenesis. Induced tumors showed upregulation of insulin receptor substrate-1 (Irs1) expression and phosphorylation of IGF1R and AKT, consistent with activated IGF signaling seen in primary human medulloblastomas (Del Valle et€al. 2002). Studies implementing anti-IGF1 and anti-IGF1R neutralizing antibodies, dominant negative mutants, and RNA interference have demonstrated efficacy of IGF1R targeting in inhibiting proliferation and tumorigenesis in€vitro (Dunn et€al. 1998; Reiss et€al. 1998; Liao and Wang 2005; Wang et€al. 2005b; Araki et€al. 2006). Currently, several highly specific anti-IGF1R antibodies are in the early phases of clinical trials. The use of a blocking monoclonal anti-IGF1R antibody has previously been shown to decrease the growth of the Daoy medulloblastoma cell line (Chin et€ al. 1996). Although this Phase I trial has not yet opened, patients with recurrent or refractory advanced solid tumors, including medulloblastoma, are expected to be enrolled in a trial to evaluate one such antibody, R1507. This study is entitled “Multiple Ascending Dose (MAD) Phase I Study of the IGF1R Antagonist R1507 Administered as an Intravenous Infusion in Children and Adolescents With Advanced Solid Tumors”. In addition, several small molecule inhibitors of IGF1R, which specifically inhibit IGF-mediated IGF1R autophosphorylation and abrogate its downstream signaling cascade, have been developed (Garcia-Echeverria et€al. 2004; Scotlandi et€al. 2005). Preclinical studies using NVP-AEW541, a pyrrolo[2,3-d]pyrimidine derivative, show growth inhibition of mouse and human medulloblastoma cell lines in€ vitro (Urbanska et€ al. 2007). Interestingly, this antitumoral effect of NVPAEW541 was enhanced in anchorage-independent culture conditions and was potentiated by inhibition of GSK3b constitutive phosphorylation (which is coincidentally observed when mouse medulloblastoma cells are deprived of attachment). The authors, therefore, proposed that the use of this inhibitor in combination with agents activating GSK3b (dephosphorylation) might be particularly effective
278
Y.-J. Cho and S.L. Pomeroy
against disseminated medulloblastoma (Urbanska et€ al. 2007). Although clinical trials have not formally been introduced for medulloblastoma specifically, many of these IGF1R tyrosine kinase inhibitors are in various phases of clinical trials for other cancers and diseases associated with increased IGFR signaling.
CXCR4 The chemokine receptor CXCR4 is essential for the normal development of the cerebellar cortex and targeted gene deletion leads to mislocalization and limited proliferation of granule neuron precursor cells (Stumm et€al. 2003; Vilz et€al. 2005). In addition, increased expression of CXCR4 has been documented in desmoplastic medulloblastomas, medulloblastomas with extensive nodularity, and medulloblastomas generated in Ptch1+/− mice (Rubin et€ al. 2003; Schuller et€ al. 2005; Yang et€al. 2007). Targeting of CXCR4 with the specific antagonists AMD3100 and AMD3465 inhibits the growth of medulloblastoma cell line xenografts, reportedly increasing apoptosis and decreasing the proliferation of tumor cells (Rubin et€al. 2003; Yang et€al. 2007). It was also shown that the CXCR4 ligand, CXCL12, mediated tumor growth through sustained inhibition of cyclic AMP (cAMP) production, and AMD3465 blocks this cAMP suppression. Given these observations, the phosphodiesterase inhibitor (PDI), Rolipram, was used to elevate pharmacologically the levels of cAMP in medulloblastoma cell line-xenografted mice. The effect was comparable to AMD3465 but only resulted in 58% of tumor regression of the medulloblastoma cell line-xenograft. However, the combined effect of another PDI and CXCR4 antagonist (caffeine and AMD3465, respectively), resulted in an 85% reduction in tumor size in this same xenograft model. These data indicate the clinical use of PDIs and CXCR4 inhibitors that might be an effective strategy in the treatment of medulloblastoma (Yang et€al. 2007). Currently, there are no clinical trials offering PDI or CXCR4 inhibitor in the treatment of medulloblastoma.
PI3K/AKT/mTOR The mammalian target of rapamycin (mTOR) pathway is a central integrator of signals from nutrients, energy status, and growth factors. This pathway regulates growth and proliferation by modulating numerous cellular functions, including autophagy, ribosome biogenesis, and metabolism (Sabatini 2006; Guertin and Sabatini 2007). mTOR is a serine/threonine kinase that is activated by various receptor tyrosine kinases via phosphorylation and activation of PI3K and AKT. Briefly, the activation of PI3K generates phosphatidylinositol (3,4,5) trisphosphate which recruits AKT to the plasma membrane. AKT phosphorylation allows for the subsequent activation of mTOR through several intermediates. Tyrosine kinase receptors such as IGFR, PDGFR, TRKC, NGFR, and EGFR, are activated in
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
279
medulloblastoma primary tumors and cell lines and phosphorylated forms of AKT and PI3K have been demonstrated (Wlodarski et€al. 2006). Together, this indicates the potential utility of mTOR inhibitors in medulloblastoma treatment. Several mTOR inhibitors, such as rapamycin, RAD001 (Everolimus), and CCI779 (Temsirolimus), are being evaluated as novel agents for cancer therapy including medulloblastoma (Geoerger et€ al. 2001; Bjornsti and Houghton 2004; Beuvink et€ al. 2005; O’Reilly et€ al. 2006). Phase I and II cancer trials have shown that mTOR inhibitors are well tolerated by patients, with minimal adverse effects being reported. In particular, chronic immunosuppression has not been reported with the dose schedules used in current clinical trials (Yee et€al. 2006). However, rapamycin alone produces only partial responses in medulloblastoma (Geoerger et€al. 2001). Indeed, most Phase II cancer trials to date have shown limited effectiveness of rapamycin and its analogs as single agent therapy (Chang et€al. 2005; Galanis et€al. 2005). Thus, it has been suggested that rapamycin analogs should be used in conjunction with conventional chemotherapeutic agents or radiation therapy to increase efficacy. One alternative approach in mouse models combined rapamycin with intratumoral injection of oncolytic viruses. This combination was effective in promoting tumor regression in cell-line derived medulloblastoma xenografts (Lun et€al. 2007). Currently, a clinical trial entitled “Phase I/II Trial of RAD001 (Everolimus) in Pediatric Patients With Recurrent Refractory Solid Tumors or Brain Tumors With Phase II Limited to Recurrent or Refractory Rhabdomyosarcomas and Non Rhabdomyosarcomatous Soft Tissue Sarcomas” is open for enrollment.
EGFR Family The EGFR family members, ERBB1–ERBB4, have been shown to be differentially expressed in medulloblastoma cell lines and primary tumors. Overexpression of ERBB2, in particular, was evident in a large number of medulloblastomas and also reported as a negative prognostic factor (Gilbertson et€ al. 1992, 1995, 1997). Several inhibitors of EGFR family members have been developed as it is an important regulator in different types of cancer (Dancey 2004; Ciardiello 2005). These include Gefitinib and Lapitinib, which specifically inhibit ERBB1 and ERBB2; Herceptin, a monoclonal antibody to ERBB2; and Erlotinib, a specific inhibitor of ERBB1. Erlotinib and Lapitinib are currently in Phase I and II trials, respectively, for patients with medulloblastoma. The first trial entitled “Phase I Studies of TARCEVA™ (ERLOTINIB HYDROCHLORIDE, OSI-774) as Single Agent in Children With Refractory and Relapsed Malignant Brain Tumors and in Combination With Irradiation in Newly Diagnosed Brain Stem Glioma” is active but no longer recruiting patients. A second Phase II clinical trial entitled “A Molecular Biology And Phase II Study Of Lapatinib (GW572016) In Pediatric Patients With Recurrent Or Refractory Medulloblastoma, Malignant Glioma Or Ependymoma” is currently open for enrollment.
280
Y.-J. Cho and S.L. Pomeroy
PDGFR Activation of PDGFRb signaling was published as a marker for metastatic medulloblastomas in studies utilizing high-throughput gene expression arrays (MacDonald et€ al. 2001; Gilbertson and Clifford 2003). PDGFR is an attractive candidate for molecular targeting as Imatinib (Gleevec), the prototypical small molecule inhibitor, inhibits both the a and b isoforms of this receptor tyrosine kinase. In preclinical studies, blocking monoclonal antibodies to PDGFR and a small molecule inhibitor to MAP2K1, MAP2K2, and MAPK1/3, mediators of the PDGFR signaling cascade, were shown to inhibit migration of medulloblastoma cells in€vitro (MacDonald et€al. 2001; Gilbertson and Clifford 2003). Currently, there are no clinical trials of PDGFR inhibitors for use in patients with medulloblastoma.
TRKC TrkC is the preferred receptor for NT3, a neurotrophin that regulates granule cell development. TrkC is expressed in granule cell neurons where it promotes axonal maturation. It is also expressed in subsets of medulloblastomas and importantly was the first receptor tyrosine kinase to be associated with clinical significance in embryonal tumours (Segal et€al. 1994; Eberhart et€al. 2004). In particular, elevated expression in medulloblastomas is associated with favorable clinical outcome, presumably through the induction of apoptosis via NT3-TrkC signaling mechanisms (Kim et€al. 1999). Currently, there are no TrkC agonists or NT3 analogs in development or clinical trials.
HDAC Inhibitors and Retinoic Acid There is some evidence in cell culture and xenograft models suggesting the utility of histone deacetylase (HDAC) inhibitors in treating primary medulloblastomas (Li et€ al. 2005; Shu et€ al. 2006; Spiller et€ al. 2006, 2008). Valproic acid (VPA) and suberoylanilide hydroxamic acid (SAHA, marketed as Vorinostat) represent two HDAC inhibitors that are currently in early phase clinical trials. The suggested antitumoral mechanism in medulloblastoma appears to be sensitization of cells to retinoic acid receptor pathways. These have independently been shown to induce terminal differentiation and apoptosis in medulloblastoma cell lines and xenograft models through upregulation of BMP2 and caspase activity (Spiller et€ al. 2006, 2008). Thus, clinical trials have been designed combining the treatment of HDAC inhibitors with 13-cis-retinoic acid. Currently, there are two clinical trials utilizing this approach. The first, entitled “A Phase I Study of Valproic Acid in Children With Recurrent/Progressive Solid Tumors Including CNS Tumors” is no longer open for enrollment, but is still active.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
281
The second, entitled “A Phase I Study of SAHA (NSC#: 701852 IND#: 71976) in Pediatric Patients With Recurrent or Refractory Solid Tumors (Including Lymphomas) and Leukemia Followed By a Phase I Study of SAHA in Combination With 13-Cis-Retinoic Acid for Patients With Selected Recurrent/Refractory Solid Tumors” is currently open for enrollment.
Future Directions and Challenges There is clear evidence that currently available targeted therapies for medulloblastoma could offer alternatives or adjuvants to chemo and radiotherapies. Moreover, high-throughput screening methods for drug discovery and development present increasing opportunities to rapidly take biological discoveries from the bench to preclinical testing and ultimately to clinical trials and the bedside. However, several challenges will need to be addressed prior to such treatments arriving at the bedside. First, the establishment of a definitive molecular taxonomy for medulloblastoma is required. A move toward a more stringent molecular definition for each subtype of medulloblastoma will be the first step in transitioning targeted therapies into real world personalized applications. This type of schema is needed to identify patients for which current and future targeted therapies will work. Although some molecular markers and cytogenetic features have been in use for several years, they are neither uniformly used nor standardized. Whether newer markers involve the use of gene “signatures” generated from array-based studies or more predictive single or sets of biomarkers remain to be seen. Validation through CLIA certification will be necessary. Refinement of the current histopathological schema for medulloblastoma is an alternative, but will be challenged with the subjective and inconsistent nature of this diagnostic methodology. Secondly, our understanding of the biology of medulloblastomas is incomplete. Much of the current literature and development of targeted therapies is heavily weighted toward the desmoplastic/nodular class of medulloblastomas due to the informativity of the Ptch1+/− mouse model. In vivo models of medulloblastoma driven by mechanisms unrelated to SHH is somewhat lacking. There are several DNA repair deficient mice that generate medulloblastomas (Lee and McKinnon 2002; Tong et€al. 2003; Holcomb et€al. 2006; Yan et€al. 2006; Frappart et€al. 2007). However, these genes have yet to be identified as pathogenetically mutated in primary human medulloblastomas other than very rare familial case reports such as Nijmegen Breakage Syndrome and Fanconi Anemia Type D (Ruud and Wesenberg 2001; Bakhshi et€al. 2003; Distel et€al. 2003; Offit et€al. 2003; Hirsch et€al. 2004; Tischkowitz et€al. 2004). Several medulloblastoma cell lines have been instrumental in early stages of drug discovery and preclinical testing. However, it is clear from genomic comparisons of these cell lines to primary medulloblastomas that they have diverged over repeated passages making them less representative of their primary state (Cho and Pomeroy, unpublished data). The use of xenografts and neurosphere
282
Y.-J. Cho and S.L. Pomeroy
cultures derived from freshly resected primary tumor is increasingly utilized and may provide an important alternative for use in preclinical testing of new therapies. Studies in such systems will be greatly benefited by our ability to rapidly assign a molecular phenotype to a tumor soon after its resection. Third, incorporating new therapies into current treatment protocols let€ alone implementing targeted therapies as first line treatment will remain challenging. As conventional multimodal therapies achieve survival rates close to 85% in some series, introducing new targeted drugs as an alternative may be difficult, perhaps unethical. Conversely, the addition of a targeted therapy to conventional treatments may confound efficacy results (Crawford et€al. 2007). Thus, in order to realize the true potential of targeted treatments, carefully designed trials with the emphasis of eradicating unacceptable toxicities of current treatment regimens will need to be developed. Finally, the collaborative, multidisciplinary care of the patient diagnosed with medulloblastoma will be particularly tested as targeted therapies evolve. The care team will need to coordinate proper acquisition and transfer of the clinical sample from the operative suite to the laboratory, proper processing of the sample in the laboratory for DNA and RNA based studies, and proper interpretation of these results by the patient’s neurologist, oncologist, and pathologist. Definitive molecular phenotyping will need to be performed in a timely manner if this process is to dictate clinical management. Despite these challenges, development and implementation of targeted therapy for medulloblastoma shall continue with the hope of eradicating this lethal childhood disease.
References Adesina, A.M., Nalbantoglu, J., and Cavenee, W.K. 1994. p53 gene mutation and mdm2 gene amplification are uncommon in medulloblastoma. Cancer Res 54: 5649–5651. Alder, J., Cho, N.K., and Hatten, M.E. 1996. Embryonic precursor cells from the rhombic lip are specified to a cerebellar granule neuron identity. Neuron 17: 389–399. Aldosari, N., Wiltshire, R.N., Dutra, A., Schrock, E., McLendon, R.E., Friedman, H.S., Bigner, D.D., and Bigner, S.H. 2002. Comprehensive molecular cytogenetic investigation of chromosomal abnormalities in human medulloblastoma cell lines and xenograft. Neuro Oncol 4: 75–85. Androutsellis-Theotokis, A., Leker, R.R., Soldner, F., Hoeppner, D.J., Ravin, R., Poser, S.W., Rueger, M.A., Bae, S.K., Kittappa, R., and McKay, R.D. 2006. Notch signalling regulates stem cell numbers in€vitro and in€vivo. Nature 442: 823–826. Araki, K., Sangai, T., Miyamoto, S., Maeda, H., Zhang, S.C., Nakamura, M., Ishii, G., Hasebe, T., Kusaka, H., Akiyama, T., et€al. 2006. Inhibition of bone-derived insulin-like growth factors by a ligand-specific antibody suppresses the growth of human multiple myeloma in the human adult bone explanted in NOD/SCID mouse. Int J Cancer 118: 2602–2608. Badiali, M., Iolascon, A., Loda, M., Scheithauer, B.W., Basso, G., Trentini, G.P., and Giangaspero, F. 1993. p53 gene mutations in medulloblastoma. Immunohistochemistry, gel shift analysis, and sequencing. Diagn Mol Pathol 2: 23–28. Baeza, N., Masuoka, J., Kleihues, P., and Ohgaki, H. 2003. AXIN1 mutations but not deletions in cerebellar medulloblastomas. Oncogene 22: 632–636.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
283
Bakhshi, S., Cerosaletti, K.M., Concannon, P., Bawle, E.V., Fontanesi, J., Gatti, R.A., and Bhambhani, K. 2003. Medulloblastoma with adverse reaction to radiation therapy in nijmegen breakage syndrome. J Pediatr Hematol Oncol 25: 248–251. Barel, D., Avigad, S., Mor, C., Fogel, M., Cohen, I.J., and Zaizov, R. 1998. A novel germ-line mutation in the noncoding region of the p53 gene in a Li-Fraumeni family. Cancer Genet Cytogenet 103: 1–6. Barker, N., and Clevers, H. 2006. Mining the Wnt pathway for cancer therapeutics. Nat Rev Drug Discov 5: 997–1014. Baserga, R., Sell, C., Porcu, P., and Rubini, M. 1994. The role of the IGF-I receptor in the growth and transformation of mammalian cells. Cell Prolif 27: 63–71. Bayani, J., Zielenska, M., Marrano, P., Kwan Ng, Y., Taylor, M.D., Jay, V., Rutka, J.T., and Squire, J.A. 2000. Molecular cytogenetic analysis of medulloblastomas and supratentorial primitive neuroectodermal tumors by using conventional banding, comparative genomic hybridization, and spectral karyotyping. J Neurosurg 93: 437–448. Berman, D.M., Karhadkar, S.S., Hallahan, A.R., Pritchard, J.I., Eberhart, C.G., Watkins, D.N., Chen, J.K., Cooper, M.K., Taipale, J., Olson, J.M., et€al. 2002. Medulloblastoma growth inhibition by hedgehog pathway blockade. Science 297: 1559–1561. Beuvink, I., Boulay, A., Fumagalli, S., Zilbermann, F., Ruetz, S., O’Reilly, T., Natt, F., Hall, J., Lane, H.A., and Thomas, G. 2005. The mTOR inhibitor RAD001 sensitizes tumor cells to DNA-damaged induced apoptosis through inhibition of p21 translation. Cell 120: 747–759. Bjornsti, M.A., and Houghton, P.J. 2004. The TOR pathway: a target for cancer therapy. Nat Rev Cancer 4: 335–348. Bruggeman, S.W., Valk-Lingbeek, M.E., van der Stoop, P.P., Jacobs, J.J., Kieboom, K., Tanger, E., Hulsman, D., Leung, C., Arsenijevic, Y., Marino, S., et€al. 2005. Ink4a and Arf differentially affect cell proliferation and neural stem cell self-renewal in Bmi1-deficient mice. Genes Dev 19: 1438–1443. CBTRUS. 2005. Central Brain Tumor Registry of the United States, 2005. Chang, S.M., Wen, P., Cloughesy, T., Greenberg, H., Schiff, D., Conrad, C., Fink, K., Robins, H.I., De Angelis, L., Raizer, J., et€al. 2005. Phase II study of CCI-779 in patients with recurrent glioblastoma multiforme. Invest New Drugs 23: 357–361. Chin, L.S., Yung, W.K., and Raffel, C. 1996. Two primitive neuroectodermal tumor cell lines require an activated insulin-like growth factor I receptor for growth in€vitro. Neurosurgery 39: 1183–1190. Ciardiello, F. 2005. Epidermal growth factor receptor inhibitors in cancer treatment. Future Oncol 1: 221–234. Clifford, S.C., Lusher, M.E., Lindsey, J.C., Langdon, J.A., Gilbertson, R.J., Straughton, D., and Ellison, D.W. 2006. Wnt/Wingless pathway activation and chromosome 6 loss characterize a distinct molecular sub-group of medulloblastomas associated with a favorable prognosis. Cell Cycle 5: 2666–2670. Cooper, M.K., Porter, J.A., Young, K.E., and Beachy, P.A. 1998. Teratogen-mediated inhibition of target tissue response to Shh signaling. Science 280: 1603–1607. Crawford, J.R., MacDonald, T.J., and Packer, R.J. 2007. Medulloblastoma in childhood: new biological advances. Lancet Neurol 6: 1073–1085. Dahmen, R.P., Koch, A., Denkhaus, D., Tonn, J.C., Sorensen, N., Berthold, F., Behrens, J., Birchmeier, W., Wiestler, O.D., and Pietsch, T. 2001. Deletions of AXIN1, a component of the WNT/wingless pathway, in sporadic medulloblastomas. Cancer Res 61: 7039–7043. Dakubo, G.D., Mazerolle, C.J., and Wallace, V.A. 2006. Expression of Notch and Wnt pathway components and activation of Notch signaling in medulloblastomas from heterozygous patched mice. J Neurooncol 79: 221–227. Dancey, J. 2004. Epidermal growth factor receptor inhibitors in clinical development. Int J Radiat Oncol Biol Phys 58: 1003–1007. De Bortoli, M., Castellino, R.C., Lu, X.Y., Deyo, J., Sturla, L.M., Adesina, A.M., Perlaky, L., Pomeroy, S.L., Lau, C.C., Man, T.K., et€ al. 2006. Medulloblastoma outcome is adversely associated with overexpression of EEF1D, RPL30, and RPS20 on the long arm of chromosome 8. BMC Cancer 6: 223.
284
Y.-J. Cho and S.L. Pomeroy
de Chadarevian, J.P., Vekemans, M., and Bernstein, M. 1985. Fanconi’s anemia, medulloblastoma, Wilms’ tumor, horseshoe kidney, and gonadal dysgenesis. Arch Pathol Lab Med 109: 367–369. Del Valle, L., Enam, S., Lassak, A., Wang, J.Y., Croul, S., Khalili, K., and Reiss, K. 2002. Insulinlike growth factor I receptor activity in human medulloblastomas. Clin Cancer Res 8: 1822–1830. Distel, L., Neubauer, S., Varon, R., Holter, W., and Grabenbauer, G. 2003. Fatal toxicity following radio- and chemotherapy of medulloblastoma in a child with unrecognized Nijmegen breakage syndrome. Med Pediatr Oncol 41: 44–48. Dovey, H.F., John, V., Anderson, J.P., Chen, L.Z., de Saint Andrieu, P., Fang, L.Y., Freedman, S.B., Folmer, B., Goldbach, E., Holsztynska, E.J., et€ al. 2001. Functional gamma-secretase inhibitors reduce beta-amyloid peptide levels in brain. J Neurochem 76: 173–181. Dreesen, O., and Brivanlou, A.H. 2007. Signaling pathways in cancer and embryonic stem cells. Stem Cell Rev 3: 7–17. Dunn, S.E., Ehrlich, M., Sharp, N.J., Reiss, K., Solomon, G., Hawkins, R., Baserga, R., and Barrett, J.C. 1998. A dominant negative mutant of the insulin-like growth factor-I receptor inhibits the adhesion, invasion, and metastasis of breast cancer. Cancer Res 58: 3353–3361. Eberhart, C.G., Kratz, J., Wang, Y., Summers, K., Stearns, D., Cohen, K., Dang, C.V., and Burger, P.C. 2004. Histopathological and molecular prognostic markers in medulloblastoma: c-myc, N-myc, TrkC, and anaplasia. J Neuropathol Exp Neurol 63: 441–449. Eberhart, C.G., Tihan, T., and Burger, P.C. 2000. Nuclear localization and mutation of beta-catenin in medulloblastomas. J Neuropathol Exp Neurol 59: 333–337. Ebinger, M., Senf, L., and Scheurlen, W. 2006. Risk stratification in medulloblastoma: screening for molecular markers. Klin Padiatr 218: 139–142. Ellison, D.W., Onilude, O.E., Lindsey, J.C., Lusher, M.E., Weston, C.L., Taylor, R.E., Pearson, A.D., and Clifford, S.C. 2005. beta-Catenin status predicts a favorable outcome in childhood medulloblastoma: the United Kingdom Children’s Cancer Study Group Brain Tumour Committee. J Clin Oncol 23: 7951–7957. Erez, A., Ilan, T., Amariglio, N., Muler, I., Brok-Simoni, F., Rechavi, G., and Izraeli, S. 2002. GLI3 is not mutated commonly in sporadic medulloblastomas. Cancer 95: 28–31. Evans, D.G., Birch, J.M., and Orton, C.I. 1991a. Brain tumours and the occurrence of severe invasive basal cell carcinoma in first degree relatives with Gorlin syndrome. Br J Neurosurg 5: 643–646. Evans, D.G., Farndon, P.A., Burnell, L.D., Gattamaneni, H.R., and Birch, J.M. 1991b. The incidence of Gorlin syndrome in 173 consecutive cases of medulloblastoma. Br J Cancer 64: 959–961. Frappart, P.O., Lee, Y., Lamont, J., and McKinnon, P.J. 2007. BRCA2 is required for neurogenesis and suppression of medulloblastoma. EMBO J 26: 2732–2742. Gajjar, A., Chintagumpala, M., Ashley, D., Kellie, S., Kun, L.E., Merchant, T.E., Woo, S., Wheeler, G., Ahern, V., Krasin, M.J., et€al. 2006. Risk-adapted craniospinal radiotherapy followed by high-dose chemotherapy and stem-cell rescue in children with newly diagnosed medulloblastoma (St Jude Medulloblastoma-96): long-term results from a prospective, multicentre trial. Lancet Oncol 7: 813–820. Galanis, E., Buckner, J.C., Maurer, M.J., Kreisberg, J.I., Ballman, K., Boni, J., Peralba, J.M., Jenkins, R.B., Dakhil, S.R., Morton, R.F., et€al. 2005. Phase II trial of temsirolimus (CCI-779) in recurrent glioblastoma multiforme: a North Central Cancer Treatment Group Study. J Clin Oncol 23: 5294–5304. Garcia-Echeverria, C., Pearson, M.A., Marti, A., Meyer, T., Mestan, J., Zimmermann, J., Gao, J., Brueggen, J., Capraro, H.G., Cozens, R., et€ al. 2004. In vivo antitumor activity of NVPAEW541-A novel, potent, and selective inhibitor of the IGF-IR kinase. Cancer Cell 5: 231–239. Geoerger, B., Kerr, K., Tang, C.B., Fung, K.M., Powell, B., Sutton, L.N., Phillips, P.C., and Janss, A.J. 2001. Antitumor activity of the rapamycin analog CCI-779 in human primitive neuroectodermal tumor/medulloblastoma models as single agent and in combination chemotherapy. Cancer Res 61: 1527–1532.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
285
Gilbertson, R.J., and Clifford, S.C. 2003. PDGFRB is overexpressed in metastatic medulloblastoma. Nat Genet 35: 197–198. Gilbertson, R.J., Jaros, E.B., Perry, R.H., and Pearson, A.D. 1992. Prognostic factors in medulloblastoma. Lancet 340: 480. Gilbertson, R.J., Pearson, A.D., Perry, R.H., Jaros, E., and Kelly, P.J. 1995. Prognostic significance of the c-erbB-2 oncogene product in childhood medulloblastoma. Br J Cancer 71: 473–477. Gilbertson, R.J., Perry, R.H., Kelly, P.J., Pearson, A.D., and Lunec, J. 1997. Prognostic significance of HER2 and HER4 coexpression in childhood medulloblastoma. Cancer Res 57: 3272–3280. Glick, R.P., Unterman, T.G., and Lacson, R. 1993. Identification of insulin-like growth factor (IGF) and glucose transporter-1 and -3 mRNA in CNS tumors. Regul Pept 48: 251–256. Goldowitz, D., and Hamre, K. 1998. The cells and molecules that make a cerebellum. Trends Neurosci 21: 375–382. Goodrich, L.V., Milenkovic, L., Higgins, K.M., and Scott, M.P. 1997. Altered neural cell fates and medulloblastoma in mouse patched mutants. Science 277: 1109–1113. Gordon, M.D., and Nusse, R. 2006. Wnt signaling: multiple pathways, multiple receptors, and multiple transcription factors. J Biol Chem 281: 22429–22433. Guertin, D.A., and Sabatini, D.M. 2007. Defining the role of mTOR in cancer. Cancer Cell 12: 9–22. Guran, S., Tunca, Y., and Imirzalioglu, N. 1999. Hereditary TP53 codon 292 and somatic P16INK4A codon 94 mutations in a Li-Fraumeni syndrome family. Cancer Genet Cytogenet 113: 145–151. Hahn, H., Christiansen, J., Wicking, C., Zaphiropoulos, P.G., Chidambaram, A., Gerrard, B., Vorechovsky, I., Bale, A.E., Toftgard, R., Dean, M., et€ al. 1996. A mammalian patched homolog is expressed in target tissues of sonic hedgehog and maps to a region associated with developmental abnormalities. J Biol Chem 271: 12125–12128. Hahn, H., Wojnowski, L., Specht, K., Kappler, R., Calzada-Wack, J., Potter, D., Zimmer, A., Muller, U., Samson, E., Quintanilla-Martinez, L., et€al. 2000. Patched target Igf2 is indispensable for the formation of medulloblastoma and rhabdomyosarcoma. J Biol Chem 275: 28341–28344. Hallahan, A.R., Pritchard, J.I., Hansen, S., Benson, M., Stoeck, J., Hatton, B.A., Russell, T.L., Ellenbogen, R.G., Bernstein, I.D., Beachy, P.A., et€al. 2004. The SmoA1 mouse model reveals that notch signaling is critical for the growth and survival of sonic hedgehog-induced medulloblastomas. Cancer Res 64: 7794–7800. Hamilton, S.R., Liu, B., Parsons, R.E., Papadopoulos, N., Jen, J., Powell, S.M., Krush, A.J., Berk, T., Cohen, Z., Tetu, B., et€al. 1995. The molecular basis of Turcot’s syndrome. N Engl J Med 332: 839–847. Hatten, M.E., Alder, J., Zimmerman, K., and Heintz, N. 1997. Genes involved in cerebellar cell specification and differentiation. Curr Opin Neurobiol 7: 40–47. Hirsch, B., Shimamura, A., Moreau, L., Baldinger, S., Hag-alshiekh, M., Bostrom, B., Sencer, S., and D’Andrea, A.D. 2004. Association of biallelic BRCA2/FANCD1 mutations with spontaneous chromosomal instability and solid tumors of childhood. Blood 103: 2554–2559. Holcomb, V.B., Vogel, H., Marple, T., Kornegay, R.W., and Hasty, P. 2006. Ku80 and p53 suppress medulloblastoma that arise independent of Rag-1-induced DSBs. Oncogene 25: 7159–7165. Huang, H., Mahler-Araujo, B.M., Sankila, A., Chimelli, L., Yonekawa, Y., Kleihues, P., and Ohgaki, H. 2000. APC mutations in sporadic medulloblastomas. Am J Pathol 156: 433–437. Johnson, R.L., Rothman, A.L., Xie, J., Goodrich, L.V., Bare, J.W., Bonifas, J.M., Quinn, A.G., Myers, R.M., Cox, D.R., Epstein, E.H., Jr., et€al. 1996. Human homolog of patched, a candidate gene for the basal cell nevus syndrome. Science 272: 1668–1671. Katsetos, C.D., Krishna, L., Frankfurter, A., Karkavelas, G., Wolfe, D.E., Valsamis, M.P., Schiffer, D., Vlachos, I.N., and Urich, H. 1995. A cytomorphological scheme of differentiating neuronal phenotypes in cerebellar medulloblastomas based on immunolocalization of class III betatubulin isotype (beta III) and proliferating cell nuclear antigen (PCNA)/cyclin. Clin Neuropathol 14: 72–81.
286
Y.-J. Cho and S.L. Pomeroy
Keeler, R.F. 1970. Teratogenic compounds of Veratrum californicum (Durand) X. Cyclopia in rabbits produced by cyclopamine. Teratology 3: 175–180. Keeler, R.F. 1978. Cyclopamine and related steroidal alkaloid teratogens: their occurrence, structural relationship, and biologic effects. Lipids 13: 708–715. Kenney, A.M., Cole, M.D., and Rowitch, D.H. 2003. Nmyc upregulation by sonic hedgehog signaling promotes proliferation in developing cerebellar granule neuron precursors. Development 130: 15–28. Kim, J.Y., Sutton, M.E., Lu, D.J., Cho, T.A., Goumnerova, L.C., Goritchenko, L., Kaufman, J.R., Lam, K.K., Billet, A.L., Tarbell, N.J., et€al. 1999. Activation of neurotrophin-3 receptor TrkC induces apoptosis in medulloblastomas. Cancer Res 59: 711–719. Kimura, H., Ng, J.M., and Curran, T. 2008. Transient inhibition of the Hedgehog pathway in young mice causes permanent defects in bone structure. Cancer Cell 13: 249–260. Koch, A., Hrychyk, A., Hartmann, W., Waha, A., Mikeska, T., Waha, A., Schuller, U., Sorensen, N., Berthold, F., Goodyer, C.G., et€al. 2007. Mutations of the Wnt antagonist AXIN2 (Conductin) result in TCF-dependent transcription in medulloblastomas. Int J Cancer 121: 284–291. Koch, A., Waha, A., Hartmann, W., Milde, U., Goodyer, C.G., Sorensen, N., Berthold, F., DigonSontgerath, B., Kratzschmar, J., Wiestler, O.D., et€ al. 2004. No evidence for mutations or altered expression of the Suppressor of Fused gene (SUFU) in primitive neuroectodermal tumours. Neuropathol Appl Neurobiol 30: 532–539. Koch, A., Waha, A., Tonn, J.C., Sorensen, N., Berthold, F., Wolter, M., Reifenberger, J., Hartmann, W., Friedl, W., Reifenberger, G., et€ al. 2001. Somatic mutations of WNT/wingless signaling pathway components in primitive neuroectodermal tumors. Int J Cancer 93: 445–449. Lam, C.W., Xie, J., To, K.F., Ng, H.K., Lee, K.C., Yuen, N.W., Lim, P.L., Chan, L.Y., Tong, S.F., and McCormick, F. 1999. A frequent activated smoothened mutation in sporadic basal cell carcinomas. Oncogene 18: 833–836. Lauth, M., Bergstrom, A., Shimokawa, T., and Toftgard, R. 2007. Inhibition of GLI-mediated transcription and tumor cell growth by small-molecule antagonists. Proc Natl Acad Sci USA 104: 8455–8460. Lee, W.H., Wang, G.M., Lo, T., Triarhou, L.C., and Ghetti, B. 1995. Altered IGFBP5 gene expression in the cerebellar external germinal layer of weaver mutant mice. Brain Res Mol Brain Res 30: 259–268. Lee, Y., Kawagoe, R., Sasai, K., Li, Y., Russell, H.R., Curran, T., and McKinnon, P.J. 2007. Loss of suppressor-of-fused function promotes tumorigenesis. Oncogene 26: 6442–6447. Lee, Y., and McKinnon, P.J. 2002. DNA ligase IV suppresses medulloblastoma formation. Cancer Res 62: 6395–6399. LeRoith, D., Werner, H., Faria, T.N., Kato, H., Adamo, M., and Roberts, C.T., Jr. 1993. Insulinlike growth factor receptors. Implications for nervous system function. Ann N Y Acad Sci 692: 22–32. Leung, C., Lingbeek, M., Shakhova, O., Liu, J., Tanger, E., Saremaslani, P., Van Lohuizen, M., and Marino, S. 2004. Bmi1 is essential for cerebellar development and is overexpressed in human medulloblastomas. Nature 428: 337–341. Li, X.N., Shu, Q., Su, J.M., Perlaky, L., Blaney, S.M., and Lau, C.C. 2005. Valproic acid induces growth arrest, apoptosis, and senescence in medulloblastomas by increasing histone hyperacetylation and regulating expression of p21Cip1, CDK4, and CMYC. Mol Cancer Ther 4: 1912–1922. Liao, H., and Wang, J.H. 2005. Biomembrane-permeable and Ribonuclease-resistant siRNA with enhanced activity. Oligonucleotides 15: 196–205. Lin, J.C., and Cepko, C.L. 1999. Biphasic dispersion of clones containing Purkinje cells and glia in the developing chick cerebellum. Dev Biol 211: 177–197. Louis, D.N., Ohgaki, H., Wiestler, O.D., Cavenee, W.K., Burger, P.C., Jouvet, A., Scheithauer, B.W., and Kleihues, P. 2007. The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol 114: 97–109. Lun, X.Q., Zhou, H., Alain, T., Sun, B., Wang, L., Barrett, J.W., Stanford, M.M., McFadden, G., Bell, J., Senger, D.L., et€al. 2007. Targeting human medulloblastoma: oncolytic virotherapy with myxoma virus is enhanced by rapamycin. Cancer Res 67: 8818–8827.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
287
Lutolf, S., Radtke, F., Aguet, M., Suter, U., and Taylor, V. 2002. Notch1 is required for neuronal and glial differentiation in the cerebellum. Development 129: 373–385. MacDonald, T.J., Brown, K.M., LaFleur, B., Peterson, K., Lawlor, C., Chen, Y., Packer, R.J., Cogen, P., and Stephan, D.A. 2001. Expression profiling of medulloblastoma: PDGFRA and the RAS/MAPK pathway as therapeutic targets for metastatic disease. Nat Genet 29: 143–152. Machold, R., and Fishell, G. 2005. Math1 is expressed in temporally discrete pools of cerebellar rhombic-lip neural progenitors. Neuron 48: 17–24. Marino, S., Vooijs, M., van Der Gulden, H., Jonkers, J., and Berns, A. 2000. Induction of Â�medulloblastomas in p53-null mutant mice by somatic inactivation of Rb in the external granular layer cells of the cerebellum. Genes Dev 14: 994–1004. McManamy, C.S., Pears, J., Weston, C.L., Hanzely, Z., Ironside, J.W., Taylor, R.E., Grundy, R.G., Clifford, S.C., and Ellison, D.W. 2007. Nodule formation and desmoplasia in medulloblastomasdefining the nodular/desmoplastic variant and its biological behavior. Brain Pathol 17: 151–164. Michiels, E.M., Weiss, M.M., Hoovers, J.M., Baak, J.P., Voute, P.A., Baas, F., and Hermsen, M.A. 2002. Genetic alterations in childhood medulloblastoma analyzed by comparative genomic hybridization. J Pediatr Hematol Oncol 24: 205–210. O’Reilly, K.E., Rojo, F., She, Q.B., Solit, D., Mills, G.B., Smith, D., Lane, H., Hofmann, F., Hicklin, D.J., Ludwig, D.L., et€al. 2006. mTOR inhibition induces upstream receptor tyrosine kinase signaling and activates Akt. Cancer Res 66: 1500–1508. Offit, K., Levran, O., Mullaney, B., Mah, K., Nafa, K., Batish, S.D., Diotti, R., Schneider, H., Deffenbaugh, A., Scholl, T., et€al. 2003. Shared genetic susceptibility to breast cancer, brain tumors, and Fanconi anemia. J Natl Cancer Inst 95: 1548–1551. Ohgaki, H., Eibl, R.H., Wiestler, O.D., Yasargil, M.G., Newcomb, E.W., and Kleihues, P. 1991. p53 mutations in nonastrocytic human brain tumors. Cancer Res 51: 6202–6205. Oliver, T.G., Grasfeder, L.L., Carroll, A.L., Kaiser, C., Gillingham, C.L., Lin, S.M., Wickramasinghe, R., Scott, M.P., and Wechsler-Reya, R.J. 2003. Transcriptional profiling of the Sonic hedgehog response: a critical role for N-myc in proliferation of neuronal precursors. Proc Natl Acad Sci USA 100: 7331–7336. Pearson, A.D., Craft, A.W., Ratcliffe, J.M., Birch, J.M., Morris-Jones, P., and Roberts, D.F. 1982. Two families with the Li-Fraumeni cancer family syndrome. J Med Genet 19: 362–365. Peterson, J.K., Tucker, C., Favours, E., Cheshire, P.J., Creech, J., Billups, C.A., Smykla, R., Lee, F.Y., and Houghton, P.J. 2005. In vivo evaluation of ixabepilone (BMS247550), a novel epothilone B derivative, against pediatric cancer models. Clin Cancer Res 11: 6950–6958. Pietsch, T., Waha, A., Koch, A., Kraus, J., Albrecht, S., Tonn, J., Sorensen, N., Berthold, F., Henk, B., Schmandt, N., et€al. 1997. Medulloblastomas of the desmoplastic variant carry mutations of the human homologue of Drosophila patched. Cancer Res 57: 2085–2088. Pomeroy, S.L., Tamayo, P., Gaasenbeek, M., Sturla, L.M., Angelo, M., McLaughlin, M.E., Kim, J.Y., Goumnerova, L.C., Black, P.M., Lau, C., et€al. 2002. Prediction of central nervous system embryonal tumour outcome based on gene expression. Nature 415: 436–442. Radtke, F., and Clevers, H. 2005. Self-renewal and cancer of the gut: two sides of a coin. Science 307: 1904–1909. Raffel, C., Jenkins, R.B., Frederick, L., Hebrink, D., Alderete, B., Fults, D.W., and James, C.D. 1997. Sporadic medulloblastomas contain PTCH mutations. Cancer Res 57: 842–845. Rao, G., Pedone, C.A., Valle, L.D., Reiss, K., Holland, E.C., and Fults, D.W. 2004. Sonic hedgehog and insulin-like growth factor signaling synergize to induce medulloblastoma formation from nestin-expressing neural progenitors in mice. Oncogene 23: 6156–6162. Reifenberger, J., Wolter, M., Weber, R.G., Megahed, M., Ruzicka, T., Lichter, P., and Reifenberger, G. 1998. Missense mutations in SMOH in sporadic basal cell carcinomas of the skin and primitive neuroectodermal tumors of the central nervous system. Cancer Res 58: 1798–1803. Reiss, K., D’Ambrosio, C., Tu, X., Tu, C., and Baserga, R. 1998. Inhibition of tumor growth by a dominant negative mutant of the insulin-like growth factor I receptor with a bystander effect. Clin Cancer Res 4: 2647–2655. Reya, T., and Clevers, H. 2005. Wnt signalling in stem cells and cancer. Nature 434: 843–850. Romer, J.T., and Curran, T. 2004. Medulloblastoma and retinoblastoma: oncology recapitulates ontogeny. Cell Cycle 3: 917–919.
288
Y.-J. Cho and S.L. Pomeroy
Romer, J.T., Kimura, H., Magdaleno, S., Sasai, K., Fuller, C., Baines, H., Connelly, M., Stewart, C.F., Gould, S., Rubin, L.L., et€al. 2004. Suppression of the Shh pathway using a small molecule inhibitor eliminates medulloblastoma in Ptc1(+/−)p53(−/−) mice. Cancer Cell 6: 229–240. Rossi, M.R., Conroy, J., McQuaid, D., Nowak, N.J., Rutka, J.T., and Cowell, J.K. 2006. Array CGH analysis of pediatric medulloblastomas. Genes Chromosomes Cancer 45: 290–303. Rotwein, P., Burgess, S.K., Milbrandt, J.D., and Krause, J.E. 1988. Differential expression of insulin-like growth factor genes in rat central nervous system. Proc Natl Acad Sci USA 85: 265–269. Rubin, J.B., Kung, A.L., Klein, R.S., Chan, J.A., Sun, Y., Schmidt, K., Kieran, M.W., Luster, A.D., and Segal, R.A. 2003. A small-molecule antagonist of CXCR4 inhibits intracranial growth of primary brain tumors. Proc Natl Acad Sci USA 100: 13513–13518. Ruiz i Altaba, A., Sanchez, P., and Dahmane, N. 2002. Gli and hedgehog in cancer: tumours, embryos and stem cells. Nat Rev Cancer 2: 361–372. Ruud, E., and Wesenberg, F. 2001. Microcephalus, medulloblastoma and excessive toxicity from chemotherapy: an unusual presentation of Fanconi anaemia. Acta Paediatr 90: 580–583. Sabatini, D.M. 2006. mTOR and cancer: insights into a complex relationship. Nat Rev Cancer 6: 729–734. Salaroli, R., Russo, A., Ceccarelli, C., Mina, G.D., Arcella, A., Martinelli, G.N., Giangaspero, F., Capranico, G., and Cenacchi, G. 2007. Intracellular distribution of beta-catenin in human medulloblastoma cell lines with different degree of neuronal differentiation. Ultrastruct Pathol 31: 33–44. Salsano, E., Croci, L., Maderna, E., Lupo, L., Pollo, B., Giordana, M.T., Consalez, G.G., and Finocchiaro, G. 2007. Expression of the neurogenic basic helix-loop-helix transcription factor NEUROG1 identifies a subgroup of medulloblastomas not expressing ATOH1. Neuro Oncol 9: 298–307. Sastre, M., Steiner, H., Fuchs, K., Capell, A., Multhaup, G., Condron, M.M., Teplow, D.B., and Haass, C. 2001. Presenilin-dependent gamma-secretase processing of beta-amyloid precursor protein at a site corresponding to the S3 cleavage of Notch. EMBO Rep 2: 835–841. Saylors, R.L., 3rd, Sidransky, D., Friedman, H.S., Bigner, S.H., Bigner, D.D., Vogelstein, B., and Brodeur, G.M. 1991. Infrequent p53 gene mutations in medulloblastomas. Cancer Res 51: 4721–4723. Schuller, U., Koch, A., Hartmann, W., Garre, M.L., Goodyer, C.G., Cama, A., Sorensen, N., Wiestler, O.D., and Pietsch, T. 2005. Subtype-specific expression and genetic alterations of the chemokinereceptor gene CXCR4 in medulloblastomas. Int J Cancer 117: 82–89. Scotlandi, K., Manara, M.C., Nicoletti, G., Lollini, P.L., Lukas, S., Benini, S., Croci, S., Perdichizzi, S., Zambelli, D., Serra, M., et€ al. 2005. Antitumor activity of the insulin-like growth factor-I receptor kinase inhibitor NVP-AEW541 in musculoskeletal tumors. Cancer Res 65: 3868–3876. Segal, R.A., Goumnerova, L.C., Kwon, Y.K., Stiles, C.D., and Pomeroy, S.L. 1994. Expression of the neurotrophin receptor TrkC is linked to a favorable outcome in medulloblastoma. Proc Natl Acad Sci USA 91: 12867–12871. Shih Ie, M., and Wang, T.L. 2007. Notch signaling, gamma-secretase inhibitors, and cancer therapy. Cancer Res 67: 1879–1882. Shu, Q., Antalffy, B., Su, J.M., Adesina, A., Ou, C.N., Pietsch, T., Blaney, S.M., Lau, C.C., and Li, X.N. 2006. Valproic Acid prolongs survival time of severe combined immunodeficient mice bearing intracerebellar orthotopic medulloblastoma xenografts. Clin Cancer Res 12: 4687–4694. Singh, S.K., Clarke, I.D., Terasaki, M., Bonn, V.E., Hawkins, C., Squire, J., and Dirks, P.B. 2003. Identification of a cancer stem cell in human brain tumors. Cancer Res 63: 5821–5828. Singh, S.K., Hawkins, C., Clarke, I.D., Squire, J.A., Bayani, J., Hide, T., Henkelman, R.M., Cusimano, M.D., and Dirks, P.B. 2004. Identification of human brain tumour initiating cells. Nature 432: 396–401. Sottile, V., Li, M., and Scotting, P.J. 2006. Stem cell marker expression in the Bergmann glia population of the adult mouse brain. Brain Res 1099: 8–17.
Targeted Therapy in Medulloblastoma in Molecularly Targeted Therapy
289
Spiller, S.E., Ditzler, S.H., Pullar, B.J., and Olson, J.M. 2008. Response of preclinical medulloblastoma models to combination therapy with 13-cis retinoic acid and suberoylanilide hydroxamic acid (SAHA). J Neurooncol 87:133–141. Spiller, S.E., Ravanpay, A.C., Hahn, A.W., and Olson, J.M. 2006. Suberoylanilide hydroxamic acid is effective in preclinical studies of medulloblastoma. J Neurooncol 79: 259–270. Stumm, R.K., Zhou, C., Ara, T., Lazarini, F., Dubois-Dalcq, M., Nagasawa, T., Hollt, V., and Schulz, S. 2003. CXCR4 regulates interneuron migration in the developing neocortex. J Neurosci 23: 5123–5130. Taylor, M.D., Liu, L., Raffel, C., Hui, C.C., Mainprize, T.G., Zhang, X., Agatep, R., Chiappa, S., Gao, L., Lowrance, A., et€al. 2002. Mutations in SUFU predispose to medulloblastoma. Nat Genet 31: 306–310. Thompson, M.C., Fuller, C., Hogg, T.L., Dalton, J., Finkelstein, D., Lau, C.C., Chintagumpala, M., Adesina, A., Ashley, D.M., Kellie, S.J., et€al. 2006. Genomics identifies medulloblastoma subgroups that are enriched for specific genetic alterations. J Clin Oncol 24: 1924–1931. Tischkowitz, M.D., Chisholm, J., Gaze, M., Michalski, A., and Rosser, E.M. 2004. Medulloblastoma as a first presentation of fanconi anemia. J Pediatr Hematol Oncol 26: 52–55. Tong, W.M., Ohgaki, H., Huang, H., Granier, C., Kleihues, P., and Wang, Z.Q. 2003. Null mutation of DNA strand break-binding molecule poly(ADP-ribose) polymerase causes medulloblastomas in p53(−/−) mice. Am J Pathol 162: 343–352. Urbanska, K., Trojanek, J., Del Valle, L., Eldeen, M.B., Hofmann, F., Garcia-Echeverria, C., Khalili, K., and Reiss, K. 2007. Inhibition of IGF-I receptor in anchorage-independence attenuates GSK-3beta constitutive phosphorylation and compromises growth and survival of medulloblastoma cell lines. Oncogene 26: 2308–2317. Uziel, T., Zindy, F., Sherr, C.J., and Roussel, M.F. 2006. The CDK inhibitor p18Ink4c is a tumor suppressor in medulloblastoma. Cell Cycle 5: 363–365. Vilz, T.O., Moepps, B., Engele, J., Molly, S., Littman, D.R., and Schilling, K. 2005. The SDF-1/ CXCR4 pathway and the development of the cerebellar system. Eur J Neurosci 22: 1831–1839. Wallace, V.A. 1999. Purkinje-cell-derived Sonic hedgehog regulates granule neuron precursor cell proliferation in the developing mouse cerebellum. Curr Biol 9: 445–448. Wang, V.Y., Rose, M.F., and Zoghbi, H.Y. 2005a. Math1 expression redefines the rhombic lip derivatives and reveals novel lineages within the brainstem and cerebellum. Neuron 48: 31–43. Wang, Y., Hailey, J., Williams, D., Wang, Y., Lipari, P., Malkowski, M., Wang, X., Xie, L., Li, G., Saha, D., et€al. 2005b. Inhibition of insulin-like growth factor-I receptor (IGF-IR) signaling and tumor cell growth by a fully human neutralizing anti-IGF-IR antibody. Mol Cancer Ther 4: 1214–1221. Ward, E.J., Shcherbata, H.R., Reynolds, S.H., Fischer, K.A., Hatfield, S.D., and Ruohola-Baker, H. 2006. Stem cells signal to the niche through the Notch pathway in the Drosophila ovary. Curr Biol 16: 2352–2358. Wechsler-Reya, R.J., and Scott, M.P. 1999. Control of neuronal precursor proliferation in the cerebellum by Sonic Hedgehog. Neuron 22: 103–114. Weiner, H.L., Bakst, R., Hurlbert, M.S., Ruggiero, J., Ahn, E., Lee, W.S., Stephen, D., Zagzag, D., Joyner, A.L., and Turnbull, D.H. 2002. Induction of medulloblastomas in mice by sonic hedgehog, independent of Gli1. Cancer Res 62: 6385–6389. Wetmore, C., Eberhart, D.E., and Curran, T. 2001. Loss of p53 but not ARF accelerates medulloblastoma in mice heterozygous for patched. Cancer Res 61: 513–516. Wiederschain, D., Chen, L., Johnson, B., Bettano, K., Jackson, D., Taraszka, J., Wang, Y.K., Jones, M.D., Morrissey, M., Deeds, J., et€ al. 2007. Contribution of polycomb homologues Bmi-1 and Mel-18 to medulloblastoma pathogenesis. Mol Cell Biol 27: 4968–4979. Williams, J.A., Guicherit, O.M., Zaharian, B.I., Xu, Y., Chai, L., Wichterle, H., Kon, C., Gatchalian, C., Porter, J.A., Rubin, L.L., et€al. 2003. Identification of a small molecule inhibitor of the hedgehog signaling pathway: effects on basal cell carcinoma-like lesions. Proc Natl Acad Sci USA 100: 4616–4621.
290
Y.-J. Cho and S.L. Pomeroy
Wlodarski, P., Grajkowska, W., Lojek, M., Rainko, K., and Jozwiak, J. 2006. Activation of Akt and Erk pathways in medulloblastoma. Folia Neuropathol 44: 214–220. Yan, C.T., Kaushal, D., Murphy, M., Zhang, Y., Datta, A., Chen, C., Monroe, B., Mostoslavsky, G., Coakley, K., Gao, Y., et€ al. 2006. XRCC4 suppresses medulloblastomas with recurrent translocations in p53-deficient mice. Proc Natl Acad Sci USA 103: 7378–7383. Yang, L., Jackson, E., Woerner, B.M., Perry, A., Piwnica-Worms, D., and Rubin, J.B. 2007. Blocking CXCR4-mediated cyclic AMP suppression inhibits brain tumor growth in€ vivo. Cancer Res 67: 651–658. Yee, K.W., Zeng, Z., Konopleva, M., Verstovsek, S., Ravandi, F., Ferrajoli, A., Thomas, D., Wierda, W., Apostolidou, E., Albitar, M., et€al. 2006. Phase I/II study of the mammalian target of rapamycin inhibitor everolimus (RAD001) in patients with relapsed or refractory hematologic malignancies. Clin Cancer Res 12: 5165–5173. Yokota, N., Nishizawa, S., Ohta, S., Date, H., Sugimura, H., Namba, H., and Maekawa, M. 2002. Role of Wnt pathway in medulloblastoma oncogenesis. Int J Cancer 101: 198–201. Zurawel, R.H., Chiappa, S.A., Allen, C., and Raffel, C. 1998. Sporadic medulloblastomas contain oncogenic beta-catenin mutations. Cancer Res 58: 896–899.
Future Treatments of Ependymoma Richard J. Gilbertson
Introduction Although ependymoma is the third most common form of central nervous system (CNS) tumor, it is a poorly understood disease. The variable patterns of histology and clinical presentation, lack of consensus regarding appropriate means to stage and grade the disease, and the separation of pediatric and adult oncology services, have each hindered efforts to advance understanding of ependymoma biology and treatment. As a result, no new therapeutic approaches have been identified to treat ependymoma during the last 20€years and up to 40% of patients remain incurable (Brandes et€al. 2005; Merchant and Fouladi 2005). The remarkable insensitivity of ependymoma to most nonsurgical conventional therapies has contributed further to this impasse in clinical management. While ependymoma remains a challenging disease to treat, recent laboratory and clinical research has resulted in important discoveries that are likely to provide new direction for patient management. The efficient integration of these discoveries into treatment protocols will require an understanding of contemporary therapies as well as novel mechanisms to target molecular alterations. This chapter will review current and emerging understanding of ependymoma biology and treatment as well strategies to integrate this knowledge within clinical management protocols.
R.J. Gilbertsonâ•›(*) Neurobiology and Brain Tumor Program, St. Jude Children’s Research Hospital, Memphis, TN, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_14, © Springer Science+Business Media, LLC 2010
291
292
R.J. Gilbertson
What Do We Know About Ependymoma? Clinical Issues Epidemiology and Histopathology Approximately 900 new cases of ependymoma are diagnosed each year in the US and Europe; most often in children aged less than 4€years (CBTRUS 2006). These CNS ependymal tumors originate from the wall of the ventricular system and the spinal canal, and display moderate cellularity, ultrastructural properties of ependymal cells, and express markers of glial differentiation (Kleihues et€al. 2002). Although ependymomas are thought to arise from cells adjacent to the ventricle, 10% are positioned distal from the ventricular system within the brain parenchyma. The incidence of ependymoma does not appear to vary by geographic region or socioeconomic class, but the disease may be more common in Caucasians (Gurney et€al. 2001; Alston et€al. 2007). Children and adults are predisposed to develop ependymoma in different parts of the CNS (Moynihan 2003): posterior fossa tumors present most often in patients aged less than 10€years; supratentorial ependymomas tend to occur in older children and adults; and spinal forms of the disease are confined largely to adults (CBTRUS 2006). There are no pathonemonic histologic features of ependymoma, but these tumors are characterized by perivascular collections of tumor cells termed pseudorosettes. Ependymoma cells typically display a glial immunophenotype and ultrastructural properties of ependymal cells. A number of ependymoma histologic subtypes are recognized. Myxopapillary and sub-ependymomas occur almost exclusively in the cauda equina and are relatively benign tumors (WHO grade I). Classic ependymomas can display papillary or clear cell features (WHO grade II), and are contrasted with anaplastic ependymoma that displays increased cellularity, cytologic atypia, and microvascular proliferation (WHO grade III). Ependymoblastomas are now considered a form of primitive neuroectodermal tumor. The degree to which ependymoma histology grade correlates with patient prognosis remains controversial and current histologic criteria are of limited use for predicting patient outcome. Therefore, collaborative efforts between European and American neuropathologists are underway to establish a more robust tumor grading system for ependymomas. There is also no official staging classification of ependymoma, but the diagnosis of metastatic disease is thought to be important for assigning appropriate treatment. However, measures to detect disseminated disease that include neuraxis MRI and CSF cytology are unreliable and are associated with significant rates of false positive and false negative results (Butler et€al. 2002; Kumar et€al. 2007).
Current Therapy Surgery and radiation have served as the mainstay of ependymoma therapy for more than 40€years (Kricheff et€al. 1964). Total tumor resection remains the most
Future Treatments of Ependymoma
293
important treatment for patients with localized ependymoma (Rousseau et€al. 1994; Duffner et€al. 1998; Robertson et€al. 1998). Indeed, data suggest that patients who achieve a complete surgical resection may require no further therapy (Hukin et€al. 1998). This approach is to be tested in one arm of the forthcoming Phase III ependymoma trial to be conducted within the Children’s Oncology Group (COG). If a gross-total resection is not achieved at first surgery, then careful consideration should be given to the use of further surgery. Second surgery may be accomplished safely, and in most cases, has been shown to be effective at achieving a gross-total resection ((Korshunov et€al. 2004) and Osterdock, Sanford, and Merchant, unpublished data). Most ependymomas are localized at diagnosis, and are therefore treated with 54 to 59.4€Gy of focal conformal adjuvant radiation targeted to the postoperativelydefined tumor bed (Merchant et€al. 2004). Although most patients can be spared whole neuraxis irradiation (Merchant et€al. 1997; Paulino et€al. 2002), seven percent of ependymomas are disseminated throughout the neuraxis at diagnosis, and craniospinal irradiation is the required treatment for these patients. Despite this aggressive therapy, disseminated ependymoma is extremely difficult to cure and long-term survival rates for this population are likely to be less than 25%. Children with recurrent ependymoma after prior irradiation have few treatment options and the majority will die within 2€ years. Treatment failure may include local progression, neuraxis dissemination, or a combination of the two. There are limited data suggesting that re-irradiation may be an effective treatment option for children with recurrent ependymoma when combined with resection of locally recurrent or metastatic disease. Although ependymoma is sensitive to conventional chemotherapy, there is no convincing evidence that this treatment improves overall patient survival. Conventional chemotherapies including cisplatin (Khan et€al. 1982; Sexauer et€al. 1985; Walker and Allen 1988), carboplatin (Gaynon et€ al. 1990; Friedman et€ al. 1992), ifosfamide (Chastagner et€ al. 1993) and etoposide (Davidson et€ al. 1993; Needle et€al. 1997) have each demonstrated modest activity against the disease; but combination chemotherapy regimens have proven ineffective at improving the cure rates achieved with surgery and radiation therapy alone (Evans et€ al. 1996; Robertson et€ al. 1998; Timmermann et€ al. 2000). While not generally effective, chemotherapy may have a role in delaying CNS irradiation of very young patients. The Pediatric Oncology Group (POG) 8,633 study, reported a 48% response rate to a combination of vincristine, cyclophosphamide, cisplatin, and etoposide among 25 children with residual tumor after initial surgery (Duffner et€ al. 1993). Similar results have been reported with other conventional chemotherapy regimens developed within the European consortia (Grill et€al. 2001; Grundy et€al. 2007). These data are important since exposure of the developing brain to radiation results in long-term endocrine and cognitive deficits (including learning, memory attention, and behavior disorders). Thus, ependymoma is a clinically heterogeneous disease for which there are few conventional treatment options. If the tumor is totally resected, then the disease can be controlled; however, disseminated or relapsed disease is usually fatal. There is general consensus that better understanding of disease biology will lead to more
294
R.J. Gilbertson
effective treatments for these patients. However, in contrast to astrocytic and primitive neuroectodermal tumors that have served as the focus of most brain tumor biology studies, very little is known about the biology of ependymoma.
Tumor Biology Genetics Progress in understanding the biology of ependymoma has been limited since most research has been conducted by isolated groups studying the sequence and expression of a handful of genes within a small number of locally and retrospectively collected tumors. Consequently, knowledge of the genetic alterations in ependymoma has been limited largely to lists of large chromosomal gains and losses, for example, +1q, −6q, +7p, −9p, −16q, and −22q (Reardon et€al. 1999; Hirose et€al. 2001; Ward et€al. 2001; Carter et€al. 2002; Dyer et€al. 2002; Grill et€al. 2002; Jeuken et€al. 2002; Koschny et€al. 2002). Studies of heritable forms of ependymoma have identified a very small number of tumor suppressor genes (TSG). Patients with Neurofibromatosis type 2 are predisposed to spinal ependymoma (Rouleau et€ al. 1993) and somatic mutations in NEUROFIBROMIN 2 (NF2, 22q12.2) occur in approximately 25% of sporadic SP tumors (Rubio et€al. 1994; Ebert et€al. 1999). Ependymomas have also been reported in two patients with Turcot’s syndrome (germline mutation in APC) (Torres et€al. 1997; Mullins et€al. 1998), and in one patient with Li Fraumeni Syndrome (germline mutation in TP53) (Metzger et€al. 1991). However, mutations in CTNNB1, APC or TP53 occur in less than 1% of sporadic ependymomas (Ohgaki et€ al. 1991; Gaspar et€ al. 2006; Onilude et€ al. 2006). In contrast to TSG, there are no known oncogenes of ependymoma, although overexpression of members of the Epidermal Growth Factor Receptor (EGFR) family has been reported in aggressive forms of the disease (Gilbertson et€ al. 2002; Mendrzyk et€al. 2006). Similarly, high tumor cell expression of the catalytic unit of telomerase (hTERT) has been associated with poor clinical outcome among patients with ependymoma (Tabori et€ al. 2006). Identifying oncogenes and other factors such as hTERT that contribute to transformation is particularly important for advancing treatment since these genes are likely to include the most useful drug targets (Druker et€al. 2001; Vogel et€al. 2002; Lynch et€al. 2004; Paez et€al. 2004).
Genomics The lack of knowledge regarding the key causative molecular alterations of ependymoma has hindered efforts to develop new treatments of the disease. However, tools that detect genome-wide patterns of gene expression and chromosomal
Future Treatments of Ependymoma
295
alteration are now being applied to the study of large cohorts of ependymoma. Of particular note, these studies have shown that ependymomas from the different regions of the CNS are molecularly distinct diseases (Korshunov et€ al. 2003; Ammerlaan et€al. 2005; Taylor et€al. 2005; Mendrzyk et€al. 2006; Modena et€al. 2006). For example, while the great majority of cerebral ependymomas activate the NOTCH signal pathway and delete the INK4A/ARF TSG locus, spinal ependymomas express high levels of members of the HOX gene family (Taylor et€al. 2005; Modena et€ al. 2006). These data are likely to be of great clinical significance; implying that ependymomas should be regarded as distinct site-specific entities that are likely to respond differentially to molecular targeted therapies. Ongoing genomic studies are identifying amplicons in ependymomas that contain oncogenes, for example, VAV1 (19p13.3) and hTERT (5p15.33) (Ammerlaan et€al. 2005; Taylor et€al. 2005; Mendrzyk et€al. 2006). These data provide proof-ofprincipal that genomic technologies are valuable mutation-discovery tools with which to study ependymoma. Most of the amplicons identified so far extend across megabases of chromosomal DNA and contain many genes (Ammerlaan et€al. 2005; Taylor et€ al. 2005; Mendrzyk et€ al. 2006). Therefore, pinpointing specific oncogenes and TSG of ependymoma will require the analysis of much larger numbers of samples with higher resolution DNA mapping tools. Efforts are now underway to apply technologies of higher resolution (500€K single nucleotide polymorphism mapping arrays) to pinpoint specific oncogenes and TSG among large cohorts of ependymoma.
Cancer Stem Cells and The Origins of Ependymoma Considerable excitement has surrounded the recent discovery of cancer stem cells (CSC) (Clarke and Fuller 2006). CSC make up just a small fraction of the total population of the malignant cells in many solid tumors and leukemias. However, evidence indicates that these self-renewing and multipotent stem cell-like cells generate all of the phenotypically diverse cells that populate tumors (Lapidot et€al. 1994; Bonnet and Dick 1997; Al-Hajj et€al. 2003; Singh et€al. 2004; Taylor et€al. 2005). The discovery of CSC has therefore provided researchers with a practical point of focus for studying the natal cellular and molecular events of tumorigenesis. The identification of CSC is likely to have important implications for the treatment of cancer. If tumors are derived entirely from CSC, then it would follow that to be curative cancer, treatments should disable or destroy these cells. Indeed, drugs that are designed to kill CSC could prove highly effective treatments of cancer. As outlined above, ependymoma subsets have been shown to exhibit distinct patterns of gene expression and regions of chromosome gain and loss that correlate with the anatomic location of the tumor (supratentorial region, posterior fossa or spine). Taylor et€al. (2005) demonstrated that the gene expression signatures that most discriminate supratentorial, posterior fossa, and spinal ependymoma included
296
R.J. Gilbertson
many genes that are known regulators of neural precursor cells in the corresponding region of the CNS. Indeed, using in situ hybridization and immunofluorescence to map the site of expression of ependymoma signature genes in the developing mouse, this study went on to identify embryonic radial glia (RG) – that are neural progenitor cells – as candidate cells of origin of ependymoma. Importantly, the study showed also that self-renewing and multipotent CSC isolated from fresh samples of ependymoma are bipolar RG-like cells; express the CD133+/Nestin+/ RC2+/BLBP+ immunophenotype of RG; and are both required and sufficient to generate tumors in€vivo. Interestingly, pilocytic astrocytomas have been shown also to share the gene expression profiles of normal primary astrocytes and neural stem cells from the corresponding region of the brain (Sharma et€ al. 2007). Together, these data suggest a new hypothesis for the origin of ependymoma and other glial tumors; progenitor cells in different parts of the CNS are predisposed to acquire distinct genetic abnormalities that transform these cells into CSC of supratentorial, posterior fossa, and spinal tumors (Gilbertson and Gutmann 2007).
Ependymoma Stem Cell Niches Stem cells of organisms across the plant and animal kingdoms reside within stem cell niches, highlighting the fundamental importance of these specialized microenvironments for normal stem cell biology (Scheres 2007). The central structural element of the neural stem cell niche is provided by capillaries (Riquelme et€al. 2007). This organization places the stem cells in close proximity to endothelial and other vascular cells, facilitating communication among these cell types. Calabrese and colleagues recently provided compelling data that stem cells from a variety of brain tumors, including ependymoma, are maintained within vascular niches that mimic the neural stem cell niche (Calabrese et€ al. 2007). Using co-immunofluorescence and multi-photon laser scanning microscopy, they showed first that CD133+, Nestin+ cells within sections of human ependymomas are located in close proximity to tumor capillaries. They then demonstrated that CD133+, Nestin+ cells, but not other cells isolated from ependymomas, migrate to interact intimately with the vascular tubes that are formed by endothelial cells in three-dimensional cultures. The self-renewal and proliferation of these ependymoma stem cells in culture was maintained by factors secreted by endothelial cells. Importantly, by co-transplanting brain tumor stem cells and endothelial cells into immunocompromised mice, the investigators showed that endothelial-derived factors also accelerate the initiation and growth of tumors in the brain. Finally, the investigators demonstrated that the anti-VEGF antibody bevacizumab can halt the growth of brain tumors by disrupting the perivascular niche and depleting the cancer stem cell fraction. Although additional studies will be required to understand better the ependymoma stem cell niche, these early data have exposed a previously unknown vulnerability in brain tumors for therapeutic attack (Gilbertson and Rich 2007).
Future Treatments of Ependymoma
297
Prior Testing of Molecular Targeted Therapies in Ependymoma Given the paucity of information regarding the mutations that cause ependymoma and the relative rarity of the disease, it is not surprising that few molecular targeted therapies have been tested in this tumor. Those drugs that have been tested have been selected empirically, often in the context of Phase I studies that have enrolled patients with a variety of recurrent or progressive CNS tumor. For example, the US Pediatric Brain Tumor Consortium (PBTC) recently tested in Phase I the farnesyltransferase inhibitor lonafarnib (SCH66336). Radiographic responses or stable disease were reported in 10 of 53 patients with recurrent or refractory brain tumors including one child with ependymoma (Kieran et€al. 2007). Although this approach is useful for gathering drug toxicity, dose and pharmacokinetic data, more efficient mechanisms will be required to develop new therapies of ependymoma. Ongoing clinical trials are targeting the EGFR family in ependymoma (e.g., PBTC016 study of Lapatinib) that has been reported to be overexpressed in poor prognosis ependymoma (Gilbertson et€al. 2002; Mendrzyk et€al. 2006); however, a causative role for EGFR family signaling in ependymoma remains to be demonstrated. Anecdotal reports of the therapeutic activity of other growth factor receptor inhibitors (e.g., Imatinib) in ependymoma have also been reported but remain to be confirmed (Fakhrai et€al. 2004).
How Can We Improve the Treatment of Ependymoma? There is a dire need for new treatment approaches of ependymoma. Identifying these treatments will require a comprehensive and collaborative research process that builds on what we know about the disease, and continually integrates new understanding of disease biology and clinical management. This process should account for a number of key issues.
Ependymoma Is not a Single Disease The observation that ependymoma comprises subgroups of clinically and molecularly distinct diseases is a critical finding of recent studies. These data suggest strongly that cerebral, posterior fossa, and spinal tumors should be treated separately. Efforts should be made to continually integrate understanding of the genetic alterations in these disease subgroups with attempts to develop molecular targeted therapies. For example, gene expression profile studies have identified high expression of the NOTCH ligands JAGGED 1 and 2 in supratentorial tumors (Taylor et€al. 2005). Binding of JAGGED ligands to NOTCH leads to g-secretase mediated cleavage of the NOTCH receptor and activation of signaling (Radtke and Raj 2003). Thus, inhibitors of g-secretase that are being considered for clinical trials in leukemia could have utility in the treatment of supratentorial ependymoma (Huntly and Gilliland 2005). Importantly, JAGGED1 and
298
R.J. Gilbertson
2 are not expressed highly in posterior fossa or spinal ependymomas, suggesting that inhibitors of g-secretase might be less effective against these tumors. Integrated analyses of molecular data with central clinical and histologic data should also transform the classification of ependymoma, and provide a new clinico-molecular disease-risk stratification system for this disease. This new disease-risk stratification system may allow greater tailoring of conventional therapies among patients as well as identify new targets for molecular based therapies. Defining ependymoma subgroups may improve understanding of the disease and the efficiency with which we use conventional treatments; but it will also convert an already uncommon disease into a group of extremely rare tumors: this will present a major challenge to the conduct of randomized clinical trials. It will only be possible to conduct these trials through larger international collaborations. Thus, we will need to forge close links between North American and European consortia. One such collaboration is already under the leadership of groups at St Jude Children’s Research Hospital and MD Anderson Cancer Center (www.CERN-Foundation.org). This privately funded consortium has established an adult and pediatric clinical trials network as well as comprehensive projects in histopathology, genomics, drug discovery, and ependymoma stem cells. The support of the national cancer institutions in the US (e.g., NCI and NINDS) and Europe (e.g., NICR and CRUK) will also play important roles in the collaborative process. The international pediatric neuro-oncology community would be wise to give early consideration to establishing the international links necessary to complete these studies.
Some Ependymomas May Be Cured with Minimal Conventional Therapy It is important to recognize that an as yet undefined group of patients with ependymoma can be cured with minimal conventional therapy. In the case of completely (macro- and microscopically) resected intracranial disease, this may involve surgery alone. Better histologic and molecular classification systems of ependymoma may allow prospective identification of these patients. In the absence of alternative therapies, postoperative focal radiation is likely to continue to be used to treat patients undergoing sub-total resections. These patients are still at risk of distantmetastatic spread and new methods are required to define residual disease at the primary site or evidence of metastatic dissemination.
Curing All Patients with Ependymoma Will Require New Treatments While improvements may be made in the use of conventional treatments, these will not cure all patients with ependymoma. There are at least three areas that could provide fertile ground for developing new treatment approaches. First, seminal studies of a variety of hematologic and solid tumors have identified
Future Treatments of Ependymoma
299
gene mutations as targets of effective new anti-cancer therapies (Druker et€al. 2001; Vogel et€al. 2002; Lynch et€al. 2004; Paez et€al. 2004; Romer et€al. 2004). Thus, ongoing genomic efforts to characterize oncogenic mutations in ependymoma could lead to the development of new treatments. At present there are relatively few widely available cell line or mouse models of ependymoma that can be used for pre-clinical drug development. Thus, efforts to understand the causative mutations of ependymoma should generate also preclinical models that can be used to develop treatments that target these alterations. Second, the recent discovery that ependymomas are derived from radial glia-like CSC that reside in perivascular niches (Taylor et€al. 2005; Calabrese et€al. 2007) has identified a series of completely new drug targets to treat this disease (Gilbertson and Rich 2007). Drugs that target CSC or their niches in ependymoma could prove highly effective treatments; however, the similarities between normal and malignant neural stem cells predict that such treatments may also possess significant toxicities. The development of anti-CSC therapies for ependymoma will require the identification of factors that maintain ependymoma CSC, but not normal neural stem cells. Third, it may prove possible to develop strategies that increase the overall sensitivity of ependymoma to conventional treatments. These strategies are likely to target the distinct populations of cells that exist within tumors, including the CSC. Glioma stem cells have been shown to be inherently radio – and chemoresistant (Bao et€al. 2006) and might be protected further from conventional therapies by factors within the vascular niche. Treatments that disrupt aberrant vascular stem cell niches might therefore enhance sensitivity to conventional treatments. Indeed, clinical trials that include a combination of the antiangiogenic drug Bevacizumab and the cytotoxic irinotecan have revealed remarkable synergy between these two drugs in the treatment of glioblastoma (Vredenburgh et€al. 2007). It is possible that this combination therapy disrupts the vascular niche, exposing both glioblastoma stem cells and the remaining tumor cells to the cytotoxic effects of the conventional chemotherapy.
Future Directions and Challenges In 2008, ependymoma remains a poorly understood disease with few treatment options. The outlook is especially dismal for patients with disseminated disease. Advances in genomics, stem cell biology, and developmental therapeutics hold great promise to break the deadlock in the hunt for new treatments of ependymoma. International collaboration between medical, scientific, and government personnel will be critical if this effort is to succeed.
References Al-Hajj, M., Wicha, M.S., Benito-Hernandez, A., Morrison, S.J., and Clarke, M.F. 2003. From the cover: prospective identification of tumorigenic breast cancer cells. Proc Natl Acad Sci U S A 100: 3983–3988.
300
R.J. Gilbertson
Alston, R.D., Rowan, S., Eden, T.O., Moran, A., and Birch, J.M. 2007. Cancer incidence patterns by region and socioeconomic deprivation in teenagers and young adults in England. Br J Cancer 96: 1760–1766. Ammerlaan, A.C., de Bustos, C., Ararou, A., Buckley, P.G., Mantripragada, K.K., Verstegen, M.J., Hulsebos, T.J., and Dumanski, J.P. 2005. Localization of a putative low-penetrance ependymoma susceptibility locus to 22q11 using a chromosome 22 tiling-path genomic microarray. Genes Chromosomes Cancer Epublication ahead of print. Bao, S., Wu, Q., McLendon, R.E., Hao, Y., Shi, Q., Hjelmeland, A.B., Dewhirst, M.W., Bigner, D.D., and Rich, J.N. 2006. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 444: 756–760. Bonnet, D., and Dick, J.E. 1997. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat Med 3: 730–737. Brandes, A.A., Cavallo, G., Reni, M., Tosoni, A., Nicolardi, L., Scopece, L., Franceschi, E., Sotti, G., Talacchi, A., Turazzi, S., et€al. 2005. A multicenter retrospective study of chemotherapy for recurrent intracranial ependymal tumors in adults by the Gruppo Italiano Cooperativo di Neuro-Oncologia. Cancer 104: 143–148. Butler, W.E., Khan, A., and Khan, S.A. 2002. Posterior fossa ependymoma with intense but transient disseminated enhancement but not metastasis. Pediatr Neurosurg 37: 27–31. Calabrese, C., Poppleton, H., Kocak, M., Hogg, T.L., Fuller, C., Hamner, B., Oh, E.Y., Gaber, M.W., Finklestein, D., Allen, M., et€al. 2007. A perivascular niche for brain tumor stem cells. Cancer Cell 11: 69–82. Carter, M., Nicholson, J., Ross, F., Crolla, J., Allibone, R., Balaji, V., Perry, R., Walker, D., Gilbertson, R., and Ellison, D.W. 2002. Genetic abnormalities detected in ependymomas by comparative genomic hybridisation. Br J Cancer 86: 929–939. CBTRUS. 2006. Statistical Report: Primary Brain Tumors in the United States, 1995–1999. Central Brain Tumor Registry of the United States, Hinsdale, IL, 2002. Chastagner, P., Sommelet-Olive, D., Kalifa, C., Brunat-Mentigny, M., Zucker, J.M., Demeocq, F., Baranzelli, M.C., Tron, P., Bergeron, C., Pein, F., et€al. 1993. Phase II study of ifosfamide in childhood brain tumors: a report by the French Society of Pediatric Oncology (SFOP). Med Pediatr Oncol 21: 49–53. Clarke, M.F., and Fuller, M. 2006. Stem cells and cancer: two faces of eve. Cell 124: 1111–1115. Davidson, A., Lewis, I., Pearson, A.D., Stevens, M.C., and Pinkerton, C.R. 1993. 21-day schedule oral etoposide in children – a feasibility study. Eur J Cancer 29A: 2223–2225. Druker, B.J., Sawyers, C.L., Kantarjian, H., Resta, D.J., Reese, S.F., Ford, J.M., Capdeville, R., and Talpaz, M. 2001. Activity of a specific inhibitor of the BCR-ABL tyrosine kinase in the blast crisis of chronic myeloid leukemia and acute lymphoblastic leukemia with the Philadelphia chromosome. N Engl J Med 344: 1038–1042. Duffner, P.K., Horowitz, M.E., Krischer, J.P., Friedman, H.S., Burger, P.C., Cohen, M.E., Sanford, R.A., Mulhern, R.K., James, H.E., Freeman, C.R., et€ al. 1993. Postoperative chemotherapy and delayed radiation in children less than three years of age with malignant brain tumors. N Engl J Med 328: 1725–1731. Duffner, P.K., Krischer, J.P., Sanford, R.A., Horowitz, M.E., Burger, P.C., Cohen, M.E., Friedman, H.S., and Kun, L.E. 1998. Prognostic factors in infants and very young children with intracranial ependymomas. Pediatr Neurosurg 28: 215–222. Dyer, S., Prebble, E., Davison, V., Davies, P., Ramani, P., Ellison, D., and Grundy, R. 2002. Genomic imbalances in pediatric intracranial ependymomas define clinically relevant groups. Am J Pathol 161: 2133–2141. Ebert, C., von Haken, M., Meyer-Puttlitz, B., Wiestler, O.D., Reifenberger, G., Pietsch, T., and von Deimling, A. 1999. Molecular genetic analysis of ependymal tumors. NF2 mutations and chromosome 22q loss occur preferentially in intramedullary spinal ependymomas. Am J Pathol 155: 627–632. Evans, A.E., Anderson, J.R., Lefkowitz-Boudreaux, I.B., and Finlay, J.L. 1996. Adjuvant chemotherapy of childhood posterior fossa ependymoma: cranio-spinal irradiation with or without adjuvant CCNU, vincristine, and prednisone: a Childrens Cancer Group study. Med Pediatr Oncol 27: 8–14.
Future Treatments of Ependymoma
301
Fakhrai, N., Neophytou, P., Dieckmann, K., Nemeth, A., Prayer, D., Hainfellner, J., and Marosi, C. 2004. Recurrent spinal ependymoma showing partial remission under Imatimib. Acta Neurochir (Wien) 146: 1255–1258. Friedman, H.S., Krischer, J.P., Burger, P., Oakes, W.J., Hockenberger, B., Weiner, M.D., Falletta, J.M., Norris, D., Ragab, A.H., Mahoney, D.H., Jr., et€ al. 1992. Treatment of children with progressive or recurrent brain tumors with carboplatin or iproplatin: a Pediatric Oncology Group randomized phase II study. J Clin Oncol 10: 249–256. Gaspar, N., Grill, J., Geoerger, B., Lellouch-Tubiana, A., Michalowski, M.B., and Vassal, G. 2006. p53 pathway dysfunction in primary childhood ependymomas. Pediatr Blood Cancer 46: 604-613. Gaynon, P.S., Ettinger, L.J., Baum, E.S., Siegel, S.E., Krailo, M.D., and Hammond, G.D. 1990. Carboplatin in childhood brain tumors. A Children’s Cancer Study Group Phase II trial. Cancer 66: 2465–2469. Gilbertson, R.J., Bentley, L., Hernan, R., Junttila, T.T., Frank, A.J., Haapasalo, H., Connelly, M., Wetmore, C., Curran, T., Elenius, K., et€ al. 2002. ERBB receptor signaling promotes ependymoma cell proliferation and represents a potential novel therapeutic target for this disease. Clin Cancer Res 8: 3054–3064. Gilbertson, R.J., and Gutmann, D.H. 2007. Tumorigenesis in the brain: location, location, location. Cancer Res 67: 5579–5582. Gilbertson, R.J., and Rich, J.N. 2007. Making a tumour’s bed: glioblastoma stem cells and the vascular niche. Nat Rev Cancer 7: 733–736. Grill, J., Avet-Loiseau, H., Lellouch-Tubiana, A., Sevenet, N., Terrier-Lacombe, M.J., Venuat, A.M., Doz, F., Sainte-Rose, C., Kalifa, C., and Vassal, G. 2002. Comparative genomic hybridization detects specific cytogenetic abnormalities in pediatric ependymomas and choroid plexus papillomas. Cancer Genet Cytogenet 136: 121–125. Grill, J., Le Deley, M.C., Gambarelli, D., Raquin, M.A., Couanet, D., Pierre-Kahn, A., Habrand, J.L., Doz, F., Frappaz, D., Gentet, J.C., et€al. 2001. Postoperative chemotherapy without irradiation for ependymoma in children under 5 years of age: a multicenter trial of the French Society of Pediatric Oncology. J Clin Oncol 19: 1288–1296. Grundy, R.G., Wilne, S.A., Weston, C.L., Robinson, K., Lashford, L.S., Ironside, J., Cox, T., Chong, W.K., Campbell, R.H., Bailey, C.C., et€al. 2007. Primary postoperative chemotherapy without radiotherapy for intracranial ependymoma in children: the UKCCSG/SIOP prospective study. Lancet Oncol 8: 696–705. Gurney, J., Smith, M., and Bunin, G. 2001. CNS and miscellaneous intra-cranial and intraspinal neoplasms. SEER Pediatric Monograph, National Cancer Institute: 51–63. Hirose, Y., Aldape, K., Bollen, A., James, C.D., Brat, D., Lamborn, K., Berger, M., and Feuerstein, B.G. 2001. Chromosomal abnormalities subdivide ependymal tumors into clinically relevant groups. Am J Pathol 158: 1137–1143. Hukin, J., Epstein, F., Lefton, D., and Allen, J. 1998. Treatment of intracranial ependymoma by surgery alone. Pediatr Neurosurg 29: 40–45. Huntly, B.J., and Gilliland, D.G. 2005. Leukaemia stem cells and the evolution of cancer-stem-cell research. Nat Rev Cancer 5: 311–321. Jeuken, J.W., Sprenger, S.H., Gilhuis, J., Teepen, H.L., Grotenhuis, A.J., and Wesseling, P. 2002. Correlation between localization, age, and chromosomal imbalances in ependymal tumours as detected by CGH. J Pathol 197: 238–244. Khan, A.B., D’Souza, B.J., Wharam, M.D., Champion, L.A., Sinks, L.F., Woo, S.Y., McCullough, D.C., and Leventhal, B.G. 1982. Cisplatin therapy in recurrent childhood brain tumors. Cancer Treat Rep 66: 2013–2020. Kieran, M.W., Packer, R.J., Onar, A., Blaney, S.M., Phillips, P., Pollack, I.F., Geyer, J.R., Gururangan, S., Banerjee, A., Goldman, S., et€al. 2007. Phase I and pharmacokinetic study of the oral farnesyltransferase inhibitor lonafarnib administered twice daily to pediatric patients with advanced central nervous system tumors using a modified continuous reassessment method: a Pediatric Brain Tumor Consortium Study. J Clin Oncol 25: 3137–3143. Kleihues, P., Louis, D.N., Scheithauer, B.W., Rorke, L.B., Reifenberger, G., Burger, P.C., and Cavenee, W.K. 2002. The WHO classification of tumors of the nervous system. J Neuropathol Exp Neurol 61: 215–225; discussion 226–219.
302
R.J. Gilbertson
Korshunov, A., Golanov, A., Sycheva, R., and Timirgaz, V. 2004. The histologic grade is a main prognostic factor for patients with intracranial ependymomas treated in the microneurosurgical era: an analysis of 258 patients. Cancer 100: 1230–1237. Korshunov, A., Neben, K., Wrobel, G., Tews, B., Benner, A., Hahn, M., Golanov, A., and Lichter, P. 2003. Gene expression patterns in ependymomas correlate with tumor location, grade, and patient age. Am J Pathol 163: 1721–1727. Koschny, R., Koschny, T., Froster, U.G., Krupp, W., and Zuber, M.A. 2002. Comparative genomic hybridization in glioma: a meta-analysis of 509 cases. Cancer Genet Cytogenet 135: 147–159. Kricheff, II, Becker, M., Schneck, S.A., and Taveras, J.M. 1964. Intracranial ependymomas: factors influencing prognosis. J Neurosurg 21: 7–14. Kumar, P., Rastogi, N., Jain, M., and Chhabra, P. 2007. Extraneural metastases in anaplastic ependymoma. J Cancer Res Therap: 102–104. Lapidot, T., Sirard, C., Vormoor, J., Murdoch, B., Hoang, T., Caceres-Cortes, J., Minden, M., Paterson, B., Caligiuri, M.A., and Dick, J.E. 1994. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 367: 645–648. Lynch, T.J., Bell, D.W., Sordella, R., Gurubhagavatula, S., Okimoto, R.A., Brannigan, B.W., Harris, P.L., Haserlat, S.M., Supko, J.G., Haluska, F.G., et€al. 2004. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non-small-cell lung cancer to gefitinib. N Engl J Med 350: 2129–2139. Epub 2004 Apr 2129. Mendrzyk, F., Korshunov, A., Benner, A., Toedt, G., Pfister, S., Radlwimmer, B., and Lichter, P. 2006. Identification of gains on 1q and epidermal growth factor receptor overexpression as independent prognostic markers in intracranial ependymoma. Clin Cancer Res 12: 2070–2079. Merchant, T.E., and Fouladi, M. 2005. Ependymoma: new therapeutic approaches including radiation and chemotherapy. J Neurooncol 75: 287–299. Merchant, T.E., Haida, T., Wang, M.H., Finlay, J.L., and Leibel, S.A. 1997. Anaplastic ependymoma: treatment of pediatric patients with or without craniospinal radiation therapy. J Neurosurg 86: 943–949. Merchant, T.E., Mulhern, R.K., Krasin, M.J., Kun, L.E., Williams, T., Li, C., Xiong, X., Khan, R.B., Lustig, R.H., Boop, F.A., et€al. 2004. Preliminary results from a phase II trial of conformal radiation therapy and evaluation of radiation-related CNS effects for pediatric patients with localized ependymoma. J Clin Oncol 22: 3156–3162. Metzger, A.K., Sheffield, V.C., Duyk, G., Daneshvar, L., Edwards, M.S., and Cogen, P.H. 1991. Identification of a germ-line mutation in the p53 gene in a patient with an intracranial ependymoma. Proc Natl Acad Sci U S A 88: 7825–7829. Modena, P., Lualdi, E., Facchinetti, F., Veltman, J., Reid, J.F., Minardi, S., Janssen, I., Giangaspero, F., Forni, M., Finocchiaro, G., et€al. 2006. Identification of tumor-specific molecular signatures in intracranial ependymoma and association with clinical characteristics. J Clin Oncol 24: 5223–5233. Moynihan, T.J. 2003. Ependymal tumors. Curr Treat Options Oncol 4: 517–523. Mullins, K.J., Rubio, A., Myers, S.P., Korones, D.N., and Pilcher, W.H. 1998. Malignant ependymomas in a patient with Turcot’s syndrome: case report and management guidelines. Surg Neurol 49: 290–294. Needle, M.N., Molloy, P.T., Geyer, J.R., Herman-Liu, A., Belasco, J.B., Goldwein, J.W., Sutton, L., and Phillips, P.C. 1997. Phase II study of daily oral etoposide in children with recurrent brain tumors and other solid tumors. Med Pediatr Oncol 29: 28–32. Ohgaki, H., Eibl, R.H., Wiestler, O.D., Yasargil, M.G., Newcomb, E.W., and Kleihues, P. 1991. p53 mutations in nonastrocytic human brain tumors. Cancer Res 51: 6202–6205. Onilude, O.E., Lusher, M.E., Lindsey, J.C., Pearson, A.D., Ellison, D.W., and Clifford, S.C. 2006. APC and CTNNB1 mutations are rare in sporadic ependymomas. Cancer Genet Cytogenet 168: 158–161. Paez, J.G., Janne, P.A., Lee, J.C., Tracy, S., Greulich, H., Gabriel, S., Herman, P., Kaye, F.J., Lindeman, N., Boggon, T.J., et€ al. 2004. EGFR mutations in lung cancer: correlation with clinical response to gefitinib therapy. Science 29: 29.
Future Treatments of Ependymoma
303
Paulino, A.C., Wen, B.C., Buatti, J.M., Hussey, D.H., Zhen, W.K., Mayr, N.A., and Menezes, A.H. 2002. Intracranial ependymomas: an analysis of prognostic factors and patterns of failure. Am J Clin Oncol 25: 117–122. Radtke, F., and Raj, K. 2003. The role of Notch in tumorigenesis: oncogene or tumour suppressor? Nat Rev Cancer 3: 756–767. Reardon, D.A., Entrekin, R.E., Sublett, J., Ragsdale, S., Li, H., Boyett, J., Kepner, J.L., and Look, A.T. 1999. Chromosome arm 6q loss is the most common recurrent autosomal alteration detected in primary pediatric ependymoma. Genes Chromosomes Cancer 24: 230–237. Riquelme, P.A., Drapeau, E., and Doetsch, F. 2007. Brain micro-ecologies: neural stem cell niches in the adult mammalian brain. Philos Trans R Soc Lond B Biol Sci 23: 23. Robertson, P.L., Zeltzer, P.M., Boyett, J.M., Rorke, L.B., Allen, J.C., Geyer, J.R., Stanley, P., Li, H., Albright, A.L., McGuire-Cullen, P., et€al. 1998. Survival and prognostic factors following radiation therapy and chemotherapy for ependymomas in children: a report of the Children’s Cancer Group. J Neurosurg 88: 695–703. Romer, J.T., Kimura, H., Magdaleno, S., Sasai, K., Fuller, C., Baines, H., Connelly, M., Stewart, C.F., Gould, S., Rubin, L.L., et€al. 2004. Suppression of the Shh pathway using a small molecule inhibitor eliminates medulloblastoma in Ptc1(+/−)p53(−/−) mice. Cancer Cell 6: 229–240. Rouleau, G.A., Merel, P., Lutchman, M., Sanson, M., Zucman, J., Marineau, C., Hoang-Xuan, K., Demczuk, S., Desmaze, C., Plougastel, B., et€al. 1993. Alteration in a new gene encoding a putative membrane-organizing protein causes neuro-fibromatosis type 2. Nature 363: 515. Rousseau, P., Habrand, J.L., Sarrazin, D., Kalifa, C., Terrier-Lacombe, M.J., Rekacewicz, C., and Rey, A. 1994. Treatment of intracranial ependymomas of children: review of a 15-year experience. Int J Radiat Oncol Biol Phys 28: 381–386. Rubio, M.P., Correa, K.M., Ramesh, V., MacCollin, M.M., Jacoby, L.B., von Deimling, A., Gusella, J.F., and Louis, D.N. 1994. Analysis of the neurofibromatosis 2 gene in human ependymomas and astrocytomas. Cancer Res 54: 45–47. Scheres, B. 2007. Stem-cell niches: nursery rhymes across kingdoms. Nat Rev Mol Cell Biol 8: 345–354. Sexauer, C.L., Khan, A., Burger, P.C., Krischer, J.P., van Eys, J., Vats, T., and Ragab, A.H. 1985. Cisplatin in recurrent pediatric brain tumors. A POG Phase II study. A Pediatric Oncology Group Study. Cancer 56: 1497–1501. Sharma, M.K., Mansur, D.B., Reifenberger, G., Perry, A., Leonard, J.R., Aldape, K.D., Albin, M.G., Emnett, R.J., Loeser, S., Watson, M.A., et€al. 2007. Distinct genetic signatures among pilocytic astrocytomas relate to their brain region origin. Cancer Res 67: 890–900. Singh, S.K., Hawkins, C., Clarke, I.D., Squire, J.A., Bayani, J., Hide, T., Henkelman, R.M., Cusimano, M.D., and Dirks, P.B. 2004. Identification of human brain tumour initiating cells. Nature 432: 396–401. Tabori, U., Ma, J., Carter, M., Zielenska, M., Rutka, J., Bouffet, E., Bartels, U., Malkin, D., and Hawkins, C. 2006. Human telomere reverse transcriptase expression predicts progression and survival in pediatric intracranial ependymoma. J Clin Oncol 24: 1522–1528. Taylor, M.D., Poppleton, H., Fuller, C., Su, X., Liu, Y., Jensen, P., Magdaleno, S., Dalton, J., Calabrese, C., Board, J., et€al. 2005. Radial glia cells are candidate stem cells of ependymoma. Cancer Cell 8: 323–335. Timmermann, B., Kortmann, R.D., Kuhl, J., Meisner, C., Slavc, I., Pietsch, T., and Bamberg, M. 2000. Combined postoperative irradiation and chemotherapy for anaplastic ependymomas in childhood: results of the German prospective trials HIT 88/89 and HIT 91. Int J Radiat Oncol Biol Phys 46: 287–295. Torres, C.F., Korones, D.N., and Pilcher, W. 1997. Multiple ependymomas in a patient with Turcot’s syndrome. Med Pediatr Oncol 28: 59–61. Vogel, C.L., Cobleigh, M.A., Tripathy, D., Gutheil, J.C., Harris, L.N., Fehrenbacher, L., Slamon, D.J., Murphy, M., Novotny, W.F., Burchmore, M., et€al. 2002. Efficacy and safety of trastuzumab as a single agent in first-line treatment of HER2-overexpressing metastatic breast cancer. J Clin Oncol 20: 719–726.
304
R.J. Gilbertson
Vredenburgh, J.J., Desjardins, A., Herndon, J.E., II, Dowell, J.M., Reardon, D.A., Quinn, J.A., Rich, J.N., Sathornsumetee, S., Gururangan, S., Wagner, M., et€ al. 2007. Phase II Trial of Bevacizumab and Irinotecan in Recurrent Malignant Glioma, pp. 1253–1259. Walker, R.W., and Allen, J.C. 1988. Cisplatin in the treatment of recurrent childhood primary brain tumors. J Clin Oncol 6: 62–66. Ward, S., Harding, B., Wilkins, P., Harkness, W., Hayward, R., Darling, J.L., Thomas, D.G., and Warr, T. 2001. Gain of 1q and loss of 22 are the most common changes detected by comparative genomic hybridisation in paediatric ependymoma. Genes Chromosomes Cancer 32: 59–66.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions of INI1/hSNF5 Tumor Suppressor Ganjam V. Kalpana and Melissa E. Smith
Introduction Rhabdoid tumors (RTs) are rare, but highly aggressive and mostly incurable �pediatric malignancies (Strother 2005; Biegel 2006). RTs occur in children younger than 5 years of age and the peak incidence is between birth and 3 years of age (Biegel 2006). RTs were originally described in kidneys and termed Malignant Rhabdoid Tumors (MRT), and were subsequently found in the central nervous system (Atypical Teratoid and Rhabdoid Tumors, AT/RT), and soft tissues (extrarenal Rhabdoid Tumors) (Biegel 2006). Irrespective of their location, all RTs are characterized by the presence of sheets or nests of rhabdoid cells. Recent molecular genetic studies have established that RTs are distinguished from other tumors by the presence of recurrent biallelic deletions and/or mutations in the INI1/hSNF5 gene, located at chromosome 22q11.2, providing a definitive diagnostic criteria (Versteege et€al. 1998; Biegel 1999; Biegel et€al. 1999; Sevenet et€al. 1999). Current therapeutic regimens for RT involve empirically selected combinations of chemotherapeutic agents that are highly toxic and rarely curative, as such, the survival rate for children with RTs remains poor (Packer et€al. 2002; Reddy 2005; Strother 2005; Biegel 2006; Yamamoto et€al. 2006). Thus, there is a dire need to develop novel therapeutic strategies for RTs, preferably based on the understanding of the molecular factors responsible for the genesis, growth, and survival of these tumors. In the following sections, we will describe features of RTs, current therapeutic practices, the molecular basis of rhabdoid tumorigenesis, and development of potential targeted therapies based on these understandings.
G.V. Kalpana (*) Department of Molecular Genetics and Albert Einstein College Cancer Center, Albert Einstein College of Medicine of Yeshiva University, 1300, Morris Park Ave., Ullman 821, Bronx, NY 10461, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_15, © Springer Science+Business Media, LLC 2010
305
306
G.V. Kalpana and M.E. Smith
General Features of Rhabdoid Tumors Pathology RTs arise most commonly in the kidney and these MRTs comprise from 1.5 to 4% of malignant renal tumors (Mitchell et€ al. 2000; Tomlinson et€ al. 2005). Additionally, AT/RTs account for approximately 1 to 2% of pediatric brain tumors and 10% of CNS tumors arising in infants (Biegel 2006). Although AT/RTs arise within the entire CNS, approximately 50% of these tumors are present within the posterior fossa (Packer et€ al. 2002). In this location, there is preference for the cerebellopontine angle and the tumor often invades adjacent structures such as meninges and ependyma (Packer et€ al. 2002). Imaging of these tumors reveals mixed cellularity, hemorrhage, necrosis, and cysts (Packer et€al. 2002). Computed tomographic (CT) images of AT/RT show a hyperdense mass with enhanced intensity (Packer et€ al. 2002). The T1-weighted magnetic resonance images (MRI) visualize the tumor as being isointense often with hyperintense foci due to intratumoral hemorrhage and the T2 images are heterogeneous with hypo-, iso-, and hyper-intense foci (Packer et€al. 2002).
Diagnosis RTs were first described as a separate entity in 1978 (Beckwith and Palmer 1978). Until that time, they were considered an aggressive “rhabdomyosarcomatoid” variant of Wilms’ tumor and were often misclassified. MRTs have been commonly misdiagnosed as rhabdomyosarcomas, Wilms’ tumors, choroid plexus carcinoma (CPC), germ cell tumors, or ependymomas while AT/RTs are mistaken for glioblastoma, medulloblastoma (MB), or primitive neuroectodermal tumors (PNET) (Biegel 2006). One reason for this misdiagnosis is that AT/RTs display nonspecific radiologic features. AT/RTs appear as sections of increased density on nonenhanced CT scan, and with MRI, there is decreased signal intensity on various images due to hypercellularity, features that do not allow for the differential diagnosis of AT/RTs versus PNET, MB, CPC, teratoma, astrocytoma, or ependymoma (Fenton and Foreman 2003). As studies of RTs progressed, advanced diagnostic methods became available based on cellular, genetic, and molecular features of these tumors. Histologically, all RTs contain sheets or nests of rhabdoid cells alongside primitive neuroepithelial, epithelial, and/or mesenchymal elements (Fenton and Foreman 2003). Rhabdoid cells are large polygonal cells with eccentric nuclei, a prominent nucleolus, and juxtanuclear eosinophilic cytoplasmic inclusions (Packer et€ al. 2002; Hoot et€ al. 2004). Though varied, the immunophenotypic profile of RT cells can support diagnosis. The most consistently expressed markers within RTs are vimentin, epithelial membrane antigen, smooth muscle actin, and cytokeratin (Bruch et€al. 2001; Hoot
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
307
et€ al. 2004; Hauser et€ al. 2001). Among these, positive staining for epithelial membrane antigen often distinguishes RTs from other CNS tumors (Biegel 2006). Defining the molecular genetics of these rare tumors (discussed further in detail below) has provided additional tools for diagnosis. In brief, identification of the INI1/hSNF5 gene, located at chromosome 22q11.2, as a critical tumor suppressor inactivated in a majority (>95%) of RTs has facilitated definitive diagnosis for identification of RTs (Versteege et€al. 1998; Biegel 2006). Immunohistochemical analysis for loss of INI1/hSNF5 expression has been shown to effectively distinguish MRTs and extra-renal RTs from other pediatric soft tissue tumors (Hoot et€al. 2004). Additional studies demonstrated that fluorescence in situ hybridization (FISH) for chromosome region 22q11.2 could be used to visualize loss of INI1/hSNF5 and to effectively distinguish AT/RTs from other CNS tumors (Bruch et€al. 2001).
Current Therapies for Rhabdoid Tumors RTs are aggressively treated with multiple modalities including surgery, chemotherapy, and radiation therapy. Recent strategies to intensify treatments for RTs have led to isolated cases of long-term survival (Madigan et€al. 2007). Despite such aggressive treatment strategies, mean survival rates remain extremely low, and it has been reported that the mean survival with surgical intervention alone is only 3 months and increases slightly to 8 months with adjuvant chemotherapy and radiotherapy (Fenton and Foreman 2003). Surgical intervention is used as a primary treatment strategy in order to remove bulk tumor mass. Chemotherapy with or without radiation is used as adjuvant therapy and some patients with advanced disease stage receive hematopoietic stem cell rescue (HSCT) (Madigan et€al. 2007). Those patients most likely to experience long-term survival present with localized disease, allowing for complete surgical resection (Madigan et€al. 2007). Since MRTs were initially misclassified as an aggressive, “rhabdomyosarcomatoid” variants of Wilms tumor, they were originally treated using Wilms tumor protocols (Madigan et€al. 2007). These protocols, involving vincristine, actinomycin-D, doxorubicin, and cyclophosphamide have proven to be ineffective in the treatment of MRT (Gururangan et€al. 1993). Some disease responses have resulted, however, with combinations of cisplatin, doxorubicin, and imidazole carboxamide (DTIC) (Roper et€al. 1981; Kent et€al. 1987). Other instances using chemotherapeutic regimens containing vincristine, dactinomycin, doxorubicin, and cyclophosphamide, however, did little to improve mortality rates (Tomlinson et€al. 2005). Such poor responses to conventional chemotherapies led to more aggressive chemotherapeutic treatments, which are often supplemented with intrathecal chemotherapy in order to reduce cerebral spinal fluid (CSF) seeding of AT/RT cells (Hilden et€al. 2004). Radiation therapy has been used in addition to surgery and chemotherapy in attempts to increase treatment efficacy. Defining the role of radiotherapy in increasing patient survival is difficult as its use is limited in younger children due to subsequent neurological defects. Therefore the population of patients likely to
308
G.V. Kalpana and M.E. Smith
receive significant doses of radiation therapy is older patients, who have a more favorable initial prognosis. Analysis of a patient group with disregard to age has shown that patients receiving 25 Gy or more of radiation appear to have a better disease outcome (Tomlinson et€ al. 2005). A more extensive study provided that patients receiving radiation doses higher than 30 Gy exhibited an increased mean survival rate of 16 months compared to 6 months survival for those treated with lower doses or no radiotherapy (Chen et€al. 2005). With and without radiation therapy, some success has been found for patients 3€years and older with treatments involving surgical resection and chemotherapy using ifosfamide, carboplatin, and etoposide (ICE) (Tekautz et€al. 2005). Successful treatment of three young patients with MRT was achieved using a similar strategy involving surgical resection, radiation, and chemotherapy using ICE alternating with rounds of vincristine, doxorubicin, and cyclophosphamide (VDCy) (Wagner et€ al. 2002; Yamamoto et€ al. 2006). VDCy was originally used to treat Ewing �sarcoma patients (Yamamoto et€al. 2006). Along with total surgical resection and 30.6 Gy radiation therapy, a similar chemotherapeutic regimen of VDCy alternating with ifosfamide and etoposide (IE) was able to successfully treat an infant with extrarenal RT of the chest wall (Hosoi et€al. 2007). Another intensive therapeutic program that was used with success on three AT/RT patients employed rounds of cisplatin, doxorubicin, vincristine, actinomycin, cyclophosphamide, cisplatin, and vincristine and adriamycin, cyclophosphamide, followed by triple intrathecal therapy and radiotherapy (Chen et€al. 2005). Ifosfamide is common to many of these therapeutic regimens. Gururangan et€al. treated patients with ifosfamide alone or in combination with carboplatin and etoposide, bleomycin, cyclophosphamide, doxorubicin, methotrexate, or prednisolone. From these studies, they concluded that transient partial responses can be observed in some patients treated with ifosfamide alone or in combination with carboplatin and etoposide (Gururangan et€al. 1993).
Ongoing Clinical Trials for RTs A number of clinical trials have been completed recently or are underway in attempts to define improved treatment strategies for RTs. An ongoing phase I study has tested the efficacy of intensive chemotherapy and peripheral blood stem cell rescue in infants with malignant brain or spinal cord tumors (including, but not limited to RTs). This study will test the efficacy of treatment with cisplatin, vincristine, cyclophosphamide, and etoposide followed by treatment with filgrastim and peripheral blood stem cell harvest. The study’s second phase involves treatment with carboplatin and thiotepa followed by peripheral blood stem cell transplantation (trial #NCT00003141). A recently completed phase I trial with a more simplistic treatment strategy tested irinotecan in pediatric patients with various malignancies including AT/RTs. This study defined the maximum tolerated dose for irinotecan and showed that four cycles of 30€ mg/m2 irinotecan was able to induce stable Â�disease in one AT/RT patient (Blaney et€al. 2001).
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
309
A number of phase II trials are currently recruiting new patients. One such study is focused specifically on treating AT/RT with intrathecal combination therapy (methotrexate, cytarabine and hydrocortisone), systemic chemotherapy (various alternating combinations of vincristine, carboplatin, cyclophosphamide, etoposide, dactinomycin, dexrazoxane, temozolomide, and filgrastim), and radiotherapy (trial #NCT00084838). Another phase II trial is focused on determining the use of myeloablative chemotherapy with peripheral blood stem cell rescue and its ability to increase survival rates of patients with newly diagnosed high-risk CNS tumors, including, but not limited too AT/RT (trial #NCT00179803). One phase II trial testing different drugs is currently underway to test the ability of antineoplastons A10 and AS2-1 in children with AT/RT (trial #NCT00003469). One phase II clinical trial that has recently been completed was focused on the use of oxaliplatin in pediatric patients with recurrent or refractory AT/RT, MB, or supratentorial PNET (trial #NCT00047177). Oxaliplatin was administered as a 2-h infusion every 21 days (one round). One out of five patients with AT/RT experienced stable disease after 17 rounds. This study provided that oxaliplatin is relatively well tolerated in children, but had limited activity in patients with recurrent CNS tumors that were treated previously with platinum compounds (Fouladi et€al. 2006). Most trials currently focus on the treatment of AT/RT; however, one phase II trial is focused solely on the treatment of high-risk kidney tumors, including MRT. This study, proposed by Children’s Oncology Group, will combine surgical resection, chemotherapy (rounds of VDCy alternating with rounds of cyclophosphamide, etoposide, and carboplatin along with filgrastim, and in some cases, additional rounds of Â�vincristine and irinotecan), and radiotherapy (trial #NCT00335556). Though these clinical trials may define more effective treatment strategies, most are based upon multi-drug combinations involving chemotherapeutics that do not necessarily target factors directly involved in the survival and/or growth of RTs. Despite some success, recurrent or progressive RT often appears, even during active courses of chemotherapy and attempts to treat progressive RT have been reported as unsuccessful with additional therapy (Tekautz et€al. 2005). Even with the use of potent chemotherapeutic, radiotherapeutic, and surgical interventions, the prognosis for RT patients remains poor, especially for children diagnosed with AT/RT before the age of 3 (Biegel et€al. 2002; Packer et€al. 2002). Current estimates suggest a 15% 2-year survival rate for children afflicted with AT/RT (Biegel 2006) and a 19 to 25% 5-year survival rate for those diagnosed with MRT (Yamamoto et€al. 2006). Factors that further predict poor survival are younger age of tumor onset and higher stage of disease at the time of diagnosis (Tomlinson et€al. 2005). One possible reason for the failure of current therapies may be because they involve treatment regimens based on other tumor types that are similar to, but are distinct from RTs. Since RTs have a unique identity, it may be necessary to develop selective therapeutic strategies that are specific to this tumor type. It is likely that understanding the molecular mechanism of rhabdoid tumorigenesis could pave the way to develop molecularly targeted therapies. Identification of INI1/hSNF5 as a critical tumor suppressor deleted in RTs has facilitated such development.
310
G.V. Kalpana and M.E. Smith
The �following sections will summarize what is known about INI1-mediated tumor �suppression in RT cells, specifically in light of developing targeted therapy based on this understanding.
Mechanisms of INI1-Mediated Tumor Suppression INI1/hSNF5, also known as BAF47 and SMARCB1, was first isolated as an interacting protein for HIV-1 integrase via the yeast two-hybrid system, and subsequently, it was demonstrated to be a component of the human chromatin remodeling SWI/ SNF complex (Kalpana et€al. 1994; Wang et€al. 1996). SWI/SNF is an evolutionarily conserved, multi-subunit, high molecular weight (>2€ MDa) complex that remodels chromatin in an ATP-dependent manner. There are two broad classes of chromatin remodeling machines (CRM), including those that remodel chromatin by covalent modifications such as acetylation, deacetylation, and methylation, and those that modify chromatin by disrupting and/or repositioning the nucleosomes in an ATP-dependent manner. Currently, four different classes of ATP-dependent chromatin remodeling complexes have been identified (SWI/SNF, ISWI, Mi-1, and Ino80), each characterized by a unique subunit composition and defined by the presence of distinct ATPases (Fig.€1) (Martens and Winston 2003). SWI/SNF, a prototypical ATP-dependent remodeling complex, consists of at least nine subunits that are conserved among eukaryotes (Fig.€1) (Narlikar et€al. 2001). ATP-dependent CRMs Key ATPase (human)
hSWI/SNF-A (BAF) BAF53 BAF155
OSA1 BAF170 b-Actin
ISWI
Mi-2
Ino80
SWI2/SNF2L BRG1 BRM
ISW1
Mi-2/CHD
Ino80
hSWI/SNF-B (PBAF)
HBrm
BAF180
BAF250 BAF60a
SWI/SNF
BRG1 Or BRM
INI1
BAF60a BAF155 BAF57 BAF170 b-Actin
HBrg1(I)
BAF250
BAF250 BAF53
BRG1
BAF60
BAF53 BAF155
BAF57 BAF170
INI1
BRM
INI1 P220
HBrg1(II)
HDAC1
Sin3A
RbAP48
BAF60 BAF155 BAF57
BAF170
BAF250 BAF155 BRG1
INI1 P220
BAF57 BAF170
BRG1
INI1
HDAC1
HDAC2
RbAP48
Sin3A
P66
RbAP48 HDAC2
BAF53
BAF53
P66
P66
Fig.€1╅ Association of INI1/hSNF5 with multiple SWI/SNF complexes in mammalian cells. CRM chromatin remodeling machine. Four ATP-dependent remodeling complexes have been identified in mammalian cells, which are distinguished by the key ATPase present within each complex. Different types of SWI/SNF complexes have been identified that differ in their subunit composition. However, INI1/hSNF5 is present in each one of these SWI/SNF complexes
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
311
Among these are four core subunits required for chromatin remodeling, including SWI2/SNF2, the key ATPase subunit, and INI1/hSNF5. The Drosophila orthologue of SWI2/SNF2 is termed Brahma, from which the two human paralogues derive their names, hBRM and BRG1 (Brahma Related Gene). There are two functionally distinct classes of SWI/SNF complexes in mammalian cells, hSWI/SNF-A or BAF and hSWI/SNF-B or PBAF. Three additional complexes that contain a mixture of components derived from histone deacetylase complex (HDAC) and SWI/SNF are known as hBRM, hBRG1(I), and hBRG1(II) complexes (Fig.€ 1) (Martens and Winston 2003). The stoichiometry, subnuclear distribution, or exact functions of SWI/SNF complexes are not well defined in mammalian cells. It is important to note that while these complexes differ greatly in subunit composition, INI1/hSNF5 is present in each of the different SWI/SNF complexes, implicating its multiple and varied functions in mammalian cells.
Regulation of Transcription by INI1 and the SWI/SNF Complex Chromatin remodeling complexes are global regulators of cellular transcription. However, the activity of each particular complex is restricted to a subset of genes. The SWI/SNF complex has been shown to effect a subset of approximately 2 to 10% of cellular genes (Martens and Winston 2003). One mechanism by which SWI/SNF regulates only a small subset of genes is by selective targeting to specific gene promoters. SWI/SNF components have no sequence-specific DNA binding ability, and therefore, are recruited to specific promoters via protein–protein interactions with transcription factors that directly bind promoter elements (Fig.€ 2). In this regard, we have found that INI1/hSNF5 interacts with cMYC, a sequence-specific transcription factor, and that the SWI/SNF complex is required for the transactivation function of cMYC (Cheng et€al. 1999). Determining the mechanism by which target specificity is achieved by INI1/hSNF5 and defining the subset of cellular genes affected will likely provide more insight into the mechanism of tumor suppression by INI1/hSNF5. SWI/SNF complexes are well characterized as transcriptional activators. Nevertheless, INI1/hSNF5 and these complexes are not only involved in activation, but also in repression of target genes (Martens and Winston 2003). The activation functions of SWI/SNF result directly from its ability to remodel chromatin at promoter regions. However, the mechanism by which SWI/SNF mediates repression is not completely understood. Recent evidence, including studies from our laboratory, indicates that repression is due to direct recruitment of INI1/hSNF5 and the SWI/ SNF complex to specific promoter regions (Fig.€2). This recruitment is mediated by SWI/SNF interaction with sequence-specific DNA-binding repressors such as E2F, co-repressors such as REST, components of the HDAC1 complex, or mammalian heterochromatin protein (HP1a) (Martens and Winston 2003). It is not clear however, how SWI/SNF complexes selectively activate some promoters while repressing others. One possible explanation involves the existence of distinct SWI/SNF
312
G.V. Kalpana and M.E. Smith
a SWI/SNF BRG1
INI1 TF
Promoter
Activation
b INI1
HDAC1
TSA
TF
Promoter
Repression
Fig.€ 2╅ SWI/SNF complexes can both activate and repress transcription. (a) Absence of a sequence-specific binding subunit requires that the SWI/SNF complex be recruited to specific promoters by interaction of its components (such as INI1/hSNF5) with sequence-specific DNA binding transcription factors (TF). Remodeling or rearrangement of nucleosomes is likely to activate transcription. (b) One mechanism of SWI/SNF-mediated repression is by its association with members of the HDAC1 complex. Trichostatin A (TSA) is an inhibitor of HDAC1 that reverses transcriptional inhibition mediated by HDAC1 complex
complexes, some of which are associated with components of the HDAC1 repressor complex as described previously (Fig.€1). Understanding the function and specificity of each of these complexes in distinct cell types is necessary to fully comprehend the mechanisms by which INI1/hSNF5 mediates tumor suppression in RT precursor cells.
Pathways Affected by Loss of INI1 in Rhabdoid Tumors Based on the association of INI1/hSNF5 with the SWI/SNF complex and the requirement of this complex for varied cellular functions, it can be surmised that INI1/hSNF5 mediates tumor suppression in part by selective transcriptional �regulation of genes important for cell cycle, differentiation, senescence, or apoptosis. A hallmark of cancer cells is the lack of regulatory circuits to maintain normal cell division and homeostasis. Tumors cells are characterized by: (1) self sufficiency in growth signals, (2) insensitivity to growth inhibitory signals, (3) evasion of apoptosis, (4)
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
313
limitless replicative potential, (5) sustained angiogenesis, and (6) potential for tissue invasion and metastasis [Hanahan and Weinberg (2000)]. These properties are often achieved by mutations in oncogenes and/or tumor suppressors; genes that directly or indirectly regulate these processes. Consistent with this hypothesis, it has been established that reintroduction of INI1/hSNF5 into RT cells leads to deregulation of various cellular pathways important for normal cell �division and homeostasis. These studies have been critical to understand the mechanism of tumor suppression by INI1/hSNF5.
Induction of G0/G1 Arrest by INI1/hSNF5 In vitro studies have established that reintroduction of INI1/hSNF5 into RT cells is sufficient to induce G0/G1 cell cycle arrest and flat cell formation, indicative of senescence (Zhang et€al. 2002; Reincke et€al. 2003; Oruetxebarria et€al. 2004; Chai et€ al. 2005). Induction of G0/G1 arrest and senescence by INI1 is correlated to transcriptional repression of cyclin D1 (Zhang et€al. 2002) and activation of p16INK4a and p21CIP (Betz et€al. 2002). Studies within our laboratory demonstrated that INI1/ hSNF5 directly represses cyclin D1 by recruiting the HDAC1 complex to its promoter (Fig.€2b) (Zhang et€al. 2002). Repression of cyclin D1 by INI1/hSNF5 specifically occurred in rhabdoid cells but not in other cell types such as HeLa (Zhang et€al. 2002). Among the three D-type cyclins (cyclin D1–D3), only cyclin D1 was repressed by INI1/hSNF5 in rhabdoid cells, while other D-type cyclins (cyclins D2 and D3) were unaffected (Tsikitis et€al. 2005). Additionally, we demonstrated that co-expression of cyclin D1 from a heterologous promoter along with INI1/hSNF5 is sufficient to overcome INI1-mediated cell cycle arrest in RT cells (Zhang et€al. 2002). These studies indicate that transcriptional regulation of a selective subset of genes in a cell-type specific manner is likely important for INI1-mediated tumor suppression in RT precursor cells and that cyclin D1 could be a critical downstream target of INI1-mediated repression.
Role of INI1 in Mitotic Spindle Check-Point INI1 is also involved in mitotic spindle checkpoint activation through the p16Cyclin D1/CDK4-pRb-E2F pathway (Vries et€al. 2005). It was demonstrated that when reintroduced into rhabdoid cells, INI1/hSNF5 decreased the number of aberrant aneuploid cells and facilitated diploidization. Interestingly, it was found that reintroduction of cancer-associated INI1-mutants, P48S and R127G, failed to generate a diploid cell population and expression of S284L and S289A mutants strongly promoted polyploidization (Vries et€ al. 2005). Gene expression profile analysis indicated that, in addition to activating p16INK4a, INI1 represses a high degree of mitotic genes including PLK1 (see below) (Morozov et€al. 2007).
314
G.V. Kalpana and M.E. Smith
INI1/hSNF5 Induces Interferon Signaling and Markers of Senescence Several cDNA microarray studies to investigate pathways downstream of INI1/ hSNF5 have been carried out. These studies revealed that expression of only a small percentage (approximately 2%) of genes is affected by INI1/hSNF5. Early studies compared gene expression patterns of RTs to other pediatric tumors (Pomeroy et€al. 2002). These studies indicated that RTs have a distinct molecular signature despite their presence in varied anatomical locations (Pomeroy et€al. 2002). Other studies have compared the changes in global expression patterns when INI1/hSNF5 was reintroduced into INI1-null rhabdoid cells (Medjkane et€al. 2004; Vries et€al. 2005; Morozov et€ al. 2007). Others have explored INI1/hSNF5-mediated expression profile changes in normal mouse embryo fibroblasts (Isakoff et€ al. 2005). Genes upregulated by INI1/hSNF5 appear to be anti-proliferative or involved in senescence or differentiation while genes repressed by INI1/hSNF5 are involved in cell cycle progression (Morozov et€ al. 2007). Genes upregulated by INI1/hSNF5 and involved in senescence include MMP1 and PAI-1 (Reincke et€al. 2003; Chai et€al. 2005; Morozov et€al. 2007). An overwhelming majority of mitosis-specific genes were downregulated by INI1/hSNF5 including PLK1, TOP2A, STK6, KIF2C, CENP-F, and Securin (Morozov et€ al. 2007). Our studies indicated that INI1/ hSNF5 induced many interferon-stimulated genes (ISGs) at early time points and that interferon (IFN) signaling through its receptor was required for this induction of ISGs (Morozov et€al. 2007). The microarray results are consistent with the functions of INI1/hSNF5 as a tumor suppressor and provided us with many pathways and genes to investigate as potential therapeutic targets for RT treatment, which will be discussed below.
Rhabdoid Tumor Cell of Origin and the Role of INI1 in Differentiation Reintroduction of INI1/hSNF5 into RT cells has an anti-proliferative effect. Contrarily, many studies have indicated that INI1/hSNF5 is required for the Â�survival of nonRT cells and normal cells based on the following reports. (1) Homozygous deletion of INI1/hSNF5 is embryonic lethal and Ini1−/− mouse fibroblasts undergo cell death in culture (Klochendler-Yeivin et€al. 2000; Guidi et€al. 2001), (2) Conditional deletions of Ini1/Snf5 in neonatal mice result in massive apoptosis and hemorrhage in the liver and other organs (Roberts et€ al. 2002), (3) Knockdown of INI1/hSNF5 in HeLa cells leads to a senescence-like phenotype and cell death (Kato et€al. 2007). These results indicate that INI1/hSNF5 has differential effects on the survival of RT cells versus nonRT or normal cells. One possible reason for this is that regulation of cellular pathways by INI1/hSNF5 is cell-type specific.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
315
Knowing the cell of origin of RTs may provide insight into the differential function and mechanism of INI1/hSNF5-mediated tumor suppression; however, the RT cell of origin remains undefined. Since RTs arise in many anatomical locations, several possible cellular progenitors have been proposed, including neural, neuroectodermal, myogenic, mesenchymal, and epithelial cell types (Ota et€ al. 1993). The hypothesis of a mesenchymal cell of origin was supported by comparative global gene expression profile analysis, which revealed a common mesenchymal origin gene signature among RTs (Pomeroy et€al. 2002). It is unclear however, if this gene signature is an acquired characteristic due to loss of INI1/hSNF5 or if it reflects the original cell lineage. Another hypothesis, based on studies using genetic crosses between TgT121 and Ini1+/− knockout mice suggested that RTs might arise from undefined neural progenitor cells (Chai et€ al. 2007). Interestingly, loss of INI1/ hSNF5 was recently associated with schwannomatosis, a tumor arising from schwann cells, and therefore, ectodermal in origin. These results suggest that loss of INI1 leads to schwannomatosis from schwan cells (Hulsebos et€al. 2007).
INI1’s Involvement in Cellular Differentiation Pathways The highly undifferentiated phenotype of RTs suggests that loss of INI1/hSNF5 impairs differentiation programs within RT precursor cells; an idea supported by various studies. Reintroduction of INI1/hSNF5 into DEV and MON RT cell lines induces differentiation into the neural lineage (Albanese et€al. 2006). Furthermore, INI1/hSNF5 was found to be essential for SWI/SNF-dependent induction of neural differentiation programs as siRNA knock-down significantly impairs the ability of rat PC12 cells to differentiate into the neural lineage upon exposure to neural growth factor (Albanese et€al. 2006). The involvement of INI1/hSNF5 in adipocyte differentiation pathways has also been demonstrated. Reintroduction of INI1/ hSNF5 into MON cells using an inducible expression system led to differentiation toward the adipocyte lineage (Caramel et€al. 2007). siRNA knock-down of INI1/ hSNF5 in murine 3T3-L1 preadipocytes and human mesenchymal stem cells eliminates the ability of these cells to differentiate into the adipocyte lineage (Caramel et€al. 2007). Mechanistic studies showed that INI1/hSNF5 cooperates with C/EBPb and PPARg2 to activate adipocyte-specific gene expression (Caramel et€al. 2007). In addition to its involvement in directing differentiation into neural and adipocyte lineages, INI1/hSNF5 is also essential for hepatocyte differentiation. Inactivation of INI1/hSNF5 in the developing liver blocked the expression of many genes normally upregulated during liver development and severely impaired glycogen storage and epithelial morphogenesis, characteristics of hepatocyte differentiation (Gresh et€al. 2005). These findings indicate that loss of INI1/hSNF5 in RT precursor cells Â�disrupts cell fate, causing the severely undifferentiated phenotype common to RT cells. These results provide useful, but limited information on the role of INI1/ hSNF5 during differentiation and on the possible cell of origin, but the relationship between its role in differentiation and tumor suppression remains unanswered.
316
G.V. Kalpana and M.E. Smith
Therapeutic Targeting of Downstream Effectors of INI1/hSNF5 for Treatment of Rhaboid Tumors Though there is still a large deficit in knowledge regarding the RT cell of origin or the specific mechanism by which loss of INI1/hSNF leads to RT formation, the abundance of information about the pathways affected by INI1/hSNF5 have provided insight that may lead to the development of novel targeted treatment strategies for these tumors. The remaining sections of this review summarize various preclinical studies that explore the development of targeted therapies for RTs with the potential for having greatly improved efficacy compared to current treatments. Development of successful targeted therapies relies on knowledge of specific cellular factors or pathways that are both deregulated and required for the genesis, survival, or progression of RTs. While many factors are deregulated in RTs, only those known to be necessary for the growth of these tumors may serve as potentially effective therapeutic targets. As will be described, Cyclin D1 is a factor with such potential. Extensive research has demonstrated that INI1/hSNF5 suppresses the RT phenotype by repressing cyclin D1 and activating p16INK4a, indicating that targeting the Cyclin/cdk-axis may be a useful strategy for developing a selective therapy for RTs. Additional in€ vitro studies have implicated other pathways downstream of INI1/hSNF5 and supported their possible use as therapeutic targets.
Hypothesis for the Mechanism of Tumor Suppression by INI1 The current observations suggest that the mechanism of tumor suppression by INI1/hSNF5 is complex, involving its ability to regulate multiple cellular pathways and affect expression of and genes. One hypothesis is that the rhabdoid precursor cell is destined to undergo terminal cell cycle exit due to stimuli leading to differentiation, and/or senescence (Fig.€3). INI1/hSNF5 may facilitate the terminal cell cycle exit, differentiation, and/or senescence by: (1) specific repression of genes important for cell cycle progression, (2) activation of cell cycle inhibitors, and (3) activation of genes involved in differentiation or senescence (Fig.€3, top panel). When INI1/hSNF5 is deleted in the precursor cell, its downstream pathways will be deregulated, thus leading to tumorigenesis by conferring unlimited proliferative capacity and by protecting the cells from senescence and apoptosis (Fig.€ 3, bottom panel). It follows that a useful treatment strategy for RTs is to mimic the effect of INI1/hSNF5. This can be done by treatment with small molecular weight inhibitors (drugs) or siRNA constructs that selectively affect pathways downstream of INI1/hSNF5 that are critically deregulated in RTs (Fig.€3, bottom panel). In the sections that follow, we will describe how we have tested this, using preclinical models.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
a
INI1/hSNF5
Rhabdoid precursor Cell
pRB
Cyclin D1/cdk4
p16/Ink4a
Senescence
b
Apoptosis
G1
E2F
S
G2
PLK1 and other Mitotic genes M
X
X
pRB
Cyclin D1/cdk4
Mitotic catastrophe Apoptosis
E2F
PLK1 and other Mitotic genes
?
X
Senescence
c
X
Apoptosis
G1
S
G2
Cyclin D1/cdk4
p16/Ink4a
pRB
E2F
?
Senescence
Apoptosis
M
X
G1 Mitotic catastrophe Apoptosis
Targeting G1 and Mitotic genes
Drugs/siRNA
X
G1
Rhabdoid Cell
INI1/hSNF5
p16/Ink4a
317
G1
S
G2
PLK1 and other Mitotic genes
M
G1 Mitotic catastrophe Apoptosis
Fig.€3â•… Mechanism of INI1/hSNF5-mediated tumor suppression. (a) In a rhabdoid precursor cell (unknown) INI1/hSNF5 regulates transcription of key cell cycle regulatory genes. It activates p16, and represses cyclin D1 and mitotic genes, leading to G0–G1 arrest and mitotic arrest. Activation of p16 is associated with senescence and control of ploidy, and repression of cyclin D1 often inhibits apoptosis. (b) In a rhabdoid cell, loss of INI1/hSNF5 leads to transcriptional deregulation of genes involved in cell cycle – p16 is not activated and cyclin D1 and PLK1 are derepressed. As a result, cells escape cell cycle control and senescence and undergo uncontrolled cell division. (c) Use of drugs or siRNA that can inhibit genes or pathways (Cyclin/cdk and mitotic genes) normally inhibited by INI1/hSNF5 could be useful as molecularly targeted therapies for RTs
Critical Role of Cyclin D1 in Rhabdoid Tumor Formation and Survival As described, INI1/hSNF5 directly and selectively represses transcription of cyclin D1 in a cell cycle-independent manner. Further studies showed that Cyclin D1 is a critical downstream target of INI1/hSNF5 and that its over-expression promotes rhabdoid tumorigenesis. In vitro studies indicated that INI1/hSNF5 transcriptionally represses cyclin D1 by recruitment of the HDAC1 complex, and that INI1/
318
G.V. Kalpana and M.E. Smith
hSNF5-mediated cell cycle arrest is at least partially overcome by co-expression of cyclin D1 from a heterologous promoter (Zhang et€al. 2002). Consistent with INI1/ hSNF5’s ability to repress transcription of cyclin D1, we found that Cyclin D1 is de-repressed/over-expressed in both mouse and human primary RTs (Zhang et€al. 2002; Fujisawa et€al. 2005; Tsikitis et€al. 2005; Donner et€al. 2007). To unequivocally establish the critical role of Cyclin D1 in rhabdoid tumorigenesis, we utilized a genetically engineered mouse model of RTs. As indicated earlier, heterozygous Ini1+/− mice develop RTs at a high frequency, which mimic the etiology and characteristics of human RTs. We generated three cohorts of Ini1+/− mice harboring different cyclin D1 genotypes. We found that Ini1+/− mice with cyclin D1+/+ or +/− genotypes were able to develop RTs at the usual frequency of about 25%, while none of the Ini1+/− mice with a cyclin D1−/− background developed RTs (Tsikitis et€ al. 2005). These striking results established that genetic ablation of cyclin D1 is sufficient to abrogate the genesis of RTs and provided proof that Cyclin D1 is a necessary and critical component for the formation and/or survival of RTs.
Development of Novel Therapeutic Strategies Against RTs by Targeting Cyclin/cdk Axis The exquisite dependence of RTs on Cyclin D1 suggests that targeting Cyclin D1 or its pathway is a practical option for developing a molecularly targeted therapy for RTs. It is important, however, to first demonstrate that Cyclin D1 is not only necessary for the genesis of RTs, but also for the survival of established RTs. This was accomplished by siRNA-mediated knockdown of cyclin D1 in RT cells, which induced G0–G1 arrest and apoptosis (Alarcon-Vargas et€al. 2006). Taken together, the in€vitro and in€vivo data defining the role of Cyclin D1 in RTs provide compelling evidence that Cyclin D1 is required for rhabdoid tumorigenesis and survival. Cyclin D1 is overexpressed in several human tumors including approximately 30% of human breast cancers, as well as pituitary tumors, MBs, head and neck carcinoma, and esophageal squamous cell carcinoma. The oncogenic potential of cyclin D1 is in part due to its ability to facilitate cell cycle progression through G1 phase. Cyclin D1 binds to and activates Cyclin-dependent kinases (cdks) 4 and 6, which in turn phosphorylate retinoblastoma protein (Rb). Phosphorylation of Rb relieves E2F, allowing for activation of E2F-target gene transcription. In addition, overexpressed Cyclin D1 sequesters p21cip1 (cdk inhibitor), effectively activating the Cyclin E-cdk2 holoenzyme and resulting in G1–S transition. Cyclin D1 exhibits additional, cdk-independent functions, which are also important for tumorigenesis (Ewen and Lamb 2004; Fu et€ al. 2004). Independent of interaction with cdks, Cyclin D1 modulates the activity of several transcription factors including estrogen receptor (ER), D1-interacting-myb-like protein 1(DMP1), androgen receptor (AR), and signal transducer and activator of transcription (STAT3). Taken together, these studies indicate that both cdk-dependent and -independent functions of Cyclin D1
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
319
INI1
p
RNAPII Cyclin D1
p16
p Cyclin D1
(pTEFb)
CDK
Fenretinide
CyclinT CDK9
Degradation p16
CyclinD
CyclinB
M
CDC2
CDK4 CDK6 p21
G2
Flavopiridol
Flavopiridol
G1 CyclinA CDC2
S
CyclinE CDK2
Fig.€4╅ Targeting the Cyclin/cdk pathway in rhabdoid tumors. INI1/hSNF5 represses cyclin D1 and activates p16 to control the Cyclin/cdk pathway. Drugs such as Fenretinide and Flavopiridol target Cyclins and cdks as part of their mechanism of action. Fenretinide is known to inhibit transcription of cyclin D1 and facilitate degradation of Cyclin D1. Flavopiridol is a pan-cdk inhibitor, which, in addition to inhibiting cdk activity also inhibits cyclin D1 transcription through inhibition of cdk9. Preclinical studies have demonstrated that these two drugs significantly inhibit RT growth in€vitro and in€vivo with their activity correlated to repression of Cyclin D1
are important for tumorigenesis, and therefore, targeting Cyclin D1 and/or the Cyclin/cdk axis may effectively inhibit RT growth. The following sections will describe preclinical studies that investigated the effects of two different chemotherapeutic agents, 4HPR (N-(4-hydroxyphenyl) retinamide, or Fenretinide) and Flavopiridol, and their ability to inhibit the Cyclin/cdk axis in RTs (Fig.€4).
4HPR as a Therapeutic Agent for Rhabdoid Tumors 4HPR is a synthetic retinoid shown to have a low toxicity profile and potent chemopreventive effects in various preclinical models. It exhibits in€vitro cytotoxicity and
320
G.V. Kalpana and M.E. Smith
suppresses tumor cell growth at low micromolar concentrations (IC50’s ranging from 1 to 10€mM) (Garaventa et€al. 2003). 4HPR is an FDA approved drug that has shown much promise in clinical trials and continues to be under phase II trials for many cancers. In pediatric neuroblastoma patients, 4HPR has induced prolonged disease stabilization in pilot clinical studies (Reynolds et€ al. 2003; Lovat et€ al. 2004, 2005; Reynolds 2004). It has been largely studied as a chemo-preventive agent in animal models of carcinogen-induced epithelial tumors and in patients at risk for breast cancer (Fontana and Rishi 2002; Formelli et€ al. 2003; Hail et€ al. 2006; Zanardi et€ al. 2006). A 15-year follow-up study of phase III trials using 4HPR to prevent second breast cancer has recently been completed. This follow-up indicated that 4HPR significantly reduces the risk of recurring breast cancer in premenopausal women (Bonanni et€al. 2007). 4HPR induces apoptosis in tumor cell lines by various mechanisms including: (1) activation of retinoid receptors RAR b and g, (2) induction of ceramide-dependent cell cytotoxicity, (3) generation of radical oxygen species, (4) increase of nitric oxide synthase (NOS) resulting in increased NO-dependent cell cytotoxicity, and (5) increase of mitochondrial permeability transition (Fontana and Rishi 2002; Reynolds et€al. 2003; Lovat et€al. 2004, 2005; Hail et€al. 2006). 4HPR also induces cell cycle arrest likely by down-modulating the expression or activity of proliferation-related factors such as c-Myc, telomerase, p34/cdc2 and cyclin D1 (Igawa et€al. 1994; Delia et€al. 1995; Bednarek et€al. 1999; Sun et€al. 1999, 2001; Reynolds and Lemons 2001; Soria et€al. 2001, 2003; Bowman et€al. 2002; Hiyama and Hiyama 2002; Christov et€al. 2003; Reynolds et€al. 2003). For example, 4HPR affects transcription and protein stability of Cyclin D1 in a concentration dependent manner (DiPietrantonio et€ al. 1998; Panigone et€ al. 2000; Christine Pratt et€ al. 2003; Dragnev et€al. 2004). Furthermore, it has been demonstrated that overexpression of Cyclin D1 sensitizes breast cancer cells to 4HPR (Pirkmaier et€al. 2003). Although 4HPR has been reported to down-modulate Cyclin D1, it was never used to specifically target the Cyclin D1 pathway in cancer cells. The effect of 4HPR on Cyclin D1 transcription and stability prompted us to carry out a series of in€vitro and in€ vivo experiments testing its efficacy on RTs (Alarcon-Vargas et€ al. 2006). Additionally, the effect of combining 4HPR with 4OH-Tamoxifen (4OH-Tam) was tested because of the reported synergism of these two drugs in€ vitro (Wang et€ al. 2003). Furthermore, suppression of tumor growth by 4OH-Tam has been associated with down-modulation of Cyclin D1 (Jang et€al. 2001). This study demonstrated that 4HPR and 4OH-Tam synergistically inhibit the survival of RT cells, induce G1 arrest, and activate caspase 3/7-mediated apoptosis (Alarcon-Vargas et€al. 2006). These cell culture studies were extended in€vivo using a xenograft model of RTs and it was found that the effect of 4HPR and 4OH-Tam is as well synergistic in inhibiting RT growth in€vivo (Alarcon-Vargas et€al. 2006). Importantly, it was found that inhibition of RT cell growth, both in€ vitro and in€ vivo was associated with reduction of Cyclin D1 levels, and in€vivo, with differentiation of xenograft tumors into bone and cartilage (Alarcon-Vargas et€ al. 2006). These preclinical results support our hypothesis that down-modulation of Cyclin D1 is an effective therapeutic strategy for RTs and demonstrate that 4HPR and 4OH-Tam are effective chemotherapeutic agents for RTs.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
321
Flavopiridol as a Therapeutic Agent for Rhabdoid Tumors Another method of inhibiting the Cyclin D1 pathway is to use drugs that inhibit cdk activity. Flavopiridol is a pan-cdk inhibitor, and one of the first broad-spectrum cdkinhibitors to enter clinical trails. Flavopiridol is a competitive inhibitor of multiple cdks including cdk 1, 2, 4, 6, 7 and 9 (Chao and Price 2001). Its ability to inhibit cdks 1, 2, 4, and 6 directly affects the function of cyclin/cdk complexes resulting in cell cycle arrest. Furthermore, it inhibits cdk9 resulting in inactivation of P-TEFb (Cyclin T1/cdk9 complex), therefore blocking RNA polymerase II-mediated transcription of many genes, including cyclin D1 (Mani et€ al. 2000; Chao and Price 2001; De Azevedo et€al. 2002). Flavopiridol also indirectly inhibits TNF-mediated induction of NF-kB, which regulates Cyclin D1 (Takada and Aggarwal 2004). Furthermore, flavopiridol inhibits the activation of Akt leading to inhibition of forkhead proteins, resulting in decreased translation of Cyclin D1 protein (Schmidt et€al. 2002; Cappellini et€ al. 2003; Liang and Slingerland 2003). Recent studies using pharmacologically based schedules of flavopiridol administration have shown promising activity in clinical trials of chronic lymphocytic leukemia (CLL) (Byrd et€al. 2007). The strong effect of flavopiridol on various aspects of the cyclin/cdk-axis have led us to test its ability to inhibit the growth of RTs (Smith et€al. 2008). In vitro studies demonstrated that nanomolar concentrations of flavopiridol inhibited RT cell growth. Treatment with flavopiridol induced G1 or G2 arrest and apoptosis in a manner dependent on drug concentration and time of exposure. Interestingly, inhibition of rhabdoid cell proliferation at concentrations required for 50% cell killing (IC50) was apparently independent of Cyclin D1 while higher levels of cell killing (IC95) were correlated to down-modulation of Cyclin D1, upregulation of p21, and induction of caspase 3/7 activities. Furthermore, in€vivo studies demonstrated that flavopiridol at 7.5€ mg/kg significantly inhibited the growth of xenografted RTs with effects correlated to upregulation of p21 and downmodulation of Cyclin D1 (Smith et€al. 2008) These results demonstrated that, in addition to 4HPR and 4OH-Tam, flavopiridol is potentially a novel chemotherapeutic agent for RTs and strongly supports our hypothesis that therapeutically targeting the Cyclin–cdk axis is effective in inhibiting RTs.
Other Pathways Amenable for Developing Targeted Therapies for RTs Gene expression profile analysis and other studies have provided clues to many other pathways, apart from the Cyclin/cdk axis, that are amenable for therapeutic targeting in RTs. These include the mitotic spindle checkpoint, IFN, and Akt pathways. Preclinical studies investigating the relevance of these additional pathways are �currently ongoing. We will provide a brief account of what is known about the strategies for targeting these additional pathways.
322
G.V. Kalpana and M.E. Smith
Targeting the Mitotic Spindle Checkpoint The mitotic spindle checkpoint ensures proper segregation of chromatids to daughter nuclei during mitosis and guards against the deleterious effects of aneuploidy. Many cancer cells have a weakened mitotic checkpoint due to overexpression of genes necessary for mitotic progression. Recent investigations have demonstrated that cancer-causing INI1/hSNF5 mutants induce aberrant chromosomal segregation and aneuploidy in RT cells (Vries et€al. 2005). Furthermore, we have demonstrated that a high frequency of genes necessary for mitotic progression, including PLK1, Aurora, TOP2A, etc., are targets of INI1-mediated repression in rhabdoid cells. Immunohistochemical analysis indicated that PLK1 is highly expressed in the cytoplasm of cells in human and mouse RTs, independent of cell cycle stage, suggesting that PLK1 is deregulated in RTs (Morozov et€al. 2007). Repression of PLK1 by INI1/hSNF5 and its overexpression in human and mouse RTs suggests that PLK1 is a potential target for therapeutic intervention for these tumors. PLK1, a crucial mitosis-promoting kinase, is currently an attractive target for developing anti-cancer therapies (McInnes et€al. 2005). Overexpression of this protein is a prognostic marker for various cancers and is associated with aggressive stage and poor survival in cancer patients (Eckerdt et€ al. 2005). Furthermore, it appears that tumor cells are sensitive to down-modulation of PLK1 as they succumb to mitotic catastrophe, while normal diploid cells tolerate its inhibition, possibly due to the presence of redundant mechanisms (Guan et€ al. 2005; Liu et€ al. 2006). Currently, there are several drugs in the developmental stage to target the activity of PLK1. As a first test of targeting PLK1 in RTs, RNA interference analysis (RNAi) was used to inhibit its expression in two rhabdoid cell lines (MON and G401). Transfection of RT cells with siRNA to PLK1 resulted in defective growth and enlarged cells with multiple and fragmented nuclei or aberrantly segregated chromatids (Morozov et€al. 2007). Further characterization indicated that inhibition of PLK1 leads to decreased cell survival, increased Caspase 3/7 activity, and apoptosis (Morozov et€al. 2007). Together, these results suggest that targeting PLK1 is potentially useful in inhibiting RTs. However, these are in€vitro studies and further in€vivo studies are required to facilitate the development of therapies by targeting PLK1 or other mitotic genes in RTs.
Targeting the Interferon Signaling Pathway Reintroduction of INI1 activates the expression of a high degree of IFN signal induced genes ISGs (Morozov et€ al. 2007). Furthermore, addition of exogenous type I or type II IFNs to RT cells also induced ISGs and resulted in G0-G1 arrest and flat cell formation, a phenotype similar to that resulting from reintroduction of INI1/hSNF5 (Morozov et€al. 2007). These results imply that addition of IFNs may
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
323
mimick the effect of INI1/hSNF5. IFN treatment is an approved therapy and has been used as adjuvant therapy for a number of neoplasms including skin cancer, renal cell cancer, leukemia, liver cancer, and myeloma. IFNs have been shown to induce anti-proliferative effects in various cancer cell lines (Decatris et€ al. 2002; Kirkwood 2002; Wall et€al. 2003). Furthermore, induction of ISGs has been associated with inhibition of immortalization and induction of a senescence-like phenotype in certain cancer cells (Xin et€al. 2003; Moiseeva et€al. 2006; Pammer et€al. 2006). Therefore, IFN therapy should be considered, at least as an adjuvant therapy for RTs. Studies on the use of IFNs as RT therapy are preliminary however, and further in€vivo studies are necessary to establish their therapeutic efficacy. Several AT/RT or MRT tumor xenograft models have been included in the panel of pediatric solid tumors in preclinical studies testing efficacy of novel anticancer drugs. Efficacy of several drugs including but not limited to a-VEGF antibody (Soffer et€ al. 2002), HDAC inhibitors (benzamide derivatives and desipeptides) (Graham et€ al. 2006; Jaboin et€ al. 2002), EGFR-tyrosine kinase inhibitors (gefitinib) (Kuwahara et€al. 2004), fungal derivatives (Irofulven) (Leggas et€al. 2002), and mTOR inhibitor (Rapamycin) (Houghton et€al. 2007) have been tested in rhabdoid xenograft tumor models. Some of these drugs have demonstrated partial or complete response and further mechanistic studies are in progress to establish the basis of therapeutic efficacy of these drugs.
Future Directions and Challenges Analysis of the molecular pathways affected by loss of INI1/hSNF5 has provided insights for developing targeted therapies for RTs. The aggressive nature of RTs and their insensitivity to many potent chemotherapy regimens pose a challenge. The aggressiveness of RTs could be a consequence of deregulation of multiple cell cycle and growth control pathways. It is clear that INI1/hSNF5 regulates multiple downstream genes and pathways, which in turn influences its ability to mediate tumor suppression. Many of these genes/pathways repressed by INI1/hSNF5 are de-repressed in RTs and several of them are necessary for the genesis and survival of these tumors (Zhang et€al. 2002; Tsikitis et€al. 2005; Alarcon-Vargas et€al. 2006; Morozov et€al. 2007). Therefore, one can envision that a multi-targeted therapy may be required to effectively inhibit the growth of RTs. It is possible that simultaneously or sequentially targeting multiple INI1/hSNF5-downstream pathways is more effective in inhibiting RTs to ensure long term survival in those affected. While some of the current therapies have demonstrated promise, the mechanism of drug action or the overall effectiveness of these drugs against RTs is not completely understood. In cases of long-term survival, lack of understanding of molecular basis for these treatment strategies makes it hard to predict future outcomes in additional patients treated in the same way. In cases where the outcome is poor, the reasons for treatment failures could be multifold including development of drug resistance, suboptimal drug delivery to the tumor, constraints of the blood brain
324
G.V. Kalpana and M.E. Smith
barrier, or drug related toxicities. None of these parameters have been studied in detail, partly due to the rarity of these tumors and lack of dedicated funding opportunities to study these tumors. Nevertheless, identification of the genetic basis of RTs with the discovery of INI1/hSNF5 as the critical tumor suppressor has provided a wealth of information to develop and investigate molecularly targeted therapies against these tumors. While preclinical studies have indicated promising leads, further clinical studies are needed to formulate these concepts into a therapeutic strategy that is both effective and that confers long lasting survival in children with RTs. Acknowledgmentsâ•… We apologize to all investigators whose work was not included in this article due to space limitation. Work in our laboratory is supported by grants from American Cancer Society, Children’s Tumor foundation, Mark Trauner faculty scholar and Irma T. Hirschl awards.
References Alarcon-Vargas, D., Zhang, Z., Agarwal, B., Challagulla, K., Mani, S., and Kalpana, G.V. 2006. Targeting cyclin D1, a downstream effector of INI1/hSNF5, in rhabdoid tumors. Oncogene 25: 722–734. Albanese, P., Belin, M.F., and Delattre, O. 2006. The tumour suppressor hSNF5/INI1 controls the differentiation potential of malignant rhabdoid cells. Eur J Cancer 42: 2326–2334. Beckwith, J.B., and Palmer, N.F. 1978. Histopathology and prognosis of Wilms tumors: results from the First National Wilms’ Tumor Study. Cancer 41: 1937–1948. Bednarek, A., Shilkaitis, A., Green, A., Lubet, R., Kelloff, G., Christov, K., and Aldaz, C.M. 1999. Suppression of cell proliferation and telomerase activity in 4-(hydroxyphenyl)retinamidetreated mammary tumors. Carcinogenesis 20: 879–883. Betz, B.L., Strobeck, M.W., Reisman, D.N., Knudsen, E.S., and Weissman, B.E. 2002. Re-expression of hSNF5/INI1/BAF47 in pediatric tumor cells leads to G1 arrest associated with induction of p16ink4a and activation of RB. Oncogene 21: 5193–5203. Biegel, J.A. 1999. Cytogenetics and molecular genetics of childhood brain tumors. Neuro Oncol 1: 139–151. Biegel, J.A. 2006. Molecular genetics of atypical teratoid/rhabdoid tumor. Neurosurg Focus 20: E11. Biegel, J.A., Kalpana, G., Knudsen, E.S., Packer, R.J., Roberts, C.W., Thiele, C.J., Weissman, B., and Smith, M. 2002. The role of INI1 and the SWI/SNF complex in the development of rhabdoid tumors: meeting summary from the workshop on childhood atypical teratoid/rhabdoid tumors. Cancer Res 62: 323–328. Biegel, J.A., Zhou, J.Y., Rorke, L.B., Stenstrom, C., Wainwright, L.M., and Fogelgren, B. 1999. Germ-line and acquired mutations of INI1 in atypical teratoid and rhabdoid tumors. Cancer Res 59: 74–79. Blaney, S., Berg, S.L., Pratt, C., Weitman, S., Sullivan, J., Luchtman-Jones, L., and Bernstein, M. 2001. A phase I study of irinotecan in pediatric patients: a pediatric oncology group study. Clin Cancer Res 7: 32–37. Bonanni, B., Lazzeroni, M., and Veronesi, U. 2007. Synthetic retinoid fenretinide in breast cancer chemoprevention. Expert Rev Anticancer Ther 7: 423–432. Bowman, R., Clarke, B., Duhig, E., Larsen, J., and Fong, K. 2002. Re: Effects of N-(4-hydroxyphenyl)retinamide on hTERT expression in the bronchial epithelium of cigarette smokers. J Natl Cancer Inst 94: 949–950; author reply 950–941
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
325
Bruch, L.A., Hill, D.A., Cai, D.X., Levy, B.K., Dehner, L.P., and Perry, A. 2001. A role for fluorescence in situ hybridization detection of chromosome 22q dosage in distinguishing atypical teratoid/rhabdoid tumors from medulloblastoma/central primitive neuroectodermal tumors. Hum Pathol 32: 156–162. Byrd, J.C., Lin, T.S., Dalton, J.T., Wu, D., Phelps, M.A., Fischer, B., Moran, M., Blum, K.A., Rovin, B., Brooker-McEldowney, M., et€al. 2007. Flavopiridol administered using a pharmacologically derived schedule is associated with marked clinical efficacy in refractory, genetically high-risk chronic lymphocytic leukemia. Blood 109: 399–404. Cappellini, A., Tabellini, G., Zweyer, M., Bortul, R., Tazzari, P.L., Billi, A.M., Fala, F., Cocco, L., and Martelli, A.M. 2003. The phosphoinositide 3-kinase/Akt pathway regulates cell cycle progression of HL60 human leukemia cells through cytoplasmic relocalization of the cyclindependent kinase inhibitor p27(Kip1) and control of cyclin D1 expression. Leukemia 17: 2157–2167. Caramel, J., Medjkane, S., Quignon, F., and Delattre, O. 2007. The requirement for SNF5/INI1 in adipocyte differentiation highlights new features of malignant rhabdoid tumors. Oncogene 27(14): 2035–2044. Chai, J., Charboneau, A.L., Betz, B.L., and Weissman, B.E. 2005. Loss of the hSNF5 gene concomitantly inactivates p21CIP/WAF1 and p16INK4a activity associated with replicative senescence in A204 rhabdoid tumor cells. Cancer Res 65: 10192–10198. Chai, J., Lu, X., Godfrey, V., Fletcher, C., Roberts, C.W., Van Dyke, T., and Weissman, B.E. 2007. Tumor-specific cooperation of retinoblastoma protein family and Snf5 inactivation. Cancer Res 67: 3002–3009. Chao, S.H., and Price, D.H. 2001. Flavopiridol inactivates P-TEFb and blocks most RNA polymerase II transcription in€vivo. J Biol Chem 276: 31793–31799. Chen, M.L., McComb, J.G., and Krieger, M.D. 2005. Atypical teratoid/rhabdoid tumors of the central nervous system: management and outcomes. Neurosurg Focus 18: E8. Cheng, S.-W.G., Davies, K.P., Yung, E., Beltran, R.J., Yu, J., and Kalpana, G.V. 1999. c-MYC interacts with INI1/hSNF5 and requires the SWI/SNF complex for transactivation function. Nat Genet 22: 102–105. Christine Pratt, M.A., Niu, M., and White, D. 2003. Differential regulation of protein expression, growth and apoptosis by natural and synthetic retinoids. J Cell Biochem 90: 692–708. Christov, K., Ikui, A., Shilkaitis, A., Green, A., Yao, R., You, M., Grubbs, C., Steele, V., Lubet, R., and Weinstein, I.B. 2003. Cell proliferation, apoptosis, and expression of cyclin D1 and cyclin E as potential biomarkers in tamoxifen-treated mammary tumors. Breast Cancer Res Treat 77: 253–264. de Azevedo, W.F. Jr., Canduri, F., and da Silveira, N.J. 2002. Structural basis for inhibition of cyclin-dependent kinase 9 by flavopiridol. Biochem Biophys Res Commun 293: 566–571. Decatris, M., Santhanam, S., and O’Byrne, K. 2002. Potential of interferon-alpha in solid tumours: part 1. BioDrugs 16: 261–281. Delia, D., Aiello, A., Formelli, F., Fontanella, E., Costa, A., Miyashita, T., Reed, J.C., and Pierotti, M.A. 1995. Regulation of apoptosis induced by the retinoid N-(4-hydroxyphenyl) retinamide and effect of deregulated bcl-2. Blood 85: 359–367. DiPietrantonio, A.M., Hsieh, T.C., Olson, S.C., and Wu, J.M. 1998. Regulation of G1/S transition and induction of apoptosis in HL-60 leukemia cells by fenretinide (4HPR). Int J Cancer 78: 53–61. Donner, L.R., Wainwright, L.M., Zhang, F., and Biegel, J.A. 2007. Mutation of the INI1 gene in composite rhabdoid tumor of the endometrium. Hum Pathol 38: 935–939. Dragnev, K.H., Pitha-Rowe, I., Ma, Y., Petty, W.J., Sekula, D., Murphy, B., Rendi, M., Suh, N., Desai, N.B., Sporn, M.B., et€ al. 2004. Specific chemopreventive agents trigger proteasomal degradation of G1 cyclins: implications for combination therapy. Clin Cancer Res 10: 2570–2577. Eckerdt, F., Yuan, J., and Strebhardt, K. 2005. Polo-like kinases and oncogenesis. Oncogene 24: 267–276. Ewen, M.E., and Lamb, J. 2004. The activities of cyclin D1 that drive tumorigenesis. Trends Mol Med 10: 158–162.
326
G.V. Kalpana and M.E. Smith
Fenton, L.Z., and Foreman, N.K. 2003. Atypical teratoid/rhabdoid tumor of the central nervous system in children: an atypical series and review. Pediatr Radiol 33: 554–558. Fontana, J.A., and Rishi, A.K. 2002. Classical and novel retinoids: their targets in cancer therapy. Leukemia 16: 463–472. Formelli, F., Camerini, T., Cavadini, E., Appierto, V., Villani, M.G., Costa, A., De Palo, G., Di Mauro, M.G., and Veronesi, U. 2003. Fenretinide breast cancer prevention trial: drug and retinol plasma levels in relation to age and disease outcome. Cancer Epidemiol Biomarkers Prev 12: 34–41. Fouladi, M., Blaney, S.M., Poussaint, T.Y., Freeman, B.B., 3rd, McLendon, R., Fuller, C., Adesina, A.M., Hancock, M.L., Danks, M.K., Stewart, C., et€al. 2006. Phase II study of oxaliplatin in children with recurrent or refractory medulloblastoma, supratentorial primitive neuroectodermal tumors, and atypical teratoid rhabdoid tumors: a pediatric brain tumor consortium study. Cancer 107: 2291–2297. Fu, M., Wang, C., Li, Z., Sakamaki, T., and Pestell, R.G. 2004. Minireview: cyclin D1: normal and abnormal functions. Endocrinology 145: 5439–5447. Fujisawa, H., Misaki, K., Takabatake, Y., Hasegawa, M., and Yamashita, J. 2005. Cyclin D1 is overexpressed in atypical teratoid/rhabdoid tumor with hSNF5/INI1 gene inactivation. J Neurooncol 73: 117–124. Garaventa, A., Luksch, R., Lo Piccolo, M.S., Cavadini, E., Montaldo, P.G., Pizzitola, M.R., Boni, L., Ponzoni, M., Decensi, A., De Bernardi, B., et€al. 2003. Phase I trial and pharmacokinetics of fenretinide in children with neuroblastoma. Clin Cancer Res 9: 2032–2039. Graham, C., Tucker, C., Creech, J., Favours, E., Billups, C.A., Liu, T., Fouladi, M., Freeman, B.B., 3rd, Stewart, C.F., and Houghton, P.J. 2006. Evaluation of the antitumor efficacy, pharmacokinetics, and pharmacodynamics of the histone deacetylase inhibitor depsipeptide in childhood cancer models in€vivo. Clin Cancer Res 12: 223–234. Gresh, L., Bourachot, B., Reimann, A., Guigas, B., Fiette, L., Garbay, S., Muchardt, C., Hue, L., Pontoglio, M., Yaniv, M., et€al. 2005. The SWI/SNF chromatin-remodeling complex subunit SNF5 is essential for hepatocyte differentiation. EMBO J 24: 3313–3324. Guan, R., Tapang, P., Leverson, J.D., Albert, D., Giranda, V.L., and Luo, Y. 2005. Small interfering RNA-mediated polo-like kinase 1 depletion preferentially reduces the survival of p53defective, oncogenic transformed cells and inhibits tumor growth in animals. Cancer Res 65: 2698–2704. Guidi, C.J., Sands, A.T., Zambrowicz, B.P., Turner, T.K., Demers, D.A., Webster, W., Smith, T.W., Imbalzano, A.N., and Jones, S.N. 2001. Disruption of Ini1 leads to peri-implantation lethality and tumorigenesis in mice. Mol Cell Biol 21: 3598–3603. Gururangan, S., Bowman, L.C., Parham, D.M., Wilimas, J.A., Rao, B., Pratt, C.B., and Douglass, E.C. 1993. Primary extracranial rhabdoid tumors. Clinicopathologic features and response to ifosfamide. Cancer 71: 2653–2659. Hail, N., Jr., Kim, H.J., and Lotan, R. 2006. Mechanisms of fenretinide-induced apoptosis. Apoptosis 11: 1677–1694. Hanahan, D., and Weinberg, R.A. 2000. The hallmarks of cancer. Cell 100: 57–70. Hauser, P., Slowik, F., Bognar, L., Babosa, M., Fazekas, I., and Schuler, D. 2001. Atypical teratoid/rhabdoid tumor or medulloblastoma? Med Pediatr Oncol 36: 644–648. Hilden, J.M., Meerbaum, S., Burger, P., Finlay, J., Janss, A., Scheithauer, B.W., Walter, A.W., Rorke, L.B., and Biegel, J.A. 2004. Central nervous system atypical teratoid/rhabdoid tumor: results of therapy in children enrolled in a registry. J Clin Oncol 22: 2877–2884. Hiyama, K., and Hiyama, E. 2002. [Telomere and telomerase in lung cancer]. Nippon Rinsho 60 Suppl 5: 737–742. Hoot, A.C., Russo, P., Judkins, A.R., Perlman, E.J., and Biegel, J.A. 2004. Immunohistochemical analysis of hSNF5/INI1 distinguishes renal and extra-renal malignant rhabdoid tumors from other pediatric soft tissue tumors. Am J Surg Pathol 28: 1485–1491. Hosoi, H., Iehara, T., Tsuchiya, K., Misawa, A., Miyaji, M., Yagyu, S., Koizumi, M., Nishimura, T., Tokiwa, K., Iwai, N., et€al. 2007. Continuous remission in an infant with chest wall malignant rhabdoid tumor after relapse. J Pediatr Surg 42: E9–E12.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
327
Houghton, P.J., Morton, C.L., Kolb, E.A., Gorlick, R., Lock, R., Carol, H., Reynolds, C.P., Maris, J.M., Keir, S.T., Billups, C.A., et€ al. 2007. Initial testing (stage 1) of the mTOR inhibitor rapamycin by the pediatric preclinical testing program. Pediatr Blood Cancer 50(4): 799–805. Hulsebos, T.J., Plomp, A.S., Wolterman, R.A., Robanus-Maandag, E.C., Baas, F., and Wesseling, P. 2007. Germline mutation of INI1/SMARCB1 in familial schwannomatosis. Am J Hum Genet 80: 805–810. Igawa, M., Tanabe, T., Chodak, G.W., and Rukstalis, D.B. 1994. N-(4-hydroxyphenyl) retinamide induces cell cycle specific growth inhibition in PC3 cells. Prostate 24: 299–305. Isakoff, M.S., Sansam, C.G., Tamayo, P., Subramanian, A., Evans, J.A., Fillmore, C.M., Wang, X., Biegel, J.A., Pomeroy, S.L., Mesirov, J.P., et€al. 2005. Inactivation of the Snf5 tumor suppressor stimulates cell cycle progression and cooperates with p53 loss in oncogenic transformation. Proc Natl Acad Sci USA 102: 17745–17750. Jaboin, J., Wild, J., Hamidi, H., Khanna, C., Kim, C.J., Robey, R., Bates, S.E., and Thiele, C.J. 2002. MS-27-275, an inhibitor of histone deacetylase, has marked in€vitro and in€vivo antitumor activity against pediatric solid tumors. Cancer Res 62: 6108–6115. Jang, T.J., Park, J.H., Cho, M.Y., and Kim, J.R. 2001. Chemically induced rat mammary tumor treated with tamoxifen showed decreased expression of cyclin D1, cyclin E, and p21(Cip1). Cancer Lett 170: 109–116. Kalpana, G.V., Marmon, S., Wang, W., Crabtree, G.R., and Goff, S.P. 1994. Binding and stimulation of HIV-1 integrase by a human homolog of yeast transcription factor SNF5 [see comments]. Science 266: 2002–2006. Kato, H., Honma, R., Sanda, T., Fujiwara, T., Ito, E., Yanagisawa, Y., Imai, J., Okamoto, T., and Watanabe, S. 2007. Knock down of hSNF5/Ini1 causes cell cycle arrest and apoptosis in a p53-dependent manner. Biochem Biophys Res Commun 361: 580–585. Kent, A.L., Mahoney, D.H., Jr., Gresik, M.V., Steuber, C.P., and Fernbach, D.J. 1987. Malignant rhabdoid tumor of the extremity. Cancer 60: 1056–1059. Kirkwood, J. 2002. Cancer immunotherapy: the interferon-alpha experience. Semin Oncol 29: 18–26. Klochendler-Yeivin, A., Fiette, L., Barra, J., Muchardt, C., Babinet, C., and Yaniv, M. 2000. The murine SNF5/INI1 chromatin remodeling factor is essential for embryonic development and tumor suppression. EMBO Rep 1: 500–506. Kuwahara, Y., Hosoi, H., Osone, S., Kita, M., Iehara, T., Kuroda, H., and Sugimoto, T. 2004. Antitumor activity of gefitinib in malignant rhabdoid tumor cells in€ vitro and in€ vivo. Clin Cancer Res 10: 5940–5948. Leggas, M., Stewart, C.F., Woo, M.H., Fouladi, M., Cheshire, P.J., Peterson, J.K., Friedman, H.S., Billups, C., and Houghton, P.J. 2002. Relation between irofulven (MGI-114) systemic exposure and tumor response in human solid tumor xenografts. Clin Cancer Res 8: 3000–3007. Liang, J., and Slingerland, J.M. 2003. Multiple roles of the PI3K/PKB (Akt) pathway in cell cycle progression. Cell Cycle 2: 339–345. Liu, X., Lei, M., and Erikson, R.L. 2006. Normal cells, but not cancer cells, survive severe Plk1 depletion. Mol Cell Biol 26: 2093–2108. Lovat, P.E., Corazzari, M., Di Sano, F., Piacentini, M., and Redfern, C.P. 2005. The role of gangliosides in fenretinide-induced apoptosis of neuroblastoma. Cancer Lett 228: 105–110. Lovat, P.E., Corazzari, M., Goranov, B., Piacentini, M., and Redfern, C.P. 2004. Molecular mechanisms of fenretinide-induced apoptosis of neuroblastoma cells. Ann NY Acad Sci 1028: 81–89. Madigan, C.E., Armenian, S.H., Malogolowkin, M.H., and Mascarenhas, L. 2007. Extracranial malignant rhabdoid tumors in childhood: the Childrens Hospital Los Angeles experience. Cancer 110: 2061–2066. Mani, S., Wang, C., Wu, K., Francis, R., and Pestell, R. 2000. Cyclin-dependent kinase inhibitors: novel anticancer agents. Expert Opin Investig Drugs 9: 1849–1870. Martens, J.A., and Winston, F. 2003. Recent advances in understanding chromatin remodeling by Swi/Snf complexes. Curr Opin Genet Dev 13: 136–142. McInnes, C., Mezna, M., and Fischer, P.M. 2005. Progress in the discovery of polo-like kinase inhibitors. Curr Top Med Chem 5: 181–197.
328
G.V. Kalpana and M.E. Smith
Medjkane, S., Novikov, E., Versteege, I., and Delattre, O. 2004. The tumor suppressor hSNF5/ INI1 modulates cell growth and actin cytoskeleton organization. Cancer Res 64: 3406–3413. Mitchell, C., Jones, P.M., Kelsey, A., Vujanic, G.M., Marsden, B., Shannon, R., Gornall, P., Owens, C., Taylor, R., Imeson, J., et€al. 2000. The treatment of Wilms’ tumour: results of the United Kingdom Children’s cancer study group (UKCCSG) second Wilms’ tumour study. Br J Cancer 83: 602–608. Moiseeva, O., Mallette, F.A., Mukhopadhyay, U.K., Moores, A., and Ferbeyre, G. 2006. DNA damage signaling and p53-dependent senescence after prolonged beta-interferon stimulation. Mol Biol Cell 17: 1583–1592. Morozov, A., Lee, S.J., Zhang, Z.K., Cimica, V., Zagzag, D., and Kalpana, G.V. 2007. INI1 induces interferon signaling and spindle checkpoint in rhabdoid tumors. Clin Cancer Res 13: 4721–4730. Narlikar, G.J., Phelan, M.L., and Kingston, R.E. 2001. Generation and interconversion of multiple distinct nucleosomal states as a mechanism for catalyzing chromatin fluidity. Mol Cell 8: 1219–1230. Oruetxebarria, I., Venturini, F., Kekarainen, T., Houweling, A., Zuijderduijn, L.M., Mohd-Sarip, A., Vries, R.G., Hoeben, R.C., and Verrijzer, C.P. 2004. P16INK4a is required for hSNF5 chromatin remodeler-induced cellular senescence in malignant rhabdoid tumor cells. J Biol Chem 279: 3807–3816. Ota, S., Crabbe, D.C., Tran, T.N., Triche, T.J., and Shimada, H. 1993. Malignant rhabdoid tumor. A study with two established cell lines. Cancer 71: 2862–2872. Packer, R.J., Biegel, J.A., Blaney, S., Finlay, J., Geyer, J.R., Heideman, R., Hilden, J., Janss, A.J., Kun, L., Vezina, G., et€al. 2002. Atypical teratoid/rhabdoid tumor of the central nervous system: report on workshop. J Pediatr Hematol Oncol 24: 337–342. Pammer, J., Reinisch, C., Birner, P., Pogoda, K., Sturzl, M., and Tschachler, E. 2006. Interferonalpha prevents apoptosis of endothelial cells after short-term exposure but induces replicative senescence after continuous stimulation. Lab Invest 86: 997–1007. Panigone, S., Debernardi, S., Taya, Y., Fontanella, E., Airoldi, R., and Delia, D. 2000. pRb and Cdk regulation by N-(4-hydroxyphenyl)retinamide. Oncogene 19: 4035–4041. Pirkmaier, A., Yuen, K., Hendley, J., O’Connell, M.J., and Germain, D. 2003. Cyclin d1 overexpression sensitizes breast cancer cells to fenretinide. Clin Cancer Res 9: 1877–1884. Pomeroy, S.L., Tamayo, P., Gaasenbeek, M., Sturla, L.M., Angelo, M., McLaughlin, M.E., Kim, J.Y., Goumnerova, L.C., Black, P.M., Lau, C., et€al. 2002. Prediction of central nervous system embryonal tumour outcome based on gene expression. Nature 415: 436–442. Reddy, A.T. 2005. Atypical teratoid/rhabdoid tumors of the central nervous system. J Neurooncol 75: 309–313. Reincke, B.S., Rosson, G.B., Oswald, B.W., and Wright, C.F. 2003. INI1 expression induces cell cycle arrest and markers of senescence in malignant rhabdoid tumor cells. J Cell Physiol 194: 303–313. Reynolds, C.P. 2004. Detection and treatment of minimal residual disease in high-risk neuroblastoma. Pediatr Transplant 8 Suppl 5: 56–66. Reynolds, C.P., and Lemons, R.S. 2001. Retinoid therapy of childhood cancer. Hematol Oncol Clin North Am 15: 867–910. Reynolds, C.P., Matthay, K.K., Villablanca, J.G., and Maurer, B.J. 2003. Retinoid therapy of highrisk neuroblastoma. Cancer Lett 197: 185–192. Roberts, C.W., Leroux, M.M., Fleming, M.D., and Orkin, S.H. 2002. Highly penetrant, rapid tumorigenesis through conditional inversion of the tumor suppressor gene Snf5. Cancer Cell 2: 415–425. Roper, M., Parmley, R.T., Crist, W.M., Kelly, D.R., Hyland, C.H., and Salter, M. 1981. Rhabdoid tumor of the kidney: complete remission induced by cis-platinum and adriamycin. Med Pediatr Oncol 9: 175–180. Schmidt, M., Fernandez de Mattos, S., van der Horst, A., Klompmaker, R., Kops, G.J., Lam, E.W., Burgering, B.M., and Medema, R.H. 2002. Cell cycle inhibition by FoxO forkhead transcription factors involves downregulation of cyclin D. Mol Cell Biol 22: 7842–7852.
Development of Targeted Therapies for Rhabdoid Tumors Based on the Functions
329
Sevenet, N., Lellouch-Tubiana, A., Schofield, D., Hoang-Xuan, K., Gessler, M., Birnbaum, D., Jeanpierre, C., Jouvet, A., and Delattre, O. 1999. Spectrum of hSNF5/INI1 somatic mutations in human cancer and genotype–phenotype correlations. Hum Mol Genet 8: 2359–2368. Smith, M.E., Cimica, V., Chinni, S., Challagulla, K., Mani, S., and Kalpana, G.V. 2008. Rhabdoid tumor growth is inhibited by flavopiridol. Clin Cancer Res 14(2): 523–532. Soffer, S.Z., Kim, E., Huang, J., McCrudden, K., Yokoi, A., Moore, J.T., Manley, C., O’Toole, K., Middlesworth, W., Stolar, C., et€ al. 2002. Resistance of a VEGF-producing tumor to antiVEGF antibody: unimpeded growth of human rhabdoid tumor xenografts. J Pediatr Surg 37: 528–532. Soria, J.C., Moon, C., Wang, L., Hittelman, W.N., Jang, S.J., Sun, S.Y., Lee, J.J., Liu, D., Kurie, J.M., Morice, R.C., et€al. 2001. Effects of N-(4-hydroxyphenyl)retinamide on hTERT expression in the bronchial epithelium of cigarette smokers. J Natl Cancer Inst 93: 1257–1263. Soria, J.C., Xu, X., Liu, D.D., Lee, J.J., Kurie, J., Morice, R.C., Khuri, F., Mao, L., Hong, W.K., and Lotan, R. 2003. Retinoic acid receptor beta and telomerase catalytic subunit expression in bronchial epithelium of heavy smokers. J Natl Cancer Inst 95: 165–168. Strother, D. 2005. Atypical teratoid rhabdoid tumors of childhood: diagnosis, treatment and challenges. Expert Rev Anticancer Ther 5: 907–915. Sun, S.Y., Yue, P., Kelloff, G.J., Steele, V.E., Lippman, S.M., Hong, W.K., and Lotan, R. 2001. Identification of retinamides that are more potent than N-(4-hydroxyphenyl)retinamide in inhibiting growth and inducing apoptosis of human head and neck and lung cancer cells. Cancer Epidemiol Biomarkers Prev 10: 595–601. Sun, S.Y., Yue, P., and Lotan, R. 1999. Induction of apoptosis by N-(4-hydroxyphenyl)retinamide and its association with reactive oxygen species, nuclear retinoic acid receptors, and apoptosisrelated genes in human prostate carcinoma cells. Mol Pharmacol 55: 403–410. Takada, Y., and Aggarwal, B.B. 2004. Flavopiridol inhibits NF-kappaB activation induced by various carcinogens and inflammatory agents through inhibition of IkappaBalpha kinase and p65 phosphorylation: abrogation of cyclin D1, cyclooxygenase-2, and matrix metalloprotease-9. J Biol Chem 279: 4750–4759. Tekautz, T.M., Fuller, C.E., Blaney, S., Fouladi, M., Broniscer, A., Merchant, T.E., Krasin, M., Dalton, J., Hale, G., Kun, L.E., et€ al. 2005. Atypical teratoid/rhabdoid tumors (ATRT): improved survival in children 3 years of age and older with radiation therapy and high-dose alkylator-based chemotherapy. J Clin Oncol 23: 1491–1499. Tomlinson, G.E., Breslow, N.E., Dome, J., Guthrie, K.A., Norkool, P., Li, S., Thomas, P.R., Perlman, E., Beckwith, J.B., D’Angio, G.J., et€ al. 2005. Rhabdoid tumor of the kidney in the National Wilms’ Tumor Study: age at diagnosis as a prognostic factor. J Clin Oncol 23: 7641–7645. Tsikitis, M., Zhang, Z., Edelman, W., Zagzag, D., and Kalpana, G.V. 2005. Genetic ablation of cyclin D1 abrogates genesis of rhabdoid tumors resulting from Ini1 loss. Proc Natl Acad Sci USA 102: 12129–12134. Versteege, I., Sevenet, N., Lange, J., Rousseau-Merck, M.F., Ambros, P., Handgretinger, R., Aurias, A., and Delattre, O. 1998. Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer. Nature 394: 203–206. Vries, R.G., Bezrookove, V., Zuijderduijn, L.M., Kia, S.K., Houweling, A., Oruetxebarria, I., Raap, A.K., and Verrijzer, C.P. 2005. Cancer-associated mutations in chromatin remodeler hSNF5 promote chromosomal instability by compromising the mitotic checkpoint. Genes Dev 19: 665–670. Wagner, L., Hill, D.A., Fuller, C., Pedrosa, M., Bhakta, M., Perry, A., and Dome, J.S. 2002. Treatment of metastatic rhabdoid tumor of the kidney. J Pediatr Hematol Oncol 24: 385–388. Wall, L., Burke, F., Smyth, J.F., and Balkwill, F. 2003. The anti-proliferative activity of interferongamma on ovarian cancer: in€vitro and in€vivo. Gynecol Oncol 88: S149–S151. Wang, H., Charles, A.G., Frankel, A.J., and Cabot, M.C. 2003. Increasing intracellular ceramide: an approach that enhances the cytotoxic response in prostate cancer cells. Urology 61: 1047–1052. Wang, W., Cote, J., Xue, Y., Zhou, S., Khavari, P.A., Biggar, S.R., Muchardt, C., Kalpana, G.V., Goff, S.P., Yaniv, M., et€al. 1996. Purification and biochemical heterogeneity of the mammalian SWI–SNF complex. EMBO J 15: 5370–5382.
330
G.V. Kalpana and M.E. Smith
Xin, H., Curry, J., Johnstone, R.W., Nickoloff, B.J., and Choubey, D. 2003. Role of IFI 16, a member of the interferon-inducible p200-protein family, in prostate epithelial cellular senescence. Oncogene 22: 4831–4840. Yamamoto, M., Suzuki, N., Hatakeyama, N., Mizue, N., Hori, T., Kuroiwa, Y., Hareyama, M., Oda, T., Kudoh, T., Nui, A., et€ al. 2006. Treatment of stage IV malignant rhabdoid tumor of the kidney (MRTK) with ICE and VDCy: a case report. J Pediatr Hematol Oncol 28: 286–289. Zanardi, S., Serrano, D., Argusti, A., Barile, M., Puntoni, M., and Decensi, A. 2006. Clinical trials with retinoids for breast cancer chemoprevention. Endocr Relat Cancer 13: 51–68. Zhang, Z.K., Davies, K.P., Allen, J., Zhu, L., Pestell, R.G., Zagzag, D., and Kalpana, G.V. 2002. Cell cycle arrest and repression of cyclin D1 transcription by INI1/hSNF5. Mol Cell Biol 22: 5975–5988.
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors Brigitte Widemann
Introduction Neurofibromatosis Type 1 (NF1), previously referred to as von Recklinghausen disease, is a relatively common (1:2,500 to 1:3,000) autosomal dominant, progressive tumor predisposition syndrome characterized by manifestations in many organ systems including neurocutaneous findings and the propensity to develop tumors of the peripheral and central nervous system (Friedman 2002; Korf 2002; Ferner 2007). The natural history of NF1 is poorly understood, and for most NF1 related tumor manifestations the only standard treatment option is surgery (Korf 2001; Ferner et€ al. 2007). Increasing knowledge of molecular and biologic pathways implied in the development of NF1 related tumors has resulted in the development of treatment trials with targeted agents (Packer et€al. 2002). This chapter reviews NF1 related tumor manifestations, pathways implied in tumor development and progression, ongoing, and planned clinical trials for NF1 related tumors. Differences in the development of agents for NF1 related tumors and refractory cancers and resulting challenges toward the development of safe and effective therapies for NF1 related tumors are highlighted.
Genetics and Diagnosis of NF1 Neurofibromatosis type 1 is caused by a mutation in the NF1 tumor suppressor gene on chromosome 17q11.2 – 350€kb, 60 exons – (Friedman 2002; Ferner 2007). The gene product neurofibromin (2,818 amino acids) contains a domain with significant homology to RAS GTPase-activating proteins (GAP) and thus regulates
B. Widemannâ•›(*) Pharmacology and Experimental Therapeutics Section, National Cancer Institute, Pediatric Oncology Branch, 10 Center Drive, 10-CRC, Room 1-5750, MSC 1101, Bethesda, MD 20892, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_16, © Springer Science+Business Media, LLC 2010
331
332
B. Widemann
RAS activity. RAS is a small GTPase and plays a central role in survival, proliferation, and differentiation by transducing responses to growth stimuli initiated in the cell surface to intracellular signaling molecules (Satoh and Kaziro 1992). RAS functions as a molecular switch in which it toggles between an inactive (GDP-bound) and active (GTP-bound) state. Neurofibromin accelerates RAS-GTP hydrolysis and thus functions as a potent negative regulator of RAS (Fig.€1). Lack of functional neurofibromin in NF1 therefore leads to dysregulated RAS and tumorigenesis (Cichowski and Jacks 2001). Neurofibromatosis type 1 has 100% penetrance, but features variable expressivity. A wide range of mutations have been described, and while mutation analysis of the NF1 gene allows identification of 95% of mutations to date, only two phenotype
Fig.€1╅ Schema of the NF1 pathway. The NF1 gene product neurofibromin functions as GTPase activating protein and thus facilitates turning RAS from the active GTP bound form to the inactive GDP bound form. Neurofibromin is dysfunctional in NF1 and thus results in increased activity of RAS
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
333
genotype correlations have been described (Upadhyaya et€al. 1998; Messiaen et€al. 2000; Upadhyaya et€al. 2007). The diagnosis of NF1 is thus based on clinical criteria including café-au-lait macules, axillary and inguinal freckling, presence of NF1 related tumors, a family history of NF1, and can typically be made by 6 years of age (NIH et€al. 1988). NF1 is a progressive disorder, and manifestations in essentially every organ (skin, bone, cardio-vascular system, cognitive system) develop throughout life, which explains the need for a multidisciplinary approach in the care for NF1.
NF1 Related Tumor Manifestations Patients with NF1 have an increased risk of developing tumors of the central and peripheral nervous system including plexiform neurofibromas (PN) (25%), dermal neurofibromas (>99%), optic pathway gliomas (OPG) (15%), brain tumors (2 to 3%), malignant peripheral nerve sheath tumors (MPNST) (8 to 13%), juvenile myelomonocytic leukemia -JMML-, pheochromocytomas (2%), rhabdomyosarcomas (1.5 to 6%), glomus tumors, and benign hamartomas of the iris called Lisch nodules (>95%) (North 1997; Korf 2000, 2002; Evans et€al. 2002). Highlighted below are tumor manifestations for which targeted clinical trials are ongoing or becoming available.
Dermal Neurofibromas The hallmark feature of NF1, typically develop in adolescence with continued development of new tumors throughout adulthood and during pregnancy (Huson et€al. 1988; Huson et€al. 1989). Surgical excision is sometimes required, but neurofibromas may grow back after surgery. No medication is available to control the growth of these tumors. While dermal neurofibromas rarely, if ever, progress to malignancy they impose a substantial cosmetic burden, which is exacerbated by the unpredictability of their growth and ultimate number.
Plexiform Neurofibromas These are benign nerve sheath tumors that grow along the length of nerves and involve multiple branches of a nerve (Korf 1999). These tumors are usually diagnosed early in life, but may develop throughout life. Morbidity from PN includes substantial disfigurement, compression of vital structures, progressive neurologic deficit, and often unremitting pain (Fig.€2). The only standard treatment for PN is surgery. However, given the location, infiltrative nature, high vascularity, and size of PN, complete surgical removal is usually not feasible, and up to 44% of tumors progress after the first surgery, most commonly in patients younger than 10 years
334
B. Widemann
Fig.€2╅ Progressive disfiguring plexiform neurofibroma in a young girl with NF1 demonstrating the need to develop effective medical interventions for young children
of age with head and neck tumors (Needle et€al. 1997). There is no known effective medical treatment for patients with PN.
Malignant Peripheral Nerve Sheath Tumors Also called neurogenic sarcomas, malignant schwannomas, and neurofibrosarcomas, are soft tissue sarcomas which arise from a peripheral nerve or show nerve sheath differentiation and are associated with a high risk of local recurrence and hematogenous metastasis (Ferner and Gutmann 2002). They account for 10% of all soft tissue sarcomas, and half of these malignancies arise in patients with NF1 with a lifetime risk in NF1 of 8 to 13% (Evans et€al. 2002). Early diagnosis of MPNSTs is crucial, as only complete surgical resection has been shown to be curative. However, the diagnosis of MPNSTs in NF1 is difficult to establish because clinical indicators of malignancy (mass and pain) may also be the features of preexisting benign PN from which most MPNSTs arise. For NF1 associated MPNSTs, younger age at diagnosis and decreased survival have been described (Evans et€al. 2002; Ferner and Gutmann 2002; Carli et€ al. 2005), and the largest retrospective analysis describes a lower response rate to chemotherapy in individuals with NF1 versus sporadic MPNST (Carli et€ al. 2005). However, the role of chemotherapy in MPNST has not been assessed prospectively, and the response to standard chemotherapy agents used to treat pediatric and adult sarcomas is unknown and the objective of an ongoing trial of standard a sarcoma chemotherapy agents for high-grade unresectable MPNSTs.
Optic Pathway Gliomas These tumors develop in 15% of children with NF1 (King et€al. 2003) and typically manifest clinically by an age of 6 years (Listernick et€al. 1994). Only approximately 50% of children with NF1 related OPG will develop related signs or symptoms.
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
335
Current recommendations are to treat NF1 associated OPG only if there is clear evidence of radiographic or ophthalmological (visual acuity, fields, color vision) progression (Listernick et€al. 2007). Chemotherapy with vincristine and cisplatin is considered standard treatment for children with progressive disease, and targeted trials for refractory tumors are in development. Radiotherapy should be avoided because of the risk of malignancy and vascular problems, including moyamoya syndrome.
Juvenile Myelomonocytic Leukemia This is a rare myeloproliferative disorder of early childhood. Activating RAS mutations have been reported in 18 to 25% of children with JMML, 10 to 14% of children with JMML have a clinical diagnosis of NF1, and NF1 mutations are found in approximately 30% of patients with JMML. JMML cells are hypersensitive to GM-CSF, and GM-CSF stimulation is associated with increased levels of RASGTP in hematopoietic cell lines (Emanuel 2004). Chemotherapy followed by hematopoietic stem cell transplantation is the only treatment resulting in extended survival, but relapse rates are high ranging from 35 to 55% (Emanuel 2004).
Molecular Features of Tumorigenesis RAS Pathway The NF1 gene encodes neurofibromin, a GTPase-activating protein (GAP) for members of the p21RAS family, which negatively regulates RAS output by accelerating the conversion of active RAS-GTP to inactive RAS-GDP (DeClue et€ al. 1991; DeClue et€ al. 1992; Cichowski and Jacks 2001). Analysis of tumors from patients with NF1 has shown biochemical evidence of hyperactive RAS as well as frequent loss of the normal NF1 allele, consistent with its role as a tumor suppressor gene (Kluwe et€al. 1999a; Kluwe et€al. 1999b; Weiss et€al. 1999; Serra et€al. 2000; Sherman et€ al. 2000; Perry et€ al. 2002). RAS has three major isoforms (K-RAS, N-RAS, and H-RAS). Studies in NF−/− astrocytes demonstrated that activation of only K-RAS and not H-RAS accounted for the proliferative advantage in these cells, and that this could be corrected by a dominant inhibitory K-RAS (Dasgupta et€ al. 2005a). Similarly, K-RAS was found to modulate the RAF/ mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase kinase (ERK) pathway and phosphatidylinositol 3¢-kinase (PI3K) pathway in wildtype NF+/− mast cells and genetic deletion of K-RAS reduced this gain of function (Khalaf et€al. 2007). Introduction of the functional NF1-GAP-related domain could partially reverse the transformation of human NF1 tumor-derived Schwann cells, but was not sufficient to decrease the in€ vitro angiogenic potential of these cells
336
B. Widemann
(Hiatt et€al. 2001; Thomas et€al. 2006). While NF1 likely serves functions outside of its role as RAS-GAP; for example, through regulation of cyclic adenosine monophosphate (Hegedus et€al. 2007), its function as activator of RAS has most clearly been linked to tumorigenesis.
Cells of Tumor Origin and Tumor Environment Nearly all individuals with NF1 develop dermal neurofibromas, and some PN, which are composed of axons, Schwann cells, fibroblasts, perineurial cells, endothelial cells, and mast cells. Schwann cells are believed to be the primary pathogenic cell in neurofibromas because they show biallelic mutation of NF1 (Cichowski and Jacks 2001). Recent data support that tumors do not arise from neural crest stem cells, but from proliferation of fully differentiated nonmyelinated Schwann cells in a microenvironment with degeneration of normal nonmyelinated axon/Schwann cell relationships and mast cell infiltration (Joseph et€al. 2008; Wu et€al. 2008; Zheng et€al. 2008) with the timing of loss of NF1 being critical in order for neurofibromas to develop (Wu et€al. 2008). However, while loss of NF1 in the Schwann cell lineage was sufficient to generate tumors, complete NF1 mediated tumorigenesis required both a loss of NF1 in cells destined to become neoplastic as well as heterozygosity in nonneoplastic cell (Zhu et€ al. 2002). For example, in in€ vitro studies NF1 deficient Schwann cells were shown to be angiogenic and invasive and to secrete Kit ligand (Kim et€ al. 1997; Yang et€ al. 2003). This was shown to stimulate mast cell migration, and NF1+/− mast cells were shown to be hypermotile in response to Kit ligand and to secrete TGF-b (Yang et€al. 2006). The addition of imatinib to these cultures blocked the activity of NF+/− mast cells on fibroblast proliferation, collagen remodeling, and fibroblast migration (Yang et€al. 2006). These findings and studies in a genetically engineered NF1 mouse model (Zhu et€al. 2002) indicate an important role of tumor microenvironment in neurofibroma development.
Angiogenesis Angiogenesis also contributes to tumor formation in NF1. RAS mutations can upregulate VEGF expression (Rak et€al. 1995; Kranenburg et€al. 2004) and vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (FGF) are also highly expressed in neurofibromas from patients with NF1 at the mRNA and protein level (Kawachi et€al. 2003). Furthermore, VEGF expression and tumor vascularization significantly increased in MPNSTs (Angelov et€ al. 1999). Use of a specific small molecular inhibitor of VEGF-receptor 2 (VEGFR2) in a mouse explant model of neurogenic sarcomas showed a reduction in tumor growth due to
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
337
decreased tumor angiogenesis with subsequent reduction in tumor cell proliferation and an increase in apoptosis (Angelov et€al. 1999). NF1 deficient Schwann cells induced FGF, platelet derived growth factor (PDGF), and midkine (MK). Midkine was found to stimulate human MPNST and fibroblastoid cells of neurofibromas (Mashour et€ al. 2001). Subsequent studies demonstrated that MK and stem cell factor, but not epidermal growth factor (EGF) were substantially increased in serum of NF1 patients compared to healthy controls (Mashour et€al. 2004).
Growth Factors and Growth Factor Receptors PNs have increased expression of growth factors and growth factor receptors, including PDGF receptors and VEGF (Kim et€al. 1997; DeClue et€al. 2000; Ingram et€al. 2001; Mashour et€al. 2001; Kawachi et€al. 2003). In addition, epidermal growth factor receptor (EGFR) has been identified as an upstream activator of RAS in NF1. EGFR is not normally expressed by normal Schwann cells, but EGFR expression was demonstrated in benign neurofibromas, in MPNST cell lines established from NF1 patients (DeClue et€al. 2000), and in tumor cell lines derived from NF1:TP53 mice. Cell lines responded to EGF by activation of the downstream signaling pathways MAPK/ERK, and phosphatidylinositol 3¢-kinase (PI3k)/AKT. The growth of these cell lines could be blocked by an EGFR antagonist (Li et€al. 2002). The potential role of EGFR in peripheral nerve tumor formation was subsequently demonstrated in transgenic mouse Schwann cells expressing EGFR, which elicited features of neurofibromas. In addition, genetic reduction of EGFR in NF1+/− p53+/− mice that develop sarcomas significantly improved survival (Ling et€al. 2005). Agents blocking Raf-MAPK-ERK and PI3K-AKT pathways in NF1 have been shown to block proliferation (Lau et€al. 2000; Li et€al. 2002; Mattingly et€al. 2006a, b; Khalaf et€al. 2007), and therefore may be rational targets in NF1 in addition to agents blocking upstream activation of RAS.
Mammalian Target of Rapamcyin (mTOR) Recent studies have demonstrated that NF1 regulates mTOR pathway activation. NF1 loss in mouse embryonic fibroblasts and primary mouse astrocytes resulted in RAS- and PI3 kinase/AKT dependent mTOR pathway activation, which could be inhibited with rapamycin (sirolimus) (Dasgupta et€ al. 2005b; Johannessen et€ al. 2005). Increased proliferation associated with loss of neurofibromin expression in human MPNST cell lines was also dramatically reduced by the treatment with sirolimus (Johannessen et€al. 2005). Using a genetic mouse model of NF1-deficient malignant peripheral nerve sheath tumor – (MPNST) development, sirolimus completely inhibited the growth of these tumors in€vivo (Johannessen et€al. 2008). In addition, sirolimus treatment of optic gliomas developing in a genetically-engineered
338
B. Widemann
NF1 mouse model resulted in attenuated mTOR signaling in€ vitro and in€ vivo (Sandsmark et€ al. 2007; Hegedus et€ al. 2008). Similarly, the mTOR inhibitor RAD001 (everolimus) decreased growth in MPNST cell lines, and prevented the growth of subcutaneously implanted MPNST in mice (Johansson et€al. 2008). A negative feedback loop between mTOR and Akt was not found in the NF1 optic glioma and MPNST preclinical models after the treatment with sirolimus (Hegedus et€al. 2008; Johannessen et€al. 2008).
Molecular Features for MPNSTs Genetic and molecular alterations in addition to NF1 loss and changes described above contribute to the progression of PN to malignancy. Immunohistochemical analysis and molecular studies (Kourea et€ al. 1999; DeClue et€ al. 2000; Li et€ al. 2002; Perry et€al. 2002; Zhou et€al. 2003) have implicated TP53, EGFR p16INK4A, and p27 as potential contributors to malignant transformation in peripheral nerve sheath tumors. Further, highlighting the unique pathogenesis of this tumor type, both NF1 deletions and homozygous p16INK4A deletions appear to be relatively restricted to MPNSTs in comparison to other spindle cell sarcomas with overlapping morphologic features (Perry et€al. 2002). Gene expression profiling, immunohistochemistry, and/or Western blot analysis of human MPNSTs showed expression/ overexpression of EGFR of human MPNSTs (Watson et€al. 2004), PDGFRa and b (Badache and De Vries 1998; Holtkamp et€al. 2006), and C-Kit (Dang et€al. 2005; Holtkamp et€ al. 2006) matrix metalloprotetinase 13 (MMP13) (Holtkamp et€ al. 2007). In order to identify events contributing to malignant transformation a number of studies were performed comparing human Schwann cells and MPNST (Miller et€al. 2006), and PN and MPNST (Skotheim et€al. 2003; Holtkamp et€al. 2004; Levy et€al. 2004; Miller et€al. 2006), and identified that MPNSTs differentially expressed neural crest stem cell markers SOX9 and TWIST1 (Miller et€ al. 2006), genes involved in cell proliferation (MKI67, TOP2A, CCNE2), apoptosis (BIRC5/Survivin, TP73), extracellular matrix remodeling (MMP13, MMP9), genes involved in the RAS signaling pathway (RASF2, HMMR/RHAMM) and the Hedgehog-Gli signaling pathway (DHH, PTCH2). In addition, high resolution array comparative genomic hybridization was used to compare, dermal neurofibromas, PN, and MPNST and identified amplification of some genes, including PDGFRa, MET, TP73, and HGF, deletions in NF1, MMP13, p16INK4A, and TP53 demonstrating the potential of array CGH in identifying markers for MPNST (Mantripragada et€al. 2008). Analysis for somatic mutations in MPNST from NF1 patients showed loss of heterozygosity (LOH) across the TP53 region and TP53 mutation in 14 of 20 tumors (Upadhyaya et€al. 2008). However, while NF1 associated MPNST appears to have worse outcome compared to sporadic tumors, gene expression profiling of NF1 associated (n╛=╛25) and sporadic (n╛=╛17) MPNSTs did not identify a molecular signature that could reliably distinguish between both groups (Watson et€al. 2004).
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
339
Rationale of Pathways A number of targeted agents under development for the treatment of adults and children with refractory cancers have a good scientific rationale to be considered for development in NF1 related tumors. These include inhibitors of RAS, angiogenesis, growth factors, mast cell proliferation, and mTOR. Farnesyltransferase inhibitors (FTI) were among the first targeted agents to be evaluated for NF1 related tumors. They were developed to prevent posttranslational RAS processing and transduction of proliferative signals, which is required for the activity of mutant and wildtype RAS, and inhibition of RAS activity was the rationale for the evaluation of FTIs in cancer and NF1 (Rowinsky et€ al. 1999). However, farnesylation is required for the activity of a number of additional proteins, including Rho-B, Rac, membrane lamins, and centromeric proteins that interact with microtubules to promote mitosis, which may contribute to the anti-tumor effects of FTIs in clinical trials (Lancet and Karp 2003; Santos et€al. 2004). In addition, N-, and K-RAS can undergo an alternate lipid modification by geranyl-geranylation and thus overcome the effect of FTI (Rowinsky et€al. 1999). This is critical in light of recent data pointing toward K-ras as a key mediator of RAS activity in NF1 tumors (Dasgupta et€al. 2005a; Khalaf et€al. 2007). Thus, additional strategies, such as novel RAS inhibitors blocking all RAS isoforms (Barkan et€al. 2006) or agents blocking downstream pathways of RAS, are being considered for the development in NF1.
Current Strategies: Development of Targeted Treatments for NF1 Tumors Completed, ongoing, and planned clinical trials with targeted agents for NF1 related tumors are summarized in Table€1. Most of these focused on PN and used oral agents, which were administered on a chronic schedule. Initial trials used similar trial designs and endpoints as used in trials for refractory cancers. However, differences between refractory cancers and NF1 required different approaches toward drug development for NF1 related tumors (Table€ 2). This includes the requirement for more prolonged preclinical toxicology studies, including reproduction toxicology for NF1 trials given the near normal life expectancy of individuals with NF1. There are also important differences between individuals with NF1 related PN and refractory cancers, which is highlighted by the development of the FTI tipifarnib for NF1. In the phase I trial of tipifarnib, which was performed simultaneously in children with refractory cancers or NF1 related PN, the median age at trial entry was 7 years for NF1 (range 5 to 16 years) compared to 15 years (range 5 to 18 years) for refractory cancers (Widemann et€al. 2006). In addition, the treatment duration for children with NF1 was longer (median cycle number 10, range 1 to 32 cycles) compared to children with refractory cancers, who were removed from treatment with tipifarnib after a median cycle number of 1 (range 1 to 4) for
FTase
Fibroblast
Immune modulation Angiogenesis
Tipifarnib
Pirfenidone
Peginterferon alfa-2b
Oralâ•›×â•›21d q 28d Oral continuous
II
Oral continuous SC q week
SC q week
II
II
I
Oral continuous
I
II
Oralâ•›×â•›21 d q 28d
I
Inoperable, symptomatic, or progressive PN
Progressive PN 3–21 years Inoperable PN 1.6–21 years
Inoperable PN 3–21 years
Progressive PN 3–25 years Inoperable or sympt. PN Adults
Inoperable PN 3–21 years
Who-3D MRI
3D MRI Clinical
TTP Response
3D MRI
3D MRI
3D MRI
3D MRI
WHO
MTD
TTP
Toxicity PK
Response
TTP
Toxicity PK, PD
Table€1╅ Completed, ongoing, and planned clinical trials with targeted agents for NF1 related tumors Mechanism of Response Drug action/target Phase Schedule Eligibility Endpoint evaluation Plexiform neurofibroma Thalidomide Angiogenesis I Oral Progressive PN Toxicity WHO continuous >5 years
Ongoing, enrollment complete (n╛=╛36) MTD 1€mcg/kg SC weekly Several clinical and 3D MRI responses Ongoing
Optimal dose 500€mg/ m2/d Median cy # at MTD 15#
Gupta et€al. (2003)
Max dose given 200€mg/day Minor response and symptomatic improvement in few pts. MTD 200€mg/m2/d po BID ×21 d q28d Med cy #10 Ongoing, enrollment complete (nâ•›=â•›62) Volume decrease in 4 patients
–
Jakacki et al. (2008)
BabovicVuksanovic et€al. (2006) BabovicVuksanovic et€al. (2007) –
Widemann et€al. (2006) –
Reference
Result
340 B. Widemann
C-Raf, B-Raf, VEGFR2, C-Kit, PDGFRb
AZD2171
Sorafenib
FTase
C-Raf, B-Raf, VEGFR2, C-Kit, PDGFRb C-Kit, PDGFRb, VEGFR1
Ranibizumab
Angiogenesis VEGF
Dermal neurofibromas
JMML Tipifarnib
Imatinib
Sorafenib
EGFR
VEGFR2
Sirolimus
MPNST Erlotinib
C-Kit, PDGFRb, VEGFR1 mTOR
Imatinib
Pilot
II window
Intratumor injection
Oralâ•›×â•›21 d q 28d, 2 cy.
Oral continuous
II
II
Oral continuous Oral continuous
Oral continuous
I
II
Oral continuous
Oral continuous
Oral continuous
II
II
II
Response
>18 years dermal neurofibromas
Newly diagnosed JMML
Tumor volume VEGF signaling
Response
Response
Response
Refractory adults
Refractory children and adults
Response
Refractory adults
TTP Inoperable or Response progressive PN children and adults Response Inoperable or progressive PN adults Inoperable PN Toxicity, 3–21 years PK, PD, DEMRI
Children and adults
Soon to open
3D MRI
WBC, Liver size
Albritton et€al. (2006) Maki et al. (2009)
–
–
–
–
Ongoing
–
Response in WBC and Castleberry et€al. organomegaly in (2005) 58% of pts.
Ongoing
Inactive TTP 2 cycles Inactive
Ongoing
3 D MRI
RECIST
Ongoing
Ongoing
3D MRI
RECIST
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors 341
342
B. Widemann
Table€2╅ Differences in clinical drug development for refractory cancers and NF1 Endpoint Type of study Objective Cancer NF1 Preclinical Safe start dose Toxicology in rodent and Requirement for chronic toxicity and nonrodent species reproduction toxicity studies for human trial Chronic dosing requires redefining Toxicity during cycle #1 Phase I MTD DLT and MTD Pharmaco- Cumulative toxicity not Monitoring for cumulative toxicity assessable kinetics Phase II
Activity
Response Time to progression
Phase III
Efficacy
Survival, QOL
Response may be unrealistic with targeted agents Progression difficult to measure with solid tumor response criteria Unknown natural history requires control group Near normal survival in NF1QOL
MTD, maximum tolerated dose; DLT, dose-limiting toxicity; QOL, quality of life; PK, pharmacokinetics
disease progression. Cumulative toxicity could thus not be assessed in children with refractory cancers, but could be assessed in NF1, which was critical for the development of a subsequent phase II trial. Standard two-dimensional measurements (WHO) to assess response (Miller et€al. 1981) were inadequate to monitor changes in NF1 related PN. For subsequent clinical trials, a method of automated volumetric MRI analysis of PN was developed to sensitively and reproducibly monitor smaller changes in PN size (Solomon et€al. 2004). This method is now used in most ongoing PN trials and defines disease progression based on much smaller changes than standard solid tumor response criteria, which allows to limit the time period exposed to potentially inactive or toxic agents (Tableâ•›3). For the phase II trial with tipifarnib time to progression (TTP) rather than response was chosen as primary trial endpoint, as the expected pharmacologic effect of a cytostatic drug, such as tipifarnib, would be a slowing of PN growth, rather than a measurable decrease in PN size. The absence of prior data about the time to disease progression of PN required a concurrent control population. Therefore, the phase II trial was designed as a randomized, placebo-controlled, double-blinded, crossover phase II trial with the primary objective of determining whether tipifarnib increases time to disease progression in patients with progressive PN. Subsequent phase II trials for PN (pirfenidone, pegintron, sirolimus) will use the placebo arm of the tipifarnib phase II trial for PN as historical control to define the effect of the agent under study on time to disease progression. Eligibility criteria for these trials are identical to ensure that the historical control from the tipifarnib arm is valid. Other endpoints incorporated in clinical trials include pharmacodynamic effects, quality of life, and pain. Targeted treatment trials have also become available for MPNSTs, and while the agent studied did not result in responses, the feasibility of performing histology specific trials within an acceptable time period has been confirmed (Albritton et€al. 2006; D’Adamo et€al. 2007).
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
343
Table€3╅ Equivalent percent increase in diameter (RECIST), product (WHO), and volume for spherical lesions. Bold typeface highlights the definition of progression by RECIST, WHO, and the ongoing tipifarnib phase II trial for NF1 PN Percent change in tumor size RECIST diameter (1D) WHO product (2D) Tipfarnib phase II volume (3D) 6% 12% 20%
13% 25% 44%
20% 40% 73%
Future Directions and Challenges Preclinical Evaluation of Agents Until recently, the evaluation of agents in preclinical models of NF1 was difficult due to the lack of models recapitulating the disease and the number of cell types involved in NF1 related tumors. However, in€vitro (Ingram et€al. 2000; Yang et€al. 2003; Yang et€al. 2006), xenograft (Johannessen et€al. 2005; Johansson et€al. 2008), and transgenic mouse models (Cichowski et€al. 1999; Reilly et€al. 2000; McClatchey and Cichowski 2001; Hegedus et€al. 2008) of NF1 have become available, and are beginning to be validated for their value in predicting response in NF1 (Hegedus et€al. 2008). These studies are resource intensive. Recently, the Children’s Tumor Foundation (former National Neurofibromatosis Foundation) proceeded to fund a preclinical NF consortium with the goal to accelerate the identification of effective therapies for NF1. Candidate drugs will be evaluated in multiple tumor models of NF1 with the goal to move the most promising agents to clinical trials similar to the preclinical pediatric testing program for pediatric cancers (Houghton et€al. 2002; Houghton et€al. 2008). This work will lead to more rational selection of agents for clinical trials. In addition, the hope is that these models will help identify the role of specific cancer genes, perform target validation, define steps in tumorigenesis, evaluate impact of the tumor environment, and identify tumor specific biomarkers and gene modifiers (Gutmann and Giovannini 2002).
Infrastructure for Clinical Trials Initial clinical trials were performed through collaboration of limited institutions with funding provided by the Department of Defense (DoD). In order to develop an effective trials program, the infrastructure to conduct clinical trials in sequence is required. Recently, the DoD initiated funding of a NF Consortium, with the goal to allow for timely development of clinical trials in order to develop more effective therapies. This Consortium has nine participating site nationwide, and just initiated the first clinical trial of sirolimus for PN.
344
B. Widemann
Development of Agents for Young Children Longitudinal volumetric MRI analysis of PN demonstrated that PNs grow more rapidly in younger children compared to older children (Dombi et€ al. 2007). This raises the potential need for age stratification in clinical trials but also emphasizing the need to develop agents for young children. This is associated with multiple challenges. The lower age limit for most currently ongoing clinical trials for PN is 3 years due to the concern of toxicity in developing organs, particularly the nervous system. However, many young children may already have substantial PN burden at that time they become 3 years old and eligible for clinical trials (Fig.€1). Availability of age appropriate drug formulations and of trial designs, which provide for careful monitoring of toxicities in children will be critical to the conduct of trials in this population. This is exemplified by the development of angiogenesis inhibitors, which may have unique toxicities in children compared to adults. Inhibition of angiogenesis in growing but not aged animals resulted in the growth retardation and expanded growth plates (Gerber et€al. 1999). While this was largely reversible, there is concern for this toxicity, particularly in young children who have near normal life expectancy as in NF1. This toxicity cannot be easily monitored in phase I trials for refractory childhood cancers, as patients are older and remain on trial for short time periods (Bender et€ al. 2008). In a soon-to-open phase in trial of sorafenib for children with NF1 patients will be carefully monitored over multiple cycles for the development of unique toxicities including bony toxicity and growth plate changes using a newly developed method of automated volumetric MRI analysis of the growth plate. Observations from this and other future trial may thus not only inform for the drug development in NF1 but also for the drug development in childhood cancers.
Evaluation of Effect of Agents on Other NF1 Manifestations NF1 is a multisystem disorder and its manifestations in many organ systems may be, at least partially, the result of abnormal RAS function. While the primary goal in clinical trials directed at NF1 tumors is to monitor the effect on the tumor, agents under study may reverse deficits resulting from excessive RAS function in other organ systems. Thus, careful consideration should be given for the evaluation of other organ systems on clinical trials for NF1 related tumors.
Target Validation Access to tumor biopsies for target validation will be limited in NF1 related tumors given that many trials enroll children, and that biopsies frequently cannot be easily performed. Development of noninvasive assays, which allow monitoring of effects on desired targets should be an important goal for future NF1 trials.
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
345
Combination Therapy Finally, progress in the treatment of childhood cancers has been made by combining chemotherapy agents with different mechanisms of action. Given that multiple pathways are implied in tumor development in NF1, the evaluation of combined therapies may be required for successful treatment, which poses another challenge for future trials.
References Albritton, K., Rankin, C., Coffin, M., Ratner, N., Budd, G., Schuetze, S., Randall, L., DeClue, J., and Borden, E. 2006. Phase II study of erlotinib in metastatic or unresectable malignant peripheral nerve sheath tumors (MPNSTs). Paper presented at: ASCO (Atlanta, GA). Angelov, L., Salhia, B., Roncari, L., McMahon, G., and Guha, A. 1999. Inhibition of angiogenesis by blocking activation of the vascular endothelial growth factor receptor 2 leads to decreased growth of neurogenic sarcomas. Cancer Res 59: 5536–5541. Babovic-Vuksanovic, D., Ballman, K., Michels, V., McGrann, P., Lindor, N., King, B., Camp, J., Micic, V., Babovic, N., Carrero, X., et€ al. 2006. Phase II trial of pirfenidone in adults with neurofibromatosis type 1. Neurology 67: 1860–1862. Babovic-Vuksanovic, D., Widemann, B.C., Dombi, E., Gillespie, A., Wolters, P.L., ToledoTamula, M.A., O’Neill, B.P., Fox, E., MacDonald, T., Beck, H., et€al. 2007. Phase I trial of pirfenidone in children with neurofibromatosis 1 and plexiform neurofibromas. Pediatr Neurol 36: 293–300. Badache, A., and De Vries, G.H. 1998. Neurofibrosarcoma-derived Schwann cells overexpress platelet-derived growth factor (PDGF) receptors and are induced to proliferate by PDGF BB. J Cell Physiol 177: 334–342. Barkan, B., Starinsky, S., Friedman, E., Stein, R., and Kloog, Y. 2006. The RAS inhibitor Farnesylthiosalicylic acid as potential therapy for neurofibromatosis type 1. Clin Cancer Res 12: 5533–5541. Bender, J.L., Adamson, P.C., Reid, J.M., Xu, L., Baruchel, S., Shaked, Y., Kerbel, R.S., CooneyQualter, E.M., Stempak, D., Chen, H.X., 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. Carli, M., Ferrari, A., Mattke, A., Zanetti, I., Casanova, M., Bisogno, G., Cecchetto, G., Alaggio, R., De Sio, L., Koscielniak, E., et€ al. 2005. Pediatric malignant peripheral nerve sheath tumor: the Italian and German soft tissue sarcoma cooperative group. J Clin Oncol 23: 8422–8430. Castleberry, R., Loh, M., Jayaprakash, N., Peterson, A., Casey, V., Chang, M., Widemann, B., and Emanuel, P. 2005. Phase II window study of the farnesyltransferase inhibitor R115777 (zarnestra) in untreated juvenile myelomonocytic leukemia (JMML): A Children’s Oncology Group Study. Paper presented at: American Society of Hematology (Atlanta, GA). Cichowski, K., and Jacks, T. 2001. NF1 tumor suppressor gene function: narrowing the GAP. Cell 104: 593–604. Cichowski, K., Shih, T.S., Schmitt, E., Santiago, S., Reilly, K., McLaughlin, M.E., Bronson, R.T., and Jacks, T. 1999. Mouse models of tumor development in neurofibromatosis type 1. Science 286: 2172–2176. Dang, I., Nelson, J., and DeVries, G. 2005. C-kit receptor expression in normal human schwann cells and scwann cell lines derived from neurofibromatosis type 1 tumors. J Neurosci Res 82: 465–471.
346
B. Widemann
Dasgupta, B., Li, W., Perry, A., and Gutmann, D.H. 2005a. Glioma formation in neurofibromatosis 1 reflects preferential activation of K-RAS in astrocytes. Cancer Res 65: 236–245. Dasgupta, B., Yi, Y., Chen, D.Y., Weber, J.D., and Gutmann, D.H. 2005b. Proteomic analysis reveals hyperactivation of the mammalian target of rapamycin pathway in neurofibromatosis 1-associated human and mouse brain tumors. Cancer Res 65: 2755–2760. DeClue, J.E., Cohen, B.D., and Lowy, D.R. 1991. Identification and characterization of the neurofibromatosis type 1 protein product. Proc Natl Acad Sci USA 88: 9914–9918. DeClue, J.E., Heffelfinger, S., Benvenuto, G., Ling, B., Li, S., Rui, W., Vass, W.C., Viskochil, D., and Ratner, N. 2000. Epidermal growth factor receptor expression in neurofibromatosis type 1-related tumors and NF1 animal models. J Clin Invest 105: 1233–1241. DeClue, J.E., Papageorge, A.G., Fletcher, J.A., Diehl, S.R., Ratner, N., Vass, W.C., and Lowy, D.R. 1992. Abnormal regulation of mammalian p21ras contributes to malignant tumor growth in von Recklinghausen (type 1) neurofibromatosis. Cell 69: 265–273. Dombi, E., Solomon, J., Gillespie, A.J., Fox, E., Balis, F.M., Patronas, N., Korf, B.R., BabovicVuksanovic, D., Packer, R.J., Belasco, J., et€al. 2007. NF1 plexiform neurofibroma growth rate by volumetric MRI: relationship to age and body weight. Neurology 68: 643–647. Emanuel, P.D. 2004. Juvenile myelomonocytic leukemia. Curr Hematol Rep 3: 203–209. Evans, D.G., Baser, M.E., McGaughran, J., Sharif, S., Howard, E., and Moran, A. 2002. Malignant peripheral nerve sheath tumours in neurofibromatosis 1. J Med Genet 39: 311–314. Ferner, R.E. 2007. Neurofibromatosis 1 and neurofibromatosis 2: a twenty first century perspective. Lancet Neurol 6: 340–351. Ferner, R.E., and Gutmann, D.H. 2002. International consensus statement on malignant peripheral nerve sheath tumors in neurofibromatosis. Cancer Res 62: 1573–1577. Ferner, R.E., Huson, S.M., Thomas, N., Moss, C., Willshaw, H., Evans, D.G., Upadhyaya, M., Towers, R., Gleeson, M., Steiger, C., et€al. 2007. Guidelines for the diagnosis and management of individuals with neurofibromatosis 1. J Med Genet 44: 81–88. Friedman, J.M. 2002. Neurofibromatosis 1: clinical manifestations and diagnostic criteria. J Child Neurol 17: 548–554; discussion 571–542, 646–551. Gerber, H.P., Vu, T.H., Ryan, A.M., Kowalski, J., Werb, Z., and Ferrara, N. 1999. VEGF couples hypertrophic cartilage remodeling, ossification and angiogenesis during endochondral bone formation. Nat Med 5: 623–628. Gupta, A., Cohen, B.H., Ruggieri, P., Packer, R.J., and Phillips, P.C. 2003. Phase I study of thalidomide for the treatment of plexiform neurofibroma in neurofibromatosis 1. Neurology 60: 130–132. Gutmann, D.H., and Giovannini, M. 2002. Mouse models of neurofibromatosis 1 and 2. Neoplasia 4: 279–290. Hegedus, B., Banerjee, D., Yeh, T.H., Rothermich, S., Perry, A., Rubin, J.B., Garbow, J.R., and Gutmann, D.H. 2008. Preclinical cancer therapy in a mouse model of neurofibromatosis-1 optic glioma. Cancer Res 68: 1520–1528. Hegedus, B., Dasgupta, B., Shin, J.E., Emnett, R.J., Hart-Mahon, E.K., Elghazi, L., BernalMizrachi, E., and Gutmann, D.H. 2007. Neurofibromatosis-1 regulates neuronal and glial cell differentiation from neuroglial progenitors in€vivo by both cAMP- and RAS-dependent mechanisms. Cell Stem Cell 1: 443–457. Hiatt, K.K., Ingram, D.A., Zhang, Y., Bollag, G., and Clapp, D.W. 2001. Neurofibromin GTPaseactivating protein-related domains restore normal growth in Nf1−/− cells. J Biol Chem 276: 7240–7245. Holtkamp, N., Atallah, I., Okuducu, A.F., Mucha, J., Hartmann, C., Mautner, V.F., Friedrich, R.E., Mawrin, C., and von Deimling, A. 2007. MMP-13 and p53 in the progression of malignant peripheral nerve sheath tumors. Neoplasia 9: 671–677. Holtkamp, N., Mautner, V.F., Friedrich, R.E., Harder, A., Hartmann, C., Theallier-Janko, A., Hoffmann, K.T., and von Deimling, A. 2004. Differentially expressed genes in neurofibromatosis 1-associated neurofibromas and malignant peripheral nerve sheath tumors. Acta Neuropathol 107: 159–168. Holtkamp, N., Okuducu, A.F., Mucha, J., Afanasieva, A., Hartmann, C., Atallah, I., Estevez-Schwarz, L., Mawrin, C., Friedrich, R.E., Mautner, V.F., et€ al. 2006. Mutation and expression of
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
347
PDGFRA and KIT in malignant peripheral nerve sheath tumors, and its implications for imatinib sensitivity. Carcinogenesis 27: 664–671. Houghton, P., Adamson, P., Blaney, S., Fine, H., Gorlick, R., Haber, M., Helman, L., Hirschfeld, S., Hollingshead, M., Israel, M., et€al. 2002. Testing of new agents in childhood preclinical models: Meeting summary. Clin Cancer Res 8: 3646–3657. Houghton, P., Morton, C., Kolb, E., Giorlick, R., Lock, R., Carol, H., Reynolds, C., Marris, J., Keir, S., Billups, C., et€al. 2008. Initial testing (stage 1) of the mTOR inhibiotr rapamycin by the pediatric preclinical testing program. Pediatr Blood Cancer 50: 799–805. Huson, S.M., Compston, D.A., Clark, P., and Harper, P.S. 1989. A genetic study of von Recklinghausen neurofibromatosis in south east Wales. I. Prevalence, fitness, mutation rate, and effect of parental transmission on severity. J Med Genet 26: 704–711. Huson, S.M., Harper, P.S., and Compston, D.A. 1988. Von Recklinghausen neurofibromatosis. A clinical and population study in south-east Wales. Brain 111 (Pt 6): 1355–1381. Ingram, D.A., Hiatt, K., King, A.J., Fisher, L., Shivakumar, R., Derstine, C., Wenning, M.J., Diaz, B., Travers, J.B., Hood, A., et€ al. 2001. Hyperactivation of p21(ras) and the hematopoieticspecific Rho GTPase, Rac2, cooperate to alter the proliferation of neurofibromin-deficient mast cells in€vivo and in€vitro. J Exp Med 194: 57–69. Ingram, D.A., Yang, F.C., Travers, J.B., Wenning, M.J., Hiatt, K., New, S., Hood, A., Shannon, K., Williams, D.A., and Clapp, D.W. 2000. Genetic and biochemical evidence that haploinsufficiency of the Nf1 tumor suppressor gene modulates melanocyte and mast cell fates in€vivo. J Exp Med 191: 181–188. Jakacki, R., Dombi, E., Goldmann, S., Allen, J., Pollack, I., Geyer, R. 2008. Phase I trial of pegylated interferon alfa-2b for children and young adults with neurofibromatosis type I and unresectable plexiformneurofibromas. Neuro Oneol 10: 463. Johannessen, C.M., Johnson, B.W., Williams, S.M., Chan, A.W., Reczek, E.E., Lynch, R.C., Rioth, M.J., McClatchey, A., Ryeom, S., and Cichowski, K. 2008. TORC1 is essential for NF1-associated malignancies. Curr Biol 18: 56–62. Johannessen, C.M., Reczek, E.E., James, M.F., Brems, H., Legius, E., and Cichowski, K. 2005. The NF1 tumor suppressor critically regulates TSC2 and mTOR. Proc Natl Acad Sci USA 102: 8573–8578. Johansson, G., Mahller, Y.Y., Collins, M.H., Kim, M.O., Nobukuni, T., Perentesis, J., Cripe, T.P., Lane, H.A., Kozma, S.C., Thomas, G., et€al. 2008. Effective in€vivo targeting of the mammalian target of rapamycin pathway in malignant peripheral nerve sheath tumors. Mol Cancer Ther 7: 1237–1245. Joseph, N.M., Mosher, J.T., Buchstaller, J., Snider, P., McKeever, P.E., Lim, M., Conway, S.J., Parada, L.F., Zhu, Y., and Morrison, S.J. 2008. The loss of Nf1 transiently promotes selfrenewal but not tumorigenesis by neural crest stem cells. Cancer Cell 13: 129–140. Kawachi, Y., Xu, X., Ichikawa, E., Imakado, S., and Otsuka, F. 2003. Expression of angiogenic factors in neurofibromas. Exp Dermatol 12: 412–417. Khalaf, W.F., Yang, F.C., Chen, S., White, H., Bessler, W., Ingram, D.A., and Clapp, D.W. 2007. K-ras is critical for modulating multiple c-kit-mediated cellular functions in wild-type and Nf1+/− mast cells. J Immunol 178: 2527–2534. Kim, H.A., et€al. 1997. Nf1-deficient mouse Schwann cells are angiogenic and invasive and can be induced to hyperproliferate: reversion of some phenotypes by an inhibitor of farnesyl protein transferase. Mol Cell Biol 17: 862–872. King, A., Listernick, R., Charrow, J., Piersall, L., and Gutmann, D.H. 2003. Optic pathway gliomas in neurofibromatosis type 1: the effect of presenting symptoms on outcome. Am J Med Genet A 122: 95–99. Kluwe, L., Friedrich, R., and Mautner, V.F. 1999a. Loss of NF1 allele in Schwann cells but not in fibroblasts derived from an NF1-associated neurofibroma. Genes Chromosomes Cancer 24: 283–285. Kluwe, L., Friedrich, R.E., and Mautner, V.F. 1999b. Allelic loss of the NF1 gene in NF1associated plexiform neurofibromas. Cancer Genet Cytogenet 113: 65–69. Korf, B.R. 1999. Plexiform neurofibromas. Am J Med Genet 89: 31–37. Korf, B.R. 2000. Malignancy in neurofibromatosis type 1. Oncologist 5: 477–485.
348
B. Widemann
Korf, B.R. 2001. Diagnosis and management of neurofibromatosis type 1. Curr Neurol Neurosci Rep 1: 162–167. Korf, B.R. 2002. Clinical features and pathobiology of neurofibromatosis 1. J Child Neurol 17: 573–577; discussion 602–574, 646–551. Kourea, H.P., Cordon-Cardo, C., Dudas, M., Leung, D., and Woodruff, J.M. 1999. Expression of p27(kip) and other cell cycle regulators in malignant peripheral nerve sheath tumors and neurofibromas: the emerging role of p27(kip) in malignant transformation of neurofibromas. Am J Pathol 155: 1885–1891. Kranenburg, O., Gebbink, M.F., and Voest, E.E. 2004. Stimulation of angiogenesis by RAS proteins. Biochim Biophys Acta 1654: 23–37. Lancet, J.E., and Karp, J.E. 2003. Farnesyltransferase inhibitors in hematologic malignancies: new horizons in therapy. Blood 102: 3880–3889. Lau, N., Feldkamp, M., Roncari, L., Loehr, A., Shannon, P., Gutmann, D., and Guha, A. 2000. Loss of neurofibromin is associated with activation of RAS/MAPK and PI3-K/AKT signaling in neurofibromatosis 1 astrocytoma. J Neuropathol Exp Neurol 59: 759–767. Levy, P., Vidaud, D., Leroy, K., Laurendeau, I., Wechsler, J., Bolasco, G., Parfait, B., Wolkenstein, P., Vidaud, M., and Bieche, I. 2004. Molecular profiling of malignant peripheral nerve sheath tumors associated with neurofibromatosis type 1, based on large-scale real-time RT-PCR. Mol Cancer 3: 20. Li, H., Velasco-Miguel, S., Vass, W.C., Parada, L.F., and DeClue, J.E. 2002. Epidermal growth factor receptor signaling pathways are associated with tumorigenesis in the Nf1:p53 mouse tumor model. Cancer Res 62: 4507–4513. Ling, B.C., Wu, J., Miller, S.J., Monk, K.R., Shamekh, R., Rizvi, T.A., Decourten-Myers, G., Vogel, K.S., DeClue, J.E., and Ratner, N. 2005. Role for the epidermal growth factor receptor in neurofibromatosis-related peripheral nerve tumorigenesis. Cancer Cell 7: 65–75. Listernick, R., Charrow, J., Greenwald, M., and Mets, M. 1994. Natural history of optic pathway tumors in children with neurofibromatosis type 1: a longitudinal study. J Pediatr 125: 63–66. Listernick, R., Ferner, R.E., Liu, G.T., and Gutmann, D.H. 2007. Optic pathway gliomas in neurofibromatosis-1: controversies and recommendations. Ann Neurol 61: 189–198. Mantripragada, K.K., Spurlock, G., Kluwe, L., Chuzhanova, N., Ferner, R.E., Frayling, I.M., Dumanski, J.P., Guha, A., Mautner, V., and Upadhyaya, M. 2008. High-resolution DNA copy number profiling of malignant peripheral nerve sheath tumors using targeted microarray-based comparative genomic hybridization. Clin Cancer Res 14: 1015–1024. Maki, R.G., D’Adamo, D.R., Keohan, M.L., Saulle, M., Schuetze, S.M., Undevia, S.D., Livingston, M.B., Cooney, M.M., Hensley, M.L., Mita, M.M., Takimoto, C.H., Kraft, A.S., Elias, A.D., Brockstein, B., Blachère, N.E., Edgar, M.A., Schwartz, L.H., Qin, L.X., Antonescu, C.R., Schwartz, G.K. 2009. Phase II study of sorafenib in patients with metastatic or recurrent sarcomas. J Clin Oncol 27: 3133–3140. Mashour, G.A., Driever, P.H., Hartmann, M., Drissel, S.N., Zhang, T., Scharf, B., FelderhoffMuser, U., Sakuma, S., Friedrich, R.E., Martuza, R.L., et€al. 2004. Circulating growth factor levels are associated with tumorigenesis in neurofibromatosis type 1. Clin Cancer Res 10: 5677–5683. Mashour, G.A., Ratner, N., Khan, G.A., Wang, H.L., Martuza, R.L., and Kurtz, A. 2001. The angiogenic factor midkine is aberrantly expressed in NF1-deficient Schwann cells and is a mitogen for neurofibroma-derived cells. Oncogene 20: 97–105. Mattingly, R., Kraniak, J., Dilworth, J., Mathieu, P., Bealmear, B., Nowak, J., Benjamins, J., Tainsky, M., and Reiners, J. 2006a. The mitogen-activated protein kinase/extracellular signal regulated kinase kinase inhibitor PD184352 (CI-1040) selectively induces apoptosis in malignant schwannoma cell lines. J Pharmacol Exp Ther 316: 456–465. Mattingly, R.R., Kraniak, J.M., Dilworth, J.T., Mathieu, P., Bealmear, B., Nowak, J.E., Benjamins, J.A., Tainsky, M.A., and Reiners, J.J., Jr. 2006b. The mitogen-activated protein kinase/extracellular signal-regulated kinase kinase inhibitor PD184352 (CI-1040) selectively induces apoptosis in malignant schwannoma cell lines. J Pharmacol Exp Ther 316: 456–465.
Development of Targeted Therapies for Neurofibromatosis Type 1 (NF1) Related Tumors
349
McClatchey, A.I., and Cichowski, K. 2001. Mouse models of neurofibromatosis. Biochim Biophys Acta 1471: M73–M80. Messiaen, L.M., Callens, T., Mortier, G., Beysen, D., Vandenbroucke, I., Van Roy, N., Speleman, F., and Paepe, A.D. 2000. Exhaustive mutation analysis of the NF1 gene allows identification of 95% of mutations and reveals a high frequency of unusual splicing defects. Hum Mutat 15: 541–555. Miller, A.B., Hoogstraten, B., Staquet, M., and Winkler, A. 1981. Reporting results of cancer treatment. Cancer 47: 207–214. Miller, S.J., Rangwala, F., Williams, J., Ackerman, P., Kong, S., Jegga, A.G., Kaiser, S., Aronow, B.J., Frahm, S., Kluwe, L., et€al. 2006. Large-scale molecular comparison of human schwann cells to malignant peripheral nerve sheath tumor cell lines and tissues. Cancer Res 66: 2584–2591. Needle, M.N., et€al. 1997. Prognostic signs in the surgical management of plexiform neurofibroma: the Children’s Hospital of Philadelphia experience, 1974–1994. J Pediatr 131: 678–682. NIH, Consensus, Development, and Conference. 1988. Neurofibromatosis statement. Arch Neurol 45: 475. North, K. 1997. Neurofibromatosis type 1 in childhood (London, Mac keith Press). Packer, R., Gutmann, D., Rubenstein, A., Viskochil, D., Zimmerman, R., Vezina, G., Small, J., and Korf, B. 2002. Plexiform neurofibromas in NF1: toward biologic-based therapy. Neurology 58: 1461–1470. Perry, A., Kunz, S.N., Fuller, C.E., Banerjee, R., Marley, E.F., Liapis, H., Watson, M.A., and Gutmann, D.H. 2002. Differential NF1, p16, and EGFR patterns by interphase cytogenetics (FISH) in malignant peripheral nerve sheath tumor (MPNST) and morphologically similar spindle cell neoplasms. J Neuropathol Exp Neurol 61: 702–709. Rak, J., Mitsuhashi, Y., Bayko, L., Filmus, J., Shirasawa, S., Sasazuki, T., and Kerbel, R.S. 1995. Mutant ras oncogenes upregulate VEGF/VPF expression: implications for induction and inhibition of tumor angiogenesis. Cancer Res 55: 4575–4580. Reilly, K., Loisel, D., Bronson, R., McLaughlin, M., and Jacks, T. 2000. Nf1;Trp53 mutant mice develop glioblastoma with evidence of strain specific effects. Nat Genet 26: 109–113. Rowinsky, E., Windle, J., and Von Hoff, D. 1999. RAS protein farnesyltransferse: a strategic target for anticancer therapeutic development. J Clin Oncol 17: 3631–3652. Sandsmark, D.K., Pelletier, C., Weber, J.D., and Gutmann, D.H. 2007. Mammalian target of rapamycin: master regulator of cell growth in the nervous system. Histol Histopathol 22: 895–903. Santos, E.S., Rosenblatt, J.D., and Goodman, M. 2004. Role of farnesyltransferase inhibitors in hematologic malignancies. Expert Rev Anticancer Ther 4: 843–856. Satoh, T., and Kaziro, Y. 1992. RAS in signal transduction. Cancer Biol 3: 169–177. Serra, E., Rosenbaum, T., Winner, U., Aledo, R., Ars, E., Estivill, X., Lenard, H.G., and Lazaro, C. 2000. Schwann cells harbor the somatic NF1 mutation in neurofibromas: evidence of two different Schwann cell subpopulations. Hum Mol Genet 9: 3055–3064. Sherman, L.S., Atit, R., Rosenbaum, T., Cox, A.D., and Ratner, N. 2000. Single cell RAS-GTP analysis reveals altered RAS activity in a subpopulation of neurofibroma Schwann cells but not fibroblasts. J Biol Chem 275: 30740–30745. Skotheim, R.I., Kallioniemi, A., Bjerkhagen, B., Mertens, F., Brekke, H.R., Monni, O., Mousses, S., Mandahl, N., Soeter, G., Nesland, J.M., et€al. 2003. Topoisomerase-II alpha is upregulated in malignant peripheral nerve sheath tumors and associated with clinical outcome. J Clin Oncol 21: 4586–4591. Solomon, J., Warren, K., Dombi, E., Patronas, N., and Widemann, B. 2004. Automated detection and volume measurement of plexiform neurofibromas in neurofibromatosis 1 using magnetic resonance imaging. Comput Med Imaging Graph 28: 257–265. Thomas, S.L., Deadwyler, G.D., Tang, J., Stubbs, E.B., Jr., Muir, D., Hiatt, K.K., Clapp, D.W., and De Vries, G.H. 2006. Reconstitution of the NF1 GAP-related domain in NF1-deficient human Schwann cells. Biochem Biophys Res Commun 348: 971–980. Upadhyaya, M., Huson, S.M., Davies, M., Thomas, N., Chuzhanova, N., Giovannini, S., Evans, D.G., Howard, E., Kerr, B., Griffiths, S., et€al. 2007. An absence of cutaneous neurofibromas associated with a 3-bp inframe deletion in exon 17 of the NF1 gene (c.2970-2972 delAAT):
350
B. Widemann
evidence of a clinically significant NF1 genotype-phenotype correlation. Am J Hum Genet 80: 140–151. Upadhyaya, M., Kluwe, L., Spurlock, G., Monem, B., Majounie, E., Mantripragada, K., Ruggieri, M., Chuzhanova, N., Evans, D.G., Ferner, R., et€ al. 2008. Germline and somatic NF1 gene mutation spectrum in NF1-associated malignant peripheral nerve sheath tumors (MPNSTs). Hum Mutat 29: 74–82. Upadhyaya, M., Ruggieri, M., Maynard, J., Osborn, M., Hartog, C., Mudd, S., Penttinen, M., Cordeiro, I., Ponder, M., Ponder, B.A., et€al. 1998. Gross deletions of the neurofibromatosis type 1 (NF1) gene are predominantly of maternal origin and commonly associated with a learning disability, dysmorphic features and developmental delay. Hum Genet 102: 591–597. Watson, M.A., Perry, A., Tihan, T., Prayson, R.A., Guha, A., Bridge, J., Ferner, R., and Gutmann, D.H. 2004. Gene expression profiling reveals unique molecular subtypes of neurofibromatosis type I-associated and sporadic malignant peripheral nerve sheath tumors. Brain Pathol 14: 297–303. Weiss, B., Bollag, G., and Shannon, K. 1999. Hyperactive RAS as a therapeutic target in neurofibromatosis type 1. Am J Med Genet 89: 14–22. Widemann, B.C., Salzer, W.L., Arceci, R.J., Blaney, S.M., Fox, E., End, D., Gillespie, A., Whitcomb, P., Palumbo, J.S., Pitney, A., et€al. 2006. Phase I trial and pharmacokinetic study of the farnesyltransferase inhibitor tipifarnib in children with refractory solid tumors or neurofibromatosis type I and plexiform neurofibromas. J Clin Oncol 24: 507–516. Wu, J., Williams, J.P., Rizvi, T.A., Kordich, J.J., Witte, D., Meijer, D., Stemmer-Rachamimov, A.O., Cancelas, J.A., and Ratner, N. 2008. Plexiform and dermal neurofibromas and pigmentation are caused by Nf1 loss in desert hedgehog-expressing cells. Cancer Cell 13: 105–116. Yang, F.C., Chen, S., Clegg, T., Li, X., Morgan, T., Estwick, S.A., Yuan, J., Khalaf, W., Burgin, S., Travers, J., et€ al. 2006. Nf1+/− mast cells induce neurofibroma like phenotypes through secreted TGF-beta signaling. Hum Mol Genet 15: 2421–2437. Yang, F.C., Ingram, D.A., Chen, S., Hingtgen, C.M., Ratner, N., Monk, K.R., Clegg, T., White, H., Mead, L., Wenning, M.J., et€ al. 2003. Neurofibromin-deficient Schwann cells secrete a potent migratory stimulus for Nf1+/− mast cells. J Clin Invest 112: 1851–1861. Zheng, H., Chang, L., Patel, N., Yang, J., Lowe, L., Burns, D.K., and Zhu, Y. 2008. Induction of abnormal proliferation by nonmyelinating schwann cells triggers neurofibroma formation. Cancer Cell 13: 117–128. Zhou, H., Coffin, C.M., Perkins, S.L., Tripp, S.R., Liew, M., and Viskochil, D.H. 2003. Malignant peripheral nerve sheath tumor: a comparison of grade, immunophenotype, and cell cycle/ growth activation marker expression in sporadic and neurofibromatosis 1-related lesions. Am J Surg Pathol 27: 1337–1345. Zhu, Y., Ghosh, P., Charnay, P., Burns, D.K., and Parada, L.F. 2002. Neurofibromas in NF1: Schwann cell origin and role of tumor environment. Science 296: 920–922.
Molecular Therapy for Neuroblastoma Yaël P. Mossé and John M. Maris
Introduction Despite the generally held notion that embryonal cancers are simpler from a genetic standpoint than adult neoplasms, these cancers show multiple chromosomal rearrangements, suggesting a complicated series of acquired alterations during malignant evolution. This is best exemplified in neuroblastoma, an enigmatic and highly heterogeneous pediatric neoplasm that accounts for 15% of pediatric cancer mortality and remains an important clinical problem in which tumor genomics correlate with disease phenotype (George et€al. 2007). Neuroblastoma represents a spectrum of diseases, and distinct patient subsets exist based on tumor biological features (Maris 2005). Low-risk neuroblastomas are likely characterized by mitotic dysfunction leading to a hyperdiploid modal karyotype with whole chromosome gains, but few, if any, structural cytogenetic rearrangements. These tumors may regress spontaneously, differentiate, or respond completely to modest doses of chemotherapy. In contrast, aggressive neuroblastomas are characterized by complex segmental chromosomal aberrations. Two main categories of unfavorable neuroblastomas exist, with a well-defined subset (approximately 40%) defined by deletion of the short arm of chromosome 1 and MYCN amplification. On the other hand, the majority of aggressive neuroblastomas show multiple other chromosomal rearrangements, most frequently involving chromosome 11q (Attiyeh et€al. 2005). All high-risk patients are treated with dose-intensive induction chemotherapy and surgery, followed by myeloablative therapy with stem cell rescue, local radiation therapy and biological response modification using retinoids. Despite this aggressive approach, survival on recent high-risk studies remains <35% (Berthold et€al. 2005; Matthay et€al. 1999) with only modest improvement over the past few
Y.P. Mossé (*) Division of Oncology, Department of Pediatrics, Children’s Hospital of Philadelphia, University of Pennsylvania, 3615 Civic Center Blvd., ARC 907C, Philadelphia, PA, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_17, © Springer Science+Business Media, LLC 2010
351
352
Y.P. Mossé and J.M. Maris 100
1990-1994 (N=575) 1995-1999 (N=345) 2000-2004 (N=856)
80 60 40 20 0
0
1
2
3
4
5
6
7
8 9 10 11 12 13 14 15 16 17 Years
Fig.€ 1â•… Overall survival probability in 5-year intervals for all high-risk neuroblastoma patients enrolled in the United States-based pediatric cooperative group clinical trials from 1990 to 2004. While the most recent Kaplan-Meier survival curve is significantly different from the others (Pâ•›<â•›0.001), the clinical significance of this improved outcome is debatable, especially with longer follow-up. Data courtesy of Dr. Wendy London, Children’s Oncology Group
years (Fig.€1). It is clear that major improvements in patient outcome will unlikely be realized by simply adding cytotoxic therapy to the already dose-intensive regimens prescribed for children with high-risk neuroblastoma. This chapter explores the potential of how we can rationally exploit known tumor-specific alterations toward current and future treatment strategies.
Risk Classification Risk assessment depends upon a number of clinical and biologic features. The Children’s Oncology Group currently stratifies patients into low-, intermediate-, or high-risk categories based upon the well-defined prognostic factors of age at diagnosis, disease stage, tumor histopathologic classification, DNA index, and MYCN amplification status. Although this system is useful, there are almost certainly misclassifications resulting in patients who are either over- or undertreated. Amplification of the MYCN oncogene is a strong prognostic factor, but over 60% of patients with high-risk neuroblastoma do not have MYCN amplification present in the diagnostic specimen. There is also very likely a subset of children currently classified as high-risk who do not require the extreme dose-intensive therapeutic approach currently prescribed. Because approaches to risk stratification have varied greatly throughout the world, efforts have recently been made to develop a consensus approach that will permit comparison of outcomes for patients with neuroblastoma
Molecular Therapy for Neuroblastoma
353
treated in centers around the globe. In an effort to develop an International Neuroblastoma Risk Group (INRG) classification system, a working group, representing the major pediatric cooperative groups around the world, met in 2005 to review data collected on 11,054 patients treated in Europe, Japan, North America, and Australia between 1974 and 2002. Considerable progress has been made in the past decade toward understanding human neuroblastoma at a cellular and molecular level. Many studies clearly indicate that somatically acquired chromosomal aberrations in diagnostic neuroblastomas are intricately related to the clinical phenotype, and suggest that tumor genomics will provide a much more sensitive and specific indication of clinical phenotype and will greatly refine, or even replace, our current risk classification system.
Current Approach to Therapy Locoregional Tumors The majority of localized neuroblastomas have favorable biological features and most are successfully treated with surgery alone (Alvarado et€ al. 2000; Berthold et€al. 1982; Evans et€al. 1996; Kushner et€al. 1996b; Matthay et€al. 1989). Studies suggest that a subset of localized tumors will spontaneously regress, and these patients can be safely observed without any treatment (Nishihira et€al. 2000; Oue et€al. 2005; Suita et€al. 1996; Yamamoto et€al. 1998). Local recurrences can typically be managed surgically. Metastatic recurrences are rare and often treated successfully with chemotherapy. The treatment of patients with localized tumors with unfavorable biological features, particularly MYCN amplification, remains controversial. Although these children have significantly worse outcome than patients with localized disease that lacks MYCN amplification, a subset may achieve longterm remission following surgery alone if a gross total resection is achieved (Alvarado et€al. 2000; Cohn et€al. 1995; Perez et€al. 2000). These rare cases require continued prospective evaluation to clarify optimal management. The management of more invasive locoregional tumors (INSS stage 3) also remains controversial (Matthay et€al. 1998b). Traditionally, patients with favorable biology INSS stage 3 tumors have received moderately intensive chemotherapy with the goal of facilitating subsequent surgical resection. Given the favorable outcome for these children, most investigators avoid the use of radical surgery or radiotherapy. In fact, there is a gradual movement toward reducing adjuvant therapy in these cases despite the presence of gross residual tumor (Kushner et€al. 1996a), although these approaches require prospective validation. The current COG intermediate-risk clinical trial utilizes additional molecular genetic features (1p36 and 11q23 allelic status) and a response-based algorithm to attempt further reduction in chemotherapy exposure to these often very young children. In contrast, for invasive locoregional tumors with unfavorable biological features, intensive multimodality therapy is often required to achieve cure.
354
Y.P. Mossé and J.M. Maris
Metastatic Tumors Treatment of older children with widely disseminated neuroblastoma (stage 4) remains one of the greatest challenges for pediatric oncologists. It is clear that very young children with metastatic disease frequently have a much less malignant disease, and again this is largely predictable by tumor biology (Schmidt et€al. 2000, 2005). In recent years, attempts have been made to improve outcomes in high-risk patients (generally children older than 1 year of age at diagnosis with metastatic disease) by delivering intensive induction therapy. Commonly used agents include cisplatin, etoposide, doxorubicin, cyclophosphamide, and vincristine (Cheung and Cheung 2001). The combination of topotecan and cyclophosphamide has been used in the relapse setting for more than a decade, and a recent pilot study has demonstrated the feasibility of integrating this combination into an aggressive induction regimen. During induction therapy, stem cells are harvested in preparation for the consolidation phase of therapy. The goal of consolidation is to eliminate any remaining tumor, usually with myeloablative cytotoxic agents and stem cell rescue. Delayed surgical resection of the primary tumor and external beam radiotherapy to the primary and major metastatic sites are provided. The concept of eliminating resistant tumor clones with supralethal chemotherapy has been studied in neuroblastoma since the early 1980s, and the results of a randomized Children’s Cancer Group (CCG) Phase III cooperative group study demonstrated improved event-free survival with autologous transplantation (Matthay et€ al. 1999). An intent-to-treat analysis of a randomized clinical trial of high-risk neuroblastoma patients conducted by the German Society of Pediatric Oncology and Hematology also demonstrated an improvement in 3-year event-free survival with myeloablative therapy and autologous stem cell rescue compared to maintenance chemotherapy (Berthold et€ al. 2005). Similar to the CCG study, however, overall survival rates were not statistically different. Prolongation of the disease free interval is clinically important, but clearly much remains to be done to improve overall survival for high-risk patients. The COG now has an active Phase III study with a topotecan-containing induction and a randomization between single versus tandem myeloablative peripheral blood transplantation. As relapse is a frequent occurrence after autologous transplantation, biological therapy to treat persistent minimal residual disease has been added to current treatment regimens. Several novel agents specifically targeted to the unique biology of neuroblastoma may be effective in eliminating minimal residual disease. The retinoids are a class of compounds known to induce terminal differentiation of neuroblastoma cells in€ vitro (Sidell 1982). The use of 13-cis-retinoic acid in the post-transplant setting was tested in a randomized Phase III trial conducted by the CCG. The cohort of patients assigned to receive post-transplant therapy with 13-cis-retinoic acid had a significantly improved event free survival and toxicity was acceptable (Matthay et€al. 1999). Thus, retinoid-based biotherapy in the posttransplant setting is now widely used. Other retinoids such as 9-cis-retinoic acid, all trans-retinoic acid and/or fenretinide may also have activity for minimal residual
Molecular Therapy for Neuroblastoma
355
disease in high-risk neuroblastoma patients and warrant further study. There has been extensive interest in monoclonal antibody-based therapy to clear minimal residual disease. Both murine and chimeric anti-GD2 agents have shown activity in refractory neuroblastoma, and the ch14.18 monoclonal antibody is currently being studied in a randomized Phase III trial in the COG in conjunction with 13-cisretinoic acid and cytokines (Berthold et€al. 2005; Kushner et€al. 2001; Ozkaynak et€al. 2000).
Strategies to Identify Molecular Targets While further intensification of therapy as proposed in the next COG high-risk study (tandem myeloablative consolidation) may result in incremental improvement, it can be reasonably argued that no substantive advances in cure rates will be achieved until we can integrate radically new treatment strategies based on the fundamental molecular alterations present in neuroblastoma. A major effort has been mounted toward utilization of genomic technology to discover neuroblastomaspecific therapeutic targets. Strategies such as the neuroblastoma TARGET (Therapeutically Applicable Research to Generate Effective Treatments) initiative have been funded to perform a comprehensive assessment of the neuroblastoma genome designed to identify the critical genetic alterations and to leverage this information to develop targeted therapies. We anticipate that the combination of genomic (including noncoding RNAs), transcriptomic, and epigenomic data from carefully designed experiments will be synergistic in terms of our understanding of the biological basis of neuroblastoma, and that integration of these datasets will lead to the identification of rational therapeutic targets. We are also taking a global approach that takes into account that many oncogenic events will be silent at the DNA and RNA level. It is likely that overexpressed and constitutively active proteins resulting from activating mutations may be ideal therapeutic targets, and that high-throughput transient siRNA knockdown screens of druggable proteins (i.e., protein kinases) in representative neuroblastoma cell lines will enable us to identify the proteins essential for cellular survival. Finally, there is the pediatric preclinical testing program (PPTP), which was designed to evaluate the potential efficacy of new drugs in a manner that will allow direct translation to the clinic.
New Approaches to High Risk Disease While there are highly effective salvage options for patients with low- and intermediate-risk disease with local relapses, refractory high-risk disease remains a significant clinical challenge with no known curative approach. Over the past several years, an expanding portfolio of novel therapeutic agents with efficacy in the relapsed setting have been developed (Table€1) and several are highlighted in this section.
356
Y.P. Mossé and J.M. Maris
Table€1╅ Investigational agents in early phase clinical development for neuroblastoma Mechanism of action/target Examples Reference Tubulin binding agents ABT-751 Fox et€al. (2006) Proteosome inhibition Bortezomib Brignole et€al. (2006), Houghton et€al. (2007) VEGF inhibition Bevacizumab Dickson et€al. (2007), Maris et€al. (2008) AZD2171 VEGF TRAP Glutathione synthetase inhibitor Buthionine Anderson and Reynolds (2002) sulfoximine Trk tyrosine kinase inhibitor Lestaurtinib Evans et€al. (1999), Marshall et€al. (2005) SRC tyrosine kinase inhibitor Dasatinib Shah et€al. (2004) Demethylation Decitabine Teitz et€al. (2000) Histone deacetylase inhibitor Depsipeptide Marks et€al. (2001), Coffey et€al. (2001) Vorinostat SAHA BCL2 antisense Genasense Banerjee (2001) Histone deacetylase inhibitor Vorinostat Keshelava et€al. (2007), Yang et€al. (2007) EGFR tyrosine kinase inhibition Iressa Ho et€al. (2005) Heat shock protein modulation 17-AAG Bagatell et€al. (2007) MYCN inhibition Bisphosphonate Zoledronic acid Sohara et€al. (2003) Mitotic spindle inhibitors Zhou et€al. (1998) Centrosome inhibition Slack et€al. (2007) 131 Norepinephrine receptor/targeted I-MIBG Matthay et€al. (2006), (2007) radiotherapeutic Retinoids Fenretinide Maurer et€al. (2000) Ceramide modulator Safingol Maurer et€al. (1999), Reynolds et€al. (2004) Anti-GD2 3F8 Berthold et€al. (2005), Kushner et€al. (2001), Ozkaynak et€al. 14.18 (2000) ch14.18 Anti-anti-GD2 (anti-id antibody) mAb1A7 Yu et€al. (2004) Humanized anti-GD2 and IL2 fusion hu14.18-IL2 Osenga et€al. (2006) protein Immune response enhancer/iC3b Beta-d-glucan Cheung and Modak (2002) leukocyte receptor Immune modulation IL12 Lode et€al. (1999), Siapati et€al. (2003) Anti-IGF-I receptor monoclonal Ab IMC-A12 Miller and Yee (2005)
Cytotoxic Agents The topoisomerase 1 inhibitors, topotecan and irinotecan, are often used early in the relapse setting because of their proven efficacy and low toxicity profile (Langler et€al. 2002; Saylors et€al. 2001; Vassal et€al. 2003; Wagner et€al. 2004). Although
Molecular Therapy for Neuroblastoma
357
both have efficacy as single agents, a recent randomized trial showed that topotecan combined with low dose cyclophosphamide was superior to topotecan alone (CR╛+╛PR 31% vs. 19%) (Frantz et€al. 2004). Irinotecan combined with temozolomide has been shown to be tolerable in the Phase I setting (Wagner et€al. 2004), and the activity is currently being tested in a COG Phase II study. The most common toxicity related to Irinotecan is diarrhea, which can potentially be prevented with the concomitant use of oral antibiotics (Alimonti et€al. 2003). The combination of irinotecan and topotecan has been studied in the Phase I setting, but unacceptable toxicity has limited further development (Rodriguez-Galindo et€al. 2006). A promising novel cytotoxic agent is ABT-751, an oral tubulin-binding agent that is not a classic multidrug resistance pump substrate. Early clinical trials suggest an activity against refractory neuroblastoma (Cho et€al. 2004), and ABT-751 is being studied in a COG Phase II trial.
Tyrosine Kinase Inhibitors Malignant transformation of neuroblasts may result in part from a failure to respond to normal differentiation signals. The factors responsible for regulating normal differentiation in the sympathetic nervous system are not completely understood, but they at least in part involve the neurotrophin receptor pathways (Brodeur 2003). Neurotrophins critical to these pathways include nerve growth factor (NGF) (Azar et€al. 1990; Baker et€al. 1989), as well as brain-derived neurotrophin growth factor (BDNF), neurotropin-3 (NT-3), and neurotrophin-4 (NT-4) (Brodeur et€al. 1996). To date, three high affinity receptor tyrosine kinases for this family of �neurotrophins have identified. The genes TrkA, TrkB, and TrkC encode the primary receptors for NGF, BDNF, and NT-3, respectively (Brodeur et€al. 1996). The primary receptor for NT-4 is not known, but it appears to function through TrkB. The expression pattern of the Trk neurotrophin receptors is correlated with biological and clinical features of neuroblastoma. High TrkA expression has been shown to be associated with younger age (<1 year), lower stage (stage 1, 2, and 4S), and a favorable outcome (Kogner et€al. 1993; Nakagawara et€al. 1993; Suzuki et€al. 1993; Tanaka et€al. 1998). In addition, Nakagawara and colleagues have demonstrated that there is an inverse correlation between TrkA expression and MYCN amplification and that the combined assessment of MYCN copy number and TrkA expression provides additional prognostic information over either variable alone (Nakagawara et€ al. 1992, 1993). Furthermore, primary neuroblastoma cells with high TrkA expression differentiate in the presence of NGF in€ vitro, whereas the same cells die in the absence of NGF (Nakagawara et€al. 1993). Thus, the NGF/ TrkA pathway may explain the propensity for some neuroblastomas to differentiate or to regress spontaneously. The studies of TrkB and TrkC expression in neuroblastomas are more limited. TrkB is expressed in about a third of neuroblastomas (Nakagawara et€al. 1994). The truncated form of TrkB, lacking the tyrosine kinase domain, is expressed predominantly
358
Y.P. Mossé and J.M. Maris
in more differentiated tumors (ganglioneuroblastomas and ganglioneuromas), whereas the full-length TrkB transcript is expressed in tumors with MYCN amplification often in association with its cognate ligand BDNF (Nakagawara et€al. 1994). Thus, the TrkB/BDNF pathway may serve as an autocrine or paracrine pathway to promote survival in MYCN-amplified, and perhaps other, high-risk tumors. In addition, constitutively active BDNF/TrkB signaling seems to confer (or at least be associated with) resistance to conventional chemotherapeutic agents (Ho et€ al. 2002; Jaboin et€al. 2002a). TrkC expression is present in about 25% of neuroblastomas tested, and its pattern of expression resembles that of TrkA (Ryden et€ al. 1996; Yamashiro et€al. 1996), and may represent an alternate or additional pathway for neuronal differentiation in these tumors. Evans and colleagues showed that small molecule inhibition of the Trk tyrosine kinases causes significant growth inhibition of neuroblastoma xenografts (Evans et€al. 2001). CEP-701, a small molecule inhibitor of Trk tyrosine kinase, has been shown to have a significant growth inhibitory effect on neuroblastoma in€vivo (Evans et€al. 1999, 2001), providing the rationale for the ongoing Phase I clinical trial of this compound. Other tyrosine kinase inhibitors, including inhibitors of the epidermal growth factor receptor, may have activity against neuroblastoma (Ho et€al. 2005), and are entering clinical trials. Imatinib mesylate has also been studied in neuroblastoma because some tumors appear to express c-KIT and/or PDGFR (Beppu et€al. 2004; Vitali et€al. 2003), but activating mutations in these receptors have not been reported. Screens of the neuroblastoma “kinome” for other potential drug targets, especially those in late stage clinical development for adult malignancies, are ongoing.
The Neuroblastoma Stem Cell Conundrum Evidence suggests that malignancies may arise or be maintained in a stem cell compartment with the attributes of limitless replication and self-renewal. The Notch, Sonic hedgehog and Wnt/b-catenin developmental programs play a critical role in stem cell determination and renewal in diverse tissues and misappropriation of these pathways appears to be a recurring theme in embryonal tumorigenesis (Allenspach et€ al. 2002; Taipale and Beachy 2001; Giles et€ al. 2003). b-Catenin signaling is involved in the maintenance and expansion of neural crest stem cells (Reya and Clevers 2005; Chenn and Walsh 2002) and neural progenitors (Zechner et€al. 2003). An engineered gain-of-function b-catenin allele targeted to neural tissues causes marked neural progenitor expansion by promoting cell cycle entry at the expense of differentiation (Chenn and Walsh 2002). It is intriguing to speculate that this pathway may contribute to maintenance of neuroblastoma stem cells as well, raising the question of whether emerging inhibitors of this pathway might have therapeutic utility. Indeed, our own unpublished data suggest that aberrant b-catenin signaling may play a role in promoting a high-risk phenotype through induction of MYC and additional advantageous target genes in neuroblastomas without MYCN amplification.
Molecular Therapy for Neuroblastoma
359
Stem cells also appear programmed to subvert many death stressors, sparing them from local insults to allow for tissue regeneration and repair. Neural stem cells promote resistance to cytokine and death receptor mediated apoptosis at least partly through developmentally down-regulated caspase-8 and upregulated PEA15. A significant proportion of primary neuroblastomas lack caspase-8 expression (Teitz et€al. 2000), yet ectopic restoration of caspase-8 in deficient cells does not fully restore apoptosis. This suggests that inhibition of apoptosis occurs at multiple levels, and it is possible that the absent caspase-8 expression reflects a developmental program or sustained stem cell feature of neuroblasts rather than a tumor-specific somatic mutation (Goldsmith and Hogarty 2005). This distinction has important implications for experimental therapeutics targeting these latent death-signaling pathways.
Targeted Delivery of Radionucleotides Because neuroblastoma is a radiation sensitive, but also systemic tumor, there has been interest in targeting the delivery of radioactive molecules that are selectively concentrated in neuroblastoma cells. Approaches have included the attachment of radionuclide to metaiodobenzylguanidine (MIBG) (Garaventa et€ al. 1999; Matthay et€ al. 1998a; Tepmongkol and Heyman 1999), somatostatin analogs (Borgstrom et€al. 1999; O’Dorisio et€al. 1994; Wiseman and Kvols 1995), and anti-GD2 antibodies (Cheung et€al. 1998; Kushner et€al. 2001; Yu et€al. 1998). The compound 131I-MIBG is furthest in clinical development. Low dose 131I-MIBG has been shown to be highly effective for disease palliation (Kang et€al. 2003), and a Phase I dose escalation trial established 12€ mCi/kg of 131I-MIBG as the maximum tolerated dose with the only significant toxicity being hematopoietic (Matthay et€ al. 1998a). This trial further showed that doses up to 18€mCi/kg are tolerable with stem cell support for the 1/3 of patients with protracted Grade 4 hematopoietic toxicity. A completed Phase II study showed an objective response rate of approximately 40% in heavily pretreated patients receiving 18€mCi/kg with acceptable toxicity (DuBois et€al. 2004; Howard et€al. 2005). Current clinical investigation is focusing on further dose intensification of MIBG, combining 131 I-MIBG with myeloablative therapies for high-risk patients with a poor response to initial treatment (Yanik et€al. 2002), as well as combining 131I-MIBG with radiosensitizing agents. The European experience combining topotecan and 131I-MIBG has excellent preclinical activity in mouse xenograft models of neuroblastoma, and had no unexpected toxicities in a pilot clinical study. The New Approaches to Neuroblastoma Therapy (NANT) Consortium is leading an ongoing Phase I trial to test the safety of irinotecan and vincristine when given with 131I-MIBG.
Immunotherapy Immunotherapeutic strategies for treating neuroblastoma were originally postulated based on the observations of spontaneous regression, as well as tumor infiltrating
360
Y.P. Mossé and J.M. Maris
lymphocytes in some favorable neuroblastomas, both suggesting a host response to tumor (Cooper et€ al. 2001). Targeted molecules directed against neuroblastomaspecific cellular antigens are a strategy that seems promising for application during the minimal residual disease phase of therapy. Murine, chimeric, and humanized antibodies specific to the cell surface ganglioside GD2, highly expressed in the majority of neuroblastomas, either alone or with cytokines, have shown the activity in preclinical models (Barker et€al. 1991; Cheung et€al. 1986; Kushner and Cheung 1989; Lode et€al. 1999), as well as in Phase I and Phase II clinical trials. Measurable responses have been observed in patients with refractory neuroblastoma (Cheung et€ al. 1998; Murray et€ al. 1994; Yu et€ al. 1998), and the monoclonal antibody directed against the GD2 antigen is now furthest in clinical development with a current Phase III study looking at the chimeric monoclonal antibody ch14.18 in the treatment of minimal residual disease. To maximize the immunogenicity of monoclonal antibodies, particularly in heavily pretreated patients who have chemotherapysuppressed function of immune effector cells, immunocytokines or tumor-specific monoclonal antibodies linked with cytokines have been developed. A humanized GD2 antibody fused to IL2 (Neal et€al. 2004; Yu et€al. 1998; King et€al. 2004) has been investigated in a COG Phase II trial, and future trials are being developed to combine these antibodies with anti-angiogenic drugs. Limiting the broad application of this strategy are the difficulties with antibody production and the development of neutralizing antibodies, even to the chimeric molecules (Neal et€al. 2004; Yu et€al. 1998; King et€al. 2004). Neuropathic pain is also a significant immediate toxicity and can be dose-limiting (Yu et€ al. 1998) (hu14.18-IL2) (Neal et€ al. 2004; King et€al. 2004). Additional immunotherapeutic strategies early in the development include DNA (Bolesta et€al. 2005), cellular (Rousseau et€al. 2003), and anti-idiotypic vaccination strategies (Yu et€al. 2004) as well as using engineered cytolytic T lymphocytes for cellular immunotherapy (Gonzalez et€al. 2004).
Retinoids Based on the results from the randomized trial of 13-cis-retinoic acid following myeloablative chemotherapy (Matthay et€al. 1999), most investigators consider this agent standard of care for high-risk neuroblastoma patients who have a complete or very good partial response following myeloablative chemotherapy and stem cell rescue. Current investigation is focused on finding the retinoid compound that gives optimal systemic exposure allowing for maximal antitumor effect without excess toxicity (Reynolds 2004). Fenretinide is currently the lead retinoid in the development because it has been shown to achieve multilog cell kill in multiple neuroblastoma cell lines, even those resistant to other retinoic acids, at least in part through ceramide upregulation (Maurer et€ al. 1999, 2000). Phase I studies done in the United States and Italy have shown that the drug is generally well-tolerated and a Phase II trial recently completed enrollment within the COG (Garaventa et€al. 2003;
Molecular Therapy for Neuroblastoma
361
Reynolds et€al. 2003). The current Fenretinide formulation, however, has low oral bioavailability requiring young children to take large numbers of capsules. Liquid and intravenous formulations are currently in phase I trials through the NANT to circumvent this issue.
Angiogenesis Inhibitors Tumor vascularity is correlated with an aggressive phenotype in neuroblastoma making angiogenesis inhibitors an attractive therapeutic option. High-risk tumors are highly vascular, and the tumor vascular index is strongly correlated with adverse prognostic features, including MYCN amplification and poor survival probability (Meitar et€al. 1996). In addition, several pro-angiogenic factors, including VEGF, have been shown to be differentially expressed in primary neuroblastomas in a pattern suggesting promotion of an angiogenic phenotype in high-risk tumors (Eggert et€al. 2000; Fotsis et€al. 1999), while low-risk neuroblastomas are characterized by a rich stromal component that produces angiogenesis inhibitors (Chlenski et€ al. 2002; Huang et€ al. 2000). Preclinical therapeutic studies, however, have shown variable efficacy in neuroblastoma (Erdreich-Epstein et€al. 2000; Katzenstein et€al. 1999; Morowitz et€al. 2005; Shusterman et€al. 2001) that is not always predicted by the activity seen in adult and other pediatric tumors, making the translation to the clinic somewhat slow. Strategies focused on neutralization of circulating VEGF are currently in early clinical trials with specific development for neuroblastoma underway. The challenge of providing neovascular inhibition strategies in young children with a developing cardiovascular system need to be considered during preclinical and clinical development.
Targeting MYCN The cancer genes most commonly altered in adult carcinogenesis (e.g., TP53, CDKN2A, RAS) are rarely aberrant in neuroblastoma. TP53 inactivating mutations are infrequent in primary tumors, although they have been documented in cell lines at relapse (Carr et€al. 2006; Keshelava et€al. 2001; Tweddle et€al. 2001). Homozygous deletion of CDKN2A (INK4A/p16) has been identified in a subset of neuroblastoma cell lines (Thompson et€al. 2001), but there is no consistent evidence for inactivation of this locus in primary tumors. Finally, though persuasive evidence supports RAS and MYC gene cooperation in tumorigenesis (Sears et€al. 2000; Yaari et€al. 2005), activation of RAS does not appear to constitute a preferred secondary pathway for neuroblastomas, even those with MYCN amplification (Dam et€ al. 2006). Thus, major oncogenic pathways governing human neoplasia do not appear deregulated in neuroblastoma with the exception of MYCN in a subset. The MYCN protein provides an obvious potential therapeutic target, but it is so massively dysregulated in neuroblastoma cells with amplification that this has been
362
Y.P. Mossé and J.M. Maris
difficult to date. Constitutive activation of the BDNF:TrkB signaling pathway in high-risk neuroblastoma provides another rationale target that might be exploited. Evans and colleagues showed that small molecule inhibition of the Trk tyrosine kinases causes significant growth inhibition of neuroblastoma xenografts (Evans et€ al. 1999, 2001), providing the rationale for an ongoing Phase I trial. Another strategy currently being explored is the use of demethylating agents such as Decitabine. The rationale for this approach is that methylation of genes critical for apoptosis such as caspase 8 appear to occur frequently in neuroblastomas, particularly in association with MYCN amplification (Teitz et€al. 2000). Histone deacetylase inhibitors have demonstrated preclinical activity against neuroblastoma (Coffey et€al. 2001; Jaboin et€al. 2002b), and there are several in clinical trials for patients with refractory solid tumors. With the relatively small number of pediatric patients available for experimental therapeutic trials and the growing number of potential drugs to be tested, it is becoming increasingly important to have firm biological rationale and evidence of efficacy in appropriate preclinical models to help prioritize drug development. Many other strategies, some of which are listed in Table€1 are under development in the laboratory or in early phase clinical trials.
Mitotic Spindle Inhibition Among several potential mechanisms for chromosome instability in cancer cells, much attention has recently been given to functional abnormality of centrosomes because of its prevalence in the majority of solid tumors as well as leukemias and lymphomas, and a strong association between centrosome abnormality and aneuploidy in cancers (D’Assoro et€al. 2002; Fukasawa 2005, 2007). The Aurora family of serine/threonine protein kinases plays a critical role in the regulation of chromosomal segregation and cytokinesis during mitotic progression. The Aurora A kinase gene is amplified and/or overexpressed in many tumors, including colon, breast, pancreatic, and bladder cancers, and its overexpression results in the transformation of normal cells, thus supporting its role as an oncogene. Preliminary data suggest that the Aurora A kinase gene is overexpressed in neuroblastoma cells and that selective small molecule inhibition of this kinase has potential application in this disease.
Other Strategies Epigenetic silencing of genes that are critical for induction of programmed cell death, such as caspase 8, appears to occur frequently in neuroblastoma (Teitz et€al. 2000). Therefore, demethylating agents such as decitabine are currently being studied. Histone deacetylase inhibitors have also demonstrated preclinical activity against neuroblastoma (Coffey et€al. 2001; Jaboin et€al. 2002b). At least three histone deacetylase inhibitors are now in clinical trials for patients with refractory solid
Molecular Therapy for Neuroblastoma
363
tumors. Heat shock protein 90 inhibitors are also of interest because these agents alter the function of molecules associated with neuroblastoma cell growth and proliferation, including the Type I insulin-like growth factor receptor AKT, and TrkB. With an expanding portfolio of potential drugs to be tested in pediatric Phase I clinical trials, it is becoming increasingly important to have a firm biological rationale and evidence of efficacy in preclinical models to help prioritize drug development.
Future Directions and Challenges Despite extensive literature correlating common genomic alterations with disease outcome, no bona fide target genes have been identified in neuroblastoma with the exception of MYCN. Hemizygous deletions are typically dozens of megabases without demonstrable biallelic inactivation of regional candidate genes, suggesting that complex multiple gene repression and/or haploinsufficiency may be involved. Whole genome copy number alterations (CNAs) detected with array-based CGH have confirmed the true complexity of chromosomal aberrations in neuroblastoma (Mosse et€ al. 2005b), identified distinct genetic subgroups (Vandesompele et€ al. 2001; Wang et€al. 2006) and inferred models for progression (Bilke et€al. 2005). It is likely that refined platforms will replace region-specific methods to detect crucial CNAs used in risk-stratification schemas, and enhance gene discovery efforts (Beheshti et€ al. 2003; Chen et€ al. 2004; Mosse et€ al. 2005a). Pathway discovery may be further facilitated by integrating genomic DNA alterations with mRNA expression profiles. Transcriptional profiles have been used to define prognostic signatures (Schramm et€ al. 2005) and regional CNAs have been correlated with transcriptional alterations (Wang et€al. 2006). Biological pathways engaged or suppressed as a result of recurring CNAs are being sought. To date, these include upregulation of genes involved in host immune response and antigen processing in high-risk neuroblastomas without MYCN amplification, as well as enrichment for sympathoadrenal developmental genes in low-risk neuroblastomas (Wang et€ al. 2006). Ultimately, improvements in survival will likely result from innovative treatment approaches based upon a better understanding of the critical biological pathways responsible for neuroblastoma initiation and progression.
References Alimonti A, Satta F, Pavese I, Burattini E, Zoffoli V, Vecchione A. 2003. Prevention of irinotecan plus 5-fluorouracil/leucovorin-induced diarrhoea by oral administration of neomycin plus bacitracin in first-line treatment of advanced colorectal cancer. Ann Oncol 14(5):805–6. Allenspach EJ, Maillard I, Aster JC, Pear WS. 2002. Notch signaling in cancer. Cancer Biol Ther 1(5):466–76. Alvarado CS, London WB, Look AT, Brodeur GM, Altmiller DH, Thorner PS, Joshi VV, Rowe ST, Nash MB, Smith EI, et€al. 2000. Natural history and biology of stage a neuroblastoma: a Pediatric Oncology Group Study. J Pediatr Hematol Oncol 22(3):197–205.
364
Y.P. Mossé and J.M. Maris
Anderson CP, Reynolds CP. 2002. Synergistic cytotoxicity of buthionine sulfoximine (BSO) and intensive melphalan (L-PAM) for neuroblastoma cell lines established at relapse after myeloablative therapy. Bone Marrow Transplant 30(3):135–40. Attiyeh EF, London WB, Mosse YP, Wang Q, Winter C, Khazi D, McGrady PW, Seeger RC, Look AT, Shimada H, et€al. 2005. Chromosome 1p and 11q deletions and outcome in neuroblastoma. N Engl J Med 353(21):2243–53. Azar C, Scavarda NJ, Reynolds CP, Brodeur GM. 1990. Multiple defects of the nerve growth factor receptor in human neuroblastomas. Cell Growth Differ 1:421–428. Bagatell R, Gore L, Egorin MJ, Ho R, Heller G, Boucher N, Zuhowski EG, Whitlock JA, Hunger SP, Narendran A, et€ al. 2007. Phase I pharmacokinetic and pharmacodynamic study of 17-N-allylamino-17-demethoxygeldanamycin in pediatric patients with recurrent or refractory solid tumors: a pediatric oncology experimental therapeutics investigators consortium study. Clin Cancer Res 13(6):1783–8. Baker DL, Reddy UR, Pleasure D, Thorpe CL, Evans AE, Cohen PS, Ross AH. 1989. Analysis of nerve growth factor receptor expression in human neuroblastoma and neuroepithelioma cell lines. Cancer Res 49:4142–6. Banerjee D. 2001. Genasense (Genta Inc). Curr Opin Investig Drugs 2(4):574–80. Barker E, Mueller BM, Handgretinger R, Herter M, Yu AL, Reisfeld RA. 1991. Effect of a chimeric anti-ganglioside GD2 antibody on cell-mediated lysis of human neuroblastoma cells. Cancer Res 51(1):144–9. Beheshti B, Braude I, Marrano P, Thorner P, Zielenska M, Squire JA. 2003. Chromosomal localization of DNA amplifications in neuroblastoma tumors using cDNA microarray comparative genomic hybridization. Neoplasia 5(1):53–62. Beppu K, Jaboine J, Merchant MS, Mackall CL, Thiele CJ. 2004. Effect of imatinib mesylate on neuroblastoma tumorigenesis and vascular endothelial growth factor expression. J Natl Cancer Inst 96(1):46–55. Berthold F, Boos J, Burdach S, Erttmann R, Henze G, Hermann J, Klingebiel T, Kremens B, Schilling FH, Schrappe M, et€al. 2005. Myeloablative megatherapy with autologous stem-cell rescue versus oral maintenance chemotherapy as consolidation treatment in patients with highrisk neuroblastoma: a randomised controlled trial. Lancet Oncol 6(9):649–58. Berthold F, Treuner J, Brandeis WE, et€al. 1982. Neuroblastoma study NBL-79 of the German Society for Pediatric Oncology; report after 2 years. Klin Padiatr 194:262–9. Bilke S, Chen QR, Whiteford CC, Khan J. 2005. Detection of low level genomic alterations by comparative genomic hybridization based on cDNA micro-arrays. Bioinformatics 21(7):1138–45. Bolesta E, Kowalczyk A, Wierzbicki A, Rotkiewicz P, Bambach B, Tsao CY, Horwacik I, Kolinski A, Rokita H, Brecher M, et€al. 2005. DNA vaccine expressing the mimotope of GD2 ganglioside induces protective GD2 cross-reactive antibody responses. Cancer Res 65(8):3410–8. Borgstrom P, Hassan M, Wassberg E, Refai E, Jonsson C, Larsson SA, Jacobsson H, Kogner P. 1999. The somatostatin analogue octreotide inhibits neuroblastoma growth in€vivo. Pediatr Res 46(3):328–32. Brignole C, Marimpietri D, Pastorino F, Nico B, Di Paolo D, Cioni M, Piccardi F, Cilli M, Pezzolo A, Corrias MV, et€al. 2006. Effect of bortezomib on human neuroblastoma cell growth, apoptosis, and angiogenesis. J Natl Cancer Inst 98(16):1142–57. Brodeur GM. 2003. Neuroblastoma: biological insights into a clinical enigma. Nat Rev Cancer 3(3):203–16. Brodeur GM, Nakagawara A, Yamashiro DJ, Ikegaki N, Liu X-G, Lee CP, Evans AE. 1996. Expression of TrkA, TrkB, and TrkC in human neuroblastomas. J Neurooncol 12:37–41. Carr J, Bell E, Pearson AD, Kees UR, Beris H, Lunec J, Tweddle DA. 2006. Increased frequency of aberrations in the p53/MDM2/p14(ARF) pathway in neuroblastoma cell lines established at relapse. Cancer Res 66(4):2138–45. Chen QR, Bilke S, Wei JS, Whiteford CC, Cenacchi N, Krasnoselsky AL, Greer BT, Son CG, Westermann F, Berthold F, et€al. 2004. cDNA array-CGH profiling identifies genomic alterations specific to stage and MYCN-amplification in neuroblastoma. BMC Genomics 5(1):70.
Molecular Therapy for Neuroblastoma
365
Chenn A, Walsh CA. 2002. Regulation of cerebral cortical size by control of cell cycle exit in neural precursors. Science 297(5580):365–9. Cheung IY, Cheung NK. 2001. Quantitation of marrow disease in neuroblastoma by real-time reverse transcription-PCR. Clin Cancer Res 7(6):1698–705. Cheung NK, Kushner BH, Yeh SDJ, Larson SM. 1998. 3F8 monoclonal antibody treatment of patients with stage 4 neuroblastoma: a phase II study. Int J Oncol 12(6):1299–306. Cheung NK, Landmeier B, Neely J, Nelson AD, Abramowsky C, Ellery S, Adams RB, Miraldi F. 1986. Complete tumor ablation with iodine 131-radiolabeled disialoganglioside GD2-specific monoclonal antibody against human neuroblastoma xenografted in nude mice. J Natl Cancer Inst 77(3):739–45. Cheung NK, Modak S. 2002. Oral (1â•›→â•›3), (1â•›→ 4)-beta-d-glucan synergizes with antiganglioside GD2 monoclonal antibody 3F8 in the therapy of neuroblastoma. Clin Cancer Res 8(5):1217–23. Chlenski A, Liu S, Crawford SE, Volpert OV, DeVries GH, Evangelista A, Yang Q, Salwen HR, Farrer R, Bray J, et€al. 2002. SPARC is a key Schwannian-derived inhibitor controlling neuroblastoma tumor angiogenesis. Cancer Res 62(24):7357–63. Cho SY, Adamson PC, Hagey AE, Widemann BC, Maris JM, Fox E, Medina D, Cui Y, Gordon GB, Balis FM. 2004. Phase I trial and pharmacokinetic (PK) study of ABT-751, an orally bioavailable tubulin binding agent, in pediatric patients with refractory solid tumors. J Clin Oncol ASCO Annual Meeting Proceedings 22(14S):2080 Coffey DC, Kutko MC, Glick RD, Butler LM, Heller G, Rifkind RA, Marks PA, Richon VM, La Quaglia MP. 2001. The histone deacetylase inhibitor, CBHA, inhibits growth of human neuroblastoma xenografts in€vivo, alone and synergistically with all-trans retinoic acid. Cancer Res 61(9):3591–4. Cohn SL, Look AT, Joshi VV, Holbrook T, Salwen H, Chagnovich D, Chesler L, Rowe ST, Valentine MB, Komuro H, et€al. 1995. Lack of correlation of N-myc gene amplification with prognosis in localized neuroblastoma: a Pediatric Oncology Group Study. Cancer Res 55:721–726. Cooper R, Khakoo Y, Matthay KK, Lukens JN, Seeger RC, Stram DO, Gerbing RB, Nakagawa A, Shimada H. 2001. Opsoclonus-myoclonus-ataxia syndrome in neuroblastoma: histopathologic features – a report from the Children’s Cancer Group. Med Pediatr Oncol 36(6):623–9. D’Assoro AB, Lingle WL, Salisbury JL. 2002. Centrosome amplification and the development of cancer. Oncogene 21(40):6146–53. Dam V, Morgan BT, Mazanek P, Hogarty MD. 2006. Mutations in PIK3CA are infrequent in neuroblastoma. BMC Cancer 6:177. Dickson PV, Hamner JB, Sims TL, Fraga CH, Ng CY, Rajasekeran S, Hagedorn NL, McCarville MB, Stewart CF, Davidoff AM. 2007. Bevacizumab-induced transient remodeling of the vasculature in neuroblastoma xenografts results in improved delivery and efficacy of systemically administered chemotherapy. Clin Cancer Res 13(13):3942–50. DuBois SG, Messina J, Maris JM, Huberty J, Glidden DV, Veatch J, Charron M, Hawkins R, Matthay KK. 2004. Hematologic toxicity of high-dose iodine-131-metaiodobenzylguanidine therapy for advanced neuroblastoma. J Clin Oncol 22(12):2452–60. Eggert A, Ikegaki N, Kwiatkowski J, Zhao H, Brodeur GM, Himelstein BP. 2000. High-level expression of angiogenic factors is associated with advanced tumor stage in human neuroblastomas. Clin Cancer Res 6(5):1900–8. Erdreich-Epstein A, Shimada H, Groshen S, Liu M, Metelitsa LS, Kim KS, Stins MF, Seeger RC, Durden DL. 2000. Integrins alpha(v)beta3 and alpha(v)beta5 are expressed by endothelium of high-risk neuroblastoma and their inhibition is associated with increased endogenous ceramide. Cancer Res 60(3):712–21. Evans AE, Kisselbach KD, Liu X, Eggert A, Ikegaki N, Camoratto AM, Dionne C, Brodeur GM. 2001. Effect of CEP-751 (KT-6587) on neuroblastoma xenografts expressing TrkB. Med Pediatr Oncol 36(1):181–4. Evans AE, Kisselbach KD, Yamashiro DJ, Ikegaki N, Camoratto AM, Dionne CA, Brodeur GM. 1999. Antitumor activity of CEP-751 (KT-6587) on human neuroblastoma and medulloblastoma xenografts. Clin Cancer Res 5(11):3594–602.
366
Y.P. Mossé and J.M. Maris
Evans AE, Silber JH, Shpilsky A, D’Angio GJ. 1996. Successful management of low-stage neuroblastoma without adjuvant therapies: a comparison of two decades, 1972 through 1981 and 1982 through 1992, in a single institution. J Clin Oncol 14(9):2504–10. Fotsis T, Breit S, Lutz W, Rossler J, Hatzi E, Schwab M, Schweigerer L. 1999. Down-regulation of endothelial cell growth inhibitors by enhanced MYCN oncogene expression in human neuroblastoma cells. Eur J Biochem 263(3):757–64. Fox E, Maris JM, Widemann BC, Meek K, Goodwin A, Goodspeed W, Kromplewski M, Fouts ME, Medina D, Cho SY, et€ al. 2006. A phase 1 study of ABT-751, an orally bioavailable tubulin inhibitor, administered daily for 7 days every 21 days in pediatric patients with solid tumors. Clin Cancer Res 12(16):4882–7. Frantz CN, London WB, Diller L, Seeger R, Sawyer K. 2004. Recurrent neuroblastoma: Randomized treatment with topotecanâ•›+â•›cyclophosphamide (Tâ•›+â•›C) vs. topotecan alone(T). A POG/ CCG Intergroup Study. J Clin Oncol ASCO Annual Meeting Proceedings 22(14S):8512 Fukasawa K. 2005. Centrosome amplification, chromosome instability and cancer development. Cancer Lett 230(1):6–19. Fukasawa K. 2007. Oncogenes and tumour suppressors take on centrosomes. Nat Rev Cancer 7(12):911–24. Garaventa A, Bellagamba O, Lo Piccolo MS, Milanaccio C, Lanino E, Bertolazzi L, Villavecchia GP, Cabria M, Scopinaro G, Claudiani F, et€al. 1999. 131I-metaiodobenzylguanidine (131I-MIBG) therapy for residual neuroblastoma: a mono-institutional experience with 43 patients. Br J Cancer 81(8):1378–84. Garaventa A, Luksch R, Lo Piccolo MS, Cavadini E, Montaldo PG, Pizzitola MR, Boni L, Ponzoni M, Decensi A, De Bernardi B, et€ al. 2003. Phase I trial and pharmacokinetics of fenretinide in children with neuroblastoma. Clin Cancer Res 9(6):2032–9. George R, Attiyeh E, Li S, Moreau L, Neuberg D, Li C, Fox EA, Meyerson M, Diller L, Fontina P, et€ al. 2007. Genome-wide analysis of neuroblastomas using high-density single nucleotide polymorphism arrays. PLoS One 2:e255. Giles RH, van Es JH, Clevers H. 2003. Caught up in a Wnt storm: Wnt signaling in cancer. Biochim Biophys Acta 1653(1):1–24. Goldsmith KC, Hogarty MD. 2005. Targeting programmed cell death pathways with experimental therapeutics: opportunities in high-risk neuroblastoma. Cancer Lett 228(1–2):133–41. Gonzalez S, Naranjo A, Serrano LM, Chang WC, Wright CL, Jensen MC. 2004. Genetic engineering of cytolytic T lymphocytes for adoptive T-cell therapy of neuroblastoma. J Gene Med 6(6):704–11. Ho R, Eggert A, Hishiki T, Minturn JE, Ikegaki N, Foster P, Camoratto AM, Evans AE, Brodeur GM. 2002. Resistance to chemotherapy mediated by TrkB in neuroblastomas. Cancer Res 62(22):6462–6. Ho R, Minturn JE, Hishiki T, Zhao H, Wang Q, Cnaan A, Maris J, Evans AE, Brodeur GM. 2005. Proliferation of human neuroblastomas mediated by the epidermal growth factor receptor. Cancer Res 65(21):9868–75. Houghton PJ, Morton CL, Tucker C, Payne D, Favours E, Cole C, Gorlick R, Kolb EA, Zhang W, Lock R, et€al. 2007. The pediatric preclinical testing program: description of models and early testing results. Pediatr Blood Cancer 49(7):928–40. Howard JP, Maris JM, Kersun LS, Huberty JP, Cheng SC, Hawkins RA, Matthay KK. 2005. Tumor response and toxicity with multiple infusions of high dose (131)I-MIBG for refractory neuroblastoma. Pediatr Blood Cancer 44(3):232–9. Huang D, Rutkowski JL, Brodeur GM, Chou PM, Kwiatkowski JL, Babbo A, Cohn SL. 2000. Schwann cell-conditioned medium inhibits angiogenesis. Cancer Res 60(21):5966–71. Jaboin J, Kim CJ, Kaplan DR, Thiele CJ. 2002a. Brain-derived neurotrophic factor activation of TrkB protects neuroblastoma cells from chemotherapy-induced apoptosis via phosphatidylinositol 3¢-kinase pathway. Cancer Res 62(22):6756–63. Jaboin J, Wild J, Hamidi H, Khanna C, Kim CJ, Robey R, Bates SE, Thiele CJ. 2002b. MS-27275, an inhibitor of histone deacetylase, has marked in€ vitro and in€ vivo antitumor activity against pediatric solid tumors. Cancer Res 62(21):6108–15.
Molecular Therapy for Neuroblastoma
367
Kang TI, Brophy P, Hickeson M, Heyman S, Evans AE, Charron M, Maris JM. 2003. Targeted radiotherapy with submyeloablative doses of 131I-MIBG is effective for disease palliation in highly refractory neuroblastoma. J Pediatr Hematol Oncol 25(10):769–73. Katzenstein HM, Rademaker AW, Senger C, Salwen HR, Nguyen NN, Thorner PS, Litsas L, Cohn SL. 1999. Effectiveness of the angiogenesis inhibitor TNP-470 in reducing the growth of human neuroblastoma in nude mice inversely correlates with tumor burden. Clin Cancer Res 5(12):4273–8. Keshelava N, Davicioni E, Wan Z, Ji L, Sposto R, Triche TJ, Reynolds CP. 2007. Histone deacetylase 1 gene expression and sensitization of multidrug-resistant neuroblastoma cell lines to cytotoxic agents by depsipeptide. J Natl Cancer Inst 99(14):1107–19. Keshelava N, Zuo JJ, Chen P, Waidyaratne SN, Luna MC, Gomer CJ, Triche TJ, Reynolds CP. 2001. Loss of p53 function confers high-level multidrug resistance in neuroblastoma cell lines. Cancer Res 61(16):6185–93. King DM, Albertini MR, Schalch H, Hank JA, Gan J, Surfus J, Mahvi D, Schiller JH, Warner T, Kim K, et€al. 2004. Phase I clinical trial of the immunocytokine EMD 273063 in melanoma patients. J Clin Oncol 22(22):4463–73. Kogner P, Barbany G, Dominici C, Castello MA, Raschella G, Persson H. 1993. Coexpression of messenger RNA for TRK protooncogene and low affinity nerve growth factor receptor in neuroblastoma with favorable prognosis. Cancer Res 53:2044–50. Kushner BH, Cheung NK. 1989. GM-CSF enhances 3F8 monoclonal antibody-dependent cellular cytotoxicity against human melanoma and neuroblastoma. Blood 73(7):1936–41. Kushner BH, Cheung NK, LaQuaglia MP, Ambros PF, Ambros IM, Bonilla MA, Gerald WL, Ladanyi M, Gilbert F, Rosenfield NS, et€ al. 1996a. Survival from locally invasive or widespread neuroblastoma without cytotoxic therapy. J Clin Oncol 14(2):373–81. Kushner BH, Cheung NK, LaQuaglia MP, Ambros PF, Ambros IM, Bonilla MA, Ladanyi M, Gerald WL. 1996b. International neuroblastoma staging system stage 1 neuroblastoma: a prospective study and literature review. J Clin Oncol 14(7):2174–80. Kushner BH, Kramer K, Cheung NK. 2001. Phase II trial of the anti-GD2 monoclonal antibody 3F8 and granulocyte-macrophage colony-stimulating factor for neuroblastoma. J Clin Oncol 19:4189–94. Langler A, Christaras A, Abshagen K, Krauth K, Hero B, Berthold F. 2002. Topotecan in the treatment of refractory neuroblastoma and other malignant tumors in childhood – a phase-IIstudy. Klin Padiatr 214(4):153–6. Lode HN, Xiang R, Duncan SR, Theofilopoulos AN, Gillies SD, Reisfeld RA. 1999. Tumortargeted IL-2 amplifies T cell-mediated immune response induced by gene therapy with singlechain IL-12. Proc Natl Acad Sci USA 96(15):8591–6. Maris JM. 2005. The biologic basis for neuroblastoma heterogeneity and risk stratification. Curr Opin Pediatr 17(1):7–13. Maris JM, Courtright J, Houghton PJ, Morton CL, Gorlick R, Kolb EA, Lock R, Tajbakhsh M, Reynolds CP, Keir ST, et€ al. 2008. Initial testing of the VEGFR inhibitor AZD2171 by the pediatric preclinical testing program. Pediatr Blood Cancer 50(3):581–7. Marks PA, Richon VM, Breslow R, Rifkind RA. 2001. Histone deacetylase inhibitors as new cancer drugs. Curr Opin Oncol 13(6):477–83. Marshall JL, Kindler H, Deeken J, Bhargava P, Vogelzang NJ, Rizvi N, Luhtala T, Boylan S, Dordal M, Robertson P, et€ al. 2005. Phase I trial of orally administered CEP-701, a novel neurotrophin receptor-linked tyrosine kinase inhibitor. Invest New Drugs 23(1):31–7. Matthay KK, DeSantes K, Hasegawa B, Huberty J, Hattner RS, Ablin A, Reynolds CP, Seeger RC, Weinberg VK, Price D. 1998a. Phase I dose escalation of 131I-metaiodobenzylguanidine with autologous bone marrow support in refractory neuroblastoma. J Clin Oncol 16(1):229–36. Matthay KK, Perez C, Seeger RC, Brodeur GM, Shimada H, Atkinson JB, Black CT, Gerbing R, Haase GM, Stram DO, et€al. 1998b. Successful treatment of stage III neuroblastoma based on prospective biologic staging: a Children’s Cancer Group Study. J Clin Oncol 16(4):1256–64. Matthay KK, Sather HN, Seeger RC, Haase GM, Hammond GD. 1989. Excellent outcome of stage II neuroblastoma is independent of residual disease and radiation therapy. J Clin Oncol 7(2):236–44.
368
Y.P. Mossé and J.M. Maris
Matthay KK, Tan JC, Villablanca JG, Yanik GA, Veatch J, Franc B, Twomey E, Horn B, Reynolds CP, Groshen S, et€al. 2006. Phase I dose escalation of iodine-131-metaiodobenzylguanidine with myeloablative chemotherapy and autologous stem-cell transplantation in refractory neuroblastoma: a new approaches to Neuroblastoma Therapy Consortium Study. J Clin Oncol 24(3):500–6. Matthay KK, Villablanca JG, Seeger RC, Stram DO, Harris RE, Ramsay NK, Swift P, Shimada H, Black CT, Brodeur GM, et€ al. 1999. Treatment of high-risk neuroblastoma with intensive chemotherapy, radiotherapy, autologous bone marrow transplantation, and 13-cis-retinoic acid. Children’s Cancer Group. N Engl J Med 341(16):1165–73. Matthay KK, Yanik G, Messina J, Quach A, Huberty J, Cheng SC, Veatch J, Goldsby R, Brophy P, Kersun LS, et€al. 2007. Phase II study on the effect of disease sites, age, and prior therapy on response to iodine-131-metaiodobenzylguanidine therapy in refractory neuroblastoma. J Clin Oncol 25(9):1054–60. Maurer BJ, Melton L, Billups C, Cabot MC, Reynolds CP. 2000. Synergistic cytotoxicity in solid tumor cell lines between N-(4-hydroxyphenyl)retinamide and modulators of ceramide metabolism. J Natl Cancer Inst 92(23):1897–909. Maurer BJ, Metelitsa LS, Seeger RC, Cabot MC, Reynolds CP. 1999. Increase of ceramide and induction of mixed apoptosis/necrosis by N-(4-hydroxyphenyl)-retinamide in neuroblastoma cell lines [see comments]. J Natl Cancer Inst 91(13):1138–46. Meitar D, Crawford SE, Rademaker AW, Cohn SL. 1996. Tumor angiogenesis correlates with metastatic disease, N-Myc amplification, and poor outcome in human neuroblastoma. J Clin Oncol 14(2):405–14. Miller BS, Yee D. 2005. Type I insulin-like growth factor receptor as a therapeutic target in cancer. Cancer Res 65(22):10123–7. Morowitz MJ, Barr R, Wang Q, King R, Rhodin N, Pawel B, Zhao H, Erickson SA, Sheppard GS, Wang J, et€al. 2005. Methionine aminopeptidase 2 inhibition is an effective treatment strategy for neuroblastoma in preclinical models. Clin Cancer Res 11(7):2680–5. Mosse YP, Greshock J, Margolin A, Naylor T, Cole K, Khazi D, Hii G, Winter C, Shahzad S, Asziz MU, et€al. 2005a. High-resolution detection and mapping of genomic DNA alterations in neuroblastoma. Genes Chromosomes Cancer 43(4):390–403. Mosse YP, Greshock J, Weber BL, Maris JM. 2005b. Measurement and relevance of neuroblastoma DNA copy number changes in the post-genome era. Cancer Lett 228(1–2):83–90. Murray JL, Cunningham JE, Brewer H, Mujoo K, Zukiwski AA, Podoloff DA, Kasi LP, Bhadkamkar V, Fritsche HA, Benjamin RS, et€al. 1994. Phase I trial of murine monoclonal antibody 14G2a administered by prolonged intravenous infusion in patients with neuroectodermal tumors. J Clin Oncol 12(1):184–93. Nakagawara A, Arima M, Azar CG, Scavarda NJ, Brodeur GM. 1992. Inverse relationship between trk expression and N-myc amplification in human neuroblastomas. Cancer Res 52:1364–1368. Nakagawara A, Arima-Nakagawara M, Scavarda NJ, Azar CG, Cantor AB, Brodeur GM. 1993. Association between high levels of expression of the TRK gene and favorable outcome in human neuroblastoma. N Engl J Med 328(12):847–54. Nakagawara A, Azar CG, Scavarda NJ, Brodeur GM. 1994. Expression and function of TRK-B and BDNF in human neuroblastomas. Mol Cell Biol 14:759–67. Neal ZC, Yang JC, Rakhmilevich AL, Buhtoiarov IN, Lum HE, Imboden M, Hank JA, Lode HN, Reisfeld RA, Gillies SD, et€ al. 2004. Enhanced activity of hu14.18-IL2 immunocytokine against murine NXS2 neuroblastoma when combined with interleukin 2 therapy. Clin Cancer Res 10(14):4839–47. Nishihira H, Toyoda Y, Tanaka Y, Ijiri R, Aida N, Takeuchi M, Ohnuma K, Kigasawa H, Kato K, Nishi T. 2000. Natural course of neuroblastoma detected by mass screening: s 5-year prospective study at a single institution. J Clin Oncol 18(16):3012–7. O’Dorisio MS, Hauger M, Cecalupo AJ. 1994. Somatostatin receptors in neuroblastoma: diagnostic and therapeutic implications. Semin Oncol 21(5 Suppl 13):33–7. Osenga KL, Hank JA, Albertini MR, Gan J, Sternberg AG, Eickhoff J, Seeger RC, Matthay KK, Reynolds CP, Twist C, et€al. 2006. A phase I clinical trial of the hu14.18-IL2 (EMD 273063)
Molecular Therapy for Neuroblastoma
369
as a treatment for children with refractory or recurrent neuroblastoma and melanoma: a study of the Children’s Oncology Group. Clin Cancer Res 12(6):1750–9. Oue T, Inoue M, Yoneda A, Kubota A, Okuyama H, Kawahara H, Nishikawa M, Nakayama M, Kawa K. 2005. Profile of neuroblastoma detected by mass screening, resected after observation without treatment: results of the Wait and See pilot study. J Pediatr Surg 40(2):359–63. Ozkaynak MF, Sondel PM, Krailo MD, Gan J, Javorsky B, Reisfeld RA, Matthay KK, Reaman GH, Seeger RC. 2000. Phase I study of chimeric human/murine anti-ganglioside G(D2) monoclonal antibody (ch14.18) with granulocyte-macrophage colony-stimulating factor in children with neuroblastoma immediately after hematopoietic stem-cell transplantation: a Children’s Cancer Group Study. J Clin Oncol 18(24):4077–85. Perez CA, Matthay KK, Atkinson JB, Seeger RC, Shimada H, Haase GM, Stram DO, Gerbing RB, Lukens JN. 2000. Biologic variables in the outcome of stages I and II neuroblastoma treated with surgery as primary therapy: a Children’s Cancer Group Study. J Clin Oncol 18(1):18–26. Reya T, Clevers H. 2005. Wnt signalling in stem cells and cancer. Nature 434(7035):843–50. Reynolds CP. 2004. Detection and treatment of minimal residual disease in high-risk neuroblastoma. Pediatr Transplant 8(Suppl 5):56–66. Reynolds CP, Matthay KK, Villablanca JG, Maurer BJ. 2003. Retinoid therapy of high-risk neuroblastoma. Cancer Lett 197(1–2):185–92. Reynolds CP, Maurer BJ, Kolesnick RN. 2004. Ceramide synthesis and metabolism as a target for cancer therapy. Cancer Lett 206(2):169–80. Rodriguez-Galindo C, Crews KR, Stewart CF, Furman W, Panetta JC, Daw NC, Cain A, Tan M, Houghton PH, Santana VM. 2006. Phase I study of the combination of topotecan and irinotecan in children with refractory solid tumors. Cancer Chemother Pharmacol 57:15–24. Rousseau RF, Haight AE, Hirschmann-Jax C, Yvon ES, Rill DR, Mei Z, Smith SC, Inman S, Cooper K, Alcoser P, et€al. 2003. Local and systemic effects of an allogeneic tumor cell vaccine combining transgenic human lymphotactin with interleukin-2 in patients with advanced or refractory neuroblastoma. Blood 101(5):1718–26. Ryden M, Sehgal R, Dominici C, Schilling FH, Ibanez CF, Kogner P. 1996. Expression of mRNA for the neurotrophin receptor trkC in neuroblastomas with favourable tumour stage and good prognosis. Br J Cancer 74(5):773–9. Saylors RL, 3rd, Stine KC, Sullivan J, Kepner JL, Wall DA, Bernstein ML, Harris MB, Hayashi R, Vietti TJ. 2001. Cyclophosphamide plus topotecan in children with recurrent or refractory solid tumors: a Pediatric Oncology Group phase II study. J Clin Oncol 19(15):3463–9. Schmidt ML, Lal A, Seeger RC, Maris JM, Shimada H, O’Leary M, Gerbing RB, Matthay KK. 2005. Favorable prognosis for patients 12 to 18 months of age with stage 4 nonamplified MYCN neuroblastoma: a Children’s Cancer Group Study. J Clin Oncol 23(27):6474–80. Schmidt ML, Lukens JN, Seeger RC, Brodeur GM, Shimada H, Gerbing RB, Stram DO, Perez C, Haase GM, Matthay KK. 2000. Biologic factors determine prognosis in infants with stage IV neuroblastoma: a prospective Children’s Cancer Group Study. J Clin Oncol 18(6):1260–8. Schramm A, Schulte JH, Klein-Hitpass L, Havers W, Sieverts H, Berwanger B, Christiansen H, Warnat P, Brors B, Eils J, et€al. 2005. Prediction of clinical outcome and biological characterization of neuroblastoma by expression profiling. Oncogene 24(53):7902–12. Sears R, Nuckolls F, Haura E, Taya Y, Tamai K, Nevins JR. 2000. Multiple Ras-dependent phosphorylation pathways regulate Myc protein stability. Genes Dev 14(19):2501–14. Shah NP, Tran C, Lee FY, Chen P, Norris D, Sawyers CL. 2004. Overriding imatinib resistance with a novel ABL kinase inhibitor. Science 305(5682):399–401. Shusterman S, Grupp SA, Barr R, Carpentieri D, Zhao H, Maris JM. 2001. The angiogenesis inhibitor TNP-470 effectively inhibits human neuroblastoma xenograft growth, especially in the setting of subclinical disease. Clin Cancer Res 7(4):977–84. Siapati KE, Barker S, Kinnon C, Michalski A, Anderson R, Brickell P, Thrasher AJ, Hart SL. 2003. Improved antitumour immunity in murine neuroblastoma using a combination of IL-2 and IL-12. Br J Cancer 88(10):1641–8. Sidell N. 1982. Retinoic acid-induced growth inhibition and morphologic differentiation of human neuroblastoma cells in€vitro. J Natl Cancer Inst 68(4):589–96.
370
Y.P. Mossé and J.M. Maris
Slack C, Overton PM, Tuxworth RI, Chia W. 2007. Asymmetric localisation of Miranda and its cargo proteins during neuroblast division requires the anaphase-promoting complex/cyclosome. Development 134(21):3781–7. Sohara Y, Shimada H, Scadeng M, Pollack H, Yamada S, Ye W, Reynolds CP, DeClerck YA. 2003. Lytic bone lesions in human neuroblastoma xenograft involve osteoclast recruitment and are inhibited by bisphosphonate. Cancer Res 63(12):3026–31. Suita S, Zaizen Y, Sera Y, Takamatsu H, Mizote H, Ohgami H, Kurosaki N, Ueda K, Tasaka H, Miyazaki S, et€al. 1996. Mass screening for neuroblastoma: quo vadis? A 9-year experience from the Pediatric Oncology Study Group of the Kyushu area in Japan. J Pediatr Surg 31(4):555–8. Suzuki T, Bogenmann E, Shimada H, Stram D, Seeger RC. 1993. Lack of high-affinity nerve growth factor receptors in aggressive neuroblastomas. J Natl Cancer Inst 85:377–384. Taipale J, Beachy PA. 2001. The Hedgehog and Wnt signalling pathways in cancer. Nature 411(6835):349–54. Tanaka T, Sugimoto T, Sawada T. 1998. Prognostic discrimination among neuroblastomas according to Ha-ras/Trk A gene expression. Cancer 83:1626–33. Teitz T, Wei T, Valentine MB, Vanin EF, Grenet J, Valentine VA, Behm FG, Look AT, Lahti JM, Kidd VJ. 2000. Caspase 8 is deleted or silenced preferentially in childhood neuroblastomas with amplification of MYCN. Nat Med 6:529–35. Tepmongkol S, Heyman S. 1999. 131I MIBG therapy in neuroblastoma: mechanisms, rationale, and current status. Med Pediatr Oncol 32(6):427–31; discussion 432. Thompson PM, Maris JM, Hogarty MD, Seeger RC, Reynolds CP, Brodeur GM, White PS. 2001. Homozygous deletion of CDKN2A (p16INK4a/p14ARF) but not within 1p36 or at other tumor suppressor loci in neuroblastoma. Cancer Res 61(2):679–86. Tweddle DA, Malcolm AJ, Bown N, Pearson AD, Lunec J. 2001. Evidence for the development of p53 mutations after cytotoxic therapy in a neuroblastoma cell line. Cancer Res 61(1):8–13. Vandesompele J, Speleman F, Van Roy N, Laureys G, Brinskchmidt C, Christiansen H, Lampert F, Lastowska M, Bown N, Pearson A, et€al. 2001. Multicentre analysis of patterns of DNA gains and losses in 204 neuroblastoma tumors: how many genetic subgroups are there? Med Pediatr Oncol 36(1):5–10. Vassal G, Doz F, Frappaz D, Imadalou K, Sicard E, Santos A, O’Quigley J, Germa C, Risse ML, Mignard D, et€al. 2003. A phase I study of irinotecan as a 3-week schedule in children with refractory or recurrent solid tumors. J Clin Oncol 21(20):3844–52. Vitali R, Cesi V, Nicotra MR, McDowell HP, Donfrancesco A, Mannarino O, Natali PG, Raschella G, Dominici C. 2003. c-Kit is preferentially expressed in MYCN-amplified neuroblastoma and its effect on cell proliferation is inhibited in€vitro by STI-571. Int J Cancer 106(2):147–52. Wagner LM, Crews KR, Iacono LC, Houghton PJ, Fuller CE, McCarville MB, Goldsby RE, Albritton K, Stewart CF, Santana VM. 2004. Phase I trial of temozolomide and protracted irinotecan in pediatric patients with refractory solid tumors. Clin Cancer Res 10(3):840–8. Wang Q, Diskin S, Rappaport E, Attiyeh E, Mosse Y, Shue D, Seiser E, Jagannathan J, Shusterman S, Bansal M, et€al. 2006. Integrative genomics identifies distinct molecular classes of neuroblastoma and shows that multiple genes are targeted by regional alterations in DNA copy number. Cancer Res 66(12):6050–62. Wiseman GA, Kvols LK. 1995. Therapy of neuroendocrine tumors with radiolabeled MIBG and somatostatin analogues. Semin Nucl Med 25(3):272–8. Yaari S, Jacob-Hirsch J, Amariglio N, Haklai R, Rechavi G, Kloog Y. 2005. Disruption of cooperation between Ras and MycN in human neuroblastoma cells promotes growth arrest. Clin Cancer Res 11(12):4321–30. Yamamoto K, Hanada R, Kikuchi A, Ichikawa M, Aihara T, Oguma E, Moritani T, Shimanuki Y, Tanimura M, Hayashi Y. 1998. Spontaneous regression of localized neuroblastoma detected by mass screening. J Clin Oncol 16(4):1265–9. Yamashiro DJ, Nakagawara A, Ikegaki N, Liu XG, Brodeur GM. 1996. Expression of TrkC in favorable human neuroblastomas. Oncogene 12(1):37–41.
Molecular Therapy for Neuroblastoma
371
Yang Q, Tian Y, Liu S, Zeine R, Chlenski A, Salwen HR, Henkin J, Cohn SL. 2007. Thrombospondin-1 peptide ABT-510 combined with valproic acid is an effective antiangiogenesis strategy in neuroblastoma. Cancer Res 67(4):1716–24. Yanik GA, Levine JE, Matthay KK, Sisson JC, Shulkin BL, Shapiro B, Hubers D, Spalding S, Braun T, Ferrara JL, et€al. 2002. Pilot study of iodine-131-metaiodobenzylguanidine in combination with myeloablative chemotherapy and autologous stem-cell support for the treatment of neuroblastoma. J Clin Oncol 20(8):2142–9. Yu A, Batova A, Strother D, Angelini P, Castleberry RP. 2004. Promising results of a pilot trial of a GD2 directed anti-idiotypic antibody as a vaccine for high risk neuroblastoma. Genoa, Italy. p 75. Yu AL, Uttenreuther-Fischer MM, Huang CS, Tsui CC, Gillies SD, Reisfeld RA, Kung FH. 1998. Phase I trial of a human-mouse chimeric anti-disialoganglioside monoclonal antibody ch14.18 in patients with refractory neuroblastoma and osteosarcoma. J Clin Oncol 16(6):2169–80. Zechner D, Fujita Y, Hulsken J, Muller T, Walther I, Taketo MM, Crenshaw EB, 3rd, Birchmeier W, Birchmeier C. 2003. Beta-catenin signals regulate cell growth and the balance between progenitor cell expansion and differentiation in the nervous system. Dev Biol 258(2):406–18. Zhou H, Kuang J, Zhong L, Kuo WL, Gray JW, Sahin A, Brinkley BR, Sen S. 1998. Tumour amplified kinase STK15/BTAK induces centrosome amplification, aneuploidy and transformation. Nat Genet 20(2):189–93.
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows Jeffrey A. Toretsky and Aykut Üren
Ewing’s Sarcoma Patients Require Improved Therapy Ewing’s Sarcoma Family of Tumors (ESFT) are highly malignant tumors of bone and soft tissue that occur in children, adolescents, and young adults (Arndt and Crist 1999; Toretsky 2003). Currently, the standard therapy for ESFT patients is a five-drug regimen that consists of alternating cycles of doxorubicin/vincristine/ cyclophosphamide and etoposide/ifosfamide over the course of approximately 9 months. Side effects include nausea, vomiting, and severe hematologic cytopenias, and patients often develop life-threatening infections while receiving chemotherapy. Patients who present with localized ESFT have approximately 70% disease-free survival. Patients who present with metastatic ESFT have a poor prognosis, reporting only 20% disease-free survival despite receiving intensive therapy (Grier et€ al. 2003). These clinical response rates have persisted for the past decade, even after patients received dose-intensifying chemotherapy and bone-marrow transplantation. Current treatment-related morbidity includes cardiac, musculoskeletal, and second malignancies (Fuchs et€al. 2003). We need to discover novel therapeutic approaches to reduce treatment-related morbidity as well as improve overall survival. Novel therapies should exploit tumor vulnerability based on ESFT ontogeny, oncogenesis, and tumor-maintenance pathways.
Ontogeny of ESFT In 1921, James Ewing, M.D., suggested an endothelial cell origin for Ewing’s Sarcoma (Ewing 1921), and there has been a debate over the matter ever since. Morphologically, Ewing’s cells appear as small round cells with a high nuclear to
J.A. Toretskyâ•›() Departments of Oncology and Pediatrics, Lombardi Comprehensive Cancer Center, Georgetown University, 3970 Reservoir Rd. N.W., Washington, DC 20057-1469, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_18, © Springer Science+Business Media, LLC 2010
373
374
J.A. Toretsky and A. Üren
cytoplasm ratio, and Ewing’s tumors contain a very limited amount of stromal tissue. These features indicate an undifferentiated malignancy and make it difficult to determine the cell or tissue of origin. Ewing’s sarcoma and peripheral primitive neuroepithelioma cells were found to share a chromosomal translocation (Turc-Carel et€al. 1984; Whang-Peng et€al. 1984) and a similar pattern of oncogene expression (McKeon et€ al. 1988). The ESFT-specific chromosomal translocations t(11;22), t(21;22), and others result in the expression of pathognomonic fusion proteins. The most common of these chimeric proteins, EWS-FLI1, is critical for ESFT survival. The cellular origins of ESFT have been investigated through different experimental approaches, including analysis of gene expression patterns, forced induction of differentiation, immunohistochemical profiles, and in€vitro transformation followed by tumor xenograft models. Studies using cDNA microarray and other related technologies found thousands of gene expression signatures for ESFT cells, including a prevalence of neuronal markers (Baer et€al. 2004; Khan et€al. 2001; Prieur et€al. 2004; Staege et€al. 2004). This discovery is consistent with the expression of neuronal and neuroectodermal markers in ESFT as detected by immunohistochemistry (Cavazzana 1994; Lipinski et€al. 1987; Noguera et€al. 1992). The potential neuroectodermal origin of ESFT is further supported by the finding that modulation of Notch signaling in ESFT cell lines induces neural differentiation (Baliko et€al. 2007). Expression of EWS-FLI1 in a variety of non-ESFT models demonstrates profound effects on cell fate through altered cellular differentiation (Eliazer et€al. 2003; Rorie et€al. 2004; Torchia et€al. 2003). Most recently, GLI1 was identified as a direct target of EWS-FLI1, and its expression likely suppresses ESFT differentiation (Beauchamp et€al. 2009; Zwerner et€ al. 2008). On the other hand, epithelial markers are expressed in a fraction of ESFT samples (Collini et€al. 2001; Gu et€al. 2000; Schuetz et€al. 2005), suggesting a more pluripotent cell of origin. Another characteristic feature of ESFT is the very common expression of the cell surface molecule CD99MIC2 (Ambros et€ al. 1991; Fellinger et€ al. 1991; Hamilton et€ al. 1989a), which was identified as an androgen inducible gene in prostate cancer cells. CD99MIC2 is expressed in normal prostate cells only during embryogenesis; however, expression of CD99MIC2 in adult cells can be observed in undifferentiated carcinomas and some hyperplasias. CD99MIC2 is also expressed in CD34+ hematopoietic precursor cells (Dworzak et€al. 1994). Therefore, CD99MIC2 expression in ESFT suggests an undifferentiated precursor cell of origin for Ewing’s Sarcoma. Expression of EWS-FLI1 in pluripotent bone marrow stromal cells was found to inhibit their ability to differentiate osteogenic and adipogenic lineages (Torchia et€al. 2003). The effects of EWS-FLI1 modulation on gene expression were investigated with the hypothesis that loss of EWS-FLI1 expression might remove the inhibition of differentiation in ESFT and revert the tumor cells back to their cell of origin. The study found that inhibition of EWS-FLI1 expression in three ESFT cell lines resulted in cDNA microarray profiles resembling those of mesenchymal stem cells (Tirode et€al. 2007).
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
375
EWS-FLI1 can transform only a subpopulation of cells in€ vitro, a result suggesting that the ESFT cell of origin is particularly susceptible to EWS-FLI1induced transformation. EWS-FLI1 can transform NIH3T3 cells, but it has failed to transform other fibroblast cell lines (May et€ al. 1993). In primary mouse embryo fibroblasts (Deneen and Denny 2001) or human foreskin fibroblasts (Lessnick et€al. 2002), EWS-FLI1 expression was shown to induce growth arrest and apoptosis. More successful models of expression were obtained when primary murine bone marrow-derived cells were transfected with EWS-FLI1, supporting a mesenchymal stem cell origin for ESFT (Castillero-Trejo et€al. 2005). When bone marrow-derived cells were enriched for mesenchymal progenitor cells and transfected with EWS-FLI1, the results were malignant transformations resembling ESFT phenotypes (Riggi et€al. 2005). These studies provide evidence to support a pluripotential mesenchymal stem cell as the potential cell of origin for ESFT. Cancer stem cells, which represent a small population of therapy-resistant cancer cells that are capable of inducing a new tumor, have been described in many types of cancer (Reya et€al. 2001). These cells are considered the primary source of tumor recurrence following a patient’s good clinical response to initial therapy. The clinical course of ESFT is consistent with this hypothesis, but no cancer stem cells have been described for ESFT. The identification, purification, and propagation of a cell origin for ESFT would help advance models of the disease, but even in the absence of this knowledge, several models have been developed to provide clues to its pathogenesis.
Models of ESFT Knowledge of ESFT has been accrued by studying cell lines established in culture from patient tumors as well as by studying these tumors directly. Spheroid models of ESFT cell lines demonstrate different gene expression and enhanced resistance to anticancer compounds compared to monolayer cultures (Kang et€ al. 2007; Lawlor et€ al. 2002). These spheroid models are postulated to closely resemble actual ESFT growth in patients; however, this conclusion requires additional supportive evidence. To achieve a xenograft model, ESFT cell lines are injected into nude or athymic mice, creating tumors that are close replicates of human tumors (Vormoor et€ al. 2001). Recently, one model was developed by injecting a human cell line into the gastrocnemius to replicate an orthotopic setting found in humans. When the orthotopic tumor grows to a certain size, the primary tumor is resected and virtually all the mice develop metastases, mostly to the lung (Merchant et€al. 2004). An additional metastatic model was developed from the cyclic re-injection of tumor cells into the tail vein of a nude mouse, followed by the re-injection of those cells that grew as pulmonary metastases (Jia et€al. 2003).
376
J.A. Toretsky and A. Üren
Additional models have been created based upon the expression of EWS-FLI1 in various cell types, and some of these derived cell lines have been used to initiate xenografts. It does not appear that all of these derived cell lines closely approximate human tumor xenografts (Braunreiter et€al. 2006), but these models could still be used to analyze the tumorigenicity of EWS-FLI1. Unlike other diseases driven by chromosomal translocations, such as myeloid leukemia (Dobson et€ al. 1999), alveolar rhabdomyosarcoma (Keller et€ al. 2004), and synovial sarcoma (Haldar et€al. 2007), there are no transgenic models of ESFT. Since the cell origin of ESFT is unknown, choosing a promoter that appropriately expresses the fusion protein EWS-FLI1 in the correct tissue and developmental stage has been challenging. In one murine model, expression of EWS-ERG, a translocation occurring in about 10% of ESFT patients and a very close homolog to EWS-FLI1, led to the development of myeloid leukemia rather than Ewing’s Sarcoma (Codrington et€ al. 2005). Recent advances in molecular biology have made it possible to generate transgenic animals with the EWS-ERG and EWS-FLI1 fusion proteins in hematopoietic stem cells (Forster et€al. 2005; Torchia et€al. 2007). These transgenic animals showed a leukemia phenotype similar to that observed in human cases, which is a promising development. A further attempt to create an EWS-FLI1 transgenic mouse, using a Prx1 promoter, designed to express in mesenchymal progenitors, resulted in animals with disrupted skeletons and poorly differentiated sarcomas (Lin et€ al. 2008). At the present time, promising therapies should be evaluated in existing models until future advances lead to a better endogenous model of ESFT.
An Ideal Molecular Target While the EWS-FLI1 fusion protein has presented challenges in modeling, it is an ideal opportunity for targeted therapy because it only exists in ESFT cells. EWS-FLI1 arises from the balanced chromosomal translocation t(11;22) (q24;q12), which occurs in >95% of ESFT. This translocation joins the Ewing’s Sarcoma (EWS) gene located on chromosome 22 to an ets family gene, most commonly either the Friend Leukemia Insertion (FLI1) gene located on chromosome 11, t(11;22), or the ERG gene located on chromosome 21, t(21;22). Unfortunately, despite the knowledge that inactivation of EWS-FLI1 causes ESFT cell death, strategies to inactivate this ideal target have not yet been brought to the clinic. Chromosomal translocations involving EWSR1 and FLI1 in ESFT may take place at different exon–intron boundaries. Exons 7, 9, or 10 from EWS are combined with exons 4, 5, 6, 7, or 8 from FLI1, resulting in a family of EWS-FLI1 translocations. The most common combination is the EWS exon 7 fused to FLI1 exon 6; this is known as a Type-1 translocation and occurs in approximately 50 to 64% of patients with ESFT. Initial retrospective analyses identified that patients who have localized tumors with the 7/6 fusion have a 70% 4-year survival rate, while
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
377
patients with the other variants have a 20% 4-year survival (de Alava et€al. 1998; Zoubek et€ al. 1996); however, follow-up investigations have not confirmed these earlier findings (Huang et€al. 2005). The finding that EWS-FLI1 is expressed only in tumor cells makes it a very promising molecular target. Therapies directed toward inactivating EWS-FLI1 might address the significant problem of recurrent disease, and eliminate nonspecific side effects for patients.
EWS-FLI1 Modulates Transcription and Splicing EWS-FLI1 has been found to modulate transcription, and some of the protein partners of EWS-FLI1 indicate a role in RNA synthesis and splicing (Fig.€1, upper). Unfortunately, the entire EWS-FLI1-containing transcriptosome has not been described. There is enough evidence, however, to conclude that EWS-FLI1 functions differently from both untranslocated EWS and untranslocated FLI1 (May et€al. 1993). One known protein partner of EWS-FLI1 is hsRPB7, the seventh largest subunit of human RNA polymerase II. In Ewing’s sarcoma cell lines, EWS-FLI1 and hsRPB7 were found in a complex by immunoprecipitation (Petermann et€al. 1998). Similarly, in€vitro translated EWS-FLI1 and hsRPB7 were found in an immunoprecipitated complex, suggesting a direct interaction between these proteins. In the same study, researchers did not detect association of hsRPB7 with either full-length endogenous
Fig.€1╅ EWS-FLI1 regulates transcription and RNA metabolism. EWS-FLI1 is a protein partner in complexes that regulate transcription and translation. The proteins shown are those that have been identified as binding to EWS-FLI1 either directly or indirectly. Proteins that are directly regulated by EWS-FLI1 are shown in the lower half of the figure
378
J.A. Toretsky and A. Üren
EWS or full-length endogenous FLI1, which implies that hsRPB7interacts with EWS-FLI1 only in its chimeric state. The functional interaction between these two proteins is further supported by the finding that hsRPB7 augmented EWS-FLI1 transcriptional activity on a reporter construct. EWS-FLI1 binds to DNA through the conserved ets binding domain. Many studies have identified transcripts that are increased or decreased as a result of EWS-FLI1 expression (Fig.€1, lower). These studies investigated changes in gene expression based on EWS-FLI1 expression in different cell lines (Braun et€al. 1995) and expression profiles of ESFT patients compared to those of related childhood tumors that lack EWS-FLI1 (Baer et€ al. 2004; Khan et€ al. 2001). The observed changes in gene expression could be a result of direct interaction of EWS-FLI1 with cell-specific coregulatory expression or a secondary effect. Genes that appear to be directly regulated by EWS-FLI1 include PTPL1 (Abaan et€ al. 2005), Id2 (Fukuma et€al. 2003; Nishimori et€al. 2002), p21 (Nakatani et€al. 2003), MK-STYX (Siligan et€al. 2005), IGFBP-3 (Prieur et€al. 2004), TGFßRII (Hahm et€al. 1999), NROB1 (Kinsey et€ al. 2006; Mendiola et€ al. 2006), NKX2.2 (Smith et€ al. 2006), Uridine Phosphorylase (Deneen et€al. 2003), PLD2 (Kikuchi et€al. 2007), caveolin-1 (Tirado et€al. 2006), and thrombospondin (Potikyan et€al. 2007). Although it is interesting to note that these are all genes whose protein products contribute to cell growth or transformation, a clear picture of how EWS-FLI1 regulates these genes has yet to emerge. Identification and evaluation of proteins that interact with EWS-FLI1 might clarify the mechanisms of transformation, and these mechanisms might also help to explain the differences in patient survival based on the breakpoint of the fusion protein. There is also evidence to suggest that EWS-FLI1 is directly involved in RNA splicing. The link between EWS-FLI1 and RNA splicing is supported by the interaction between EWS-FLI1 and a small nuclear ribonucleoprotein (snRNP), which is part of the spliceosome complex involved in mRNA splicing. U1C, a member of the U1 snRNP-specific protein family, was identified as an EWS-FLI1 protein partner, using a yeast two-hybrid system with the EWS domain of EWSFLI1 as bait (Knoop and Baker 2000). In vitro, translated U1C and EWS-FLI1 expressed in bacteria were found to coprecipitate in GST pull-down experiments. This interaction was mapped to the amino terminus of EWS, revealing that both EWS-FLI1 and full-length EWS can bind to U1C. Transcription by an EWSFLI1-dependent reporter construct was inhibited when the cells were transfected with U1C expression plasmids. As promising as these in€vitro results may be, a direct interaction between endogenous EWS-FLI1 and U1C in ESFT, as well as the involvement of EWS-FLI1 in U1C driven RNA splicing, remains to be investigated. Functionally, EWS-FLI1 can alter mRNA by alternatively splicing an E1A transcript (Knoop and Baker 2001). There are no literature reports of altered mRNA splicing in ESFT cells, which would manifest as functional consequences such as loss of growth regulation or transformation, but it is possible that some of the genes thought to be modulated by EWS-FLI1 are, in fact, regulated by mRNA splicing
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
379
rather than direct transcriptional regulation. Additional evidence suggests that EWS-FLI1 inhibits splicing modulated by RNA polymerase and YB-1 (Chansky et€al. 2001; Yang et€al. 2000).
Elimination of EWS-FLI1 Reduces ESFT Cell and Tumor Growth EWS-FLI1 is required to maintain the growth of ESFT cell lines. When EWS-FLI1 expression is reduced, ESFT cell lines die and tumors in nude mice regress. Reduction of EWS-FLI1 by antisense oligodeoxynucleotides (ODN) (Tanaka et€al. 1997; Toretsky et€ al. 1997a) and by antisense RNA expressed from a vector (Kovar et€al. 1996; Maksimenko et€al. 2003a; Ouchida et€al. 1995) inhibited not only the proliferation of ESFT cell lines, but also the growth of xenografted tumors in nude mice. Encouraging results were also obtained when EWS-FLI1 was inhibited by siRNA, a type of RNA interference that serves as a sequence-specific posttranscriptional silencing mechanism. This evolutionarily conserved mechanism defends the cell from viruses by degrading double-stranded RNA fragments. siRNA uses 18 to 21 nucleotide double-stranded RNA to induce sequence-specific gene silencing. Targeting the junction point of EWS-FLI1 fusion protein by siRNA significantly diminished the expression of EWS-FLI1 in ESFT cell lines (Chansky et€al. 2004; Dohjima et€al. 2003; Kovar et€al. 2003). Like the antisense oligonucleotides, siRNA inhibited the proliferation of ESFT cell lines in€vitro and stopped xenografted tumor growth in nude mice (Hu-Lieskovan et€al. 2005). The ODN, antisense RNA, and siRNA studies have each established that EWSFLI1 is critical to ESFT cells, but effective strategies of treatment for ESFT patients remain to be developed. The pharmacokinetics and pharmacodynamics of siRNA and antisense ODN have been less than satisfactory in humans. Recent advances in nanotechnology have improved delivery and controlled the release of siRNA and antisense ODN in animal models (Lambert et€al. 2000; Maksimenko et€al. 2003b), but even with current technologies, reduction of EWS-FLI1 with antisense ODN and siRNA remains far from possible in humans. An alternative approach to reducing EWS-FLI1 involves blocking EWS-FLI1 function with a dominant negative protein, the Kruppel-associated box (KRAB) (Kovar et€al. 1996). KRAB is a 75 amino acid protein domain that can function as a transcriptional repressor. When linked to the DNA-binding domains of different transcription factors, KRAB inhibited promoter–reporter constructs. Similarly, a KRAB-FLI1 fusion protein was found to inhibit EWS-FLI1 oncogenic function, most likely by competing for the same DNA-binding sites in the cells (Chan et€al. 2003). Based on multiple models of biologic intervention, EWS-FLI1 is critical for ESFT; therefore, inhibition of EWS-FLI1 should be evaluated as a specific cancer therapy for ESFT.
380
J.A. Toretsky and A. Üren
Molecular Targeting of ESFT The development of targeted molecular agents for patients with ESFT is necessary in order to improve patient survival. The most obvious target for these agents is to disable or eliminate EWS-FLI1. As a regulator of transcription, EWS-FLI1 serves as an Achilles’ heel to the tumor when it is inhibited. As discussed earlier, developments in nanotechnology may improve the delivery of siRNA and antisense ODN to eliminate EWS-FLI1 in tumor cells; however, the technology has not been tested in humans. Alternatives to eliminating EWS-FLI1 are the inhibition of its function by disrupting the binding of EWS-FLI1 to key functional protein partners, finding small molecules to mimic EWS-FLI1 reduction, using single chain antibodies, and inhibiting specific key transcriptional targets of EWS-FLI1. Additional molecular targeting could be directed toward key growth factor pathways or other aspects of tumor biology that present a reasonable therapeutic window. The overall goal of targeted therapy is to eliminate the tumor and spare normal function as much as possible in terms of both acute and long-term effects. Knowledge gained by studying and targeting EWS-FLI1 would have broad benefits beyond ESFT. EWS and FLI1 homologs are partners in translocations that occur in a wide range of sarcomas and leukemias (Helman and Meltzer 2003). EWS and its homolog TLS are involved in chromosomal translocations of clear cell sarcoma, mixoid liposarcoma, desmoplastic small round cell tumors, chondrosarcoma, and acute myeloid leukemia. The FLI1 homolog ERG is translocated in acute myeloid leukemia. The importance of EWS and FLI1 homologs in so many forms of cancer suggests that EWS-FLI1 may serve as a model system, possibly shedding further light on a family of diseases that are related by translocation partners.
EWS-FLI1: The Perfect Target Small-Molecule Protein–Protein Interaction Inhibitors EWS-FLI1 is clearly established as a molecular target for ESFT, but strategies to reduce protein levels are not ready for clinical use. Small-molecule inhibitors of EWS-FLI1 function may lead to targeted therapeutics; however, since EWS-FLI1 lacks intrinsic enzymatic function, targeted inhibition represents a significant challenge. Disruption of key EWS-FLI1 protein–protein interactions, a strategy that has been suggested to regulate other ets family proteins (Li et€al. 2000a), might lead to novel therapeutic agents. Computer-assisted design of small molecules that fit into functional pockets has been shown to prevent max from binding to myc, and has thereby inhibited myc-modulated transcription (Yin et€ al. 2003). While threedimensional structures have been critical to the optimization of targeted smallmolecule inhibitors, EWS-FLI1 presents a challenge based upon its biochemical characteristics, described below (Uren et€al. 2004a).
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
381
The novelty of EWS-FLI1 may cause it to interact with other proteins, such as basal transcriptional regulators, in unique three-dimensional protein–protein interactions. The separation of EWS-FLI1 from the key proteins required for EWS-FLI1 oncogenesis might cause ESFT cell death. Structural information about EWS-FLI1 as a fusion protein is limited. EWS-FLI1 has the potential to both homodimerize and heterodimerize with EWS; however, it is not clear that this happens in ESFT cells (Spahn et€al. 2003; Watson et€al. 1997). While crystallographic and NMR structures of ets DNA-binding domains are available, no data exist for EWS-FLI1. Furthermore, due to changes in the overall amino acid composition and charge distribution in the fusion protein, EWS and FLI1 domains in EWS-FLI1 may have a unique three-dimensional structure that is different from both wild-type protein partners (Uren et€ al. 2004a). An analysis of the primary amino acid sequence of EWS-FLI1 suggests that the secondary structure of EWS-FLI1 is largely disordered under native conditions due to low overall hydrophobicity. This hypothesis was supported by the discovery that EWS-FLI1 lost function following mutation of its aromatic residues (Ng et€al. 2007). These analyses suggest that EWS-FLI1 will be in considerable motion in solution and, as a result, may not be crystallized easily. The overall size of EWS-FLI1 and its poor solubility make it difficult to study with current magnetic resonance techniques (Uren et€al. 2004a). Strategies used to identify EWS-FLI1 protein partners include yeast two-hybrid, column chromatography, promoter analysis, and phage display. Yeast two-hybrid and column chromatography have identified partners that are components of transcription, such as RNA Pol II (Petermann et€al. 1998) and TFIID (Bertolotti et€al. 1998). Scrutiny of the uridine phosphorylase promoter, an EWS-FLI1-regulated transcript, revealed AP-1 as a critical EWS-FLI1 cofactor (Kim et€al. 2006). Recently, we generated a functional recombinant EWS-FLI1, which was then used to screen bacterial phage libraries to identify peptide sequences that bind to EWS-FLI1. These peptides led to the identification of proteins that directly bind to EWS-FLI1 as well as small molecules whose homologous structure might prevent EWS-FLI1 from binding to key proteins. We identified RNA helicase A (RHA) as a significant protein partner of EWS-FLI1, whose interaction with EWS-FLI1 represents a therapeutic target (Toretsky et€al. 2006).
YK-4-279 is a Novel Small-Molecule Inhibitor of EWS-FLI1 Disrupting the interaction between EWS-FLI1 and its key protein partner RHA holds particular promise as a targeted therapy because of the specific protein interaction of the two partners. The region of RHA (aa 647-1075) that interacts with EWS-FLI1 uniquely binds to EWS-FLI1; it has yet to be identified as a binding partner with the basal transcription apparatus (Anderson et€al. 1998; Aratani et€al. 2001; Nakajima et€al. 1997; Tang and Wong-Staal 2000). A ten amino acid peptide that mimics the interaction of RHA blocks the interaction with EWS-FLI1 (Erkizan et€al. 2009). A library of small molecules led to the discovery of a lead compound,
382
J.A. Toretsky and A. Üren
YK-4-279, that also binds to EWS-FLI1 and, in addition, prevents RHA binding (Erkizan et€ al. 2009). Moreover, this novel compound inhibits EWS-FLI1 transcriptional activity, reduces ESFT cell-line viability, induces apoptosis, and inhibits xenograft growth (Erkizan et€al. 2009). While still early in the development process, YK-4-279 demonstrates a novel approach to the inhibition of tumor-specific fusion proteins and promises to be a future therapeutic for patients with ESFT.
Alternate Approaches to Small Molecule Identification Recently, a novel approach to the inactivation of EWS-FLI1 was developed through cDNA expression arrays that established a signature for EWS-FLI1-expressing cells (Lessnick et€al. 2002). An ESFT cell line was treated with a panel of compounds, most of which are already in clinical use. cDNA signatures were established for each of the compounds used on ESFT cells, and these signatures were compared with those that resembled the siRNA reduction of EWS-FLI1. The results of this comparison identify cytosine arabinoside (Ara-C) as a molecule that caused ESFT cells to have an expression profile most similar to that of EWS-FLI1 reduction (Stegmaier et€al. 2007). Unfortunately, a recent Children’s Oncology Group phase II study for patients with recurrent ESFT demonstrated that Ara-C did not have clinical activity (DuBois et€al. 2009).
Single-Chain Antibodies The use of single-chain antibodies to inactivate proteins was developed to better study the antigen–antibody interaction (Denzin et€al. 1991) and, in part, to develop recombinant therapies (Begent et€al. 1996). The use of a specific antibody to target EWS-FLI1 has been explored, but the most recent result was an antibody that recognizes wild-type EWS but not the fusion protein in its native form (Aryee et€al. 2006).
Targeting the Targets Knowing that EWS-FLI1 regulates many transcripts that result in ESFT, one clear challenge is to decide which of these is a worthy therapeutic target, in terms of both specificity and drug ability. Since the targets of EWS-FLI1 include genes that might be important for nontumor cells, including hematopoietic cells, the specificity gained by targeted therapy may be lost. Alternatively, there may be a key target of EWS-FLI1 that is not important for nontumor physiology and would thus make a good anticancer target. The ideal targets would have unique biology, from kinases
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
383
to phosphatases to nonenzymatic proteins. In each case, the biology of the target needs to be considered and validated as critical to the growth of ESFT tumors. This validation would include reducing the level of the target gene to demonstrate reduced ESFT growth (Kinsey et€ al. 2006; Potikyan et€ al. 2007). For example, caveolin-1 is a putative EWS-FLI1 target (Tirado et€al. 2006), but a general targeting of caveolin-1 might disrupt signaling in all cells, thus reducing the specificity of a potential caveolin-1-directed therapy. Given that many of the current anticancer therapies are not targeted and that agents like proteosome inhibitors or anti-heat shock agents appear to convey some sense of therapeutic index, the approach of inhibiting the targets of EWS-FLI1 may be fruitful. However, targeting less ESFTspecific molecules may lead to significant and intolerable toxicity.
Disruption of EWS-FLI1 DNA Binding Another targeted approach for disabling EWS-FLI1 might stem from the disruption of DNA binding. There are two challenges to this approach. The first is that DNA binding may not be necessary for EWS-FLI1 transformation (Welford et€al. 2001), although DNA binding may still be necessary for tumor maintenance. Second, the DNA sequence that EWS-FLI1 binds is a well-recognized ets-binding site and demonstrates significant promiscuity in transcription factor binding (Watson et€al. 1997). While specificity of novel compounds may overcome these challenges, disrupting EWS-FLI1 and DNA binding is not a feasible therapeutic strategy at this point.
Non-EWS-FLI1 ESFT Targets In order to improve survival for ESFT patients, many agents are being used in phase I and phase II studies (Table€1). These agents are supported by varying levels of rationale. Those with specific evidence in ESFT are described here.
Modulation of Apoptotic Tendency Apoptosis signaling pathways can be categorized into two classes: the intrinsic pathway involving the mitochondria and the extrinsic pathway involving cell surface receptors. These death receptors are members of the tumor necrosis factor (TNF) receptor family, which can induce apoptosis independently of p53 (Ashkenazi 2002). One subgroup of this family includes TNF receptor 1 (TNFR1), FAS, and death receptors 4 and 5 (DR4 and DR5). These receptors contain an intracellular death domain that, upon ligand binding, can recruit the FAS-associated death domain (FADD) and activate caspase 8. The ligand for DR4 and DR5 is
Table€1╅ Current clinical trials using targeted therapies in ESFT Agent Other names Function Phase Manufacturer 17-AAG 17-AAG Geldanamycin analog, HSP-90 inhibitor Phase I CP-751,871 CP-751,871 IGF-IR inhibitor monoclonal antibody Phase I Pfizer R-1507 IGF-IR inhibitor monoclonal antibody Phase II Roche Imatinib Gleevec, STI-571 TX inhibitor BCR-ABL, c-KIT, PDGFRb Phase II Novartis Depsipeptide FK-228, NSC-630176 HDAC inhibitor Phase I/Il Reolysin Reolysin Wild-type human reovirus with anti-ras Phase II Oncolytics Biotech Inc. activity Bevacizumab Avastin Monoclonal antibody against VEGF Phase II Genentech Topotecan Hycamtin Topoisomerase 1 inhibitor Phase III GlaxoSmithKline Irininotecan Camptosar Topoisomerase 1 inhibitor Phase II Pfizer Cytarabine Ara-C, Cytosar-U DNA polymerase inhibitor Phase II Pfizer Trabectedin ET-743, Yondelis Directly binds to DNA, inhibits replication Phase II PharmaMar, Johnson & Johnson and repair Desatinib BMS-354825, Sprycel TK inhibitor BCR-ABL, SRC, c-KIT, EPHA2, Phase II Bristol-Myers Squibb and PDGFRb Phase II Bayer Sorafenib Nexavar TK inhibitor Raf, VEGFR, PDGFR, Fit-3, c-Kit, RET Sunitinib Sutent TK inhibitor VEGFR, PDGFR, Fit-3, c-Kit Phase II Pfizer Key to abbreviations: 17-AAG 17-(Allylamino)-17-demethoxygeldanamycin, IGF-IR insulin-like growth factor-I receptor, PDGFRb platelet derived growth factor receptor-b, HDAC histonedeacetylase, VEGF vascular endothelial cell growth factor, VEGFR vascular endothelial cell growth factor receptor, TK tyrosine kinase, FLT Fms-like tyrosine kinase
384 J.A. Toretsky and A. Üren
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
385
TNF-related, apoptosis-inducing ligand (TRAIL); DR4 and DR5 are also called TRAIL receptor 1 and TRAIL receptor 2, respectively. TRAIL receptors function as homotrimers that can activate the extrinsic apoptosis pathway by converting inactive procaspase 8 to active caspase 8. Once activated, caspase 8 can activate caspase 3 and other downstream caspases to initiate apoptosis. Caspase 8 can also cleave BID and activate intrinsic apoptosis pathway through the mitochondria (Duiker et€al. 2006). Since TRAIL receptors are not expressed in normal cells but are highly expressed in many tumor cells, they have been evaluated as a potential therapeutic target for cancer (Fesik 2005; Takeda et€al. 2007). A common approach has been to activate these death receptors either by agonistic antibodies or recombinant soluble TRAIL proteins. Both approaches activate the TRAIL receptors and induce apoptosis in tumor cells with little to no toxic effects on normal cells. However, some tumor cells are able to resist TRAIL-induced apoptosis by expressing decoy receptors, which can bind TRAIL but cannot propagate the signal due to their lack of a cytoplasmic death domain. In these cases, use of the antibody approach may have benefits over using recombinant ligands. Agonistic antibodies not only activate the death receptor to initiate apoptosis, but they can also initiate an immune response through their Fc domain. An antibody attached to the surface of a tumor cell can activate both complement-dependent cytotoxicity and antibody-dependent cellular cytotoxicity. ESFT cell lines express both FAS and TRAIL receptors (Kontny et€ al. 2001; Kontny et€al. 1998). Furthermore, application of ligands to Ewing’s cells induces apoptosis in most cell lines in both in€ vitro and xenograft models. Interestingly, resistant cells become sensitive when they are pretreated with interferon-g (IFN-g) (Abadie and Wietzerbin 2003; Merchant et€ al. 2004). This cooperation between IFN-g and TRAIL or TNF-a appears to stem from the stimulation of ESFT cells with IFN-g as well as TNF-a-induced expression of TRAIL (Abadie et€al. 2004). IFN-g also induces expression of caspase 8. Induction of caspase 8 by either IFN-g or an expression vector was enough to convert a TRAIL-resistant ESFT cell line into a sensitive one (Lissat et€al. 2007). These findings provide accumulating evidence toward a rationale for using TRAIL and TNF-a in combination with IFN-g.
Cytokine-Regulated Growth Pathways The insulin-like growth factor type 1 receptor (IGF-IR) pathway, which leads to transformation, cell division, and survival, is activated by the ligands IGF-I or IGF-II. The receptor is a heterotetrameric membrane-bound tyrosine kinase that, upon autophosphorylation, directly binds a number of molecules to advance signaling, including the docking protein insulin-receptor substrate-1 (IRS-1), thus initiating multiple growth and survival pathways (Butler et€al. 1998; Singleton et€al. 1996; Tanaka et€al. 1996; Valentinis and Baserga 1996). The IGF-IR pathway is required for transformation by many oncogenes; therefore, it is thought to be a critical pathway for cells to become malignant. (Baserga 1995).
386
J.A. Toretsky and A. Üren
IGF-IR was first recognized as important in ESFT based on its modulation of the antigen HBA-71 (CD99, p30/32MIC2), effects upon cell-line growth, and IGF-I ligand expression (Hamilton et€al. 1989b, 1991; van Valen et€al. 1992; Yee et€al. 1990). An antibody bound to the IGF-IR reduced ESFT cell-line growth and enhanced the effects of chemotherapy (Hofbauer et€al. 1993), suggesting that IGF-I acts as an autocrine factor for the growth of ESFT. Whether autocrine, paracrine, or endocrine, IGF-I appears to activate a significant pathway in ESFT. The expression of the IGF-IR was found to be necessary for the EWS-FLI1 oncogene to transform fibroblasts (Toretsky et€al. 1997b). Xenograft studies that blocked the IGF-IR with antisense RNA and antibodies further demonstrated the importance of IGF-IR signaling for ESFT growth (Benini et€al. 2001; Scotlandi et€al. 1998). In addition, the IGF-I ligand has been shown to enhance tumor cell survival, and a pilot study suggested that a high ratio of IGF-I to IGFBP-3 led to reduced overall survival in ESFT patients (Toretsky et€al. 1999, 2001). Recently, the IGF-binding protein-3 (IGFBP3) was identified as a transcriptional target suppressed by EWS-FLI1 (Prieur et€al. 2004), which is an interesting finding because IGFBP-3 may directly or indirectly enhance apoptosis (Gill et€al. 1997; Rajah et€al. 1997). The inhibition of the IGF-IR pathway in patients with ESFT has been a research goal for almost 20 years. The first small-molecule inhibitors of the IGF-IR kinase activity were effective at reducing the growth of ESFT cell lines and xenografts (Scotlandi et€al. 2005), but there are currently no clinical trials for these inhibitors in ESFT patients. The experience with the EGFR suggests that the small-molecule kinase inhibitors are most useful when patients have constitutively activated mutations of the EGFR (Lynch et€al. 2004; Pao et€al. 2004). ESFT have not yet been systematically screened for IGF-IR-activating mutations. In fact, no IGF-IRactivating mutations have been described in any naturally occurring tumor. Tumors that appear to rely on IGF-IR signaling generally have a significant overexpression of the receptor, leading to a more constitutive activation. The EWS-WT1 fusion protein, which results from a t(11;22) with a different fusion partner and is found in desmoplastic small round blue cell tumors, directly activates expression of the IGF-IR, while EWS-FLI1 does not directly increase IGF-IR expression in ESFT (Karnieli et€al. 1996). Despite these caveats, there is significant interest in targeting the IGF-IR pathway in ESFT. A number of commercially developed antibodies are currently available in phase I and II trials. These clinical trials will answer whether IGF-IR is both targetable and a worthwhile target in ESFT. Another potential therapeutic target for ESFT is the platelet-derived growth factor (PDGF), a secreted glycoprotein with important mitogenic and chemoattractant functions (Heldin et€ al. 1986). There are two well-characterized PDGF ligands: PDGF-A and PDGF-B, which form heterodimers as PDGF-AB and homodimers as PDGF-AA and PDGF-BB (Williams 1989). Recently, two new members of this family have been discovered: PDGF-C and PDGF-D (LaRochelle et€al. 1993; Li et€al. 2000b). PDGF-A and PDGF-B chains are secreted from the cell as mature ligands, while PDGF-C and PDGF-D require proteolytic cleavage outside the cell in order to become active. PDGF ligands exert their effect on target cells through specific dimeric receptors. The two different subunits of PDGF receptors (PDGFRs)
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
387
have different ligand-binding specificity; a-PDGFR binds PDGF-A, B, and C chains, but b-PDGFR binds only PDGF-B and D chains. PDGF-C has affinity toward a-PDGFR similar to PDGF-A, and PDGF-D has affinity toward b-PDGFR similar to PDGF-B. Ligand binding initiates dimerization and autophosphorylation of the specific receptors. PDGF is important in physiologic processes, including embryonic development, wound healing, angiogenesis, hemostasis, and platelet aggregation. Over-activity of PDGF signaling contributes to the pathogenesis of many diseases, including lung fibrosis, glomerulonephritis, atherosclerosis, and cancer (Ostman and Heldin 2001). The PDGF-B homolog, v-sis, was one of the earliest discovered oncogenes (Doolittle 1983, Waterfield 1983). Overexpression of PDGF ligands and receptors has been reported for many solid tumors, including glioblastoma, meningioma, melanoma, ovarian cancer, prostate cancer, lung cancer, pancreatic cancer, and gastric carcinoma (Ostman and Heldin 2001). Analysis of ESFT cell lines and tumor samples provided evidence that PDGF signaling could be used as a molecular target. ESFT cell lines express b-PDGFR, which has a significantly higher transforming capacity than a-PDGFR (Merchant et€ al. 2002; Uren et€ al. 2003). Expression of b-PDGFR was also detected in a majority of tumor samples from ESFT patients (Bozzi et€al. 2007; Uren et€al. 2003). Activation of b-PDGFR in ESFT cell lines induces proliferation and chemotaxis, so it may contribute to both tumor development and metastasis. Screening of a cDNA library from EWS-FLI1-transformed fibroblasts revealed PDGF-C as an EWS-FLI1-induced gene (Zwerner and May 2001). Later, PDGF-C expression was found to have a strong correlation with EWS-FLI1’s ability to transform different fibroblast cell lines (Zwerner et€ al. 2003). The ability of EWS-FLI1 to induce PDGF-C expression suggests a potential autocrine signaling pathway in ESFT cells. However, unlike fibroblasts that express both PDGF receptor isoforms, ESFT cells mainly express b-PDGFR, which cannot bind PDGF-C (Merchant et€al. 2002; Uren et€al. 2003). A dominant negative PDGF-C inhibited the growth of ESFT cell lines (Zwerner and May 2002), but in this study, a potential cross reactivity of dominant negative PDGF-C was not ruled out. Since PDGF ligands can form heterodimers, a dominant negative PDGF-C may inhibit PDGF-B function as well. Additionally, clinical samples also showed strong phosphorylation of b-PDGFR. This paradoxical finding may be explained by a paracrine signaling mechanism. PDGF-BB from serum and stromal cells can activate b-PDGFR in ESFT cells and induce their growth and motility. PDGF-C secreted from ESFT cells can act on tumor stroma to provide vascularity. KIT is a cell surface receptor glycoprotein found in the Type III class of receptor tyrosine kinases, the same class as PDGFRs (Fantl et€al. 1993). This class of receptor tyrosine kinases is characterized by five extracellular immunoglobulin-like domains and an intracellular tyrosine kinase domain that is interrupted by a large insert sequence termed the “kinase insert” region. The ligand for KIT is stem cell factor, which induces dimerization and autophosphorylation of the receptor. KIT plays an important role in erythropoiesis, lymphopoiesis, mast cell development, megakaryopoiesis, gametogenesis, and melanogenesis (Ronnstrand 2004). KIT
388
J.A. Toretsky and A. Üren
gain-of-function mutations have been described in gastrointestinal stromal tumors (GIST), mastocytomas, sinonasal T-cell lymphomas, dysgerminomas, and acute myelogenous leukemias (Heinrich et€al. 2002). Autocrine and paracrine activation of wild-type KIT has been implemented in several types of tumors as well. Expression of KIT in ESFT cell lines and tumor samples has been shown in multiple studies (Bozzi et€al. 2007; Merchant et€al. 2002; Tamborini et€al. 2003), but expression of KIT in ESFT did not correlate with patient survival and there were no gain-of-function mutations of KIT. There is still some evidence, however, of autocrine or paracrine activation of KIT in ESFT tumor samples (Bozzi et€al. 2007; Tamborini et€al. 2003). Imatinib (STI-571, Gleevec®) is a specific protein tyrosine kinase inhibitor that inhibits Abl, KIT, and PDGFR kinases with high specificity. Imatinib found clinical use for treating chronic myelogenous leukemia (CML), which expresses the BcrAbl chimeric oncoprotein as the result of a specific chromosomal translocation (Philadelphia chromosome). Another successful clinical application of Imatinib came in GIST, which contain KIT gain-of-function mutations in over 70% of patients. Clinical trials of Imatinib in ESFT patients did not demonstrate activity as a single agent in a Phase II trial (Bond et€al. 2007). Other cytokine-activated pathways found to be important in cell-based models of ESFT include bFGF (Girnita et€ al. 2000; Schweigerer et€ al. 1987; Westwood et€al. 2002; Williamson et€al. 2004), VEGF-R (Fuchs et€al. 2004; Guan et€al. 2005; Lee et€al. 2006), and wnt (Uren et€al. 2004b). Although all these findings require tissue validation, the pathways identified may be critical for the growth of ESFT. The CD99MIC2, first identified as HBA-71 (Hamilton et€al. 1989a), is relatively specific for ESFT (Weidner and Tjoe 1994), and crosslinking of this antigen is toxic to ESFT cells (Scotlandi et€al. 2006). CD99MIC2 may be involved in metastatic growth of ESFT through downstream effects on the potassium channel KCMF1, leading to enhanced cell motility (Kreppel et€ al. 2006). Although the function of CD99MIC2 remains cryptic in ESFT and its possible role in therapy is uncertain, it could potentially allow for reasonably targeted antibodies to ESFT cells.
Molecular Targets from Other Tumor Models Some therapeutic agents are active against ESFT with a broad mechanism; others are developed for mechanisms in non-ESFT cancers that might also apply to ESFT. Agents that inhibit topoisomerase, irinotecan, and topotecan are active in recurrent ESFT in combination with alkylating therapy, temozolomide, and cytoxan, respectively (Saylors et€al. 2001; Wagner et€al. 2007). By their nature, these combinations do not act toward specific molecular targets. Additional agents have undergone preclinical and early stage clinical development in adult cancers, and they may hold less-specific promise – but promise nonetheless – for patients with ESFT. Some agents, such as ET-743, demonstrate good activity against ESFT, but the mechanisms they affect remain cryptic (Scotlandi et€al. 2002). Other targeted therapies for
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
389
patients with ESFT include agents that affect protein binding to DNA, protein chaperoning, and protein degradation. Among these potentially promising agents for ESFT are the histone deacetylase (HDAC), heat shock protein 90 (HSP90), and proteosome inhibitors. These agents do not specifically act upon EWS-FLI1, but there is hope that they may provide ESFT patients more general therapy with tolerable side-effects. The HDAC family of proteins regulates the acetylation of histone proteins, thereby regulating transcription factor access to DNA. The recent discovery of nonhistone acetylation of proteins suggests novel posttranslational protein regulation mechanisms. This class of anticancer agents has yielded a long list of molecular classes and clinical trial results. For instance, cutanteous T-cell lymphoma has been one of the most responsive malignancies to HDAC inhibitors. The bicyclic peptide HDAC inhibitor FK-228 (depsipeptide, NSC-630176) induces apoptosis in a series of ESFT cell lines and reduces xenograft tumor growth (Sakimura et€al. 2005). Phase I studies of depsipeptide have been completed in children, and phase II studies are currently underway. MS-275, a benzamide HDAC inhibitor, demonstrated reduced ESFT cell and xenograft growth (Jaboin et€al. 2002). Both HDAC studies showed that ESFT cells re-expressed transcripts that are putatively suppressed by EWSFLI1, suggesting a potential mechanism of compound action. FK-228 also appeared to reduce the expression of EWS-FLI1. Heat shock proteins (HSP) are both specific and general targets of anticancer therapeutics. HSP are specifically inactivated by benzoquinonoid ansamycins, suggesting that they are specific molecular targets of HSP90. Significant reduction of HSP90 leads to a dramatic reduction of many client proteins, including Akt, IGF-IR, telomerase, and other key kinases. ESFT cells are very sensitive to reduction of xenograft growth and treatment with geldanamycin (IC50â•›=â•›5 to 8€nM) in a 3 H-thymidine incorporation assay (Whitesell et€al. 1992). The key client proteins of HSP90 have not been determined in ESFT, but IGF-IR or Akt are logical possibilities. Two pediatric phase I studies have been performed with 17AAG, a more water soluble form of geldanamycin. Phase II studies are pending the development of more potent compounds with better “drug-like” qualities. HSP90 inhibitors may be helpful to further elucidate a series of key protein targets in ESFT, given the cell lines’ significant sensitivity to HSP90 depletion. The potential role for HSP inhibitors as targeted therapy in the clinical practice is still under investigation. Protein synthesis and degradation are balanced modulators of protein levels. They are important because they lead to regulation of cell functions via the presence or absence of key proteins. In cancer cells, proteins required for cell division are synthesized rapidly while proteins that slow the cell cycle are degraded at a high rate in order to maintain tumor growth. Both of these processes can be the molecular targets of cancer therapy. For example, Rapamycin has been shown to bind to FKBP and block the binding of raptor to the mammalian target of Rapamycin (mTOR) (Hara et€ al. 1998, 2002). This binding prevents the activation of key proteins needed to enhance protein translation. Used in isolation, Rapamycin moderately decreases ESFT cell growth, but in combination with EWS-FLI1-directed antisense oligonucleotides, it
390
J.A. Toretsky and A. Üren
significantly reduces xenograft growth (Mateo-Lozano et€ al. 2006). Rapamycin reduces protein levels of EWS-FLI1, suggesting that a high level of protein synthesis is necessary to maintain EWS-FLI1 in the tumor cells. This implies that Rapamycin may be a compound synergistic to agents that disrupt EWS-FLI1 function by causing dissociation of key protein partners. Protein degradation is regulated by ubiquitination followed by proteosomemediated catabolism. While the standard proteosome inhibitors have not yet been tested in ESFT cell lines, blocking proteosomal activity causes an increase in levels of the cell cycle inhibitor p27, leading to ESFT cell senescence (Matsunobu et€al. 2004). In combination with antisense therapy directed to EWS-FLI1, these results suggest a cooperative mechanism of potential novel therapeutics. There are no published phase II studies of agents that regulate protein levels with useful numbers of ESFT patients.
Conclusion New and improved therapies are needed for more effective treatment of ESFT, especially in patients with metastatic disease. The discovery of the importance of EWS-FLI1 in ESFT has sparked a significant amount of research in the past two decades, but the translation of this basic science to the bedside has been slow. Although ESFT contains a perfect molecular target in EWS-FLI1, therapeutic development has not yet yielded a successful drug. Still, there has been some progress regarding ESFT. Advances in modern molecular biology techniques have led to a long list of genes that are modulated, at some level, by EWS-FLI1. The identification of EWS-FLI1 protein–protein interactions, as well as the role of EWS-FLI1 in RNA processing, has broadened the understanding of EWS-FLI1-mediated oncogenesis. However, improved resolution of the molecular mechanisms of EWS-FLI1 oncogenesis is required. The recently developed small-molecule and peptide probes will clearly help to advance understanding and therapy of EWS-FLI1. Additional challenges to effective targeted therapeutics include the lack of a definitive cell of origin and a clinically relevant animal model of ESFT. More recent knowledge points toward a mesenchymal cell; however, earlier data supported a neural crest origin cell of origin. This cell of origin might be clarified by identification of a promoter that will allow the proper timing and tissue expression of EWS-FLI1 to recapitulate ESFT in a mouse. An experimental transgenic animal model will be of great value both for studying ESFT biology and for testing new chemotherapeutic agents. The progression toward clinical EWS-FLI1 inhibitors has been challenged by a general scientific belief that such inhibitors were not possible, and if they were possible, pharmaceutical development would not occur due to the potentially small numbers of affected patients. The recent demonstration of a small-molecule inhibitor of EWS-FLI1 represents a breakthrough for a new discovery method as well as a class of compounds as therapeutic agents for ESFT. Novel approaches to therapeutic
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
391
development for rare diseases, such as ESFT, will clearly require partnership from philanthropy, industry, academia, and NIH resources. If we can join these forces, we can make the translation of molecular biology to bedside practice much more imminent. Acknowledgmentsâ•… The authors would like to thank Audrey Kubetin for her skillful assistance with this manuscript.
References Abaan, O. D., Levenson, A., Khan, O., Furth, P. A., Uren, A. and Toretsky, J. A., 2005. PTPL1 is a direct transcriptional target of EWS-FLI1 and modulates Ewing’s sarcoma tumorigenesis. Oncogene 24:2715–2722. Abadie, A., Besancon, F. and 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. Abadie, A. and Wietzerbin, J., 2003. Involvement of TNF-related apoptosis-inducing ligand (TRAIL) induction in interferon gamma-mediated apoptosis in Ewing tumor cells. Ann N Y Acad Sci 1010:117–120. Ambros, I. M., Ambros, P. F., Strehl, S., Kovar, H., Gadner, H. and Salzer-Kuntschik, M., 1991. MIC2 is a specific marker for Ewing’s sarcoma and peripheral primitive neuroectodermal tumors. Evidence for a common histogenesis of Ewing’s sarcoma and peripheral primitive neuroectodermal tumors from MIC2 expression and specific chromosome aberration. Cancer 67:1886–1893. Anderson, S. F., Schlegel, B. P., Nakajima, T., Wolpin, E. S. and Parvin, J. D., 1998. BRCA1 protein is linked to the RNA polymerase II holoenzyme complex via RNA helicase A. Nat Genet 19:254–256. Aratani, S., Fujii, R., Oishi, T., Fujita, H., Amano, T., Ohshima, T., Hagiwara, M., Fukamizu, A. and Nakajima, T., 2001. Dual roles of RNA helicase A in CREB-dependent transcription. Mol Cell Biol 21:4460–4469. Arndt, C. A. and Crist, W. M., 1999. Common musculoskeletal tumors of childhood and adolescence. N Engl J Med 341:342–352. Aryee, D. N., Kreppel, M., Bachmaier, R., Uren, A., Muehlbacher, K., Wagner, S., Breiteneder, H., Ban, J., Toretsky, J. A. and Kovar, H., 2006. Single-chain antibodies to the EWS NH2 terminus structurally discriminate between intact and chimeric EWS in Ewing’s sarcoma and interfere with the transcriptional activity of EWS in€vivo. Cancer Res 66:9862–9869. Ashkenazi, A., 2002. Targeting death and decoy receptors of the tumour-necrosis factor superfamily. Nat Rev Cancer 2:420–430. Baer, C., Nees, M., Breit, S., Selle, B., Kulozik, A. E., Schaefer, K. L., Braun, Y., Wai, D. and Poremba, C., 2004. Profiling and functional annotation of mRNA gene expression in pediatric rhabdomyosarcoma and Ewing’s sarcoma. Int J Cancer 110:687–694. Baliko, F., Bright, T., Poon, R., Cohen, B., Egan, S. E. and Alman, B. A., 2007. Inhibition of notch signaling induces neural differentiation in Ewing sarcoma. Am J Pathol 170:1686–1694. Baserga, R., 1995. The insulin-like growth factor I receptor: a key to tumor growth? Cancer Res 55:249–252. Beauchamp, E., Bulut, G., Abaan, O., Chen, K., Merchant, A., Matsui, W., Endo, Y., Rubin, J. S., Toretsky, J. and Uren, A., 2009. GLI1 is a direct transcriptional target of EWS-FLI1 oncoprotein. J Biol Chem 284:9074–9082. Begent, R. H., Verhaar, M. J., Chester, K. A., Casey, J. L., Green, A. J., Napier, M. P., Hope-Stone, L. D., Cushen, N., Keep, P. A., Johnson, C. J., Hawkins, R. E., Hilson, A. J. and Robson, L.,
392
J.A. Toretsky and A. Üren
1996. Clinical evidence of efficient tumor targeting based on single-chain Fv antibody selected from a combinatorial library. Nat Med 2:979–984. Benini, S., Manara, M. C., Baldini, N., Cerisano, V., Massimo, S., Mercuri, M., Lollini, P. L., Nanni, P., Picci, P. and Scotlandi, K., 2001. Inhibition of insulin-like growth factor I receptor increases the antitumor activity of doxorubicin and vincristine against Ewing’s sarcoma cells. Clin Cancer Res 7:1790–1797. Bertolotti, A., Melot, T., Acker, J., Vigneron, M., Delattre, O. and Tora, L., 1998. EWS, but not EWS-FLI-1, is associated with both TFIID and RNA polymerase II: interactions between two members of the TET family, EWS and hTAFII68, and subunits of TFIID and RNA polymerase II complexes. Mol Cell Biol 18:1489–1497. Bond, M., Bernstein, M. L., Pappo, A., Schultz, K. R., Krailo, M., Blaney, S. M. and Adamson, P. C., 2007. A phase II study of imatinib mesylate in children with refractory or relapsed solid tumors: a children’s oncology group study. Pediatr Blood Cancer 50(2):254–258 Bozzi, F., Tamborini, E., Negri, T., Pastore, E., Ferrari, A., Luksch, R., Casanova, M., Pierotti, M. A., Bellani, F. F. and Pilotti, S., 2007. Evidence for activation of KIT, PDGFRalpha, and PDGFRbeta receptors in the Ewing sarcoma family of tumors. Cancer 109:1638–1645. Braun, B. S., Frieden, R., Lessnick, S. L., May, W. A. and Denny, C. T., 1995. Identification of target genes for the Ewing’s sarcoma EWS/FLI fusion protein by representational difference analysis. Mol Cell Biol 15:4623–4630. Braunreiter, C. L., Hancock, J. D., Coffin, C. M., Boucher, K. M. and Lessnick, S. L., 2006. Expression of EWS-ETS fusions in NIH3T3 cells reveals significant differences to Ewing’s sarcoma. Cell Cycle 5:2753–2759. Butler, A. A., Yakar, S., Gewolb, I. H., Karas, M., Okubo, Y. and LeRoith, D., 1998. Insulin-like growth factor-I receptor signal transduction: at the interface between physiology and cell biology. Comp Biochem Physiol B Biochem Mol Biol 121:19–26. Castillero-Trejo, Y., Eliazer, S., Xiang, L., Richardson, J. A. and Ilaria, R. L., Jr., 2005. Expression of the EWS/FLI-1 oncogene in murine primary bone-derived cells Results in EWS/FLI-1dependent, Ewing sarcoma-like tumors. Cancer Res 65:8698–8705. Cavazzana, A., 1994. [The recent progress of biology in pediatric oncology (editorial)]. Pediatr Med Chir 16:197–199. Chan, D., Wilson, T. J., Xu, D., Cowdery, H. E., Sanij, E., Hertzog, P. J. and Kola, I., 2003. Transformation induced by Ewing’s sarcoma associated EWS/FLI-1 is suppressed by KRAB/ FLI-1. Br J Cancer 88:137–145. Chansky, H. A., Barahmand-Pour, F., Mei, Q., Kahn-Farooqi, W., Zielinska-Kwiatkowska, A., Blackburn, M., Chansky, K., Conrad, E. U., III, Bruckner, J. D., Greenlee, T. K. and Yang, L., 2004. Targeting of EWS/FLI-1 by RNA interference attenuates the tumor phenotype of Ewing’s sarcoma cells in€vitro. J Orthop Res 22:910–917. Chansky, H. A., Hu, M., Hickstein, D. D. and Yang, L., 2001. Oncogenic TLS/ERG and EWS/ Fli-1 fusion proteins inhibit RNA splicing mediated by YB-1 protein. Cancer Res 61:3586–3590. Codrington, R., Pannell, R., Forster, A., Drynan, L. F., Daser, A., Lobato, N., Metzler, M. and Rabbitts, T. H., 2005. The Ews-ERG fusion protein can initiate neoplasia from lineage-committed haematopoietic cells. PLoS Biol 3:e242. Collini, P., Sampietro, G., Bertulli, R., Casali, P. G., Luksch, R., Mezzelani, A., Sozzi, G. and Pilotti, S., 2001. Cytokeratin immunoreactivity in 41 cases of ES/PNET confirmed by molecular diagnostic studies. Am J Surg Pathol 25:273–274. de Alava, E., Kawai, A., Healey, J. H., Fligman, I., Meyers, P. A., Huvos, A. G., Gerald, W. L., Jhanwar, S. C., Argani, P., Antonescu, C. R., Pardo-Mindan, F. J., Ginsberg, J., Womer, R., Lawlor, E. R., Wunder, J., Andrulis, I., Sorensen, P. H., Barr, F. G. and Ladanyi, M., 1998. EWS-FLI1 fusion transcript structure is an independent determinant of prognosis in Ewing’s sarcoma. J Clin Oncol 16:1248–1255. Deneen, B. and Denny, C. T., 2001. Loss of p16 pathways stabilizes EWS/FLI1 expression and complements EWS/FLI1 mediated transformation. Oncogene 20:6731–6741.
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
393
Deneen, B., Hamidi, H. and Denny, C. T., 2003. Functional analysis of the EWS/ETS target gene uridine phosphorylase. Cancer Res 63:4268–4274. Denzin, L. K., Whitlow, M. and Voss, E. W., Jr., 1991. Single-chain site-specific mutations of fluorescein-amino acid contact residues in high affinity monoclonal antibody 4-4-20. J Biol Chem 266:14095–14103. Dobson, C. L., Warren, A. J., Pannell, R., Forster, A., Lavenir, I., Corral, J., Smith, A. J. and Rabbitts, T. H., 1999. The mll-AF9 gene fusion in mice controls myeloproliferation and specifies acute myeloid leukaemogenesis. EMBO J 18:3564–3574. Dohjima, T., Lee, N. S., Li, H., Ohno, T. and Rossi, J. J., 2003. Small interfering RNAs expressed from a Pol III promoter suppress the EWS/Fli-1 transcript in an Ewing sarcoma cell line. Mol Ther 7:811–816. Doolittle, R. F., Hunkapiller, M. W., Hood, L. E., Devare, S. G., Robbins, K. C., Aaronson, S. A. and Antoniades, H. N., 1983. Simian sarcoma virus onc gene, v-sis, is derived from the gene (or genes) encoding a platelet-derived growth factor. Science 221(4607):275–277. DuBois, S. G., Krailo, M. D., Lessnick, S. L., Smith, R., Chen, Z., Marina, N., Grier, H. E. and Stegmaier, K., 2009. Phase II study of intermediate-dose cytarabine in patients with relapsed or refractory Ewing sarcoma: a report from the Children’s Oncology Group. Pediatr Blood Cancer 52:324–327. Duiker, E. W., Mom, C. H., de Jong, S., Willemse, P. H., Gietema, J. A., van der Zee, A. G. and de Vries, E. G., 2006. The clinical trail of TRAIL. Eur J Cancer 42:2233–2240. Dworzak, M. N., Fritsch, G., Buchinger, P., Fleischer, C., Printz, D., Zellner, A., Schollhammer, A., Steiner, G., Ambros, P. F. and Gadner, H., 1994. Flow cytometric assessment of human MIC2 expression in bone marrow, thymus, and peripheral blood. Blood 83:415–425. Eliazer, S., Spencer, J., Ye, D., Olson, E. and Ilaria, R. L., Jr., 2003. Alteration of mesodermal cell differentiation by EWS/FLI-1, the oncogene implicated in Ewing’s sarcoma. Mol Cell Biol 23:482–492. Erkizan, V., Kong, Y., Merchant, M. S., Schlottman, S., Barber, J., Abaan, O. D., Chou, T., Dakshanamurthy, S., Brown, M. L., Uren, A. and Toretsky, J., 2009. Small molecule selected to disrupt mutant transcription factor EWS-FLI1 interaction with RNA Helicase A inhibits Ewing’s Sarcoma. Nature Medicine 15(7):750–756. Ewing, J., 1921. Diffuse endothelioma of bone. Proc N Y Pathol Soc 21:17. Fantl, W. J., Johnson, D. E. and Williams, L. T., 1993. Signalling by receptor tyrosine kinases. Annu Rev Biochem 62:453–481. Fellinger, E. J., Garin-Chesa, P., Triche, T. J., Huvos, A. G. and Rettig, W. J., 1991. Immunohistochemical analysis of Ewing’s sarcoma cell surface antigen p30/32MIC2. Am J Pathol 139:317–325. Fesik, S. W., 2005. Promoting apoptosis as a strategy for cancer drug discovery. Nat Rev Cancer 5:876–885. Forster, A., Pannell, R., Drynan, L. F., Codrington, R., Daser, A., Metzler, M., Lobato, M. N. and Rabbitts, T. H., 2005. The invertor knock-in conditional chromosomal translocation mimic. Nat Methods 2:27–30. Fuchs, B., Inwards, C. Y. and Janknecht, R., 2004. Vascular endothelial growth factor expression is up-regulated by EWS-ETS oncoproteins and Sp1 and may represent an independent predictor of survival in Ewing’s sarcoma. Clin Cancer Res 10:1344–1353. Fuchs, B., Valenzuela, R. G., Inwards, C., Sim, F. H. and Rock, M. G., 2003. Complications in long-term survivors of Ewing sarcoma. Cancer 98:2687–2692. Fukuma, M., Okita, H., Hata, J. and Umezawa, A., 2003. Upregulation of Id2, an oncogenic helix-loophelix protein, is mediated by the chimeric EWS/ets protein in Ewing sarcoma. Oncogene 22:1–9. Gill, Z. P., Perks, C. M., Newcomb, P. V. and Holly, J. M., 1997. Insulin-like growth factor-binding protein (IGFBP-3) predisposes breast cancer cells to programmed cell death in a non-IGFdependent manner. J Biol Chem 272:25602–25607. Girnita, L., Girnita, A., Wang, M., Meis-Kindblom, J. M., Kindblom, L. G. and Larsson, O., 2000. A link between basic fibroblast growth factor (bFGF) and EWS/FLI-1 in Ewing’s sarcoma cells. Oncogene 19:4298–4301.
394
J.A. Toretsky and A. Üren
Grier, H. E., Krailo, M. D., Tarbell, N. J., Link, M. P., Fryer, C. J., Pritchard, D. J., Gebhardt, M. C., Dickman, P. S., Perlman, E. J., Meyers, P. A., Donaldson, S. S., Moore, S., Rausen, A. R., Vietti, T. J. and Miser, J. S., 2003. Addition of ifosfamide and etoposide to standard chemotherapy for Ewing’s sarcoma and primitive neuroectodermal tumor of bone. N Engl J Med 348:694–701. Gu, M., Antonescu, C. R., Guiter, G., Huvos, A. G., Ladanyi, M. and Zakowski, M. F., 2000. Cytokeratin immunoreactivity in Ewing’s sarcoma: prevalence in 50 cases confirmed by molecular diagnostic studies. Am J Surg Pathol 24:410–416. Guan, H., Zhou, Z., Wang, H., Jia, S. F., Liu, W. and Kleinerman, E. S., 2005. A small interfering RNA targeting vascular endothelial growth factor inhibits Ewing’s sarcoma growth in a xenograft mouse model. Clin Cancer Res 11:2662–2669. Hahm, K. B., Cho, K., Lee, C., Im, Y. H., Chang, J., Choi, S. G., Sorensen, P. H., Thiele, C. J. and Kim, S. J., 1999. Repression of the gene encoding the TGF-beta type II receptor is a major target of the EWS-FLI1 oncoprotein. Nat Genet 23:222–227. Haldar, M., Hancock, J. D., Coffin, C. M., Lessnick, S. L. and Capecchi, M. R., 2007. A conditional mouse model of synovial sarcoma: insights into a myogenic origin. Cancer Cell 11:375–388. Hamilton, G., Havel, M. and Mallinger, R., 1989. Expression of a new human THY-1 related antigen in Ewing’s sarcoma and peripheral neuroectodermal tumors. Immunol Lett 22: 205–209. Hamilton, G., Mallinger, R. and Havel, M., 1989. Ewing’s-sarcoma-associated HBA-71 tumor antigen represents a new differentiation marker of human thymocytes. J Cancer Res Clin Oncol 115:592–596. Hamilton, G., Mallinger, R., Hofbauer, S. and Havel, M., 1991. The monoclonal HBA-71 antibody modulates proliferation of thymocytes and Ewing’s sarcoma cells by interfering with the action of insulin-like growth factor I. Thymus 18:33–41. Hara, K., Maruki, Y., Long, X., Yoshino, K., Oshiro, N., Hidayat, S., Tokunaga, C., Avruch, J. and Yonezawa, K., 2002. Raptor, a binding partner of target of rapamycin (TOR), mediates TOR action. Cell 110:177–189. Hara, K., Yonezawa, K., Weng, Q. P., Kozlowski, M. T., Belham, C. and Avruch, J., 1998. Amino acid sufficiency and mTOR regulate p70 S6 kinase and eIF-4E BP1 through a common effector mechanism. J Biol Chem 273:14484–14494. Heinrich, M. C., Blanke, C. D., Druker, B. J. and Corless, C. L., 2002. Inhibition of KIT tyrosine kinase activity: a novel molecular approach to the treatment of KIT-positive malignancies. J Clin Oncol 20:1692–1703. Heldin, C. H., Johnsson, A., Wennergren, S., Wernstedt, C., Betsholtz, C. and Westermark, B., 1986. A human osteosarcoma cell line secretes a growth factor structurally related to a homodimer of PDGF A-chains. Nature 319:511–514. Helman, L. J. and Meltzer, P., 2003. Mechanisms of sarcoma development. Nat Rev Cancer 3:685–694. Hofbauer, S., Hamilton, G., Theyer, G., Wollmann, K. and Gabor, F., 1993. Insulin-like growth factor-I-dependent growth and in€ vitro chemosensitivity of Ewing’s sarcoma and peripheral primitive neuroectodermal tumour cell lines. Eur J Cancer 29A:241–245. Hu-Lieskovan, S., Heidel, J. D., Bartlett, D. W., Davis, M. E. and Triche, T. J., 2005. Sequencespecific knockdown of EWS-FLI1 by targeted, nonviral delivery of small interfering RNA inhibits tumor growth in a murine model of metastatic Ewing’s sarcoma. Cancer Res 65:8984–8992. Huang, H. Y., Illei, P. B., Zhao, Z., Mazumdar, M., Huvos, A. G., Healey, J. H., Wexler, L. H., Gorlick, R., Meyers, P. and Ladanyi, M., 2005. Ewing sarcomas with p53 mutation or p16/ p14ARF homozygous deletion: a highly lethal subset associated with poor chemoresponse. J Clin Oncol 23:548–558. Jaboin, J., Wild, J., Hamidi, H., Khanna, C., Kim, C. J., Robey, R., Bates, S. E. and Thiele, C. J., 2002. MS-27-275, an inhibitor of histone deacetylase, has marked in€vitro and in€vivo antitumor activity against pediatric solid tumors. Cancer Res 62:6108–6115.
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
395
Jia, S. F., Zhou, R. R. and Kleinerman, E. S., 2003. Nude mouse lung metastases models of osteosarcoma and Ewing’s sarcoma for evaluating new therapeutic strategies. Methods Mol Med 74:495–505. Kang, H. G., Jenabi, J. M., Zhang, J., Keshelava, N., Shimada, H., May, W. A., Ng, T., Reynolds, C. P., Triche, T. J. and Sorensen, P. H., 2007. E-cadherin cell-cell adhesion in Ewing tumor cells mediates suppression of anoikis through activation of the ErbB4 tyrosine kinase. Cancer Res 67:3094–3105. Karnieli, E., Werner, H., Rauscher, F. J., III, Benjamin, L. E. and LeRoith, D., 1996. The IGF-I receptor gene promoter is a molecular target for the Ewing’s sarcoma-Wilms’ tumor 1 fusion protein. J Biol Chem 271:19304–19309 Keller, C., Arenkiel, B. R., Coffin, C. M., El-Bardeesy, N., DePinho, R. A. and Capecchi, M. R., 2004. Alveolar rhabdomyosarcomas in conditional Pax3:Fkhr mice: cooperativity of Ink4a/ ARF and Trp53 loss of function. Genes Dev 18:2614–2626. Khan, J., Wei, J. S., Ringner, M., Saal, L. H., Ladanyi, M., Westermann, F., Berthold, F., Schwab, M., Antonescu, C. R., Peterson, C. and Meltzer, P. S., 2001. Classification and diagnostic prediction of cancers using gene expression profiling and artificial neural networks. Nat Med 7:673–679. Kikuchi, R., Murakami, M., Sobue, S., Iwasaki, T., Hagiwara, K., Takagi, A., Kojima, T., Asano, H., Suzuki, M., Banno, Y., Nozawa, Y. and Murate, T., 2007. Ewing’s sarcoma fusion protein, EWS/Fli-1 and Fli-1 protein induce PLD2 but not PLD1 gene expression by binding to an ETS domain of 5¢ promoter. Oncogene 26:1802–1810. Kim, S., Denny, C. T. and Wisdom, R., 2006. Cooperative DNA binding with AP-1 proteins is required for transformation by EWS-Ets fusion proteins. Mol Cell Biol 26:2467–2478. Kinsey, M., Smith, R. and Lessnick, S. L., 2006. NR0B1 is required for the oncogenic phenotype mediated by EWS/FLI in Ewing’s sarcoma. Mol Cancer Res 4:851–859. Knoop, L. L. and Baker, S. J., 2000. The splicing factor U1C represses EWS/FLI-mediated transactivation. J Biol Chem 275:24865–24871. Knoop, L. L. and Baker, S. J., 2001. EWS/FLI alters 5¢-splice site selection. J Biol Chem 276:22317–22322. Kontny, H. U., Hammerle, K., Klein, R., Shayan, P., Mackall, C. L. and Niemeyer, C. M., 2001. Sensitivity of Ewing’s sarcoma to TRAIL-induced apoptosis. Cell Death Differ 8:506–514. Kontny, H. U., Lehrnbecher, T. M., Chanock, S. J. and Mackall, C. L., 1998. Simultaneous expression of Fas and nonfunctional Fas ligand in Ewing’s sarcoma. Cancer Res 58:5842–5849. Kovar, H., Aryee, D. N., Jug, G., Henockl, C., Schemper, M., Delattre, O., Thomas, G. and Gadner, H., 1996. EWS/FLI-1 antagonists induce growth inhibition of Ewing tumor cells in€vitro. Cell Growth Differ 7:429–437. Kovar, H., Ban, J. and Pospisilova, S., 2003. Potentials for RNAi in sarcoma research and therapy: Ewing’s sarcoma as a model. Semin Cancer Biol 13:275–281. Kreppel, M., Aryee, D. N., Schaefer, K. L., Amann, G., Kofler, R., Poremba, C. and Kovar, H., 2006. Suppression of KCMF1 by constitutive high CD99 expression is involved in the migratory ability of Ewing’s sarcoma cells. Oncogene 25:2795–2800. Lambert, G., Bertrand, J. R., Fattal, E., Subra, F., Pinto-Alphandary, H., Malvy, C., Auclair, C. and Couvreur, P., 2000. EWS fli-1 antisense nanocapsules inhibits Ewing sarcoma-related tumor in mice. Biochem Biophys Res Commun 279:401–406. LaRochelle, W. J., Jensen, R. A., Heidaran, M. A., May-Siroff, M., Wang, L. M., Aaronson, S. A. and Pierce, J. H., 1993. Inhibition of platelet-derived growth factor autocrine growth stimulation by a monoclonal antibody to the human alpha platelet-derived growth factor receptor. Cell Growth Differ 4:547–553. Lawlor, E. R., Scheel, C., Irving, J. and Sorensen, P. H., 2002. Anchorage-independent multicellular spheroids as an in€ vitro model of growth signaling in Ewing tumors. Oncogene 21:307–318. Lee, T. H., Bolontrade, M. F., Worth, L. L., Guan, H., Ellis, L. M. and Kleinerman, E. S., 2006. Production of VEGF165 by Ewing’s sarcoma cells induces vasculogenesis and the incorporation of CD34+ stem cells into the expanding tumor vasculature. Int J Cancer 119:839–846.
396
J.A. Toretsky and A. Üren
Lessnick, S. L., Dacwag, C. S. and Golub, T. R., 2002. The Ewing’s sarcoma oncoprotein EWS/ FLI induces a p53-dependent growth arrest in primary human fibroblasts. Cancer Cell 1:393–401. Li, R., Pei, H. and Watson, D. K., 2000. Regulation of Ets function by protein–protein interactions. Oncogene 19:6514–6523. Li, X., Ponten, A., Aase, K., Karlsson, L., Abramsson, A., Uutela, M., Backstrom, G., Hellstrom, M., Bostrom, H., Li, H., Soriano, P., Betsholtz, C., Heldin, C. H., Alitalo, K., Ostman, A. and Eriksson, U., 2000. PDGF-C is a new protease-activated ligand for the PDGF alpha-receptor. Nat Cell Biol 2:302–309. Lin, P. P., Pandey, M. K., Jin, F., Xiong, S., Deavers, M., Parant, J. M. and Lozano, G., 2008. EWS-FLI1 induces developmental abnormalities and accelerates sarcoma formation in a transgenic mouse model. Cancer Res 68:8968–8975. Lipinski, M., Braham, K., Philip, I., Wiels, J., Philip, T., Goridis, C., Lenoir, G. M. and Tursz, T., 1987. Neuroectoderm-associated antigens on Ewing’s sarcoma cell lines. Cancer Res 47:183–187. Lissat, A., Vraetz, T., Tsokos, M., Klein, R., Braun, M., Koutelia, N., Fisch, P., Romero, M. E., Long, L., Noellke, P., Mackall, C. L., Niemeyer, C. M. and Kontny, U., 2007. Interferongamma sensitizes resistant Ewing’s sarcoma cells to tumor necrosis factor apoptosis-inducing ligand-induced apoptosis by up-regulation of caspase-8 without altering chemosensitivity. Am J Pathol 170:1917–1930. Lynch, T. J., Bell, D. W., Sordella, R., Gurubhagavatula, S., Okimoto, R. A., Brannigan, B. W., Harris, P. L., Haserlat, S. M., Supko, J. G., Haluska, F. G., Louis, D. N., Christiani, D. C., Settleman, J. and Haber, D. A., 2004. Activating mutations in the epidermal growth factor receptor underlying responsiveness of non-small-cell lung cancer to gefitinib. N Engl J Med 350:2129–2139. Maksimenko, A., Lambert, G., Bertrand, J. R., Fattal, E., Couvreur, P. and Malvy, C., 2003. Therapeutic potentialities of EWS-Fli-1 mRNA-targeted vectorized antisense oligonucleotides. Ann N Y Acad Sci 1002:72–77. Maksimenko, A., Malvy, C., Lambert, G., Bertrand, J. R., Fattal, E., Maccario, J. and Couvreur, P., 2003. Oligonucleotides targeted against a junction oncogene are made efficient by nanotechnologies. Pharm Res 20:1565–1567. Mateo-Lozano, S., Gokhale, P. C., Soldatenkov, V. A., Dritschilo, A., Tirado, O. M. and Notario, V., 2006. Combined transcriptional and translational targeting of EWS/FLI-1 in Ewing’s sarcoma. Clin Cancer Res 12:6781–6790. Matsunobu, T., Tanaka, K., Matsumoto, Y., Nakatani, F., Sakimura, R., Hanada, M., Li, X., Oda, Y., Naruse, I., Hoshino, H., Tsuneyoshi, M., Miura, H. and Iwamoto, Y., 2004. The prognostic and therapeutic relevance of p27kip1 in Ewing’s family tumors. Clin Cancer Res 10:1003–1012. May, W. A., Lessnick, S. L., Braun, B. S., Klemsz, M., Lewis, B. C., Lunsford, L. B., Hromas, R. and Denny, C. T., 1993. The Ewing’s sarcoma EWS/FLI-1 fusion gene encodes a more potent transcriptional activator and is a more powerful transforming gene than FLI-1. Mol Cell Biol 13:7393–7398. McKeon, C., Thiele, C. J., Ross, R. A., Kwan, M., Triche, T. J., Miser, J. S. and Israel, M. A., 1988. Indistinguishable patterns of protooncogene expression in two distinct but closely related tumors: Ewing’s sarcoma and neuroepithelioma. Cancer Res 48:4307–4311. Mendiola, M., Carrillo, J., Garcia, E., Lalli, E., Hernandez, T., de Alava, E., Tirode, F., Delattre, O., Garcia-Miguel, P., Lopez-Barea, F., Pestana, A. and Alonso, J., 2006. The orphan nuclear receptor DAX1 is up-regulated by the EWS/FLI1 oncoprotein and is highly expressed in Ewing tumors. Int J Cancer 118:1381–1389. Merchant, M. S., Woo, C. W., Mackall, C. L. and Thiele, C. J., 2002. Potential use of imatinib in Ewing’s sarcoma: evidence for in€vitro and in€vivo activity. J Natl Cancer Inst 94:1673–1679. Merchant, M. S., Yang, X., Melchionda, F., Romero, M., Klein, R., Thiele, C. J., Tsokos, M., Kontny, H. U. and Mackall, C. L., 2004. Interferon gamma enhances the effectiveness of tumor necrosis factor-related apoptosis-inducing ligand receptor agonists in a xenograft model of Ewing’s sarcoma. Cancer Res 64:8349–8356.
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
397
Nakajima, T., Uchida, C., Anderson, S. F., Lee, C. G., Hurwitz, J., Parvin, J. D. and Montminy, M., 1997. RNA helicase A mediates association of CBP with RNA polymerase II. Cell 90:1107–1112. Nakatani, F., Tanaka, K., Sakimura, R., Matsumoto, Y., Matsunobu, T., Li, X., Hanada, M., Okada, T. and Iwamoto, Y., 2003. Identification of p21WAF1/CIP1 as a direct target of EWS-Fli1 oncogenic fusion protein. J Biol Chem 278:15105–15115. Ng, K. P., Potikyan, G., Savene, R. O., Denny, C. T., Uversky, V. N. and Lee, K. A., 2007. Multiple aromatic side chains within a disordered structure are critical for transcription and transforming activity of EWS family oncoproteins. Proc Natl Acad Sci U S A 104:479–484. Nishimori, H., Sasaki, Y., Yoshida, K., Irifune, H., Zembutsu, H., Tanaka, T., Aoyama, T., Hosaka, T., Kawaguchi, S., Wada, T., Hata, J., Toguchida, J., Nakamura, Y. and Tokino, T., 2002. The Id2 gene is a novel target of transcriptional activation by EWS-ETS fusion proteins in Ewing family tumors. Oncogene 21:8302–8309. Noguera, R., Triche, T. J., Navarro, S., Tsokos, M. and Llombart, B. A., 1992. Dynamic model of differentiation in Ewing’s sarcoma cells. Comparative analysis of morphologic, immunocytochemical, and oncogene expression parameters [see comments]. Lab Invest 66:143–151. Ostman, A. and Heldin, C. H., 2001. Involvement of platelet-derived growth factor in disease: development of specific antagonists. Adv Cancer Res 80:1–38. Ouchida, M., Ohno, T., Fujimura, Y., Rao, V. N. and Reddy, E. S., 1995. Loss of tumorigenicity of Ewing’s sarcoma cells expressing antisense RNA to EWS-fusion transcripts. Oncogene 11:1049–1054. Pao, W., Miller, V., Zakowski, M., Doherty, J., Politi, K., Sarkaria, I., Singh, B., Heelan, R., Rusch, V., Fulton, L., Mardis, E., Kupfer, D., Wilson, R., Kris, M. and Varmus, H., 2004. EGF receptor gene mutations are common in lung cancers from “never smokers” and are associated with sensitivity of tumors to gefitinib and erlotinib. Proc Natl Acad Sci U S A 101:13306–13311. Petermann, R., Mossier, B. M., Aryee, D. N., Khazak, V., Golemis, E. A. and Kovar, H., 1998. Oncogenic EWS-Fli1 interacts with hsRPB7, a subunit of human RNA polymerase II. Oncogene 17:603–610. Potikyan, G., Savene, R. O., Gaulden, J. M., France, K. A., Zhou, Z., Kleinerman, E. S., Lessnick, S. L. and Denny, C. T., 2007. EWS/FLI1 regulates tumor angiogenesis in Ewing’s sarcoma via suppression of thrombospondins. Cancer Res 67:6675–6684. Prieur, A., Tirode, F., Cohen, P. and Delattre, O., 2004. EWS/FLI-1 silencing and gene profiling of Ewing cells reveal downstream oncogenic pathways and a crucial role for repression of insulin-like growth factor binding protein 3. Mol Cell Biol 24:7275–7283. Rajah, R., Valentinis, B. and Cohen, P., 1997. Insulin-like growth factor (IGF)-binding protein-3 induces apoptosis and mediates the effects of transforming growth factor-beta1 on programmed cell death through a p53- and IGF-independent mechanism. J Biol Chem 272:12181–12188. Reya, T., Morrison, S. J., Clarke, M. F. and Weissman, I. L., 2001. Stem cells, cancer, and cancer stem cells. Nature 414:105–111. Riggi, N., Cironi, L., Provero, P., Suva, M. L., Kaloulis, K., Garcia-Echeverria, C., Hoffmann, F., Trumpp, A. and Stamenkovic, I., 2005. Development of Ewing’s sarcoma from primary bone marrow-derived mesenchymal progenitor cells. Cancer Res 65:11459–11468. Ronnstrand, L., 2004. Signal transduction via the stem cell factor receptor/c-Kit. Cell Mol Life Sci 61:2535–2548. Rorie, C. J., Thomas, V. D., Chen, P., Pierce, H. H., O’Bryan, J. P. and Weissman, B. E., 2004. The Ews/Fli-1 fusion gene switches the differentiation program of neuroblastomas to Ewing sarcoma/peripheral primitive neuroectodermal tumors. Cancer Res 64:1266–1277. Sakimura, R., Tanaka, K., Nakatani, F., Matsunobu, T., Li, X., Hanada, M., Okada, T., Nakamura, T., Matsumoto, Y. and Iwamoto, Y., 2005. Antitumor effects of histone deacetylase inhibitor on Ewing’s family tumors. Int J Cancer 116:784–792. Saylors, R. L., III, Stine, K. C., Sullivan, J., Kepner, J. L., Wall, D. A., Bernstein, M. L., Harris, M. B., Hayashi, R. and Vietti, T. J., 2001. Cyclophosphamide plus topotecan in children with recurrent or refractory solid tumors: a Pediatric Oncology Group phase II study. J Clin Oncol 19:3463–3469.
398
J.A. Toretsky and A. Üren
Schuetz, A. N., Rubin, B. P., Goldblum, J. R., Shehata, B., Weiss, S. W., Liu, W., Wick, M. R. and Folpe, A. L., 2005. Intercellular junctions in Ewing sarcoma/primitive neuroectodermal tumor: additional evidence of epithelial differentiation. Mod Pathol 18:1403–1410. Schweigerer, L., Neufeld, G. and Gospodarowicz, D., 1987. Basic fibroblast growth factor as a growth inhibitor for cultured human tumor cells. J Clin Invest 80:1516–1520. Scotlandi, K., Benini, S., Nanni, P., Lollini, P. L., Nicoletti, G., Landuzzi, L., Serra, M., Manara, M. C., Picci, P. and Baldini, N., 1998. Blockage of insulin-like growth factor-I receptor inhibits the growth of Ewing’s sarcoma in athymic mice. Cancer Res 58:4127–4131. Scotlandi, K., Manara, M. C., Nicoletti, G., Lollini, P. L., Lukas, S., Benini, S., Croci, S., Perdichizzi, S., Zambelli, D., Serra, M., Garcia-Echeverria, C., Hofmann, F. and Picci, P., 2005. Antitumor activity of the insulin-like growth factor-I receptor kinase inhibitor NVPAEW541 in musculoskeletal tumors. Cancer Res 65:3868–3876. Scotlandi, K., Perdichizzi, S., Bernard, G., Nicoletti, G., Nanni, P., Lollini, P. L., Curti, A., Manara, M. C., Benini, S., Bernard, A. and Picci, P., 2006. Targeting CD99 in association with doxorubicin: an effective combined treatment for Ewing’s sarcoma. Eur J Cancer 42:91–96. Scotlandi, K., Perdichizzi, S., Manara, M. C., Serra, M., Benini, S., Cerisano, V., Strammiello, R., Mercuri, M., Reverter-Branchat, G., Faircloth, G., D’Incalci, M. and Picci, P., 2002. Effectiveness of Ecteinascidin-743 against drug-sensitive and -resistant bone tumor cells. Clin Cancer Res 8:3893–3903. Siligan, C., Ban, J., Bachmaier, R., Spahn, L., Kreppel, M., Schaefer, K. L., Poremba, C., Aryee, D. N. and Kovar, H., 2005. EWS-FLI1 target genes recovered from Ewing’s sarcoma chromatin. Oncogene 24:2512–2524. Singleton, J. R., Dixit, V. M. and Feldman, E. L., 1996. Type I insulin-like growth factor receptor activation regulates apoptotic proteins. J Biol Chem 271:31791–31794. Smith, R., Owen, L. A., Trem, D. J., Wong, J. S., Whangbo, J. S., Golub, T. R. and Lessnick, S. L., 2006. Expression profiling of EWS/FLI identifies NKX2.2 as a critical target gene in Ewing’s sarcoma. Cancer Cell 9:405–416. Spahn, L., Siligan, C., Bachmaier, R., Schmid, J. A., Aryee, D. N. and Kovar, H., 2003. Homotypic and heterotypic interactions of EWS, FLI1 and their oncogenic fusion protein. Oncogene 22:6819–6829. Staege, M. S., Hutter, C., Neumann, I., Foja, S., Hattenhorst, U. E., Hansen, G., Afar, D. and Burdach, S. E., 2004. DNA microarrays reveal relationship of Ewing family tumors to both endothelial and fetal neural crest-derived cells and define novel targets. Cancer Res 64:8213–8221. Stegmaier, K., Wong, J. S., Ross, K. N., Chow, K. T., Peck, D., Wright, R. D., Lessnick, S. L., Kung, A. L. and Golub, T. R., 2007. Signature-based small molecule screening identifies cytosine arabinoside as an EWS/FLI modulator in Ewing sarcoma. PLoS Med 4:e122. Takeda, K., Stagg, J., Yagita, H., Okumura, K. and Smyth, M. J., 2007. Targeting death-inducing receptors in cancer therapy. Oncogene 26:3745–3757. Tamborini, E., Bonadiman, L., Albertini, V., Pierotti, M. A. and Pilotti, S., 2003. Re: Potential use of imatinib in Ewing’s sarcoma: evidence for in€vitro and in€vivo activity. J Natl Cancer Inst 95:1087–1088; author reply 1088–1089 Tanaka, K., Iwakuma, T., Harimaya, K., Sato, H. and Iwamoto, Y., 1997. EWS-Fli1 antisense oligodeoxynucleotide inhibits proliferation of human Ewing’s sarcoma and primitive neuroectodermal tumor cells. J Clin Invest 99:239–247. Tanaka, S., Ito, T. and Wands, J. R., 1996. Neoplastic transformation induced by insulin receptor substrate-1 overexpression requires an interaction with both Grb2 and Syp signaling molecules. J Biol Chem 271:14610–14616. Tang, H. and Wong-Staal, F., 2000. Specific interaction between RNA helicase A and Tap, two cellular proteins that bind to the constitutive transport element of type D retrovirus. J Biol Chem 275:32694–32700. Tirado, O. M., Mateo-Lozano, S., Villar, J., Dettin, L. E., Llort, A., Gallego, S., Ban, J., Kovar, H. and Notario, V., 2006. Caveolin-1 (CAV1) is a target of EWS/FLI-1 and a key determinant of
Ewing’s Sarcoma Family of Tumors: Molecular Targets Need Arrows
399
the oncogenic phenotype and tumorigenicity of Ewing’s sarcoma cells. Cancer Res 66:9937–9947. Tirode, F., Laud-Duval, K., Prieur, A., Delorme, B., Charbord, P. and Delattre, O., 2007. Mesenchymal stem cell features of Ewing tumors. Cancer Cell 11:421–429. Torchia, E. C., Boyd, K., Rehg, J. E., Qu, C. and Baker, S. J., 2007. EWS/FLI-1 induces rapid onset of myeloid/erythroid leukemia in mice. Mol Cell Biol 27:7918–7934. Torchia, E. C., Jaishankar, S. and Baker, S. J., 2003. Ewing tumor fusion proteins block the differentiation of pluripotent marrow stromal cells. Cancer Res 63:3464–3468. Toretsky, J. A., 2003. Ewing sarcoma and primitive neuroectodermal tumors. January 17, 2003. http://www.emedicine.com/ped/topic2589.htm Toretsky, J. A., Connell, Y., Neckers, L. and Bhat, N. K., 1997. Inhibition of EWS-FLI-1 fusion protein with antisense oligodeoxynucleotides. J Neurooncol 31:9–16. Toretsky, J. A., Erkizan, V., Levenson, A., Abaan, O. D., Parvin, J. D., Cripe, T. P., Rice, A. M., Lee, S. B. and Uren, A., 2006. Oncoprotein EWS-FLI1 activity is enhanced by RNA helicase A. Cancer Res 66:5574–5581. Toretsky, J. A., Kalebic, T., Blakesley, V., LeRoith, D. and Helman, L. J., 1997. The insulin-like growth factor-I receptor is required for EWS/FLI-1 transformation of fibroblasts. J Biol Chem 272:30822–30827. Toretsky, J. A., Steinberg, S. M., Thakar, M., Counts, D., Pironis, B., Parente, C., Eskenazi, A., Helman, L. and Wexler, L. H., 2001. Insulin-like growth factor type 1 (IGF-1) and IGF binding protein-3 in patients with Ewing sarcoma family of tumors. Cancer 92:2941–2947. Toretsky, J. A., Thakar, M., Eskenazi, A. E. and Frantz, C. N., 1999. Phosphoinositide 3-hydroxide kinase blockade enhances apoptosis in the Ewing’s sarcoma family of tumors. Cancer Res 59:5745–5750. Turc-Carel, C., Philip, I., Berger, M.-P., Philip, T. and Lenoir, G. M., 1984. Chromosome study of Ewing’s sarcoma (ES) cell lines. Consistency of a reciprocal translocation t(11;22)(q24;q12). Cancer Genet Cytogenet 12:1–19. Uren, A., Merchant, M. S., Sun, C. J., Vitolo, M. I., Sun, Y., Tsokos, M., Illei, P. B., Ladanyi, M., Passaniti, A., Mackall, C. and Toretsky, J. A., 2003. Beta-platelet-derived growth factor receptor mediates motility and growth of Ewing’s sarcoma cells. Oncogene 22:2334–2342. Uren, A., Tcherkasskaya, O. and Toretsky, J. A., 2004. Recombinant EWS-FLI1 oncoprotein activates transcription. Biochemistry 43:13579–13589. Uren, A., Wolf, V., Sun, Y. F., Azari, A., Rubin, J. S. and Toretsky, J. A., 2004. Wnt/Frizzled signaling in Ewing sarcoma. Pediatr Blood Cancer 43:243–249. Valentinis, B. and Baserga, R., 1996. The IGF-I receptor protects tumor cells from apoptosis induced by high concentrations of serum. Biochem Biophys Res Commun 224:362–368. van Valen, F., Winkelmann, W. and Jurgens, H., 1992. Type I and type II insulin-like growth factor receptors and their function in human Ewing’s sarcoma cells. J Cancer Res Clin Oncol 118:269–275. Vormoor, J., Baersch, G., Decker, S., Hotfilder, M., Schafer, K. L., Pelken, L., Rube, C., van Valen, F., Jurgens, H. and Dockhorn-Dworniczak, B., 2001. Establishment of an in€vivo model for pediatric Ewing tumors by transplantation into NOD/scid mice. Pediatr Res 49:332–341. Wagner, L. M., McAllister, N., Goldsby, R. E., Rausen, A. R., McNall-Knapp, R. Y., McCarville, M. B. and Albritton, K., 2007. Temozolomide and intravenous irinotecan for treatment of advanced Ewing sarcoma. Pediatr Blood Cancer 48:132–139. Waterfield, M. D., Scrace, G. T., Whittle, N., Stroobant, P., Johnsson, A., Wasteson, A., Westemark, B., Heldin, C.-H., Huang, J. S. and Deuel, T. F., 1983. Platelet-derived growth factor is structurally related to the putative transforming protein p28sis of simian sarcoma virus. Nature 304:35–39. Watson, D. K., Robinson, L., Hodge, D. R., Kola, I., Papas, T. S. and Seth, A., 1997. FLI1 and EWS-FLI1 function as ternary complex factors and ELK1 and SAP1a function as ternary and quaternary complex factors on the Egr1 promoter serum response elements. Oncogene 14:213–221.
400
J.A. Toretsky and A. Üren
Weidner, N. and Tjoe, J., 1994. Immunohistochemical profile of monoclonal antibody O13: antibody that recognizes glycoprotein p30/32MIC2 and is useful in diagnosing Ewing’s sarcoma and peripheral neuroepithelioma. Am J Surg Pathol 18:486–494. Welford, S. M., Hebert, S. P., Deneen, B., Arvand, A. and Denny, C. T., 2001. DNA binding domain-independent pathways are involved in EWS/FLI1-mediated oncogenesis. J Biol Chem 276:41977–41984. Westwood, G., Dibling, B. C., Cuthbert-Heavens, D. and Burchill, S. A., 2002. Basic fibroblast growth factor (bFGF)-induced cell death is mediated through a caspase-dependent and p53independent cell death receptor pathway. Oncogene 21:809–824. Whang-Peng, J., Triche, T. J., Knutsen, T., Miser, J., Douglass, E. C. and Israel, M. A., 1984. Chromosome translocation in peripheral neuroepithelioma. N Engl J Med 311:584–585. Whitesell, L., Shifrin, S. D., Schwab, G. and Neckers, L. M., 1992. Benzoquinonoid ansamycins possess selective tumoricidal activity unrelated to src kinase inhibition. Cancer Res 52:1721–1728. Williams, L. T., 1989. Signal transduction by the platelet-derived growth factor receptor. Science 243:1564–1570. Williamson, A. J., Dibling, B. C., Boyne, J. R., Selby, P. and Burchill, S. A., 2004. Basic fibroblast growth factor (bFGF)-induced cell death is effected through sustained activation of p38MAPK and up-regulation of the death receptor p75NTR. J Biol Chem 279(46):47912–47928 Yang, L., Chansky, H. A. and Hickstein, D. D., 2000. EWS/Fli-1 fusion protein interacts with hyperphosphorylated RNA polymerase II and interferes with serine-arginine protein-mediated RNA splicing. J Biol Chem 275:37612–37618. Yee, D., Favoni, R. E., Lebovic, G. S., Lombana, F., Powell, D. R., Reynolds, C. P. and Rosen, N., 1990. Insulin-like growth factor I expression by tumors of neuroectodermal origin with the t(11;22) chromosomal translocation. A potential autocrine growth factor. J Clin Invest 86:1806–1814. Yin, X., Giap, C., Lazo, J. S. and Prochownik, E. V., 2003. Low molecular weight inhibitors of Myc-Max interaction and function. Oncogene 22:6151–6159. Zoubek, A., Dockhorn-Dworniczak, B., Delattre, O., Christiansen, H., Niggli, F., Gatterer-Menz, I., Smith, T. L., Jurgens, H., Gadner, H. and Kovar, H., 1996. Does expression of different EWS chimeric transcripts define clinically distinct risk groups of Ewing tumor patients? J Clin Oncol 14:1245–1251. Zwerner, J. P., Guimbellot, J. and May, W. A., 2003. EWS/FLI function varies in different cellular backgrounds. Exp Cell Res 290:414–419. Zwerner, J. P., Joo, J., Warner, K. L., Christensen, L., Hu-Lieskovan, S., Triche, T. J. and May, W. A., 2008. The EWS/FLI1 oncogenic transcription factor deregulates GLI1. Oncogene 27:3282–3291. Zwerner, J. P. and May, W. A., 2001. PDGF-C is an EWS/FLI induced transforming growth factor in Ewing family tumors. Oncogene 20:626–633. Zwerner, J. P. and May, W. A., 2002. Dominant negative PDGF-C inhibits growth of Ewing family tumor cell lines. Oncogene 21:3847–3854.
Molecular Targeted Therapy for Wilms’ Tumor James I. Geller and Jeffrey S. Dome
Introduction Wilms’ tumor (nephroblastoma, WT) is the most common pediatric primary renal malignancy, originating from aberrant differentiation of a pluripotent renal stem cell derived from embryogenic metanephric blastema (Beckwith et€al. 1990). In the United States, WT has an annual incidence of 7.6 cases per million children, with approximately 500 new cases diagnosed each year, accounting for 6% of all childhood cancers (Bernstein et€al. 1999). The incidence rate is slightly higher in girls (female:male is 1.09), is slightly higher in black children, and significantly lower in Asian children (Dome et€ al. 2006a). Over 77% of WT patients are diagnosed prior to age 5, with girls and boys presenting at a median age of 3 and 2€ years, respectively (Pastore et€al. 2006).
Current Treatment Strategies Treatment for WT utilizes a combination of surgery, chemotherapy, and radiation. Traditionally, the National Wilms Tumor Study Group (NWTSG) has favored upfront surgical resection followed by adjuvant chemotherapy and radiation as needed, whereas the International Society of Pediatric Oncology (SIOP) has favored neoadjuvant chemotherapy followed by resection and additional chemotherapy and radiation as needed. Both approaches have yielded good results. Antimitotic agents, anthracyclines, topoisomerase II inhibitors, oxazophosphorines, platinum analogs, and now camptothecins have all shown clinical activity in WT patients and form the basis of current therapeutic strategies, which aim to maximize cure and simultaneously minimize toxicity.
J.S. Domeâ•›(*) Division of Oncology, Center for Cancer and Blood Disorders, Children’s National Medical Center, 111 Michigan Avenue NW, Washington, DC 20010, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_19, © Springer Science+Business Media, LLC 2010
401
402
J.I. Geller and J.S. Dome
The current SIOP-2001 protocol employs a treatment algorithm built upon (1) randomization of stage II and stage III patients to upfront vincristine/dactinomycin versus vincristine/dactinomycin/doxorubicin; and (2) postoperative chemotherapy dependent on histologic features assessed at the time of postneoadjuvant chemotherapy nephrectomy (Metzger and Dome 2005). The COG Renal Tumor Committee (RTC), which has replaced the NWTSG, has recently analyzed the results of prior NWTSG and SIOP studies and designed new biologic and clinical pediatric renal tumor trials. At the helm of this new design is a risk stratification protocol termed AREN03B2, whereby pathology slides, tumor cytogenetics (loss of heterozygosity at chromosomes 1p and16q), operative reports, and radiology studies are centrally reviewed by national experts. Based on a central risk assignment, patients are enrolled in one of four therapeutic studies: AREN0532 (Very low and standard-risk favorable histology (FH) Wilms tumor), AREN0533 (Higher-risk favorable histology Wilms tumor), AREN0321 (high-risk renal tumors), and AREN0534 (Bilateral Wilms tumor or bilaterally-predisposed). AREN03B2 also serves to maintain a biological samples bank for future research and monitors outcomes for patients not eligible for a COG therapeutic study. Ninety percent of WT patients have favorable histology, of which 10 to 15% relapse, 10% have anaplastic histology, of which 40% relapse, and 7% present with or develop bilateral disease, of which 30 to 40% relapse (Green 2004; Dome et€al. 2006b; Pastore et€ al. 2006). Thus, approximately 15% of WT patients develop recurrent disease and are at risk for significant morbidity and mortality. These patients require more effective and tolerable treatment approaches.
Molecular Pathways in Wilms Tumor Although Wilms tumor was one of the original models for Knudson’s two-hit hypothesis, (Knudson and Strong 1972) the genetically heterogeneous nature of WT pathogenesis is now undeniable. Over the last two decades, preclinical and clinical investigation have led to significant advances in our understanding of WT histologic differentiation and pathogenesis, WT predisposition, genetics and epigenetics, putative WT tumor suppressor genes and oncogenes, and of WT signal transduction pathway integrity and function. From such research, clinically meaningful prognostic variables have emerged enabling the application of risk-stratified therapy based on WT histology and genetic features. In addition, such advances provide a preliminary understanding of the cancer’s heterogeneous molecular profile, forming the basis for the development and application of new molecular targeted therapies. The complexity of Wilms tumor genetics is reflected by the number of Wilms tumor susceptibility genes and loci that have been identified (Table€1). Rather than focusing on individual genes, this review discusses Wilms tumor biology in the context of molecular pathways. We first discuss pathways that are defective in Wilms tumor and then review preclinical efforts to test novel therapeutic agents.
20% 9%
Deletion Mutation 3p21 CTNNB1 (b-Catenin) Mutation 17q12-q21 FWT1 Mutation-germline 19q13.3-q13.4 FWT2 Mutation-germline LOH 7p14-15 POU6F2 Mutation-germline LOH 1p13 DCLRE1B LOH 1q25.3 CACNA1E Amplification 1q25-q32 HPRT (Parafibromin) LOH 5q21-q22 APC LOH MSI 6q21 HACE Hypermethylation 9q P16INK4a LOH Methylation 11q23.2 SKCG-1 Loss (PCR); LOH? 15q IGFR1 Gain 16q E-Cadherin LOH; mutation
DCC
18q21.1
LOH MSI
21.6% 7.8% 15% Unknown Unknown Unknown 2/24 16% Unknown 8% 4/10 30% 16% 4/5 11.8% 23.7% 38% 12% 2/20; 8/35
WTX
Xq11.1
Perotti et€al. (2004)
Reference Gessler et€al. (1994) Baudry et€al. (2000) Diller et€al. (1998) Grundy et€al. (1994, 1996) Ravenel et€al. (2001, 2002) Pritchard-Jones and Vujanic (2006) Grundy et€al. (1994, 1996) Rivera et€al. (2007) Major et€al. (2007) Koesters et€al. (1996) Rahman et€al. (1996, 1998) McDonald et€al. 1998
Tumor suppressor; apoptosis
Tumor suppressor? Oncogene; growth, Anti-apoptosis Adhesion; b-catenin
Ubiquitination CDK inhibition
Fernandeaz et€al. (2001) Natrajan et€al. (2007b) Arcellana-Panlilio et€al. (2000) Singh and Roy (2006) Natrajan et€al. (2007c) Schulz et€al. (2000) Safford et€al. (2005) Ramburan et€al. (2004)
DNA repair; telomere Natrajan et€al. (2007a) Cav2.3 – PKC; MEK/ERK; EGR; b-Catenin Natrajan et€al. (2006a) RNA processing; b-catenin Zhao et€al. (2007) b-Catenin Ramburan et€al. (2005)
Unknown
Oncogene; b-catenin Unknown Tumor suppressor?
Tumor suppressor; b-catenin
Table€1â•… Putative genetic loci/genes involved in WT pathogenesis Locus Gene Genetic event Frequency Function/pathway 11p13 WT1 Mutation 10% Tumor suppressor Sporadic 10% Germline 4% LOH 30% 11p15.5 WT2 (IGF2/H19) LOI 25–40% Oncogene – growth; Anti-apoptosis LOH 35%
Molecular Targeted Therapy for Wilms’ Tumor 403
404
J.I. Geller and J.S. Dome
IGF Signaling The role of IGF signaling in cancer has received significant attention concurrent with the evolution of IGFR1 inhibitors that are now in early phase clinical trials. In general, via presumed pleiotropic mechanisms, IGF signaling provides an antiapoptotic signal, mediates proliferation, and promotes tumorigenesis (Werner and Le Roith 1997; Kurmasheva and Houghton 2006; Reidemann and Macaulay 2006). The net effect of alterations in the IGF signaling pathway is challenging to predict, (Samani et€ al. 2007) however, and growth factor receptor cross talk, particularly with the EGFR pathway, is adding complexity and possibly opportunity. IGF2, a well characterized fetal mitogen, has long been implicated in the pathogenesis of Wilms tumor. IGF2 is an imprinted gene that resides at chromosome 11p15.5, the locus for Beckwith-Wiedemann Syndrome (BWS), a cancer predisposition syndrome associated with a 5% risk of WT. (Koufos et€al. 1989; Ping et€al. 1989) Because 11p15.5 contains several imprinted genes that are implicated in BWS, the precise gene that incites WT is not completely elucidated, though IGF2 has emerged as a leading candidate (Cooper et€al. 2005; Prawitt et€al. 2005; Alger et€al. 2007). IGF2 is expressed at increased levels in WT and in embryonic kidney compared with adjacent normal kidney or adult kidney, respectively (Scott et€ al. 1985; Reeve et€al. 1985). A common molecular defect in WT and BWS is loss of imprinting (LOI) of IGF2, which involves reactivation of a normally silent maternally-derived allele (Ogawa et€al. 1993). Increased IGF2 protein expression mediated through LOI is found concurrent with downregulation of H19, another gene at the 11p15 locus, via parental-origin specific DNA methylation of the H19 promoter (Steeman et€al. 1994; Moulton et€al. 1994). Hypermethylation of H19 blocks binding of the insulator protein CTCF, permitting IGF2 activation via an enhancer shared by IGF2 and H19 (Mummert et€al. 2005; Hancock et€al. 2007). The frequency of IGF2 LOI in sporadic WT is reported to be between 25% and 40%, indicating that defects in IGF2 signaling are not restricted to BWS-related WT (Ravenel et€al. 2001; Pritchard-Jones and Vujanic 2006). Another mechanism reported to increase the effective gene dosage or protein product of IGF2 is maternal allele specific loss of heterozygosity (LOH) at 11p15. LOH of 11p15, present in approximately 35% of WTs (Grundy et€al. 1996), is characterized by selective inactivation of the maternal allele and duplication of the paternal allele, yielding a relative increase in the IGF2 “oncogene” (Dome et€al. 2006a). Aberrant imprinting of the IGF2 receptor gene (IGFR2) also can increase the IGF2 dosage because the IGFR2 receptor targets IGF2 for degradation (Xu et€al. 1997). Decreased expression of paternal IGFR2 was noted in 7/16 (44%) of kidneys and associated WTs. IGF2 released by WT cells can send a growth-promoting autocrine signal via its receptor, IGFR1 (Zumkeller et€ al. 1993; Qing et€ al. 1996; Schmitt et€ al. 1997). Interestingly, the protein product of WT1, the first Wilms tumor suppressor gene, has been demonstrated to suppress IGF2 transcription, possibly important in negatively regulating blastemal cell proliferation at various developmental stages (Drummond et€al. 1992).
Molecular Targeted Therapy for Wilms’ Tumor
405
IGFR1, a transmembrane heterotetramer tyrosine kinase that is structurally and functionally related to the insulin receptor, transmits a growth-promoting signal via the PI3kinase/AKT/mTOR and RAS/RAF/MAPK pathways when bound to IGF2 (Reidemann and Macaulay 2006). Like IGF2, IGF1R has been found to be increased in WT at the transcript, as well as protein levels, (Gansler et€al. 1988) and it is similarly subject to transcriptional repression by WT1 (Werner et€ al. 1993). Recent investigation has documented gain of IGFR1 gene copy number and increased IGFR1 protein expression in WT, a finding associated with 15q gain and with relapse (Natrajan et€al. 2006b, 2007c). In summary, data supporting IGF2/IGFR1 as a key pathway to target in WT include its frequent upregulation via various genetic and epigenetic events, its known growth promoting effects, its ability to stimulate autocrine WT growth, and its association with relapse. IGF signaling also activates numerous downstream targets amenable to therapeutic targeting (Fig.╛1) and activates additional oncogenic pathways such as the b-catenin pathway, the latter via multiple signaling cascades including RAS/RAF activation and GSK-3b inhibition (Desbois-Mouthon et€al. 2001).
IGF and mTOR Signaling Pathways IGF1 IGFR
RAS
IGFR PI3K Akt
IRS1
Raf GSK-3b
Raptor 4EBP1
MEKK
mTOR
Foxo
eIF2B ERK
eIF4E
p70S6K
Protein Synthesis
Fig.€ 1╅ IGF and mTOR signaling pathways: upon binding its receptor IGFR1, IGF transmits a signal via IRS1 downstream through the RAS and AKT/mTOR pathways. IGFR and its downstream pathway components subject to pharmaceutical targeting are shaded
406
J.I. Geller and J.S. Dome
WNT-b-catenin Pathway The WNT-b-catenin signaling pathway plays an important role during embyrogenesis, tissue regeneration, and tumorigenesis. b-catenin-driven transcription regulates multiple cellular processes including survival/apoptosis, differentiation, proliferation, and motility. Overactive WNT signaling has been linked to both tumorigenesis and cancer progression in numerous cancer models, making this pathway a desirable target for molecular therapeutic intervention (Takahashi-Yanaga and Sasaguri 2007; Herbst and Koligs 2007). WNT-4 is known to autoinduce mesenchymal to epithelial transition of normal nephrons in utero. Dysregulation of this pathway has been postulated to be involved in WT transformation. Mutations in b-catenin, downstream of WNT-4, and other WNT receptors, have been found in approximately 15% of WTs, and such mutations lead to a loss of critical b-catenin phosphorlyation sites leading to overactive b-catenin signaling (Koesters et€ al. 1996). There exists a striking association between b-catenin mutation and WT1 mutation, whereby approximately 75% of WTs with WT1 mutations also harbor b-catenin mutations (Maiti et€al. 2000; Li et€al. 2004). The b-catenin pathway was further implicated in WT with the recent discovery of the WTX gene. Approximately 30% of WT have deletions or point mutations of WTX, which encodes a protein that forms a complex with b-catenin and other proteins, ultimately promoting ubiquitination and degradation of b-catenin, thereby attenuating TCF-mediated transcription (Major et€al. 2007). A defective WNT-b-catenin pathway (overactivated), due to either b-catenin or WTX mutation, is found in up to 45% of WTs (Major et€al. 2007; Nusse 2007) (Fig.╛2). Other proteins found to be dysregulated in WT have been linked to b-catenin signaling. APC protein, a binding partner that targets b-catenin for proteasomal degradation, has been found to be subject to LOH and microsatellite instability in 30% and 16% of WT tested, respectively (Ramburan et€al. 2005). These findings further widen the spectrum of WTs likely to be affected by b-catenin signaling. In WT, b-catenin expression has been demonstrated to concentrate in the membranous component (93 to 95% of WT), and less so in the cytoplasmic component (58% of WT), of epithelial, blastema, and stromal cells (Alami et€ al. 2003a; Ramburan et€al. 2005, 2006). In one large study, b-catenin and E-cadherin expression was noted in 133/140 (95%) and 75/140 (54%) of WTs tested, respectively. Given its high prevalence, b-catenin expression in WT, in general, lacks prognostic value (Ramburan et€al. 2006). However, WT with a loss of membranous staining and shift toward cytoplasmic staining, demonstrated a statistically significant shorter survival (Ramburan et€al. 2005). Nuclear accumulation is notably increased in metastatic WT (Alami et€ al. 2003a) and in WT with associated activating b-catenin mutations (Koesters et€ al. 2003). E-cadherin, a binding partner of b-catenin at the cellular membrane, was found to be increased in pretreated WT specimens (Ramburan et€ al. 2006) and is expressed at low levels in metastatic tumors with nuclear b-catenin localization (Alami et€al. 2003a). cDNA microarray studies of b-catenin-mutated WT versus nonmutated WT suggest that PITX2, APCDD1, and two endothelin-related proteins EDN3 and EDNRA are
Molecular Targeted Therapy for Wilms’ Tumor
407
Wnt /Beta-Catenin Signaling Wnt
E-Cadherin
Frizzled b-Catenin Cytoplasm
GSK3
Axin
b-Catenin
APC
WTX
Ub b-Catenin Nucleus Proteasomal Degradation b-Catenin CBP Survivin, c-Myc, Cyclin D1, p21, MMP-7, Axin-2, CD44, PPAR-g TCF-1, etc
LEF/TCF
Fig.€2╅ WNT/b-Catenin signaling: upon binding its receptor FRIZZLED, WNT transmits a signal via b-catenin that leads to b-catenin/TCF dependent transcription. Pathway components subject to genetic or epigenetic modulation in WT have hatched lines. DiSHEVELED and its downstream pathway components subject to potential pharmaceutical targeting are shaded
activated by b-catenin in WT, as are several likely compensatory upstream inhibitors of WNT signaling such as WIF1 and PRDC. Muscle-related genes are highly upregulated in b-catenin-mutated tumors, perhaps explaining the common myogenic changes found histologically in such tumors (Zirn et€al. 2006). There has been much attention by the pharmaceutical industry in efforts to find b-catenin pathway inhibitors (Barker and Clevers 2006; Takahashi-Yanaga and Sasaguri 2007; Herbst and Koligs 2007). Inhibitors have been developed that block WNT-b-catenin signaling at the level of the Disheveled (Dvl) protein, which transduces Wnt signals from the receptor Frizzled (Fz) (Shan et€al. 2005; DeAlmeida et€al. 2007), at the putative binding site for b-catenin and Tcf (Lepourcelet et€al. 2004), and at the site of cAMP response element-binding protein (CREB) binding protein (CBP) (Emami et€al. 2004; Eguchi et€al. 2005; Ma et€al. 2005). ICG-001,
408
J.I. Geller and J.S. Dome
which downregulates b-catenin/Tcf signaling via this latter pathway, has been shown to downregulate survivin expression and induce selective apoptosis in transformed but not normal colon cells, as well as demonstrate in€ vivo activity in the SW620 colon cancer xenograft model (Emami et€ al. 2004; Barker and Clevers 2006). Additional compounds proposed to be b-catenin signal inhibitors include quercitin (Park et€al. 2005), lithium, flavinoids, and curcumin (Takahashi-Yanaga and Sasaguri 2007; Ohori et€al. 2006). Networking between the WNT-b-catenin signaling pathway with the FGF, Notch, BMP, and Hedgehog signaling pathways is important for stem cell and progenitor cell homeostasis, disruption of which can result in cancer (Katoh 2007). b-catenin signaling pathway overlap has also been demonstrated for the EGF and PDGF pathways via EGF or PDGF-dependent b-catenin phosphorylation (Takahashi-Yanaga and Sasaguri 2007). b-catenin downstream effectors such as VEGF, Cyclin D1, survivin, and COX-2 are also potential targets for molecular inhibition (Herbst and Koligs 2007). Thus, further preclinical investigation of inhibiting multiple signaling pathways such as b-catenin in combination with EGFR, PDGR, IGF, or VEGF in WT is of interest. Similarly, simultaneous inhibition of both b-catenin transcription as well as downstream effectors such as VEGF, survivin, Cox-2, or Cyclin D1 (CDKs) may prove clinically meaningful.
EGFR Pathway Limited ex vivo data are available on EGF signaling in WT. Downregulation of EGFR from WT1-mediated transcriptional activity has been demonstrated in PC-12 neuronal cells as well as in osteosarcoma lines, with WT1-mediated apoptosis in the latter mitigated by constitutive EGFR expression, suggesting a protective role of EGFR in these models. However, EGFR signaling is normal in patients with Denys Drash Syndrome with germline WT1 mutations. (Liu et€ al. 2000; Englert et€al. 1995; Vicanek et€al. 1997) EGFR, and specifically HER2 expression, has been shown in 45% of blastema WT, 53% of WT with epithelial differentiation, and 14% of stromal predominant WT. (Ghanem et€al. 2001a; Salem et€al. 2006) Importantly, anti-EGFR monoclonal antibodies have shown effect in€vivo in WT xenograft models, (Pinthus et€al. 2004; Yokoi et€al. 2003) and in the COG Phase I study of the small molecule EGFR inhibitor gefitinib, two WT patients derived clinical benefit in the form of stable disease for 8 to ³60€weeks (Daw et€al. 2005).
Angiogenesis The role of angiogenesis, and specifically, the role of vascular endothelial growth factor (VEGF) has drawn much attention in the oncology arena, and VEGF inhibitors are now readily available for clinical use. VEGF is induced by many cellular stimuli including HIF-1a, and functions as an endothelial cell mitogen. VEGF and
Molecular Targeted Therapy for Wilms’ Tumor
409
HIF-1a have been shown to be co-expressed in all WTs, with one study demonstrating equivalent distribution of VEGF expression between the blastemal and epithelial components (Karth et€al. 2000). Elevated levels of the proangiogenic protein basic fibroblast growth factor (bFGF) in the urine of pediatric WT patients as well as elevated levels of VEGF in the sera of WT patients have each been shown to correlate with tumor stage, to decrease following surgery, and postoperative increasing levels suggest relapse or persistent disease (Lin et€ al. 1995; Blann et€ al. 2001; Skoldenberg et€al. 2001). Similarly, high WT microvessel density (MVD), a marker of tumor vascularity, has been shown to correlate with a worse prognosis (Ozluk et€al. 2006). WTs characterized to have a decreased WT1 exon 5+/− ratio are associated with downregulation of VEGF, demonstrating differential levels of VEGF expression based on WT1 isoform expression pattern (Baudry et€al. 2002). VEGF and VEGFR isoforms and WT subcomponent analysis have also demonstrated that VEGF (VEGF-A) and its receptor VEGFR-1 (Flt-1) are expressed in 52% and 47% of blastemal cells, respectively, and positive expressing cells were associated with higher MVD and risk for clinical progression (Ghanem et€ al. 2003). Similarly, VEGF-C expression is noted in 100% and 30% of WT blastema and stromal cells, respectively. VEGFR-2 is highly expressed in stroma and epithelial cells, but absent from WT blastema tested. While VEGF-C was shown to be associated with clinical progression by Nowicki et€al. (2007), all WTs studied were pretreated and multivariate analysis was not conducted to rule out independence of this purported prognostic factor from the confounding factor of posttreatment residual blastema. The finding that blastema lacks VEGFR-2 is of clinical importance as some antiVEGF targeted agents are VEGF receptor-specific.
Apoptotic Pathways Conflicting data have emerged regarding the relative prognostic value of various apoptosis activators or inhibitors in WT. In pediatric renal tumors, a high survivin:Fas ratio demonstrated a positive predictive value for tumor recurrence of 85.7% and negative predictive value of 71.4% (Takamizawa et€al. 2001). However, subsequent investigation in 92 primary WTs showed survivin mRNA increased in WT compared with expression levels in adjacent kidney, but neither survivin nor caspase 8, SMAC, Bid, and Fas (CD95) were independently, or in ratio, associated with stage or risk for tumor recurrence (Miller et€ al. 2005). Both antiapoptotic BCL-2 and pro-apoptotic BAK have been demonstrated to be under WT1-mediated transcriptional control, and therefore suggest a possible role in WT tumorigenesis (Fukuzawa et€al. 2004; Morrison et€al. 2005). Additionally, BCL-2 expression has been found in nephrogenic rests (Wunsch et€al. 2001), possibly providing protection from apoptosis during early tumor development. However, several investigations have not found any prognostic value of BCL family member expression, either antiapoptotic (BCL-2, BCL-X(L/S)) or proapoptotic (BAK), in WT (Re et€al. 1999; Tanaka et€al. 1999). One study demonstrated increased BCL-2 and decreased
410
J.I. Geller and J.S. Dome
BAX in WT blastema of increasing pathologic stage, with an increasing BCL-2/ BAX ratio indicative for tumor progression (Ghanem et€al. 2001b). Nonetheless, the compromised integrity of the apoptotic program via abnormal expression of proteins involved in the apoptotic pathway in WT is supported by activation of the anti-apoptotic effects of the IGF pathway, (Werner and Le Roith 1997; Kurmasheva and Houghton 2006; Reidemann and Macaulay 2006) and by TP53 mutation inducing attenuated apoptosis in anaplastic histology WT (Bardeesy et€al. 1995). Additionally, WT blastema commonly overexpresses the MYCN oncogene, known to have both proliferative and pro-apoptotic effects, the latter typically inhibited in cancer cells via perturbation of other cellular proteins required for apoptosis execution (Shaw et€al. 1988). Similar to MYCN amplified neuroblastoma and other tumor models, pro-apoptotic caspase-8 is silenced via promoter methylation in greater than 40% of WTs. (Morris et€al. 2003) WT also commonly express DR5, a receptor for both TNF-related apoptosis inducing ligand (TRAIL) and agonist DR5 monoclonal antibodies, each of which transmit a pro-apoptotic signal by directly activating the apoptotic pathway and each currently in early phase clinical development (Takamizawa et€al. 2000). A biomarker of active apoptosis, cytokeratin-18-related tissue polypeptide-specific (TPS) antigen, typically released by epithelial cells undergoing apoptosis, is elevated in nearly all epithelial and blastemal WT specimens tested, suggesting that it may be a reasonable biomarker of tumorspecific apoptosis, easily assessable in the serum of WT patients who undergo future therapies targeting the apoptotic program (Rebhandl et€al. 2001).
Cell Cycle WT1 has been shown to induce G1 phase arrest, possibly via induction of p21CIP1, abrogated by overexpression of cyclin/cdk complexes (Kudoh et€al. 1995; Englert et€al. 1997). Suppressed p16 expression either via promoter methylation or alternative genetic or epigenetic events has also been well documented in WT (ArcellanaPanlilio et€ al. 2000; Natrajan et€ al. 2007b). Cyclin D1 is a well established transcriptional endpoint of overactive b-catenin signaling. Thus, WT1 mutation, p16 silencing, or activated b-catenin signaling all could lead to uninhibited cyclin/ cdk activity, enabling rapid cell cycle transitioning and proliferation. A preliminary association between CDK4 overexpression and relapse in WT has been suggested (Faussillon et€al. 2005). These data support further exploration of cell cycle inhibition, perhaps CDK4 inhibition, in the treatment of WT.
HGF/c-MET Hepatocyte growth factor (HGF) and its receptor c-MET are involved in cell migration, differentiation, cell growth, angiogenesis, apoptosis inhibition, and increased cancer metastases (Christensen et€al. 2005; Sattler and Salgia 2007). HGF binding
Molecular Targeted Therapy for Wilms’ Tumor
411
to c-MET has also been shown to induce c-MET/b-catenin disassociation, leading to increased free b-catenin, analogous to the effects of EGF on E-cadherin/bcatenin. The complexity of such pathway cross-talk is exemplified by the findings that HGF treatment of cancer cells increases b-catenin/TCF transcription, b-catenin fosters HGF ligand-independent c-MET mediated cell scattering, and suppression of b-catenin inhibits the HGF-induced motile phenotype (Rasola et€al. 2007). HGF has been shown to be expressed in the serum of WT patients at levels three times higher than controls (Skoldenberg et€al. 2001). Additionally, HGF and c-MET colocalize in WT and are associated with an increased proliferative index (Alami et€ al. 2002). Further investigation of HGF/c-met signaling and inhibition is warranted in WT. A number of therapeutics, both small molecule inhibitors (ARQ-197, SU11271, SU11274, SU11606, SGX523, PHA665752, and PF2341066), or antibodies are currently in early clinical development in adults. Both humanized (AV299) and fully human antibodies (AMG102) against HGF, as well as MET decoys (CGEN214) that prevent ligand binding, or receptor dimerization, and onearmed monoclonal antibodies against the MET receptor are in development.
Other Pathways Heat-shock proteins (HSP) 27 and 70 have been found to be co-expressed with one another as well as with glutathione peroxidase in WT, with HSP70 expression confined to blastema and epithelial components (Stammler and Volm 1996; Efferth et€al. 2001a). High HSP70 expression, as well as HSP90a expression, was found in WT predominantly from children who survived (Efferth et€al. 2001a; Yang et€al. 2006). WT1 mediated inhibition of cellular proliferation has been shown to require association with HSP70 (Maheswaran et€al. 1998). From these data, it is not likely that HSP inhibition will yield an overall positive effect on WT patients and could, in fact, be utilized to discourage such clinical investigation. Similarly, c-KIT has not been found to be strongly nor diffusely overexpressed in WT, with one study showing 0/6 WTs expressing c-KIT, diminishing enthusiasm for the employment of single agent c-kit inhibitors for WT (Smithey et€al. 2002; Miliaras et€ al. 2004). TRK A and B receptors are found predominantly in WT stromal compartments, whereas TRK C receptors are found within WT epithelial structures (Donovan et€al. 1994). However, high levels of full-length TRKB transcripts in WT are associated with a worse outcome, suggesting a possible role for TRKB inhibition (Eggert et€al. 2001). Epigenetic methylation of various genetic loci in WT, potentially amenable to demethylation via various clinically available demethylating agents, notably includes the tumor suppressor caspase-8 (methylated in 43% of WT). However, demethylating agents could have potentially undesirable effects at other loci, including the locus for O-6 methylguanine DNA methyl transferase (MGMT, methylated in 30% of WT) (Morris et€al. 2003). Upregulation of MGMT expression may mediate cellular resistance to alkylator-based therapy. Interestingly, 60% of
412
J.I. Geller and J.S. Dome
WTs are characterized by general hypomethylation, particularly in centromeric satellite alpha DNA, possibly contributing to karyotypic instability and WT oncogenesis or tumor progression (Ehrlich et€al. 2002, 2003). The net effect of demethylating agents on WT awaits preclinical and clinical exploration. The multidrug-resistance protein (MRP) P-glycoprotein is expressed in the majority of WT at various levels, predominantly in blastemal and epithelial compartments. In WT, MRP levels have been associated with TP53, HSP70, and LRP/ MVP expression, but have not been associated with tumor grade, stage, or clinical course (Efferth et€al. 2001b; Camassei et€al. 2002).
Preclinical Identification of New Agents for Wilms Tumor Given the overall excellent prognosis for WT, few patients are candidates for phase I and II clinical trials. With a plethora of agents coming through the drug development pipeline, oncologists need a rational approach to prioritize agents or combinations of agents for clinical study. The identification of molecular pathways that are dysregulated in Wilms tumor is a starting point to identify drugs of interest. However, as illustrated in the sections above, these molecular pathways are complex and interactive, so it is difficult to predict the physiological effect of targeting a component of a particular pathway. To that end, investigators have developed preclinical WT models with which to test the efficacy of novel agents.
In Vitro Investigation of WT Attempts at establishing reliable WT cell lines have unfortunately met limited success. The T3/73 cell line was established from a WT resected from a 9€month old boy with aniridia. Initial reports demonstrated growth factor independent rapid growth of desmin positive fusiform cells, capable of undergoing over 100 passages (Kumar et€al. 1987). Subsequent investigation showed this cell line to be dependent on co-exposure of IGF2 for in€vitro proliferation (Zumkeller et€al. 1993). Similarly, WTCL-1 – WTCL-4 cells established using human WT were used to demonstrate that IGF2 is released by WT cells and that WT-secreted IGF2 sends a growth-promoting autocrine signal via IGFR1 (Qing et€al. 1996; Schmitt et€al. 1997). Brown et€al. (1989) established five cell lines produced from five Wilms’ tumors; however, all cultures were defined by a finite lifespan, no cells showed the expected characteristics of the putative WT stem cell, and all failed to show a “transformed” phenotype. Garvin et€ al. (1985, 1987) report specialized media and conditions that permit the growth of blastemal or skeletal components of WT in€vitro, though limited passage and restrictive conditions are somewhat prohibitive for general use. However, the blastemal WT cell line, W13, established in the Garvin laboratory,
Molecular Targeted Therapy for Wilms’ Tumor
413
was used to demonstrate the growth suppressive effects of suramin and ATRA in WT, possibly due to inhibition of IGF2 mediated crosslinking of IGFR1 and IGFR2, and by modulation of MYCN, IGF2, and IGF-binding proteins, respectively (Vincent et€ al. 1996a, b). The G401 and SKNEP-1 cell lines, utilized in numerous in€vitro and in€vivo published studies, have now been shown to represent rhabdoid tumor of the kidney and Ewing’s Sarcoma, respectively (Garvin et€ al. 1993; Smith et€al. 2006). In a similar manner, one must question the origin of the “sarcomatoid” WT model, WCCS-1 (Talts et€al. 1993). Most recently, investigators at The Hospital for Sick Children in Toronto have established the WiT 49 cell line from an anaplastic histology (AH) WT lung metastasis. The cell line harbors a TP53 mutation, co-expresses HGF and c-MET, lacks E-cadherin, expresses high levels of b-catenin co-immunolocalized to membranous c-MET, and expresses moderate levels of g-catenin and ezrin. Constitutive activation of MMP9 and latent MMP2 suggest active b-catenin-driven transcription (Alami et€al. 2003b). WiT 49 has subsequently been used by investigators to further demonstrate (1) podocalyxin (PODXL), an antagonist of cellular adhesion possibly via interaction with ezrin, is positively regulated by WT1 and transcriptionally repressed by TP53 in WT, (Stanhope-Baker et€al. 2004), (2) that siRNA-mediated downregulation of C/EBPB, a transcription factor involved in proliferation and differentiation and implicated in lymphoma tumorigenesis (Piva et€al. 2006) induces apoptosis in WT, (Li et€al. 2005) and (3) that STAT1 is a prosurvival factor in WT pathogenesis (Timofeeva et€al. 2006).
In Vivo Xenograft Testing in WT The dearth of Wilms tumor cell lines prompted the development of WT xenograft models, in which human Wilms tumors are transplanted into immunodeficient mice. Several groups have successfully employed WT xenografts to study novel agents. In 1989, investigators at the Medical University of South Carolina tested the effect of intraperitoneal injection of a human IGFR1 antibody (aIR-3) against nude mice bearing human WT heterotransplants. Models were assessed, indirectly, for IGFR1 function. Following 1€week of growth after subcutaneous inoculation, when greater than 2/3 of mice from each treatment cohort (7 to 10 mice) had developed measurable flank tumor growth, mice were treated with intraperitoneal injections of 0.5€ mg antibody three times weekly. Mice were sacrificed between 30 and 50€days. Overall, at 3€weeks into treatment, the tumor diameters of treated and control mice were significantly different, with approximately 80% growth inhibition. Mice crossed over from control to aIR-3 showed initial response, and mice crossed over from aIR-3 to control showed rapid growth of tumor (Gansler et€al. 1989). In this elaborate study, which employed methods that parallel current techniques used for xenograft testing, two FH WT and one AH WT showed statistically significant response to anti-IGFR1 antibody but not to control IgG, providing additional rationale for the testing of anti-IGFR1 therapy in WT patients.
414
J.I. Geller and J.S. Dome
Preliminary Anti-EGFR and Angiogenesis Testing Two groups of investigators have confirmed the in-vivo activity of anti-erbB2 monoclonal antibodies in human WT xenografts. Pinthus et€ al. (2004) demonstrated that 3 intraperitoneal injections of N29 anti-erb2 monoclonal antibody could prevent WT growth in€ vivo. ErbB2 expression was documented in 13/14 human WT (the one negative being AH WT). Yokoi et€al. (2003) showed that anti-her2/neu monoclonal antibody could suppress tumor growth in her2/neu(+) WT xenografts, but not in her2/neu(−) WT xenografts. Suppression of angiogenesis in the responding WT xenograft was mechanistically important. Decreased antiangiogenesis as a physiologically important mechanism of anti-WT activity was also demonstrated in two xenograft models of FH WT treated with the angiogenesis and collagen type I synthesis inhibitor halofuginone. Decreased angiogenesis leading to tumor growth inhibition was accompanied by a reduction of collagen synthesis, reduced HGF receptor (c-MET), and increased levels of WT1 (Pinthus et€al. 2005). Additional anti-angiogenic agents have been tested in xenograft models of SK-NEP-1. However, SK-NEP-1 has now been shown to be a Ewing’s sarcoma, not a WT (Huang et€al. 2004; Frischer et€al. 2004; Smith et€al. 2006).
Pediatric Preclinical Testing Program WT Xenograft Testing The majority of preclinical in€vivo drug testing in WT xenografts has come out of the lab of Peter Houghton at St. Jude Children’s Research Hospital, either via internal testing or via the National Cancer Institute-sponsored Pediatric Preclinical Testing Program (PPTP). Outside the auspices of the PPTP, the Houghton laboratory has published in€ vivo activity in WT with the histone deacetylase inhibitor depsipeptide and with the antimitotic agents ABT-751 and ixabepilone (Peterson et€al. 2005; Graham et€al. 2006; Morton et€al. 2007). Response criteria, in general, were complementary to that used formally by the PPTP (Houghton et€al. 2007a). ABT-751, a novel oral antimitotic agent that binds tubulin at the colchicine binding site, demonstrated objective tumor regression in 1/6 WT models (Morton et€ al. 2007). Depsipeptide induced tumor regression and tumor stabilization in 1 WT model each, accounting for 2/7 of the sensitive models out of 39 tested models (11 kidney) (Graham et€al. 2006). Ixabepilone induced WT regressions in 5/6 models at 10€mg/kg (mouse MTD) and in 1/5 models at 4.4€mg/kg, given every fourth day, demonstrating a steep dose response. Accompanying pharmacokinetic studies showed that a dose of 10€mg/kg provided an AUC/dose of 5.8€mmol/L- h, approximating that achieved in adult patients receiving 40€ mg/m2/dose (Peterson et€ al. 2005). Both ABT-751 and Ixabepilone are currently undergoing single agent Phase II study in WT through the COG. Two WT patients were treated on the COG Phase I trial of depsipeptide and like all other patients on study, neither responded (Fouladi et€al. 2006). Given preclinical data suggesting WT susceptibility to HDAC inhibition in€ vivo, it is worth considering more formal exploration of the more potent HDAC inhibitor, SAHA, in WT patients.
Molecular Targeted Therapy for Wilms’ Tumor
415
The PPTP utilizes four WT xenografts (and four rhabdoid tumor xenografts) in its panel, all which were derived from human WTs obtained at diagnosis. Two tumors (KT-10 and KT-11) are favorable histology, one tumor has diffusely anaplastic histology (KT-13) and one tumor is characterized as a favorable histology WT with nuclear unrest (KT-5). The efforts and methods of the PPTP are designed to perform xenograft experiments in as controlled a fashion as possible, and all models have undergone microarray characterization to supplement in€vivo activity with possible scientific insight. Thus far, the PPTP has investigated 13 agents for which data have been released. The in€ vivo activity of cisplatin, vincristine, and cyclophosphamide was broad, including significant tumor regressions in WT models (Houghton et€al. 2007a; Tajbakhsh et€al. 2007). Other agents tested showed limited activity and include bortezomib (proteasome inhibitor), (Houghton et€ al. 2008) dasatinib (src inhibitor), (Kolb et€ al. 2007) 17-DMAG (Hsp 90 inhibitor), (Smith et€ al. 2007) ispinesib (anti-mitotic), (Houghton et€ al. 2006) vorinostat (HDAC inhibitor), (Houghton et€ al. 2007b) rapamycin (mTOR inhibitor), (Houghton et€ al. 2007c) AZD2171 and sunitinib (anti-angiogenic agents targeting VEGF), (Maris et€ al. 2007; Houghton et€ al. 2007d) lapatinib (anti-EGFR/AKT), (Houghton et€al. 2007e) and SCH717454 (an anti-IGFR1 inhibitor) (Houghton et€al. 2007f). Unfortunately, tested agents thus far have not provided a large “response signal” in the PPTP WT models. Tumor growth delay in WT models was noted from anti-VEGF treatment, antimitotic agents, and HDAC inhibition. The above data provide the preclinical rationale for continued exploration of novel antimitotic agents in WT. There is also a suggestion that antiangiogenic agents, histone deacetylase inhibitors, and anti-EGFR agents may be worth clinical exploration in WT. Early data with IGFR1 inhibition in€ vivo combined with the molecular characterization of WT suggest that testing IGFR1 inhibitors, now in clinical development, is a worthwhile pursuit in WT patients. b-catenin pathway activation in a majority of WT samples strongly supports further investigation of this pathway as well. Additional ex vivo WT preclinical data suggest that demethylating agents, cell cycle modifiers (CDK 4 inhibitors), and apoptosis activating or modulating agents such as agonist DR5/TR2 or TRAIL analogs may have a role in the future treatment of WT patients. It is difficult to ignore the activity of mTOR inhibitors in other renal neoplasms, (Hudes et€al. 2007) and either single agent or combination testing in WT, either with conventional chemotherapeutic agents or with inhibitors of upstream targets such as IGFR1 is justifiable.
Future Directions and Challenges Barriers to new agent development in WT exist at both the preclinical and clinical level. A major challenge is the lack of reliable, validated WT cell culture systems readily available to interested investigators. Several laboratories are actively working on the establishment of such WT cell lines, as it would both greatly facilitate
416
J.I. Geller and J.S. Dome
more rapid signaling pathway interrogation as well as provide cost-effective drug activity screening. A second issue requiring refinement is the identification of more robust prognostic tools predictive of treatment-specific response and outcome. Current markers (LOH of 1p and 16q) while useful in general, are nonetheless, rudimentary (Grundy et al. 2005). In addition, in the era of molecular targeted therapy, the identification of biomarkers of disease responsiveness must also be addressed. The genetic and molecular heterogeneity of WT suggests that the application of novel molecular therapies will not be a “one size fits all” situation. The nonuniformity of IGFR1 expression, b-catenin activation, and heterogeneous makeup of additional signaling pathways of current and future interest presents the challenge of matching patient to drug. Thus, in this new era of targeted therapy, we not only continue to battle issues of under- or over-treatment due to poor markers of disease response to conventional therapy, but we must focus our awareness of individual patient tumor biology and consider this in molecular targeted therapy options. Therefore, inclusion of individual tumor biology characterization and pharmacodynamic endpoint monitoring, in conjunction with traditional clinical endpoints, becomes all the more critical. Remaining challenges include the solicitation of pharmaceutical industry support, acquisition of additional supportive preclinical data, establishment of clinical pathways for trial implementation, and patient enrollment. To achieve these ends, the COG RTC has developed tight collaboration with the PPTP and COG Developmental Therapeutics Committee (DVL) to channel both relevant WT models and targets of interest into the screening and testing pipeline provided by those two entities, respectively. The establishment of these collaborations holds much promise for more rapid transit of new agents into the clinical arena, with appropriate correlative investigation necessary to refine assessments of clinical utility based on specific tumor biology. It is noteworthy that the recently opened WT strata of ADVL0524 (Phase II study of Ixabepilone) completed accrual in under 1€ year, providing proof that “when built, the patients do come.” Via this pathway, the COG RTC has recently endorsed Phase I and II study of the monoclonal anti-IGFR1 antibody IMC-A12 (Imclone) as well as a Phase II study of the novel oral antimitotic, ABT-751 (Abbot Laboratories). Phase III investigation of new agents or new treatment strategies requires significant patient numbers, remedied by either prolongation of accrual time or by maximizing patient access. The latter has the advantage of facilitating more rapid progress, and can be achieved via international cooperation. Through improved model discovery and refinement of our understanding of relevant signaling pathways in WT, careful patient tumor biological profiling, design of trials with thorough pharmacodynamic and correlative studies, and international collaboration, molecular targeted agents hold promise of enabling maximization of cure with minimization of toxicity for children with Wilms’ tumor.
Molecular Targeted Therapy for Wilms’ Tumor
417
References Alami J, Williams BR, Yeger H (2003a) Differential expression of E-cadherin and beta catein in primary and metastatic Wilms tumours. Mol Pathol 56:218–225. Alami J, Williams Br, Yeger H (2003b) Derivation and characterization of a Wilms tumor cell line, WiT 49. Int J Cancer 107:365–374. Alami J, Williams BR, Yeger H (2002) Expression and localization of HGF and met in Wilms’ tumours. J Pathol 196:76–84. Alger EM, Heaps L, Darmanian A, Dagar V, Prawitt D, Peter GB, et€al. (2007) Paternally inherited submicroscopic dupliction at 11p15.5 implicated insulin-like growth factor II in overgrowth and Wilms tumorigenesis. Cancer Res 67:2360–2363. Arcellana-Panlilio MY, Egeler RM, Ujack E, Pinto A, Demetrick DJ, Robbins SM, et€al. (2000) Decreased expression of the INK4 family of cyclin-dependent kinase inhibitor in Wilms tumor. Genes Chromosomes Can 29:63–69. Bardeesy N, Beckwith B, Pelletier J ( 1995) Clonal expansion and attenuated apoptosis in Wilms tumors are associated with p53 gene mutations. Cancer Res 55:215–219. Barker N, Clevers H (2006) Mining the Wnt pathway for cancer therapeutics. Nat Rev Drug Discov 5:997–1014. Baudry D, Faussillin M, Cabanis MO, Rigolet M, Zucker JM, Patte C, et€al. (2002) Changes in WT1 splicing are associated with a specific gene expression profile in Wilms tumour. Oncogene 36:5566–5573. Baudry D, Hamelin M, Cabanis M, Fournet J, Tournade M, Sarnacki S, et€al.(2000) WT1 splicing alterations in Wilms tumors. Clin Can Res 6:3957–3965. Beckwith JB, Kiviat NB, Bonadio JF (1990) Nephrogenic rests, nephroblastomatosis, and the pathogenesis of Wilms’ tumor. Pediatr Pathol 10:1–36. Bernstein L, Linet M, Smith M, Olshan AF (1999) Surveillance epidemiology end results – renal tumors. Nat Can Inst 91:1382–1396. Blann AD, Li JL, Li C, Kumar S (2001) Increased serum VEGF in 13 children with Wilms’ tumour falls after surgery but rising levels predict poor prognosis. Cancer Lett 28:183–186. Brown KW, Shaw AP, Poirier V, Tyler SJ, Berry PJ, Mott MG, et€al. (1989) Loss of chromosome 11p alleles I cultured cell derived from Wilms tumors. Br J Can 60:25–29. Camassei FD, Arancia G, Cianfriglia M, Bosman C, Francalanci P, Rava L, et€ al. (2002) Nephroblastoma: multidrug-resistance P-glycoprotein expression in tumor cells and intratumoral capillary endothelial cells. Am J Clin Pathol 117:484–490. Christensen JG, Burrows J, Salgia R (2005) c-Met as a target for human cancer and characterization of inhibitors for therapeutic intervention. Cancer Lett 8:1–26. Cooper WN, Luharia A, Evans GA, Raza H, Haire AC, Grundy R, et€al. (2005) Molecular subtypes an dphenotypic expression of Beckwith-Wiedeman syndrome. Eur J Hum Genet 13:1025–1032. Daw NC, Furman WL, Stewart CF, Iacono LC, Krailo M, Bernstein ML, et€al. (2005) Children’s Oncology Group Study. J Clin Oncol 23:6172–6180. DeAlmeida VI, Miao L, Ernst JA, Koeppen H, Polakis P, Rubinfeld B (2007) The soluble wnt receptor Frizzled8CRD-hFc inhibits the growth of teratocarcinomas in€ vivo. Cancer Res 1:5371–5379. Desbois-Mouthon C, Cadoret A, Vlivet-Van Eggelpoel M-J, Bertrand F, Cherqui G, Perret C, et€al. (2001) Insulin and IGF-1 stimulate the b-catenin pathway through two signalling cascades involving GSK-3b inhibition and Ras activation. Oncogene 20:252–9. Diller L, Ghahremani M, Morgan J, Grundy P, Reeves C, Breslow N, et€al. (1998) Constitutional WT1 mutations in Wilms tumor patients. J Clin Oncol 16:3634–3640. Dome JS, Perlman EJ, Ritchey ML, et€al. (2006a) Renal Tumors. In Principles and Practice of Pediatric Oncology 5th Ed., Poplack DG and Pizzo PA (eds), p 904. Philadelphia: Lippincott Williams and Wilkins.
418
J.I. Geller and J.S. Dome
Dome J, Cotton C, Perlman E, Breslow N, Kalapurakal J, Ritchey M, et€al. (2006b) Treatment of anaplastic histology Wilms’ tumor: results from the fifth national Wilms’ tumor study. J Clin Oncol 24:2352–2358. Donovan MJ, Hempstead B, Huber LJ, Kaplan D, Tsoulfas P, Chao M, et€al. (1994) Identification of the neurotrophin receptors p75 and trk in a series of Wilms’ tumors. Am J Pathol 145:792–801. Drummond IA, Madden SL, Rohwer-Nutter P, Bell GI, Sukhatme VP, Rauscher FJ 3â•›rd (1992) Repression of the insulin-like growth factor II gene by the Wilms tumor suppressor WT1. Science 31:674–678. Efferth T, Schulten HG, Thelen P, Bode ME, Beniers AJ, Granzen B, et€al. (2001a) Differential expression of the heat shock protein 70 in the histological compartments of nephroblastomas. Anticancer Res 21:2915–2920. Efferth T, Thelen P, Schulten HG, Bode ME, Granzen B, Benier AJ, et€al. (2001b) Differential expression of the multirug resistance-related proteins MRP1 in the histological compartments of nephroblastomas. Int J Oncol 19:367–371. Eggert A, Grotzer MA, Ikegaki N, Zhao H, Cnaan A, Brodeur GM, et€al.(2001) Expression of the neurotrophin receptor TrkB is associated with unfavorable outcome in Wilms’ tumor. J Clin Oncol 1:689–696. Eguchi M, Nguyen C, Lee SC, Kahn M (2005) ICG-001, a novel small moleculre regulator of TCF/beta-catenin transcription. Med Chem 1:467–472. Ehrlich M, Hopkins NE, Jiang G, Dome JS, Yu MC, Woods CB, et€ al. (2003) Satellite DNA hypomethylation in karyotyped Wilms tumors. Can Genet Cytogenet 141:97–105. Ehrlich M, Jiang G, Fiala E, Dome JS, Yu MC, Long TI, et€ al. (2002) Hypomethylation and hypermethylation of DNA in Wilms tumors. Oncogene 26:6694–6702. Emami KH, Nguyen C, Ma H, Kim DH, Jeong KW, Eguchi M, et€al. (2004) A small molecule inhibitor of beta-catenin/CREB-binding protein transcription [corrected]. Proc Natl Acad Sci USA 101:12682–12687. Englert C, Hou X, Maheswaran S, Bennett P, Ngwu C, Re GG, et€al. (1995) WT1 suppresses synthesis of the epidermal growth factor receptor and induces apoptosis. EMBO J 14:4662–4675. Englert C, Maheswaran S, Garvin AJ, Kreidberg J, Haber DA (1997) Induction of p21 by the Wilms’ tumor suppressor gene WT1. Cancer Res 15:1429–1434. Faussillon M, Moonier L, Junien C, Jeanpierre C (2005) Frequent overexpression of cyclin D2/ cyclin-dependent kinase 4 in Wilms tumor. Cancer Lett 18:67–75. Fernandeaz CV, Lestou VS, Wildish J, Lee CI Soreansen PH (2001) Detection of a novel t(6;15) (q21;q21) in a pediatric Wilms tumor. Canr Genet Cytogenet 129:165–167. Fouladi M, Furman WL, Chin T, Freeman BB 3rd, Dudkin L, et€al. (2006) Phase I study of depsipeptide in pediatric patients with refractory solid tumors: a Children’s Oncology Group report. J Clin Oncol 1:3678–3685. Frischer JS, Huang J, Serur A, Kadenhe-Chiweshe A, McCrudden KW, O’Toole K, et€al. (2004) Effects of potent VEGF blockade on experimental Wilms tumor and its persisting vasculature. Int J Oncol 25:549–553. Fukuzawa R, Heathcott RW, Sano M, Morison IM, Yun K, Reeve AE (2004) Myogenesis in Wilms tumors is associated with mutations of the WT1 gene and activation of Bel-2 and the Wnt signaling pathway. Pediatr Dev Pathol 7:668–669. Gansler T, Allen KD, Burant CF, Inabnett T, Scott A, Buse MG, et€al. (1988) Detection of type 1insuline-like growth factor (IGF) receptors in Wilms tumors. Am J Pathol 130:431–435. Gansler T, Furlanetto R, Gramling TS, Robinson KA, Blocker N, Buse MG, et€al. (1989) Antibody to type I insulinlike growth factor receptor inhibits growth of Wilms’ tumor in culture and athymic mice. Am J of Path 135:961–966. Garvin AJ, Re GG, Tarnowski Bi, Hazen-Martin DJ, Sens DA (1993) The G401 cell line utilized for studies of chromosomal changes in Wilms tumor, is derived from a rhabdiod tumor of the kidney. Am J Pathol 142:375–380. Garvin AJ, Sullivan JL, Bennett DD, Stanley WS, Inabnett T, Sens DA (1987) The in€vitro growth, heterotransplantation and immonohistochemical characterization of the blastemal component of Wilms Tumor. Am J Pathol 129:353–363.
Molecular Targeted Therapy for Wilms’ Tumor
419
Garvin AJ, Surrette F, Hintz DS, Rudisill MT, Sens MA, Sens DA (1985) The in€vitro growth and charaterization of the skeletal muscle component of Wilms tumor. Am J Pathol 121:298–310. Gessler M, Konig A, Arden K, Grundy P, Orkin S, Sallan S, et€al. (1994) Infrequent mutation of the WTI gene in 77 Wilms Tumors. Hum Mutat 3:212–222. Ghanem MA, Van Der Kwast TH, Den Hollander JC, Sudaryo MK, Mathoera RB, Van den Heuvel MM, et€al. (2001a) Expression and prognostic value of epidermal growth factor receptor, transforming growth factor-alpha, and c-erb B-2 in nephroblastoma. Cancer 15;92(12):3120–3129. Ghanem MA, Van der Kwast TH, Den Hollander JC, Sudaryo MK, Van den Heuvel MM, Noordzij MA, et€al. (2001b) The prognostic significance of apoptosis-associated proteins BCL-2, BAX and BCL-X in clinical nephroblastoma. Br J Can 16:1557–1563. Ghanem MA, van Steenbrugge GJ, Sudary MK, Mathoera RB, Nijman JM, van der Kwast TH (2003) Expression and prognostic relevance of vascular endothelial growth factor (VEGF) and its receptor (FLT-1) in nephroblastoma. J Clin Pathol 56:107–113. Graham C, Tucker C, Creech J, Favours E, Billups CA, Liu T, et€ al. (2006) Evaluation of the antitumor efficacy, pharmacokinetics, and pharmacodynamics of the histone deacetylase inhibitor depsipeptide in childhood cancer models in€vivo. Clin Can Res 1:223–234. Green D (2004) The treatment of stages I-IV favorable histology Wilms tumor. J of Clin Oncol 22(8):1366–72. Grundy P, Telzerow P, Breslow N, Moksness J, Huff V, Paterson M, et€al. (1994) Loss of heterozygosity for chromosomes 16q and 1 p in Wilms tumors predicts an adverse outcome. Cancer Res 54:2331–2333. Grundy P, Telzerow P, Moksness J, Beslow NE (1996) Clinicopathologic correlates of loss of heterozygosity in Wilms tumor: a preliminary analysis. Med Pediatr Oncol 27:429–433. Grundy PE, Breslow NE, LI S, Perlman E, Beckwith JB, Ritchey ML, et€al. (2005) Loss of heterozygosity for chromosomes 1p and 16q is an adverse prognostic factor in favorable-histology Wilms tumor: a report from the National Wilms Tumor Study Group. J Clin Oncol 23:7312–7321. Hancock AL, Brown KW, Moorewood K, Moon H, Holmgren C, Mardikar SH, et€al. (2007) A CTCF-binding silencer regulates the imprinted genes AWT1 and exjobots seqiemtoa; epigenetic defects during Wilms Tumorigenesis. Hum Mol Genet 1:343–354. Herbst A, Koligs FT (2007) Wnt signaling as a therapeutic target for cancer. Meth in Mol Biol 361:63–91. Houghton PJ, Maris JM, Courtright J, Friedman HS, Keir ST, Lock RB, et€ al. (2007b) Initial Testing of the Histone Deacetylase Inhibitor Vorinostat by the Pediatric Preclinical Testing Program. Proceedings of the American Association for Cancer Research Annual Meeting 48:126, Apr 2007. (Abstract C226). Houghton PJ, Maris JM, Courtright J, Friedman HS, Keir ST, Lock RB, et€al. (2007e) Pediatric preclinical testing program (PPTP) evaluation of the EGFR and ErbB2 inhibitor Lapatinib. Proceedings of the American Association for Cancer Research Annual Meeting 48:126, Apr 2007. (Abstract B118). Houghton PJ, Maris JM, Friedman HS, Keir ST, Lock RB, Carol H, et€al. (2007d) Pediatric preclinical testing program (PPTP) evaluation of the multi-targeted kinase inhibitor Sunitinib. Proceedings of the American Association for Cancer Research Annual Meeting 48:126, Apr 2007. (Abstract 527). Houghton PJ, Maris JM, Friedman HS, Keir ST, Lock RB, Carol H, et€al. (2007f) Pediatric preclinical testing program (PPTP) evaluation of the fully human anti-IGF-1R Antibody SCH 717454. (Abstract A212). Houghton PJ, Maris JM, Friedman HS, Keir ST, Lock RB, Gorlick R, et€al. (2006) Pediatric preclinical testing program (PPTP) evaluation of the KSP inhibitor ispinesib (SB-715992). European Journal of Cancer Supplements 4(12):98. (Abstract 313). Houghton PJ, Morton CL, Kolb EA, Gorlick R, Lock R, Carol H, et€ al.(2007c) Initial testing (stage 1) of the mTOR inhibitor rapamycin by the pediatric preclinial testing program. Ped Blood Can [Epub ahead of print].
420
J.I. Geller and J.S. Dome
Houghton PJ, Morton CL, Kolb EA, Lock R, Carol H, Reynolds CP, et€al. (2008) Initial testing (stage 1) of the proteasome inhibitor bortezomib by the pediatric preclinical testing program. Ped Blood Can 50:37–45. Houghton PJ, Morton CL, Tucker C, Payne D, Favours E, Cole C, et€al. (2007a) The pediatric preclinical testing program: description of models and early testing results. Ped Blood Can 49:928–940. Huang J, Frischer JS, New T, Kim ES, Serur A, Lee A, et€al. (2004) TNP-470 promotes initial vascular sprouting in xenograft tumors. Mol Cancer Ther 3:335–343. Hudes G, Carducci M, Tomczak P, Dutcher J, Figlin R, Kapoor A, et€al. (2007) Temsirolimus, interferon alfa, or both for advanced renal-cell carcinoma. N Engl J Med 31:2271–81. Karth J, Ferrer FA, Perlman E, Hanrahan C, Simons JW, Gearhart JP, et€al. (2000) Coexpression of hypoxia-inducible factor 1-alpha and vascular endothelial growth factor in Wilms’ tumor. J Pediatr Surg 35:1749–1753. Katoh M (2007) Networking of WNT, FGF, Notch, BMP, and Hedgehog signaling pathways during carcinogenesis. Stem Cell Rev 3:30–38. Knudson AG, Strong LL (1972) Mutation and cancer: a model for Wilms’ tumor of the kidney. J Natl Can Inst 48: 313–234. Koesters R, Niggli F, von Knebel Doeberitz M, Stallmach T (2003) Nuclear accumulation of betacatenin protein in Wilms’ tumours. J Pathol 1:68–76. Koesters R, Ridder R, Kopp-Schneider A, Betts D, Adams V, Niggli F, et€al. (1996) Mutational activation of the B-Catenin proto-oncogene is a common event in the development of Wilms tumors. Cancer Res 59:3880–3882. Kolb EA, Gorlick R, Houghton PJ, Morton CL, Lock RB, Tajbakhsh M, et€al. (2007) Initial testing of dasatinib by the pediatric preclinical testing program. Ped Blood Can [Epub ahead of print]. Koufos A, Grundy P, Morgan K, Aleck KA, Hadro T, Lampkin BC, et€ al. (1989) Familial Weidemann-Beckwith syndrome and a second Wilms tumor locus both map to 11p15.5. Am J Hum Genet 44:711–719. Kudoh T., Ishidate T., Moriyama M., Toyoshima K., Akiyama T. (1995) G1 phase arrest induced by Wilms tumor protein WT1 is abrogated by cyclin/CDK complexes. Proc Natl Acad Sci USA 92:4517–4521. Kumar S, Harrison CJ, Heighway J, Marsden HB, West DC, Jones PM (1987) A cell line from Wilms tumor with deletion in short arm of chromosome II. Int J Cancer 15:499–504. Kurmasheva RT, Houghton PJ (2006) IGF-I mediated survival pathways in normal and malignant cells. Biochim Biophy Acta 1766:1–22. Lepourcelet M, Chen YN, France DS, Wang H, Crews P, Peterson F (2004) Small-molecule antagonists of the oncogenic Tcf/beta-catenin protein complex. Cancer Cell 5:91–102. Li C, Kim C, Margolin A, Guo M, Zhu J (2004) CTNNB1 Mutations and Overexpression of Wnt/B-Catenin Target Genes in WT1-Mutant Wilms Tumors. Am J Pathol 165: 1943–1953. Li W, Kessler P, Yeger H, Alami J, Reeve AE, Heathcott R, Skeen J, et€al. (2005) A gene expression signature for relapse of primary wilms tumors. Cancer Res 65:2592–2601. Lin RY, Argenta PA, Sullivan KM, Adzick NS (1995) Diagnostic and prognostic role of basic fibroblast growth factor in Wilms’ tumor patients. Clin Cancer Res 1:327–331. Liu XW, Gong LJ, Guo LY, Katagiri Y, Jiang H, Wang ZY, et€al.(2000) The Wilms’ tumor gene product WT1 mediates the down-regulation of the rat epidermal growth factor receptor by nerve growth factor in PC12 cells. J Biol Chem 16:5068–5073. Ma H, Nguyen C, Lee KS, Kahn M (2005) ICG-001, a novel small molecule regulator of TCF/ beta-catenin transcription. Med Chem 1(5):467–72. Maheswaran S, Englert C, Zheng G, Lee SB, Wong J, Harkin DP, et€al. (1998) Inhibition of cellular proliferation by the Wilms tumor suppressor WT1 requires association with the inducible chaperone Hsp70. Genes Dev 15:1108–1120. Maiti S, Alam R, Amos C, Huff V (2000) Frequent association of B-catenin and WT1 Mutations in Wilms tumors. Cancer Res 60:6288–6292.
Molecular Targeted Therapy for Wilms’ Tumor
421
Major MB, Camp ND, Berndt JD, Yi X, Goldenberg SJ, Hubbert C (2007) Wilms tumor suppressor WTX negatively regulates WNT/beta-catenin signaling. Science 316:1043–1046. Maris JM, Courtright J, Houghton PJ, Morton CL, Gorlick R, Kolb EA, et€al. (2007) Initial testing of the VEGFR inhibitor AZD2171 by the pediatric preclinical testing program. Ped Blood Can [Epub ahead of print]. McDonald JM, Douglass EC, Fisher R, Geiser Cf, Krill CE, Strong LC, et€al. (1998) Linkage of familial Wilms tumor predisposition to chromosome 19 and two-locus model for the etiology of familial tumors. Cancer Res 1:1387–1390. Metzger ML, Dome JS (2005) Current therapy for Wilms’ tumor. Oncologist 10:815–826. Miliaras D, Karasavvidou F, Papanikolaou A, Sioutopoulou D (2004) KIT expression in fetal, normal adult, and neoplastic renal tissues. J Clin Pathol 57:463–466. Miller MA, Karacay B, Breslow NE, Li S, O’Dorisio MS, Grundy PE, Sandler AD (2005) Prognostic value of quantifying apoptosis factor expression in favorable histology wilms tumors. J Pediatr Hematol 27:11–14. Morris MR, Hesson LB, Wagner KL, Morgan NV, Astuti D, Lees RD, et€ al. (2003) Muligene methylation analysis of Wilms tumour and adult renal cell carcinoma. Oncogene 22:6794–6801. Morrison DJ, English MA, Licht JD (2005) WT1 induces apoptosis through transcriptional regulation of the proapoptotic Bel-2 family member Bak. Cancer Res 15:8174–8182. Morton CL, Favours EG, Mercer KS, Boltz CR, Crumpton JC, Tucker C, et€al. (2007) Evaluation of ABT-751 against childhood cancer models in€vivo. Invest New Drugs 25:285–295. Moulton T, Crenshaw T, Hao Y, Moosikauwan J, Lin N, Demitzer F, et€ al. (1994) Epigenetic lesions at the H19 locus in Wilms tumour patients. Nat Genet 7:440–447. Mummert SK, Lobanenkov VA, Feinberg AP (2005) Association of chromosome arm 16q loss with loss of imprinting of insulin-like growth factor-II in Wilms tumor. Genes Chromosomes Can 43:155–161. Natrajan R, Little S, Reis-Filho J, Hing L, Messahel B, Grundy P, et€al. (2006a) Amplification and overexpression of cacna1e correlates with relapse in favorable histology Wilms tumors. Clin Can Res 12:7284–7293. Natrajan R, little SE, Sodha N, Reis-Filho JS, Mackay A, Fenwick K et€al. (2007c) Analysis by array CGH of genomic changes associated with the progression or relapse of Wilms tumor. J Pathol 21:52–59. Natrajan R, Reis-Filho JS, Little SE, Messahel B, Brundler MA, Dome JS, et€ al. (2006b) Blastemal expression of type I insulin-like growth factor receptor in Wilms’ tumors is driven by increased copy number and correlates with relapse. Cancer Res 66: 11148–55. Natrajan R, Warren W, Messahel B, Reis-Filho JS, Brundler MA, Dome JS, et€ al. (2007b) Complex patterns of chromosome 9 alterations including the p16INK4A locus in Wilms tumours. J Clin Pathol 10:1–20. Natrajan R, Williams RD, Grigoriadis A, Mackay A, Fenwick K, Ashworth A, et€ al. (2007a) Delineation of a 1€MB breakpoint region at 1p13 in Wilms tumor by fine –tilling oligonucleotide array CGH. Genes Chromosome Can 46:607–615. Nowicki M, Ostalska-Nowicka D, Kaczmarek M, Miskowiak B Witt M (2007) The significance of VEGF-C/VEGFR-2 interaction in the neovascularization and prognosis of nephroblastoma (Wilms’ tumour). Histopathology 50:358–364. Nusse R (2007) Converging on beta-catenin in Wilms tumor. Science 316:988–989. Ogawa O, Eccles MR, Szeto J, McNoe LA, Yun K, Maw MA, et€al. (1993) Relaxation of insulinlike growth factor II gene imprinting implicated in Wilms tumour. Nature 22: 749–751. Ohori H, Yamakoshi H, Tomizawa M, Shibuya M, Kakudo Y, Takahashi A, et€al. (2006) Synthesis and biological analysis of new curcumin analogues bearing an enhanced potential for the medicinal treatment of cancer. Mol Cancer Ther 5:2563–2571. Ozluk Y, Kilicaslan I, Gulluoglu MG, Ayan I, Uysal V (2006) The prognostic significance of angiogenesis and the effect of vascular endothelial growth factor on angiogenic process in Wilms’tumour. Pathology 38:408–414.
422
J.I. Geller and J.S. Dome
Park CH, Chang JY, Hahm ER, Park S, Kim HK, Yang CH (2005) Quercetin, a potent inhibitor against beta-catenin/Tcf signaling in SW480 colon cancer cells. Biochem Biophys Res Commun 328(1):227–34. Pastore G, Znaor A, Spreafico F, Graf N, Pritchard-Jones K, Steliarova-Foucher E (2006) malignant renal tumors incidence and survival in European children (1978–1997): Report from the Automated Childhood Cancer Information System Project. European Journal of Cancer 42: 2103–2114. Perotti D, De Vecchi G, Testi MA, Ludldi E, Modena P, Mondini P, et€al. (2004) Germline mutations of the POU6F2 gene in Wilms tumors with loss of heterozygosity on chromosome 7p14. Hum Mutat 24:400–407. Peterson JK, Tucker C, Favours E, Cheshire PJ, Creech J, Billups CA, et€al. (2005) In vivo evaluation of exabepilone (BMS247550), a novel epothilone B derivative, against pediatric cancer models. Clin Cancer Res 1:6950–6958. Ping AJ, Reeve AE, Law DJ, Young MR, Boehnke M, Feinberg AP (1989) Genetic linkage of Beckwith-Wiedemann syndrome to 11p15. Am J Hum Genet 44(5):720–723. Pinthus JH, Fridman E, Dekel B, Goldberg I, Kaufman-Francis K, Eshhar Z, et€al. (2004) ErbB2 is a tumor associated antigen and a suitable therapeutic target in Wilms tumor. J Urol 172:1644–1548. Pinthus JH, Sheffer Y, Nagler A, Fridman E, Mor Y, Genina O, et€al. (2005) Inhibition of Wilms tumor xenograft progression by halofuginone is accompanied by activation of WT-1 gene expression. J Urol 174:1527–1531. Piva R, Pellergrion E, Mattioli M, Agnelli L, Lombard L, Boccalatte F, et€al. (2006) Functional validation fo the ana plastic lymphoma kinase signature identifies CEBPB and BCL2A1 as critical target genes. J Clin Invest 6:3171–3182. Prawitt D, Enklaar T, Gartner-Rupprecht B, Spangengerg C, Oswald M, Lausch E, et€al. (2005) Microdeletion of target sites for insulator protein CTCF in a chromosome 11p15 imprinting center in Beckwith-Wiedemann syndrome and Wilms’ tumor. Proc Natl Acad Sci15: 4085–4090. Pritchard-Jones K, Vujanic G (2006) Multiple Pathways to Wilms Tumor: How Much is Genetic? Ped Blood and Can 47:232–234. Qing RQ, Schmitt S, Ruelicke T, Stallmach T, Schooenle EJ (1996) Autocrine regulation of growth by insulin-like growth factor (IFG)-II mediated by type I IGF-receptor in Wilms tumor cells. Pediatr Res 39(1):160–5. Rahman N, Abidi F, Ford D, Arbour L, Rapley E, Tonin P, Barton D, et€al. (1998) Confirmation of FWT1 as a Wilms tumour susceptibility gene and phenotypic characteristic of Wilms tumour attributable to FWT1. Hum Genet 103:547–556. Rahman N, Arbour L, Tonic P, Renshaw J, Pelletier J, Baruchel S, Pritchard-Jones, et€al. (1996) Evidence for a familial wills tumour gene (FWT1) on chromosome 17q12-q21. Nat Genet 13:461–463. Ramburan A, Chetty R, Hadley G P, Naidoo R, Govender D ( 2004) Microsatellite analysis of the DCC gene in nephroblastomas: pathologic correlations and prognostic implications. Mod Pathol 17: 89–95. Ramburan A, Hadley GP, Govender D (2006) Expresson of E-cadherin, cadherin-11, alpha-, betaand gamma-catenins in nephroblastomas: relatonship with clinicopathological parameters, prognostic factors and outcome. Pathology 38(1):39–44. Ramburan A, Oladiran F, Smith C, Hadley GP, Goverder D (2005) Microsatellite analysis of the adenomatous polyposis coli (APC) gene and imunoexpression of beta catenin in nephroblastoma: a study including 83 cases treated with preoperative chemotherapy. J Clin Pathol 58:44–50. Rasola A, Fassetta M, De Bacco F, D’Alessandro L, Gramaglia D, Di Renzo MF, et€al. (2007) A positive feedback loop between hepatocyte growth factor receptor and beta-catenin sustains colorectal cancer cell invasive growth. Oncogene 15:1078–1087. Ravenel J, Broman K, Perlman E, Niemitz E, Jayawardena T, Bell D, et€ al. (2001) Loss of imprintng of insulin-like growth factor-II (IGF2) Gene in distinguishing specific biologic subtypes of Wilms tumor. J Nat Can Inst 93:1698–1703.
Molecular Targeted Therapy for Wilms’ Tumor
423
Ravenel J, Perlman E, Broman K, Feinberg A (2002) Re: loss of imprinting of insulin-like growth factor-II (IGF2) gene in distinguishing specific biologic subtypes of Wilms tumor. J Nat Canr Inst 94:1809–1810. Re GG, Hazen-Martin DJ, El Bahtimi R, Brownlee NA, Willingham MC, Garvin AJ (1999) Prognostic significance of Bcl-X(L) in rare tumor cases. Int J Cancer 20:192–200. Rebhandl W, Handisurya A, Memaran N, Felberbauer FX, Aberle J, Paya K, et€ al. (2001) Expression of cytokeratin-18-related tissue polypeptide-specific (TPS) antigen in Wilms tumor. Med Pediatr Oncol 37:357–364. Reeve AE, Eccles MR, Wilkins RJ Bell GI, Millow LJ (1985) Expression of insulin-like growth factor-II transcriptions in Wilms tumour. Nature 19:258–260. Reidemann J, Macaulay VM (2006) IGF1R signaling and its inhibition. Endocrin-Related Cancer 13:33–43. Rivera M, Kim WJ, Wells J, Driscoll DR, Brannigan BW, Han M, et€al. (2007) An X chromosome gene, WTX is commonly inactivated in Wilms tumor. Science 315:642–645. Safford SD, Freemerman AJ, Langdon S, Bentley R, Goyeau D, Grundy PE (2005) Decreased E-cadherin expression correlates with higher stage of Wilms’ tumors. J Pediatr Surg 40: 341–348. Salem M, Kinoshita Y, Tajiri T, Souzaki R, Tatsuta K, Higashi M, et€ al.(2006) Association between the HER2 expression and histological differentiation in Wilms tumor. Pediatr Surg Int (11):891–896. Samani AA, Yakar S, LeRoith D, Brodt P (2007) The role of the IGF system in cancer growth and metastasis: overview and recent insights. Endocr Rev 28:20–47. Sattler M, Salgia R (2007) c-Met and hepatocyte growth factor: potential as novel targets in cancer therapy. Curr Oncol Rep 9:102–108. Schmitt S, Ren-Qiu Q, Torresani T, Doebeli M, Zapf J, Schoenle J (1997) High molecular weight forms of IGF-II (big-IGF-II) released by Wilms tumor cells. Eur J of Endo 137:396–401. Schulz S,. Becker KF, Braungart E, Reichmuth C, Klamt B, Becker I, et€ al.(2000) Molecular analysis of E-cadherin and cadherin-11 in Wilms tumour. J Pathol 191:162–169. Scott J, Cowell J, K Robertson ME, Priestley LM, Wadey R, Hopkins B, et€al.(1985) Insulin-like growth factor-II gene expression in Wilms tumour and embryonic tissues. Nature 19:260–262. Shan J, Shi DL, Wang J, Zheng J (2005) Identification of a specific inhibitor of the dishevelled PDZ domain. Biochemistry 29:15495–15503. Shaw AP, Poirier V, Tyler S, Mott M, Berry Y, Maitland NJ (1988) Expression of the N-myc oncogene in Wilms’ tumour and related tissues. Oncogene 3(2):143–9. Singh KP, Roy D (2006) SKCG-1: a new candidate growth regulatory gene at chromosome 11q23.2 in human sporadic Wilms tumours. Br J Can 94:1524–1532. Skoldenberg EG, Christiansson J, Sandstedt B, Larsson A, Lackgren G, Christofferson R (2001) Angiogenesis and angiogenic growth factors in Wilms tumor. J Urol 165:2274–2279. Smith MA, Maris JM, Keir ST, et€al. (2007) Pediatric preclinical testing program (PPTP) efficacy and pharmacodynamic evaluation of the Hsp90 inhibitor 17-DMAG. J Clin Oncol 25 (Abstract 3575). Smith MA, Morton CL, Phelps D, Girtman K, Neale G, Houghton PJ. (2006) SK-NEP-1 and Rh1 are Ewing family tumor lines. Ped Blood Can [Epub ahead of print]. Smithey BE, Pappo AS, Hill DA (2002) C-kit expression in pediatric solid tumors: a comparative immunohistochemical study. Am J Surg Pathol 26:486–492. Stammler G, Volm M (1996) Expression of heat shock proteins, glutatahione peroxidase and catalase in childhood acute lymphoblastic leukemia and nephroblastoma. Cancer Lett 19:35–42. Stanhope-Baker P, Kessler PM, LI W, Agarwal ML, Williams BR (2004) The Wilms tumor suppressor-1 target gene podacalyxin is transcriptionally repressed by p53. J Biol Chem 6:33357–33385. Steeman M, Rainier S, Dobry CJ, Grundy P, Horon I, Feingerg AP (1994) Loss of imprinting of IGF2 is linked to reduced expression and abnormal methylation of H19 in Wilms tumour. Nat Genet 7:433–439. Tajbakhsh M, Houghton PJ, Morton CL, Kolb EA, Gorlick R, Maris JM, et€al. (2007) Initial testing of cisplatin by the pediatric preclinical testing program. Ped Blood Can [Epub ahead of print].
424
J.I. Geller and J.S. Dome
Takahashi-Yanaga F, Sasaguri T (2007) The Wnt/B-catein signaling pathway as a target in drug discovery. J Pharmacol 104:293–302. Takamizawa S, Okamoto S, Bishop W, Wen J, Kimura K, Sandler A (2000) Differential apoptosis gene expression in pediatric tumors of the kidney. J Pediatr Surg 35:390–395. Takamizawa S, Scott D, Wen J, Grundy P, Bishop W, Kimura K, et€al.(2001) The survivin:fas ratio in pediatric renal tumors. J Pediatr Surg 36:37–42. Talts JF, Aufderheide E, Sorokin L, Ocklind G, Mattson R, Ekblom P (1993) Induction of mouse tenascin expression by a human sarcomatiod Wilms tumor cell line growing in nude mice. Int J Cancer 54:868–874. Tanaka K, Granata C, Wang Y, O’Briain DS, Puri P (1999) Apoptosis and bcl-2 oncogene expression in Wilms’ tumor. Pediatr Surg Int 15:243–247. Timofeeva QA, Plisov S, Evseev AA, Peng S, Jose-Kampfner M, Lovvorn HN, et€ al. (2006) Serine-phosphorylated STAT1 is a prosurvival factor in Wilms tumor pathogenesis. Oncogene 25:7555–7564. Vicanek C, Ferretti E, Goodyer C, Torban E, Moffett P, Pelletier J, et€al. (1997) Regulation of renal EGF receptor expression is normal in Denys-Drash syndrome. Kidney Int 52(3):614–619. Vincent TS, Hazen-Martin DJ, Garvin AJ (1996a) Inhibition of insulin like growth factor II autocrin growth of Wilms tumor by suramin in€vitro and vivo. Cancer Lett 15:49–56. Vincent TS, Re GG, Hazen-Martin DJ, Tarnowsk BI, Willingham MC, Garvin AJ (1996b) Alltrans-retinoic acid-induced growth Suppression of blastemal Wilms’ tumor. Pediatr Pathol Lab Med 16:777–789. Werner H, Gian G Re, Iain A, Drummond, Vikas P, Sukhatme, Frank J, et€ al.(1993) Increase expression of the insulin-like growth factor I receptor gene, IGF1R, in Wilms tumor is correlated with modulation of IGF1R promoter activity by the WT1 Wilms tumor gene product. Proc Nat Acad Sci 90:5828–5832. Werner H, Le Roith D (1997) The insulin-like growth factor-I receptor signaling pathways are important for tumorigenesis and inhibition of apoptosis. Crit Rev Oncol 8:71–92. Wunsch L, Flemming P, Gluer S (2001) Expression of MIB and BCL-2 in patients with nephrogenic rest with and without associated Wilms tumors. Eur J Ped Surg 11:105–109. Xu YQ, Grundy P, Polychronakos C (1997) Aberrant imprinting of the insulin-like growth factor II receptor gene in Wilms’ tumor. Oncogene 14(9):1041–6. Yang Y, Niu ZB, Hou Y, Wang CL (2006) The expression of HSP70 and HSP90alpha in children with Wilms tumor. J Pediatr Surg 41:1062–1066. Yokoi A, McCrudden KW, Huang J, Kim ES, Soffer SZ, Frischer JS, et€al.(2003) Human epidermal growth factor receptor signaling contributes to tumor growth via angiogenesis in her2/ neu-expressing experimental Wilms’ tumor. J Pediatr Surg 38 1569–1573. Zhao J, Yart A, Frigerio S, Perren A, Schraml P, Weisstanner C, et€al. (2007) Sporadic human renal tumors display frequent allelic imbalances and novel mutations of the HRPT2 gene. Oncogene 26:334–3449. Zirn B, Samans B, Wittmann S, Pietsch T, Leuschner I, Graf N, et€al. (2006) Target genes of the WNT/beta-catenin pathway in Wilms tumors. Genes Chromosomes Can 45:565–574. Zumkeller W, Schwander J, Mitchell CD, Morrell DJ, Scholfield PN, Preece MA (1993) Insulinlike growth factor (IGF)-I, II and IGF binding protein-2 (IGFBP-2) in the plasma of children with Wilms tumor. Eur J Cancer 14:1973–1977.
Molecular Therapy for Rhabdomyosarcoma Raushan T. Kurmasheva, Hajime Hosoi, Ken Kikuchi, and Peter J. Houghton
Introduction Rhabdomyosarcoma (RMS) is the most common sarcoma of childhood representing about 23% of all sarcomas, and approximately 7% of all pediatric malignancies (Arndt and Crist 1999). Histologically, RMS presents as three major variants, embryonal (ERMS), representing about 60%, botryoid that are usually combined with embryonal tumors, and alveolar RMS (ARMS) that comprise 30% of RMS. The pleiomorphic variant is an unrelated high-grade sarcoma with various degrees of muscle differentiation and is rarely (if ever) diagnosed in children. Embryonal RMS is not associated with any chromosomal translocation, but loss of heterozygosity at 11p15.5 is a common feature (Scrable et€ al. 1987; Scrable et€ al. 1989) associated with loss of imprinting of the IGF-2 locus (Anderson et€ al. 1999). Alveolar RMS is characterized by specific translocations. Approximately 60 to 70% of histologically diagnosed ARMSs involve translocations of t(2;13)(q35;q14) leading to PAX3-FKHR gene fusion (Galili et€ al. 1993), whereas 10% have the t(1;13)(q36;q14) translocation that encodes the PAX7-FKHR fusion (Davis et€ al. 1994). Both translocations generate in-frame fusion between the PAX gene DNA binding domain and the transactivation domain of FKHR. Interestingly, both ERMS and translocation-positive ARMS may have loss of imprinting at 11p15.5 (Anderson et€al. 1999), suggesting that dysregulation of IGF2 may be common to both histologies. Recent studies using expression profiling suggest that histological ARMS that are translocation negative cluster with ERMS, hence the molecular classification of these tumors may differ from the histopathologic classification. Further, a small signature comprising as few as ten genes differentiates translocation-positive from translocation-negative RMS (Lae et€al. 2007). Thus, the understanding of genetic events characteristic of RMS is at a point where these may be applied to differential diagnosis, and potentially to novel treatment strategies. It is well established that P.J. Houghtonâ•›(*) Department of Molecular Pharmacology, St. Jude Children’s Research Hospital, 332 N. Lauderdale Street, Memphis, TN 38105-2794, USA e-mail:
[email protected]
P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_20, © Springer Science+Business Media, LLC 2010
425
426
R.T. Kurmasheva et al.
ARMSs are more aggressive than ERMS, and have a poorer prognosis (Qualman and Morotti 2002; Breneman et€al. 2003). Further, the prognosis for patients with metastatic disease at diagnosis is significantly worse for ARMS having the t(2;13) (q35;q14) translocation compared to those having the t(1;13)(q36;q14) variant (Sorensen et€ al. 2002). Thus, the genetic alterations impact on chemo- or radiosensitivity. However, it is also clear that any advanced stage RMS still presents a clinical challenge. Essentially, the cure rate for metastatic disease has not changed significantly in 30€years, despite intensification of cytotoxic therapy and introduction of novel cytotoxic agents (Pappo et€al. 2007). Thus, while introduction of novel cytotoxic agents, such as the camptothecins that target topoisomerase 1 (Furman et€al. 1999; Pappo et€al. 2007), may ultimately improve outcome, it is probable that it will be at the expense of additional toxicity or necessitate reduced dose intensity of other agents used in the treatment of RMS. Alternative approaches that exploit the molecular characteristics of RMS conceptually appear to offer potential benefit with lower toxicity and with reduced sequellae. The obvious example is that of imatinib mesylate in the treatment of chronic myelogenous leukemia (Druker 2003; Druker et€al. 2006) or gastrointestinal stromal tumors (Rubin et€al. 2007; Siehl and Thiel 2007). However, the complexities of RMS biology suggest that a single “genetic driver” is unlikely, and that combinations of agents that target different molecular abnormalities will be necessary to eradicate these tumors using these rational approaches. Indeed, the development of therapeutic strategies that exploit synthetic lethal interactions that are consequential to the molecular aberrations in these tumors will be the major challenge in developing curative approaches to childhood RMS. Here, we review the reported molecular characteristics of RMS as they relate to the development of molecularly targeted therapy.
What Is the Cell of Origin? It has been proposed that tumor growth may be sustained by a rare population of cells, termed cancer stem cells (Hope et€al. 2003, 2004). These cells, analogous to stem cells in normal self-renewing tissues, have the ability to self-perpetuate, and it has been proposed that they may give rise to recurrent tumor following surgery or chemo-radiation therapy. It is also postulated that failure of conventional treatments is because the stem cell is resistant, and that therapeutics should be targeted to this population. For example, in AML only a small proportion of cells (approximately 1â•›×â•›10−6) in human samples resulted in tumor growth when inoculated immune deficient mice. However, these studies with human AML cells have relied extensively on xenotransplantation techniques where human cancer cells have been inoculated into irradiated nonobese severe combined immunodeficient mice. In contrast, transplantation of syngeneic Em-myc or Em-N-RAS lymphomas, or PU.1−/− AML indicate that inoculation of as few as 10 cells results in lethal disease (Kelly et€al. 2007). Thus, while the cancer stem cell hypothesis is of interest, it will be critical to demonstrate its role in RMS. Currently, there is considerable focus on
Molecular Therapy for Rhabdomyosarcoma
427
identifying the putative RMS stem cell. Although the histogenesis of RMS is not certain, skeletal muscle characteristics of these tumors sugges t that they originate from fetal rhabdomyoblasts that fail to undergo normal differentiation. In fusion negative RMS, PAX7 is upregulated and may contribute to tumorigenesis in the absence of chimeric PAX transcription factors. In contrast, PAX3 expression is independent of translocations (Tiffin et€ al. 2003). In normal myogenesis, expression of PAX7 is concurrent with PAX3 in the dermomyotome. Expression of PAX3 suppresses myogenic differentiation in migrating myoblasts, whereas PAX7 expression is downregulated in these cells. Thus, high expression of PAX7 but not PAX3 in translocation deficient RMS suggests that these tumors are not derived from proliferating cells in the dermomyotome. PAX7 expression is essential (Asakura et€al. 2002) for specification and maintenance of myogenic satellite cells (Seale and Rudnicki 2000; Seale et€al. 2000), and as myogenic satellite cells are known to express PAX7 (Asakura et€al. 2002) it has been proposed that translocation-negative RMS may be derived from the myogenic satellite cell lineage (Tiffin et€al. 2003). This is consistent with the finding of the hepatocyte growth factor (HGF) receptor, MET, expression in all translocation-negative RMS as MET is expressed in quiescent satellite cells (Cornelison and Wold 1997). That the type of tumor induced by altered pathways in RMS may be dependent on the stage of development of the cell of origin is demonstrated by the studies of Linardic et€al. (2005). Less differentiated human skeletal muscle precursors gave rise to tumors with variable histologies when stably expressing T/t-Ag, hTERT and H-RAS, whereas committed human skeletal muscle myoblasts formed RMS-like tumors of embryonal morphology.
Whole Genome Expression Profiling As discussed briefly above, childhood RMS can be classified histologically as ERMS, ARMS or botryoid. More recently, there have been a series of reports using genome-wide expression that have sought to identify expression signatures that distinguish ERMS from ARMS, and potentially identify drug targets (Wachtel et€al. 2004; Davicioni et€al. 2006; De Pitta et€al. 2006; Goldstein et€al. 2006; Romualdi et€al. 2006; Lae et€al. 2007). Despite differences in approach (for example different platforms for assessing expression), a gene set found to be differentially expressed between fusion-positive and fusion-negative tumors has been identified. The overlap between the top 50 genes that distinguish between ARMS and ERMS in three independent studies (Wachtel et€al. 2004; Davicioni et€al. 2006; Lae et€al. 2007) identified twelve genes that were common. The ranked list overlap analysis from these three studies is shown in Fig.€1. Interestingly, the canabinoid receptor (CNR1) is highly expressed in ARMS compared to ERMS, and may be a potential therapeutic target in several human cancers (Flygare et€ al. 2005; McAllister et€ al. 2005; Sarfaraz et€ al. 2005, 2006). A small molecule agonist (WIN-55,212-2) of the canabinoid receptor has been shown to induce apoptosis. An alternative approach may be to develop antibodies that act as agonists, or that selectively deliver a cytotoxic
428
R.T. Kurmasheva et al. Watchtel et al. (204)
Lae et al. (115) OLIG2 KCNN3
NELL1 NHLH1 ADRA2C ALK NRN1 KIAA0746 C1orf45
ELA1
PIPOX CNR1 TFAP2B ELA1 ASS PODXL FOXF1 PGBD5 KIAA0523 ABAT DKFZp547PO42 DKZp762M127 GARNL4 TNRC9 JAKMIP2 DCX TSC22D2 SYN2
Davicioni et al. (38)
Fig.€1╅ Ranked list overlap analysis. The overlap between the top 50 genes (probe sets) distinguishing between ARMS and ERMS in three independent studies. Lae et€al. (2007), Wachtel et€al. (2004) and Davicioni et€al. (2006) study is shown graphically. (From reference Lae et€al. 2007 with permission)
agent to tumor. The largest expression profiling study (Davicioni et€ al. 2006) reported analysis of 139 RMS, and revealed a distinct common expression profile for PAX3-FKHR expressing tumors compared to those that were fusion negative, Fig.╛2. Thus, by this analysis there is a common profile irrespective of the fusion (PAX3 or PAX7), and this differs from fusion-negative tumors, Fig.€ 2. In another cytogenetic and molecular analysis of primary tumors AURKA, the gene encoding aurora kinase A required for chromatid axial shortening (Mora-Bermudez et€ al. 2007), was found to be overexpressed in all RMS studied. Whether or not overexpression signifies this mitotic kinase as a tumor-selective drug target needs to be established. Several aurora kinase A inhibitors are in clinical development (Hoar et€al. 2007; Manfredi et€al. 2007), and are currently being evaluated against childhood cancer preclinical models.
Chimeric Transcription Factors and Oncogenes PAX3-FKHR While expression profiling has been of some utility in identifying potential therapeutic targets, the development of novel anticancer agents to treat patients with tumors such as ARMS needs elucidation of the pathogenesis based on the abnormal gene and protein expression and aberrant signaling pathways in the tumor. Various
Molecular Therapy for Rhabdomyosarcoma
429
b
a Y
Y
Alveolar Mixed Alveolar/Embryonal Embryonal Spindle/Botryoid
Z
X
Z
X
Fig.€2╅ PAX-FKHR expression in ARMS is associated with a unique gene expression profile that is independent of tumor histology. (a) Multidimensional scaling analysis of 139 primary rhabdomyosarcoma tumors based on semisupervised analysis reveals tight clustering of most alveolar tumors, in contrast to the heterogeneous distribution of embryonal and spindle/botryoid tumors included in the main alveolar cluster are three mixed histology alveolar/embryonal tumors.The legend indicates the histologic diagnoses. (b) Replot of (a) based on normalized QRT-PCR expression levels of PAX3-FKHR and PAX7-FKHR for 59 tumors with alveolar or mixed alveolar/ embryonal histology. Relative PAX-FKHR mRNA levels are depicted by colored dots as indicated in the color scale. Note that the two tumors with low but detectable PAX-FKHR expression (blue dots) had median PAX-FKHR expression levels 1,000-fold less than alveolar tumors expressing high levels of PAX-FKHR (orange to pink dots). Rhabdomyosarcoma samples for which no QRTPCR data were available are also indicated (gray dots). Note that all nonARMS tumors were shown at diagnosis to be fusion negative by conventional RT-PCR. (From reference Davicioni et€al. 2006 with permission)
genes have been proposed as the downstream candidates for transcriptional activation by the chimeric PAX3-FKHR gene. Earlier studies focused mainly on the observation that the PAX3-FKHR fusion protein exerts greater transcriptional activity than that of wild type PAX3 protein (Fredericks et€al. 1995). Several studies, in particular, used microarrays and expression profiling to determine downstream consequences of the wild type PAX3 and PAX3-FKHR fusion proteins. In a cell culture-based approach to identify targets, the wild type gene or fusion gene was transfected and ectopically expressed in a cell line. A set of target genes, which were similarly regulated by both the wild type and fusion proteins, were identified. However, another set of target genes, which were regulated only by PAX3-FKHR, or were repressed by PAX3 but induced by PAX3-FKHR, were also identified. Of importance is that the specific downstream target genes identified were dependent on the cell line in which the fusion gene was ectopically expressed, Table╛1. An example of this cell type specificity is the downregulation of BMP4 by PAX3-FKHR in SaOS-2 cells and the upregulation of BMP4 by PAX3-FKHR when expressed in RD ERMS cells. In addition, all these results were obtained from experiments using the method of forced expression, and using cells other than alveolar rhabdomyosarcoma cells, such as mouse fibroblast cells (Zhang and Wang 2003), osteosarcoma cells (Begum et€al. 2005) or embryonal rhabdomyosarcoma
430
R.T. Kurmasheva et al.
Table€1â•… Genes regulated by ectopic expression of PAX3 or PAX3-FKHR in nonARMS cells Gene PAX3 PAX3-FKHR Cell type References → ↑ Human osteosarcoma, SaOS-2 Begum et€al. (2005) PLN CNR1 ↑ ↑ Human osteosarcoma, SaOS-2 CNR1 ↑ ↑ Human ERMS, RD cells PTHLH ↓ ↑ Human osteosarcoma, SaOS-2 SLIT2 → ↑ Human osteosarcoma, SaOS-2 SLIT2 → Human ERMS, RD cells EPHA4 ↑ Human osteosarcoma, SaOS-2 EFNB2 ↑ Human osteosarcoma, SaOS-2 MSX2 ↓ Human osteosarcoma, SaOS-2 BPM4 ↓ ↓ Human osteosarcoma, SaOS-2 BPM4 ↑ Human ERMS, RD cells PAX3 ↑ Human ERMS, RD cells Tomescu et€al. (2004) CXCR4 ↑ Human ERMS, RD cells Tomescu et€al. (2004), Begum et€al. (2005) CXCR4 → Human osteosarcoma, SaOS-2 Begum et€al. (2005) PAX7 Human ERMS, RD cells Tomescu et€al. (2004) ↑ Human ERMS, RD cells, and Nabarro et€al. (2005) H2-K 76-9 cells (MHC class 1) PAI1 ↓ Human ERMS, RD cells STAT1 ↓ Human ERMS, RD cells STAT3 ↓ Human ERMS, RD cells TAP1 ↓ Human ERMS, RD cells CXCL10 ↓ Human ERMS, RD cells CCL5 ↓ Human ERMS, RD cells IGF2 ↑ Human ERMS, 76-9 cells TRAM1 ↑ Human ERMS, 76-9 cells COLSA1 ↑ Human ERMS, 76-9 cells SKP2 ↑ Murine NIH3T3 fibroblast cells Zhang and Wang (2003)
cells (Tomescu et€al. 2004; Begum et€al. 2005; Nabarro et€al. 2005). Therefore, the issue must be critically assessed as to the degree to which any given ectopic cell culture system provides a model of the actual ARMS tumor environment. In other words, in these studies, direct effects of PAX3-FKHR in alveolar rhabdomyosarcoma cells have not yet been conducted. Up- and downregulated genes by ectopic expression of PAX3 or PAX3-FKHR reported in the literature are summarized in Table€1. To elucidate targeted genes downstream of PAX3-FKHR, targeted knockdown of PAX3-FKHR expression in ARMS cells by antisense oligonucleotides and siRNA have also been attempted (Bernasconi et€ al. 1996; Ebauer et€ al. 2007). However, again these studies have not addressed the specific effects of selectively downregulating PAX3-FKHR. For example, the antisense oligonucleotides and the siRNAs used were designed against the PAX3 gene and were not specific for PAX3-FKHR. Thus, while this approach suppressed PAX3-FKHR, it also suppressed the expression
Molecular Therapy for Rhabdomyosarcoma
431
of PAX3 itself. Hence, the phenotypes reported may be derived from PAX3 gene suppression, and not necessarily from the suppression of PAX3-FKHR. Recently, we established an in€ vitro system using several siRNAs designed against the PAX3FKHR fusion point where PAX3-FKHR fusion gene expression is specifically and efficiently depleted with no effects on expression of either PAX3 or FKHR. We have examined the effects of specific PAX3-FKHR suppression in several ARMS cell lines. Our results differ from previous and recent studies that used antisense oligonucletides and siRNA approaches to regulating PAX3-FKHR. Specifically, we found that the downregulating PAX3-FKHR causes inhibition of the rate of cellular proliferation, and an accumulation of cells in G1 phase of the cell cycle, but no increase in cell death through apoptosis. Downregulation of PAX3-FKHR led to reduced levels of the MET receptor, and reduced motility in response to stimulation with HGF, its ligand. Suppression of PAX3-FKHR induced the myogenic differentiation gene, myogenin (Fig.€ 3), and muscle differentiation (morphologic change and the expression of muscle specific proteins, desmin and myosin heavy chain). These results suggest that PAX3-FKHR has important functions in ARMS cells to promote malignant phenotypes such as proliferation, motility and to suppress differentiation. Candidate genes regulated by PAX3-FKHR that may be relevant to maintenance of the malignant phenotype, include SKP2, CXCR4, HDAC4, myogenin (myogenic factor 4), MRF5 (myogenic factor 5). Potential drug targets upregulated by PAX3FKHR include MET, the cytokine receptor CXCR4, fibroblast growth factor receptor 4 (FGFR4) and the cyclin-dependent kinase CDC2 (CDK1). PD173074, is a synthetic compound of the pyrido(2,3-d)pyrimidine class, that inhibits tyrosine kinase activities of FGF receptors including FGFR4 (Ezzat et€al. 2006). Crystal structure elucidation has identified PD173074 in complex with the tyrosine kinase domain of FGFR1 with a high degree of surface complementarity with the hydrophobic, ATPbinding pocket of FGFR1 (Mohammadi et€al. 1998). Systemic administration of this compound effectively blocks FGF-induced angiogenesis and neurotrophic actions (Skaper et€al. 2000). The staurosporine derivative PKC412 that inhibits FGFR/FLT3/ PKC has entered clinical trials (Stone et€al. 2004), and has shown some benefit in patients with cholangiocarcinoma. Nausea/vomiting and diarrhea were the most common toxicities in adults (Propper et€al. 2001). CXCR4 antagonists are in clinical development as “entry fusion inhibitors” for the treatment of type 1 human immunodeficiency virus and include small peptide agents (ALX40-4C) (Doranz et€ al. 2001) and small molecule inhibitors (AMD 3100; Plerixafor). Plerixafor is a bicyclam derivative that acts as a stem cell mobilizer by blocking the CXCR4 chemokine receptor, which triggers the rapid movement of stem cells out of the bone marrow and into circulating blood. Plerixafor is in phase III clinical trials in stem cell transplantation among cancer patients. Inhibition of CDK1 has been proposed as a potential therapy for tumors that overexpress MYC, based on the selective killing of MYC-overexpressing cells by purvanolol or RNAi targeting MYC (Goga et€ al. 2007). This may be particularly of interest if the same synthetic lethal interaction occurs in RMS that overexpress MYCN independent of p53 functional status. Agents that target CDK1 are in preclinical testing (Lin et€al. 2007), and agents such as roscovitine or R547 (DePinto et€al. 2006) are in clinical trials.
432
R.T. Kurmasheva et al.
a
.
b Western blotting
120%
PAX3FKHR
PAX3-FKHR 60% /GAPDH
MET
0%
Myogenin 100%
Met /GAPDH
IGF-IRβ
50% 25%
β-actin 400%
Myogenin /GAPDH
control
mock
siCON
si P P--F
c
200%
120% 100%
0%
100.0%
PAX3-FKHR/GAPDH PAX3/GAPDH FKHR/GAPDH
80%
IGF-IRβ /GAPDH
60%
100%
40%
50%
17.1%
20% 0%
0%
control mock si CON si P-F 10nM 10nM
3.18% 0.27%
siCON
2.22% 0.39%
siP-F
Fig.€3â•… Specificity of RNAi designed against the fusion point (siP-F) in suppressing PAX3-FKHR mRNA and the impact from PAX3-FKHR knockdown on MET, myogenin, and IGF-IRb mRNAs and proteins. (a) The relative amounts of PAX3-FKHR mRNA relative to GAPDH mRNA 24€hours after transfection. The levels of the PAX3-FKHR mRNA were not changed in the mediumonly control, mock transfection, and si-CON-transfected (scrambled control). (b) Western blot of PAX3-FKHR, MET, Myogenin, IGF-1Rb protein, and b-actin 48€hours after transfection in the medium-only control, mock transfection and si-CON-transfected (scrambled control). (c) The relative abundance of transcripts of PAX3-FKHR, PAX3, and FKHR in Rh30 cells transfected with si CON, and transfected with siP-F. In siCON-transfected cells, the mRNA levels of both PAX3 and FKHR were very low compared to that of PAX3-FKHR, being about 1/30 and 1/300 the level of PAX3-FKHR, respectively. Although siP-F decreased the mRNA level of PAX3-FKHR by 83%, but the mRNA levels of both PAX3 and FKHR were not largely affected by this siP-F (adapted from Kikuchi K, Tsuchiya K, Otabe O, Gotoh T, Tamura S, Katsumi Y, Yagyu S, TsubaiShimizu S, Miyachi M, Iehara T, Hosoi H. 2008 Effects of PAX3-FKHR on malignant phenotypes in alveolar rhabdomyosarcoma. Biochem Biophys Res Commun 365:568–74.)
MYCN Dysregulation Amplification of the MYCN oncogenic transcription factor locus is associated with poor prognosis in neuroblastoma. However, amplification and overexpression of MYCN has been reported to occur frequently in RMS, predominantly in ARMS
Molecular Therapy for Rhabdomyosarcoma
433
(Dias et€al. 1990; Driman et€al. 1994; Hachitanda et€al. 1998). Cumulatively, these relatively small studies observed amplification of MYCN in 16 of 28 ARMS and 0 of 27 ERMS. However, whether amplification correlated with outcome was controversial. Larger studies (Gordon et€al. 2000; Lu et€al. 2001) indicated that the 2p24 MYCN locus is frequently amplified, and even more frequently genes in this region are overexpressed in RMS. A report where expression (nâ•›=â•›113 cases) and copy number (nâ•›=â•›92) were determined concluded (Chesler et€ al. 2006) that increased copy number of MYCN was a feature of both ERMS and ARMS. In ERMS high expression occurred even without increased copy number. Further, in patients with ARMS, overexpression, greater than the median, or gain of MYCN copy number were significantly associated with poor outcome (Williamson et€ al. 2005). Thus, MYCN, or genes downstream of this transcription factor appear to be of importance in the genesis or progression of RMS and may reveal potential targets for therapy (Lu et€ al. 2003; Pession and Tonelli 2005; Tonelli et€ al. 2005; Morgenstern and Anderson 2006). However, whether dysregulation of MYCN is critically involved in the pathogenesis of RMS has yet to be demonstrated. Several approaches to downregulating MYCN have been proposed, including antisense oligonucleotides that retard the growth of the MYCN-driven transgenic model of neuroblastoma in mice (Burkhart et€ al. 2003) and peptide nucleic acids (Tonelli et€ al. 2005) that induce apoptosis in some cell lines. Whether such approaches, including therapeutic inhibitory RNA (siRNA) will ultimately have therapeutic utility remains a matter of conjecture at this time. Further, the effect of downregulating MYCN in RMS models has not been reported. MYC proteins form heterodimers with MAX through their HLHZip domains and the heterodimer activates transcription of multiple genes by binding the E-box motif (CACGTG). Berg et€al (2002) have reported a small molecule inhibitor of MYC/MAX dimerization; however, this does not discriminate between MYC family members, hence may have pleiotropic effects against MYCdependent tissues. Of importance for developing therapeutic approaches is that MYC proteins have a relatively short half-life of about 20 to 30€min. MYCN is also expressed only in early development, hence may represent a tumor-specific target. In neuroblastoma cells, MYCN transcription is stimulated by IGF-1 (Misawa et€al. 2000) through the mitogen activated protein kinase pathway. Inhibition of phosphoinositol 3¢-kinase (PI3K) reduces MYCN protein and decreases transcription of two downstream genes (mdm2 and mcm7). MYCN induction in response to IGF-1 is decreased by PI3K inhibitors putatively through a combination of effects on translation and MYCN protein stability (Chesler et€al. 2006). The proposed mechanism for downregulation of MYCN protein is that inhibition of PI3K leads to the activation of glycogen synthase kinase 3b (GSK3b) a kinase that phosphorylates MYCN and causes its destabilization, and induces apoptosis in neuroblastoma cells. MYCN amplified neuroblastoma cell lines are also hypersensitive to a small molecule inhibitor of translation elongation (Radhakrishnan and Gartel 2006) that causes rapid downregulation of MYCN protein (Radhakrishnan et€al. 2008). Thus, in addition to inducing selective cytotoxicity through targeting CDK1 in the context of MYC overexpression, there appear to be numerous approaches to regulate MYCN. Inhibitors of PI3K (BEZ235, SF1126), GSK3b (LiCl) and Akt (RX-0201, VQD-
434
R.T. Kurmasheva et al.
002, XL418, GSK609693) are in development, and several agents are in adult clinical trials. GSK609693 is currently undergoing the evaluation against pediatric solid tumors and acute lymphocytic leukemia models in the Pediatric Preclinical Testing Program (Houghton et€al. 2007).
TP53 Dysregulation The tumor suppressor P53 (TP53) is mutated in approximately 50% of human cancer; however, its status in RMS is somewhat unclear. In part, this is because the function of TP53 can be attenuated by overexpression of HDM2, the human homolog of MDM2, which binds TP53 preventing its transcriptional activity (Bottger et€al. 1997). HDM2 also has ubiquitin ligase activity, which targets TP53 for proteasomal degradation (Honda et€al. 1997). The reported incidence of TP53 mutations in RMS is relatively low (approximately 5%) and similarly amplification of HDM2 is found in 10% of clinical samples (Taylor et€ al. 2000). Immuno� histochemical staining for TP53 and HDM2 has been variable; clinical RMS samples showed both TP53 and HDM2 to be expressed at low levels in both ERMS and ARMS (Leuschner et€ al. 2003), although overexpression of TP53 (interpreted as indicative of stabilized mutated TP53) was postulated to be a crucial step in metastatic disease for patients with ERMS. In contrast, the study of Takahashi et€al. (2004) found overexpression of TP53 in 30% of cases, 22% had gene abnormalities, and HDM2 was amplified in 13%. However, determination of the functional status of TP53 is further complicated as TP53 is a member of a multigene family that includes TAP63 and TAP73 (Flores et€al. 2002, 2005), which encode proteins that have overlapping function with TP53, for example induction of apoptosis in response to DNA damage (Flores et€al. 2002; Melino et€al. 2004; Flores 2007). Multiple isoforms of TAp73 can be produced by alternative use of intronic promoters, alternative splicing, or use of an internal ribosome entry site (IRES) (Melino et€al. 2002; Stiewe and Putzer 2002; Sayan et€al. 2007). Notably, N-terminal truncated forms (DNp73) act as a potent transdominant negative regulators of all p53-members (Grob et€ al. 2001; Zaika et€ al. 2002), and suppress multiple developmental programs including muscle differentiation (HuttingerKirchhof et€al. 2006). Of importance is that DNp73 is expressed in many clinical samples of RMS and cell line models (Cam et€al. 2006), and confers chemoresistance through suppressing TP53-induced CD95 and proapoptotic BCL2 members (Gressner et€al. 2005; Muller et€al. 2005). Thus, it is probable that TP53 function is attenuated through HDM2 and DNp73, as well as mutations within the TP53 gene itself. As p53-family members regulate stress-induced apoptosis (Vogelstein et€al. 2000), and loss of function confers drug or radiation resistance, strategies to reactivate TP53 function appear to have some application for the treatment of RMS. The most developed approach is to attenuate the activity of HDM2, and several chemical classes of inhibitor that block HDM2-p53 interactions have been described (Vassilev 2007). This may be particularly effective in tumors where the
Molecular Therapy for Rhabdomyosarcoma
435
HDM2 gene is amplified or overexpressed (Tovar et€ al. 2006; Vassilev 2007), however, to date no studies with RMS cell lines have been reported. Nutlin 3a, a small molecule inhibitor of HDM2-p53 interactions may induce G1 cell cycle arrest and apoptosis (Logan et€ al. 2007) consistent with recruitment of p53 to chromatin and increased expression of p53-responsive genes, CDKN1A (encoding p21CIP1), PUMA, GADD45, and HDM2. Nutlin 3a can also induce p53-dependent senescence (Efeyan et€al. 2007). Interestingly, nutlin 3a can enhance chemotherapy-induced apoptosis in p53-mutant cell lines through an E2F1-dependent induction of the proapoptotic proteins TAp73a and NOXA (Ambrosini et€al. 2007). An alternative strategy proposed (Kranz and Dobbelstein 2006) is to use the HDM2 inhibitor to selectively block cells with wild type TP53 in G1 phase. In contrast, cells mutant for TP53 progress into S-phase and are selectively killed by S-phase specific agents such as gemcitabine, a strategy equally applicable to topotecan or irinotecan used in treatment of RMS. HDM2 antagonists may also have antiangiogenic activity through disrupting the HDM2-hypoxia inducible factor 1a (HIF1a) interaction that enhances transcription of vascular endothelial growth factor (VEGF) (Bardos et€al. 2004; LaRusch et€al. 2007; Secchiero et€al. 2007). However, the strategy for use of HDM2 inhibitors will be successful only if the apoptotic machinery downstream of TP53 is functional. For example, overexpression of antiapoptotic proteins of the BCL2 family or MCL1 could antagonize cell death induced by these agents. Similarly, overexpression of DNp73 would be anticipated to antagonize HDM2 inhibitors, and abrogate approaches such as introduction of wild type p53 using an adenoviral vector (Ganjavi et€ al. 2005), shown to sensitize RMS cells to DNA damaging cytotoxic drugs.
Aberrant Growth Factor Signaling in RMS Insulin-Like Growth Factor Signaling RMSs show the presence of both active type-1insulin-like growth factor receptor (IGF-1R) and the autocrine production of its ligand IGF-2 (Minniti et€ al. 1994). IGF-1 and -2 and IGF-1R regulate all aspects of the malignant phenotype with IGF-1R being activated by its ligands and also by steroid hormones (Kaleko et€al. 1990; Stewart and Rotwein 1996; Sepp-Lorenzino 1998). The activated IGF-1R is capable of phosphorylating other tyrosine-containing substrates of which the insulin receptor substrates (IRS-I-4) link the receptor to a cascade of enzyme activations via PI3K-Akt-mTOR and RAF-MAPK systems (LeRoith et€al. 1995). Activation of these pathways results in a variety of responses, such as cell proliferation, differentiation, migration, and in mouse models allows maintenance of the malignant phenotype (Bohula et€ al. 2003; Baserga 2004). IGF-1R signaling transcriptionally upregulates antiapopototic genes and suppresses proapoptotic genes, through a multitude of pathways as shown in Fig.╛4. IGF-1R signaling is
436
R.T. Kurmasheva et al. cFLIP BCL2 BAX BNIP3
Survivin
PAR4
?
cIAP-2 XIAP
CREB p53
TWIST
ERK1/2
GLUT1
RAD51 IRS-1
GSK3 b
IGF-IR
AKT
p38 NF k B
FKHR
BIM
AIB1 JNK
JIP-1
TDAG51
?
NF k B
MEF2D ?
TELOMERASE
Fig.€4╅ IGF-IR-mediated survival through transcriptional regulation. IGF-IR signaling via IRS-1, AKT, or MAP kinases (ERK1/2, p38) positively regulates transcription of antiapoptotic genes (red ) and negatively regulates proapoptotic genes (green). IGF-IR signaling regulates transcription factors that themselves regulate antiapoptotic programs. (From reference Kurmasheva and Houghton 2006 with permission)
tightly regulated since high IGF-1 levels result in a decline in IGF-1R, and IGFs may also act as negative feedback signals to repress expression of IGF-1R (Yang et€al. 1996; Hernandez-Sanchez et€ al. 1997). It has been recently shown that IGF-1R mediates Akt activation in response to inhibition of mTOR through a feedback loop in which IRS-1 becomes stabilized in Rh-30 and RD cells (Petricoin et€al. 2007). In contrast to the effect of IGFs, other growth factors, including bFGF, PDGF, and EGF, as well as estrogens, glucocorticoids, GH, FSH, luteinizing hormone, thyroid hormones stimulate IGF-1R expression (LeRoith et€ al. 1995). IGF-2 is consistently overexpressed in both embryonal and alveolar RMS, where it also cooperates with PAX3-FKHR (El-Badry et€ al. 1990; Minniti et€ al. 1994; Wang et€al. 1998). At the transcriptional level, PAX3-FKHR is able to transactivate the IGF-1R promoter in sarcoma-derived cell lines, whereas PAX3 exhibited a reduced potency in comparison to the fusion protein (Ayalon et€ al. 2001). Low concentrations of IGF-1 protein in media can be detected in RMS cells in€vitro by sensitive methods such as ELISA, although the molar ratio of IGF-2:IGF-1 is approximately 500 (RTK, unpublished data). Thus, constitutive activation of IGF-1R by its ligands (mainly IGF-2) can disrupt the delicate balance between proliferation and differentiation, creating the expanded pool of cycling myogenic progenitors that become targets of further mutagenesis and potential seeds of
Molecular Therapy for Rhabdomyosarcoma
437
future RMS. In animal models, IGF-2 is essential for the development and malignant behavior of RMS (Wang et€al. 1998; Hahn et€al. 2000). Interestingly, wild type TP53 inhibits transcription of the IGF-2 gene (Zhang et€al. 1996, 1998a). As the function of p53 appears to be suppressed in RMS cells through mutation, overexpression of HDM2 or expression of DNp73, an N-terminal truncated isoform of TAp73, this could help explain overexpression in RMS. In normal development, IGF-2 mRNA is most abundant during the fetal development and rapidly declines after birth. The activity of the IGF-2 gene in normal mouse and human muscle tissues is regulated by genomic imprinting, which is a form of nonMendelian inheritance in mammals, where the imprinted genes are expressed uniquely from one allele. The expressed allele, either paternal or maternal, is constant for each imprinted gene, unless a genetic or epigenetic alteration has occurred. Almost all imprinted genes identified to date can be classified as regulators of embryonic growth, placental growth, or adult metabolism (Morison et€ al. 2005). The IGF-2 signaling pathway has implications in all three of these functions. Moreover, it has been documented that upregulation of IGF-2, a common outcome of loss of imprinting (LOI), occurs at high frequency in a large variety of human tumors including childhood cancers. In normal tissues, the IGF-2 gene is exclusively silent at the maternal allele. Zhan et€ al. (1994) reported that normally imprinted allele of the IGF-2 gene was activated in alveolar RMS tumors as well as in Rh-28 cells. Embryonal RMS are usually characterized by LOI, overexpression of IGF-2 and paternal disomy of the IGF-2 locus. It appears that both alveolar and embryonal histologies of RMS have alterations of the normal IGF-2 locus. Loss of heterzygosity (LOH) that is caused by the loss of maternal chromosome has only been observed in tumors belonging to the embryonal subtype (Scrable et€al. 1989). Zhan et€al. (1994) have suggested that LOI represents the functional equivalent of LOH with paternal disomy in RMS, and possibly both events are related to the increased expression of IGF-2 seen in both histologic subtypes of RMS. In addition, uniparental paternal disomy (UPD) of the 11p15 locus have been described in predisposition to the development of tumors of embryonic origin (Weksberg et€al. 1993). Most of the imprinted genes known so far are organized in clusters. Two genes with in€vitro growth inhibitory capacity, namely H19 and p57KIP2 (Hao et€al. 1993; Matsuoka et€al. 1995), and growth promoting IGF-2 gene, are members of such a cluster with the first two genes being paternally imprinted. Low H19 mRNA was associated with the loss of the maternal allele or with relaxation of IGF-2 imprinting, although LOI of the IGF-2 gene was not necessarily coupled to low H19 expression in both ERMS and ARMS (Casola et€al. 1997). Low H19 mRNA levels relative to normal tissues have been found in the great majority of Wilms’ tumor samples as well as in RMS, suggesting the role of a tumor suppressor for this gene (Steenman et€al. 1994; Casola et€al. 1997). It was shown that forced expression of the transfected H19 gene inhibited cell growth of a RMS cell line and suppressed tumorigenicity of Wilms’ tumor cells (Hao et€al. 1993). The mechanism of regulation of H19/IGF-2 expression has become clearer with identification of a new component in the regulatory system. On the maternal chromo-
438
R.T. Kurmasheva et al.
some, an enhancer-blocking protein CCCTC-binding factor CTCF (recognizes the DNA sequence CCCTC) binds to the differentially methylated regions (DMR), preventing the enhancers from interacting with the promoter of IGF-2, and instead favoring H19 expression (Schoenherr et€al. 2003). But on the paternal chromosome, the DNA of the boundary element is methylated, the blocking protein cannot bind, and IGF-2 can be expressed while H19 is not. Physiologically, CTCF repeats are required to maintain the hypomethylated state of the entire DMR during the period of active de€novo methylation that occurs after implantation. Possibly, the presence of CTCF engenders a tight engagement between the DMR, the H19 promoter and the downstream enhancers, thus excluding access of the IGF-2 gene to the enhancers that allows maintenance of monoallelic IGF-2 expression (Engel et€al. 2006). It is not presently known if CTCF is the only protein capable of this type of imprinted gene regulation. IGF-2 promoters play different roles throughout development and in different tissues: they are normally active during fetal growth, but silent in postnatal life, and become reactivated in cancer. The IGF-2 gene is transcribed from four different promoters (P1–P4). P2–P4 contains CpG islands, and transcription from these promoters is subject to imprinting while biallelic expression is observed from the other alternative promoter (P1) (Pavelic et€al. 2007). The developmentally regulated transcription factor AP-2 is expressed at high expression levels in RMS and human fetal skeletal muscle cells but not in adult skeletal muscle cells. This suggests that AP-2 may also contribute to the high expression of IGF-2 in RMS cells (Zhang et€al. 1998b). The role of IGF-1R signaling in the pathogenesis of RMS, and its role in preventing apoptosis induced by a multitude of cellular stresses including cytotoxic drugs, radiation, and hypoxia (Kurmasheva and Houghton 2006) indicate that targeting this pathway may have considerable utility for therapy of RMS. As dysregulated IGF-1 signaling is common to several adult malignancies, targeting IGF-1R has become a major focus for therapeutic development (Braczkowski et€ al. 2002; Cohen et€ al. 2005; Wang et€ al. 2005; Guerreiro et€ al. 2006; Kurmasheva and Houghton 2006; Sachdev and Yee 2006, 2007). Currently, there are both small molecule drugs and fully human or humanized antibodies directed at the IGF-1R. Several classes of small molecule inhibitors of IGF-1R have been described, including tyrphostins, pyrrolopyrimidines (Mitsiades et€al. 2004; Scotlandi et€al. 2005), substituted benzimidazole derivatives (Haluska et€ al. 2006) and diarylureas (Gable et€ al. 2006). Small molecule ATP-competitive IGF-1R inhibitors do not induce downregulation of the receptor, and in most cases do not discriminate between the insulin receptor and the IGF-1R (Haluska et€al. 2006; Wittman et€al. 2007). As a consequence, these inhibitors may have significant effects on glucose homeostasis (causing hyperglycemia and hyperinsulinemia). Several small molecule inhibitors of IGF-1R. Five fully human (CP751871, AMD479, R150 (OSI-906 and BMS-754807) are inclinical trials. 7, IMC-A12, SCH717454) or humanized antibodies (H7C10) are in adult phase-I or -II clinical trials. These agents show specificity for the IGF-1R although they may inhibit chimeric receptors formed through heterodimerization with the insulin receptor. In preclinical models of childhood cancers, the prototypical antiIGF-1R antibody, a-IR3 mediated downregulation of IGF-1R significantly retards the€growth of many cell lines in€vitro (El-Badry et€al. 1990), and retards the growth of RMS xenografts (Kalebic et€al. 1994). SCH717454 significantly inhibits the growth of
Molecular Therapy for Rhabdomyosarcoma
439
RMS xenografts and induces regressions in several sarcoma histotypes notably osteosarcoma and Ewing sarcoma (Kolb et€al. 2008). Molecular characterization of these sensitive models where IGF-1R signaling appears to be critical could identify subsets of tumors that have become “addicted” to this pathway (Weinstein and Joe 2006). In other preclinical models, blocking IGF-1R signaling results in significant retardation of tumor growth although in a clinical setting this response would be scored as progressive disease. In these models with intermediate sensitivity, such as RMS, combinations of signaling inhibitors would potentially be a more effective antitumor therapy. One strategy that is being evaluated in preclinical models is the combination of the mTOR inhibitor, rapamycin, with IGF-1R inhibitors. The basis for this combination is that inhibition of mTOR upregulates IGF-1R signaling through stabilization of IRS-1 (Easton et€al. 2006), and IGF-1R signaling blocks rapamycin-induced apoptosis (Huang et€al. 2003; Thimmaiah et€al. 2003). Alternative approaches to inhibiting IGF-1R signaling include the development of ligand binding antibodies. Targeting receptor ligand rather than the receptor per se has proven a valuable approach for the antiangiogenic antibody, bevacizumab, and high affinity fully human antibodies have been developed against IGF-II (Feng et€ al. 2006). Another approach to limiting IGF-1R signaling is to administer IGF-binding proteins as therapeutics. The activity of IGFs is regulated by a family of six structurally related high affinity IGF-binding proteins (IGFBPs) (Hwa et€al. 1999). In the plasma, 99% of IGFs are complexed to IGFBPs that modulate the availability of free IGF-1 to the tissues. At the same time, binding of IGFBPs to IGFs protects IGFs from proteolytic degradation increasing their bioavailability. Normally, more than 90% of the circulating IGFs are bound to IGFBP-3. Of the IGFBPs, IGFBP-6 (approximately 30€ kDa) binds IGF-2 with the greatest affinity, having a 20- to 100-fold greater binding affinity for IGF-2 than for IGF-1 (Bach et€al. 1995). High levels of IGFBP-6 expression in cells are associated with nonproliferative and differentiated states, both in€vitro and in€vivo. In particular, IGFBP-6 expression is correlated with quiescence in myoblast culture and IGFBP-6 protein was not identified in media conditioned by RMS cells, and expression is strongly downregulated in spontaneously metastatic RMS cells compared with nonmetastatic parental RMS cells (Scholl et€ al. 2000). Additionally, IGFBP-6 inhibits IGF-2 mediated differentiation and proliferation of myoblasts (Bach et€ al. 1994, 1995). Unlike in normal cells, lower IGFBP-6 expression in RMS than in myoblasts would therefore be expected to enhance the actions of IGF-2. Gallicchio et€al. (2001) reported that IGFBP-6 significantly inhibited monolayer RD and Rh-30 cell proliferation in a dose-dependent manner. Further, overexpression of IGFBP-6 resulted in a 74 to 88% reduction in Rh-30 tumor size in€vivo after 18€days, showing that IGFBP-6 can be a potent antitumor agent.
Hepatocyte Growth Factor/MET Autocrine Signaling Increased expression of the MET receptor or its ligand hepatocyte growth factor/ scatter factor (HGF) occurs frequently in human cancer, and increases with tumor progression in some malignancies (Longati et€al. 2001). In myogenic precursors,
440
R.T. Kurmasheva et al.
PAX3 activates transcription of several genes, including MYOD and MET. As discussed above, MET is consistently downregulated when PAX3-FKHR is downregulated by antisense or siRNA, and is overexpressed in cells expressing PAX3-FKHR. MET encodes a cell-surface receptor transcriptionally activated by both PAX3 and PAX7 that mediates delamination of myoblast precursors from the epithelial dermomyotome and their migration in the limb bud in response to HGF. MET has been shown to be highly overexpressed in ARMS (Epstein et€ al. 1996; Ferracini et€ al. 1996; Ginsberg et€al. 1998), but while the gene is not mutated or amplified (Chen et€al. 2007) the expression level of MET was found to be significantly higher in patients who died of disease. MET is highly expressed in cell lines derived from ARMS (Epstein et€ al. 1996) and HGF induces motility and confers drug resistance (Jankowski et€al. 2003). Conditioned media from bone marrow derived fibroblasts chemoattracts RMS cells in an HGF-dependent manner and RMS cells with higher expression of MET increased homing/seeding to bone marrow of RMS cells inoculated intravenously into mice. In contrast to ARMS, ERMS have more variable expression of MET, but appear to express the ligand HGF at higher levels than in ARMS. Analysis of RMS xenografts indicates phosphorylation of MET (Y1234/1235), in both ARMS and some ERMS suggesting auto-activation. As murine HGF does not activate the human receptor, these data imply autocrine activation, at least under conditions of RMS growth in the mouse. Transcripts for human HGF are also detected in both ERMS (2/2) and ARMS (2/5) xenografts (unpublished data). Consistent with the model data, MET was detected in all of 68 RMS specimens and 62% of tumors coexpressed HGF, however, neither high MET nor HGF expression correlated with metastatic disease (Rees et€ al. 2006). Somewhat unexpectedly, MET was a consistent feature of ERMS and not ARMs. MET appears to be an essential mediator of PAX3-FKHR-induced oncogenesis, as exogenous HGF rescued the ability of PAX3-FKHR-transduced cells to form colonies in soft agar under low serum-containing conditions. Further, PAX3-FKHR was capable of transforming wild type murine embryo fibroblasts but not those derived from mutant (inactive) MET (Taulli et€al. 2006). Together with data showing the downregulation of MET inhibits proliferation (although induction of apoptosis has been variable between reports) and migration of RMS cells in€ vitro and in€ vivo, the evidence points to the MET receptor being a potential target for therapy of both ARMS and ERMS. Relatively specific inhibitors of MET (SU11271, SU11274, SU11606 (Wang et€ al. 2003), SGX523, PHA665752 (Ma et€ al. 2005), and PF2341066) as well as multitargeted kinase inhibitors that are potent against MET (XL880 and MP470) have been reported. Both humanized (AV299) and fully human antibodies (AMG102) against HGF, as well as MET decoys (CGEN214) that prevent ligand binding, or receptor dimerization, and one-armed monoclonal antibodies against the MET receptor are in development. Preliminary clinical results in adult trials with ARQ197, a small molecule MET inhibitor, suggest that this agent is well tolerated, with fatigue, diarrhea, and elevated ALP and AST as the most frequent toxicities (Garcia et€al., 2007), and antitumor activity has been observed with oral dosing. XL880, a multitargeted kinase/MET inhibitor, has demonstrated reversible Grade 3 elevations in lipase, hepatic transaminases and proteinurea, with objective responses
Molecular Therapy for Rhabdomyosarcoma
441
in four of 40 adult patients (Eder et€al., 2007). Common adverse events included hypertension and fatigue with the loss of concentration. These data from early clinical trials in adults are encouraging, and suggest these agents should be evaluated against relevant models of RMS, and potentially advanced to clinical trials.
Hedgehog Signaling Pathway Gorlin syndrome, also known as nevoid basal cell carcinoma syndrome (NBCCS) is caused by mutations in the hedgehog receptor PTCH gene, and is characterized by the developmental defects and predisposition to the development of tumors including fetal rhabdomyomas and ERMS (Tostar et€al. 2006). Binding of hedgehog (Hh) to PTCH suppresses smoothened (SMO) the direct activator of the transcription factor GLI1. GLI1 amplification occurs in some RMS, and overexpression of GLI1 and GLI2 suppresses skeletal muscle differentiation (Gerber et€al. 2007) through binding to and suppressing MYOD heterodimerization with E12 proteins. Losses in the PTCH locus 9q22 have been identified in 12 to 30% of ERMS (Bridge et€al. 2000; Calzada-Wack et€al. 2002; Tostar et€al. 2006). Genomic loss at 10q24 a locus encoding the GLI-suppressor SUFU occurs in 18% of ERMS (Bridge et€al. 2002). Thus, cumulative loss of negative regulatory components in the Hh pathway approaches 50% in ERMS. Interestingly, a positive regulator of GLI transcription (Mao et€al. 2002), the survival kinase DYRK1/MIRK is significantly expressed in most RMS (Mercer et€ al. 2006). Several classes of small molecule inhibitors of SMO have been reported (Chen et€ al. 2002; Williams et€ al. 2003). CUR61414, showed quite impressive activity against transgenic models of basal cell carcinoma in PTCH+/− mice (Williams et€ al. 2003) and a precursor molecule HhAntag691 prevented the development of medulloblastoma in PTCH+/− p53−/− mice (Romer et€al. 2004). CUR61414 advanced to clinical trials for the treatment of basal cell carcinoma, but failed to downregulate pharmacodynamic markers essential for establishing activity in the phase 1 clinical trial. More recently GCD-0449, another Hedgehog inhibitor has shown robust activity in treatment of basal cell carcinoma.
Other Pathways (RAS/MAPK/c-KIT) Proliferating myoblasts express the myogenic regulatory factors MYOD and MYF5, whereas myogenin and MRF4 are expressed as terminal differentiation progresses. Despite expression of these myogenic regulatory factors in myoblasts, mitogenic stimulation or activated oncogenes repress the differentiation program (Perry et€al. 2001). Activation of RAS genes has been reported in ERMS with varying frequency (3 to 30%) (Stratton et€al. 1989; Chen et€al. 2006). However, oncogenic RAS induced the formation of RMS in zebrafish, concordant with expression
442
R.T. Kurmasheva et al.
profiles consistent with an activated RAS pathway. Extension of these studies indicated that RAS pathway activation appears to occur frequently in human RMS (Langenau et€al. 2007). Activated RAF and MEK1, downstream of RAS can also prevent myogenic differentiation (Dorman and Johnson 2000; Perry et€ al. 2001), and an inhibitor of MEK1 induces growth arrest, myogenic differentiation (in some cell lines) and downregulation of c-MYC. Thus, targeting the RAS/MAPK signaling pathway may offer some therapeutic potential. Targeting activated RAS through the use of farnesyltransferase inhibitors has utility limited to the treatment of hematologic diseases (Santos et€al. 2004). However, inhibitors of RAF (Sorefenib) have clinical utility in the treatment of adult carcinomas, and MEK1 inhibitors (CI-1040, PD325901, and AZD6244) are in clinical trials for adults, and may have utility for the treatment of RMS. However, testing of AZD6244 against the PPTP panel in€vivo did not demonstrate robust antitumor activity as a single agent (unpublished data). Overexpression of the stem cell factor receptor c-KIT has been reported in 10% of spindle cell RMS (Komdeur et€al. 2003), but is absent in ARMS and only 2 of 13 (15%) of ERMS demonstrated strong, diffuse staining similar to that in GIST. Consistent with these data, imatinib mesylate has not demonstrated significant antitumor activity against RMS models in€vivo (unpublished results) or in the treatment of childhood solid tumors (Bond et€al. 2008).
Differentiation As discussed previously, RMS is characterized by arrested myogenic differentiation. It has been a goal of many studies to understand this process in light of molecular events that characterize normal myogenesis, and to devise approaches to induce terminal differentiation in these tumors. Failure to differentiate may be a consequence of activating mutations in oncogenes such as RAS, FOS, and SRC (Alema and Tato 1994), or epigenic changes in gene promoter methylation and silencing, as for PAX3 in ERMS (Kurmasheva et€ al. 2005), or silencing through altered histone acetylation, as for the CDKN1A gene that encodes the cyclin-cdk inhibitor p21CIP1 (Jaboin et€al. 2002; Kutko et€al. 2003). Approaches to induce differentiation include the use of DNA methyltransferase inhibitors, such as 5-azacytidine (Lollini et€ al. 1989), and the use of histone deacetylase inhibitors (Kutko et€ al. 2003) that induce reexpression of p21Cip1. Numerous clinical trials have been carried out with demethylating agents, but unfortunately these trials highlight the need to understand better the complete pathways leading to LOI, since global demethylation, resulting from these treatments generated additional complications (Issa et€al. 2005; Yoo and Jones 2006). However, as molecular detail of myogenic differentiation has emerged, it has been possible to determine why differentiation in RMS is attenuated. Many hypotheses have been proposed to explain how mitogenic signals suppress muscle differentiation. Phosphorylation of myogenic regulatory proteins (Li et€al. 1992), induction of dominant helix-loop-helix (HLH) proteins that block E protein heterodimerization (Benezra et€ al. 1990), cyclin D-CDK4-mediated
Molecular Therapy for Rhabdomyosarcoma
443
inhibition of the differentiation program (Rao et€al. 1994; Rao and Kohtz 1995; Skapek et€ al. 1995, 1996), and MEK1-mediated inactivation of the transactivation domain of the myogenic bHLH proteins (Perry et€al. 2001) have been implicated in arrested differentiation. Cyclin D-CDK4 complexes associate with the C-terminus of MYOD attenuating its interaction with DNA (Zhang et€al. 1999), can indirectly inhibit the activity of myogenin (Skapek et€al. 1996), and suppress the activity of RB (Novitch et€al. 1996) that is required to promote later stages of skeletal muscle differentiation. Cyclin D-CDK4 activity also blocks the association of MEF2C with the coactivator protein GRIP1, thereby inhibiting MEF2 activity (Lazaro et€ al. 2002). Expression of PAX3-FKHR also increases expression of the transcription factor SIX1 (Khan et€al. 1999) that in addition to activating the myogenic program, activates a set of protumorigenic genes, including those encoding cyclin D1, c-MYC, and Ezrin (Yu et€al. 2006). As cyclin D-CDK complexes are upregulated in both ERMS and ARMS (Knudsen et€al. 1998; Saab et€al. 2006), drive proliferation and suppress differentiation, targeting these complexes appears to be a valid therapeutic strategy. Relatively selective small molecule inhibitors of cyclin D-CDK4/6 have entered clinical trials for the treatment of adults with cancer (Fry et€al. 2004; DePinto et€al. 2006). PD0331992 is a potent inhibitor of cyclin-dependent kinases 4 and 6 and induced regressions of Colo-205 human colon tumor xenografts, eliminating phospho-RB and the proliferative marker Ki67 (Fry et€al. 2004). However, the effects of PD332991 against RMS cell lines were less robust. Although it induced G1 phase accumulation in most RMS cell lines, and some degree of myogenic differentiation, it did not promote terminal differentiation. In part, this may be a consequence of incomplete cyclin D-CDK inhibition, as residual phosphorylation of RB (S780) was detected in treated cells. PD331992 induced some growth inhibition of RMS xenografts in€vivo, although the effects were not particularly impressive (Saab et€ al. 2006). These studies support the contention that elevated CDK2/CDK4 contribute to the inability of RMS cells to growth arrest when cultured in low-mitogen medium, but that these activities alone are not sufficient to explain the failure of cells to differentiate (Knudsen et€ al. 1998). As DNp73 is expressed in many RMS and suppresses the myogenic program, it suggests additional approaches that will be required to develop effective therapies designed to induce terminal differentiation.
Angiogenesis Neovascularization of solid tumors is one component of the solid tumor microenvironment that is essential for sustenance and metastasis. Numerous approaches to decreasing tumor angiogenesis or targeting tumor vasculature are currently in preclinical or clinical trials. These include antibodies and peptides targeting the VEGF, including bevacizumab, and VEGF-Trap, small molecule inhibitors of VEGF receptors, as well as multitargeted tyrosine kinase inhibitors that inhibit both growth factor receptors and pathways downstream of receptor signaling. RMS cell
444
R.T. Kurmasheva et al.
lines secrete VEGF (Gee et€ al. 2005; Kurmasheva et€ al. 2007) as well as other angiogenic factors such as basic fibroblastic growth factor (bFGF) and interleukin 8 (Pavlakovic et€al. 2001) as well as other potential angiogenic factors (De Giovanni et€al. 1995). In most RMS cell lines, VEGF stimulates proliferation or activates the PI3K/Akt pathway (Gee et€al. 2005; Kurmasheva et€al. 2007), hence acts both as an autocrine growth factor as well as a paracrine factor involved in angiogenesis. The activity of bFGF may also be enhanced by glypican-5, a gene amplified and overexpressed in RMS (Williamson et€al. 2007). RMS cells also secrete predominantly IGF-2, which in turn signals to activate HIF1a, and upregulation of VEGF (Beckert et€al. 2006). Thus, the use of agents that blocks either secretion of VEGF or its interaction with receptors on vascular endothelial cells to inhibit RMS-driven angiogenesis appears to offer therapeutic potential. VEGF has emerged as the key stimulatory molecule for promoting angiogenesis in a variety of human malignancies (Ferrara 2002). This was first translated into successful application of an antiangiogenic strategy when the anti-VEGF monoclonal antibody bevacizumab showed a survival benefit for patients with metastatic colorectal cancer when combined with fluorouracil-based combination chemotherapy (Hurwitz et€ al. 2005). Clearly, antibodies that bind and neutralize VEGF have already defined a role in the treatment of adult cancer, and are in clinical evaluation for the treatment of pediatric cancer. In preclinical studies, a human VEGF-specific antibody inhibited the growth of numerous tumors including A673 rhabdomyosarcoma (actually a Ewing sarcoma cell line, Barlow et€al. 2006), although inhibition of both tumorderived (human) as well as mouse derived VEGF was required for optimal inhibition of tumor growth (Gerber et€ al. 2000). Thus, in a xenograft model both circulating mouse-derived VEGF as well as tumor-derived factor contributes to tumor progression. Small molecule inhibitors of the VEGF receptors such as AZD2171 are in adult and pediatric clinical trials. AZD2171 can not only inhibit VEGFR1, VEGFR2, and VEGFR3 kinase activity at low nanomolar concentrations but it also inhibits c-KIT and platelet derived growth factor receptor (PDGFR) (a and b) at similar concentrations (Wedge et€al. 2005). In preclinical testing, this agent has shown some utility against RMS xenografts, significantly retarding growth in each of five RMS models (Maris et€al. 2008a); however, all RMS xenografts demonstrated progressive disease. Other multitargeted kinase inhibitors such as sunitinib (VEGFR1-3, PDGF-a and -b, FLT-3, c-KIT, CSR-1) also inhibit angiogenesis, and inhibit the growth of RMS xenografts to a similar extent as AZD2171 (Maris et€al. 2008b). An alternative approach is to inhibit production of VEGF by tumor cells. Rapamycin, a selective inhibitor of mTOR downstream of PI3K/Akt regulates VEGF transcription and translation leading to decreased tumor angiogenesis. The antiangiogenic activities of rapamycin have been linked both to a decrease in VEGF production and to inhibited the response of vascular endothelial cells to stimulation by VEGF (Guba et€al. 2002, 2004, 2005). Rapamycin also targets vascular mesenchymal cells (pericytes, smooth muscle cells, adventitial fibroblasts) inhibiting sustained VEGF and hepatocyte growth factor (HGF) expression, via inhibition of the PDGFRa-p70S6K pathway (Tsutsumi et€al. 2004). Data that support a role for
Molecular Therapy for Rhabdomyosarcoma
445
mTOR signaling in VEGF production include regulation of HIF-1a by mTOR signaling and increased VEGF in cells deficient in the tuberous sclerosis complex (TSC1/2) that negatively regulates mTOR via Rheb (Hudson et€al. 2002; Brugarolas et€al. 2003, 2004). In growth factor-deprived neuroblastoma cells in€vitro, serum or IGF-1 induced increases in HIF-1a protein that temporally paralleled increases in VEGF mRNA. VEGF and HIF-1a levels were blocked by inhibitors of PI3K and mTOR, and to a lesser extent by the MEK1 inhibitor PD98059. In these cells, it was estimated that HIF-1a mediates approximately 40% of the growth factor activity stimulating VEGF protein expression (Beppu et€ al. 2005). Further, inhibition of signaling via the type 1 insulin-like growth factor receptor (IGF-1R) resulted in decreased VEGF in€vitro, and in€vivo and decreased tumor volume and weight, vessel density, tumor cell proliferation and increased tumor cell apoptosis (Stoeltzing et€ al. 2003). Rapamycin analogs (temsirolimus, everolimus) have recently completed phase 1 trials in children, and phase 2 trials are ongoing. In preclinical models, rapamycin significantly inhibited the growth of most solid tumors and induced regression of sarcomas including RMS and acute lymphocytic leukemia models (Houghton et€al. 2008). Although mTOR signaling is implicated in the regulation of VEGF, there are data to suggest that VEGF may be regulated through PI3K/Akt signaling independent of mTOR (Arsham et€al. 2002). Further, inhibition of mTOR in many cancer cell lines results in the anecdotal activation of Akt (Shi et€al. 2005; O’Reilly et€ al. 2006), as a consequence of inhibiting p70S6K1 phosphorylation, thus suppressing S6K1-induced phosphorylation of IRS-1 that leads to its proteasomal degradation (Haruta et€al. 2000). In vitro an Akt inhibitor (A443654) was significantly more effective in suppressing RMS-derived VEGF than rapamycin, but the combination of these agents essentially blocked increases in VEGF induced by hypoxia (Kurmasheva et€al. 2007). Whether inhibitors of Akt can be combined with rapamycin with acceptable toxicity remains to be determined. However, it is probable that to adequately suppress tumor angiogenesis, a combination of approaches will be required.
Future Directions and Challenges Childhood RMS is a genetically complex disease, with mutations and activation of multiple genes and signaling pathways. Consequently, it seems unlikely that inhibition of a single aberrant gene product, as in BCR-ABL “driven” CML will have such a dramatic effect on RMS proliferation and survival. However, novel targets, including MET, IGF-1R, VEGF, and indirectly MYCN, are now amenable to pharmacological intervention. Perhaps with the exception of MYCN inhibition, the effects of inhibiting the other signaling pathways would be anticipated to be largely cytostatic rather than cytotoxic. While the next step will be integrating these agents with conventional cytotoxic therapy, the real challenge will be in developing combinations of molecularly targeted agents that induce selective tumor cell death based on genetic changes that lead to tumor-specific dysregulation of signaling
446
R.T. Kurmasheva et al.
pathways. Such synthetic lethal interactions, first described in yeast systems, can be identified in mammalian cells by a variety of new techniques. Foremost would be whole genome approaches using RNAi to identify genes that encode products required for survival only in the context of the oncogenic change (PAX3-FKHR expression, or MYCN overexpression). Development of drugs that inhibit these gene products would potentially be tumor-selective. Pharmacologic inhibition of CDK1 resulting in death only of cells overexpressing MYC serves as a paradigm. Alternatively, identifying genes essential for survival in the presence of inhibitors of specific signaling pathways only in the context of a transforming gene may offer the opportunity to develop novel agents that in combination induce cell death selectively in RMS cells. Inhibition of IGF-1R signaling in combination with inhibition of mTOR in cells with attenuated p53 function, for example, results in shifting the cellular response to rapamycin from cytostasis to cell death, and shows quite promising in€vivo preclinical results, particularly in sarcoma models (Kurmasheva et al. 2009). Pediatric trials of IGF-1R-directed antibodies alone and in combination with rapamycin are in the planning stage, and many of the agents discussed in this chapter are being evaluated against preclinical models of RMS. It is anticipated that many of these agents will enter clinical trials for the treatment of childhood RMS in the next few years. Acknowledgments╅ Original work reported here was supported by PHS awards CA23099, CA96696, CA77776, and CA21675 (Cancer Center Support Grant), and by ALSAC.
References Alema, S., and Tato, F. 1994. Oncogenes and muscle differentiation: multiple mechanisms of interference. Semin Cancer Biol 5: 147–156. Ambrosini, G., Sambol, E.B., Carvajal, D., Vassilev, L.T., Singer, S., and Schwartz, G.K. 2007. Mouse double minute antagonist Nutlin-3a enhances chemotherapy-induced apoptosis in cancer cells with mutant p53 by activating E2F1. Oncogene 26: 3473–3481. Anderson, J., Gordon, A., McManus, A., Shipley, J., and Pritchard-Jones, K. 1999. Disruption of imprinted genes at chromosome region 11p15.5 in paediatric rhabdomyosarcoma. Neoplasia 1: 340–348. Arndt, C.A., and Crist, W.M. 1999. Common musculoskeletal tumors of childhood and adolescence. N Engl J Med 341: 342–352. Arsham, A.M., Plas, D.R., Thompson, C.B., and Simon, M.C. 2002. Phosphatidylinositol 3-kinase/Akt signaling is neither required for hypoxic stabilization of HIF-1 alpha nor sufficient for HIF-1-dependent target gene transcription. J Biol Chem 277: 15162–15170. Asakura, A., Seale, P., Girgis-Gabardo, A., and Rudnicki, M.A. 2002. Myogenic specification of side population cells in skeletal muscle. J Cell Biol 159: 123–134. Ayalon, D., Glaser, T., and Werner, H. 2001. Transcriptional regulation of IGF-I receptor gene expression by the PAX3-FKHR oncoprotein. Growth Horm IGF Res 11: 289–297. Bach, L.A., Hsieh, S., Brown, A.L., and Rechler, M.M. 1994. Recombinant human insulin-like growth factor (IGF)-binding protein-6 inhibits IGF-II-induced differentiation of L6A1 myoblasts. Endocrinology 135: 2168–2176. Bach, L.A., Salemi, R., and Leeding, K.S. 1995. Roles of insulin-like growth factor (IGF) receptors and IGF-binding proteins in IGF-II-induced proliferation and differentiation of L6A1 rat myoblasts. Endocrinology 136: 5061–5069.
Molecular Therapy for Rhabdomyosarcoma
447
Bardos, J.I., Chau, N.M., and Ashcroft, M. 2004. Growth factor-mediated induction of HDM2 positively regulates hypoxia-inducible factor 1alpha expression. Mol Cell Biol 24: 2905–2914. Barlow, J.W., Wiley, J.C., Mous, M., Narendran, A., Gee, M.F., Goldberg, M., Sexsmith, E., and Malkin, D. 2006. Differentiation of rhabdomyosarcoma cell lines using retinoic acid. Pediatr Blood Cancer 47: 773–784. Baserga, R. 2004. Targeting the IGF-1 receptor: from rags to riches. Eur J Cancer 40: 2013–2015. Beckert, S., Farrahi, F., Perveen Ghani, Q., Aslam, R., Scheuenstuhl, H., Coerper, S., Konigsrainer, A., Hunt, T.K., and Hussain, M.Z. 2006. IGF-I-induced VEGF expression in HUVEC involves phosphorylation and inhibition of poly(ADP-ribose)polymerase. Biochem Biophys Res Commun 341: 67–72. Begum, S., Emani, N., Cheung, A., Wilkins, O., Der, S., and Hamel, P.A. 2005. Cell-type-specific regulation of distinct sets of gene targets by Pax3 and Pax3/FKHR. Oncogene 24: 1860–1872. Benezra, R., Davis, R.L., Lockshon, D., Turner, D.L., and Weintraub, H. 1990. The protein Id: a negative regulator of helix-loop-helix DNA binding proteins. Cell 61: 49–59. Beppu, K., Nakamura, K., Linehan, W.M., Rapisarda, A., and Thiele, C.J. 2005. Topotecan blocks hypoxia-inducible factor-1alpha and vascular endothelial growth factor expression induced by insulin-like growth factor-I in neuroblastoma cells. Cancer Res 65: 4775–4781. Berg, T., Cohen, S.B., Desharnais, J., Sonderegger, C., Maslyar, D.J., Goldberg, J., Boger, D.L., and Vogt, P.K. 2002. Small-molecule antagonists of Myc/Max dimerization inhibit Myc-induced transformation of chicken embryo fibroblasts. Proc Natl Acad Sci USA 99: 3830–3835. Bernasconi, M., Remppis, A., Fredericks, W.J., Rauscher, F.J., 3rd, and Schafer, B.W. 1996. Induction of apoptosis in rhabdomyosarcoma cells through down-regulation of PAX proteins. Proc Natl Acad Sci USA 93: 13164–13169. Bohula, E.A., Playford, M.P., and Macaulay, V.M. 2003. Targeting the type 1 insulin-like growth factor receptor as anti-cancer treatment. Anticancer Drugs 14: 669–682. Bond, M., Bernstein, M.L., Pappo, A., Schultz, K.R., Krailo, M., Blaney, S.M., and Adamson, P.C. 2008. A phase II study of imatinib mesylate in children with refractory or relapsed solid tumors: a Children’s Oncology Group study. Pediatr Blood Cancer 50:254–258. Bottger, A., Bottger, V., Garcia-Echeverria, C., Chene, P., Hochkeppel, H.K., Sampson, W., Ang, K., Howard, S.F., Picksley, S.M., and Lane, D.P. 1997. Molecular characterization of the hdm2-p53 interaction. J Mol Biol 269: 744–756. Braczkowski, R., Schally, A.V., Plonowski, A., Varga, J.L., Groot, K., Krupa, M., and Armatis, P. 2002. Inhibition of proliferation in human MNNG/HOS osteosarcoma and SK-ES-1 Ewing sarcoma cell lines in€vitro and in€vivo by antagonists of growth hormone-releasing hormone: effects on insulin-like growth factor II. Cancer 95: 1735–1745. Breneman, J.C., Lyden, E., Pappo, A.S., Link, M.P., Anderson, J.R., Parham, D.M., Qualman, S.J., Wharam, M.D., Donaldson, S.S., Maurer, H.M., et€ al. 2003. Prognostic factors and clinical outcomes in children and adolescents with metastatic rhabdomyosarcoma – a report from the Intergroup Rhabdomyosarcoma Study IV. J Clin Oncol 21: 78–84. Bridge, J.A., Liu, J., Weibolt, V., Baker, K.S., Perry, D., Kruger, R., Qualman, S., Barr, F., Sorensen, P., Triche, T., et€al. 2000. Novel genomic imbalances in embryonal rhabdomyosarcoma revealed by comparative genomic hybridization and fluorescence in situ hybridization: an intergroup rhabdomyosarcoma study. Genes Chromosomes Cancer 27: 337–344. Bridge, J.A., Liu, J., Qualman, S.J., Suijkerbuijk, R., Wenger, G., Zhang, J., Wan, X., Baker, K.S., Sorensen, P., and Barr, F.G. 2002. Genomic gains and losses are similar in genetic and histologic subsets of rhabdomyosarcoma, whereas amplification predominates in embryonal with anaplasia and alveolar subtypes. Genes Chromosomes Cancer 33: 310–321. Brugarolas, J.B., Vazquez, F., Reddy, A., Sellers, W.R., and Kaelin, W.G., Jr. 2003. TSC2 regulates VEGF through mTOR-dependent and -independent pathways. Cancer Cell 4: 147–158. Brugarolas, J., Lei, K., Hurley, R.L., Manning, B.D., Reiling, J.H., Hafen, E., Witters, L.A., Ellisen, L.W., and Kaelin, W.G., Jr. 2004. Regulation of mTOR function in response to hypoxia by REDD1 and the TSC1/TSC2 tumor suppressor complex. Genes Dev 18: 2893–2904. Burkhart, C.A., Cheng, A.J., Madafiglio, J., Kavallaris, M., Mili, M., Marshall, G.M., Weiss, gonucleotide administration on tumorigenesis in a murine model of neuroblastoma. J Natl Cancer Inst 95: 1394–1403.
448
R.T. Kurmasheva et al.
Calzada-Wack, J., Schnitzbauer, U., Walch, A., Wurster, K.H., Kappler, R., Nathrath, M., and Hahn, H. 2002. Analysis of the PTCH coding region in human rhabdomyosarcoma. Hum Mutat 20: 233–234. Cam, H., Griesmann, H., Beitzinger, M., Hofmann, L., Beinoraviciute-Kellner, R., Sauer, M., Huttinger-Kirchhof, N., Oswald, C., Friedl, P., Gattenlohner, S., et€al. 2006. p53 family members in myogenic differentiation and rhabdomyosarcoma development. Cancer Cell 10: 281–293. Casola, S., Pedone, P.V., Cavazzana, A.O., Basso, G., Luksch, R., d’Amore, E.S., Carli, M., Bruni, C.B., and Riccio, A. 1997. Expression and parental imprinting of the H19 gene in human rhabdomyosarcoma. Oncogene 14: 1503–1510. Chen, J.K., Taipale, J., Young, K.E., Maiti, T., and Beachy, P.A. 2002. Small molecule modulation of Smoothened activity. Proc Natl Acad Sci USA 99: 14071–14076. Chen, Y., Takita, J., Hiwatari, M., Igarashi, T., Hanada, R., Kikuchi, A., Hongo, T., Taki, T., Ogasawara, M., Shimada, A., et€al. 2006. Mutations of the PTPN11 and RAS genes in rhabdomyosarcoma and pediatric hematological malignancies. Genes Chromosomes Cancer 45: 583–591. Chen, Y., Takita, J., Mizuguchi, M., Tanaka, K., Ida, K., Koh, K., Igarashi, T., Hanada, R., Tanaka, Y., Park, M.J., et€al. 2007. Mutation and expression analyses of the MET and CDKN2A genes in rhabdomyosarcoma with emphasis on MET overexpression. Genes Chromosomes Cancer 46: 348–358. Chesler, L., Schlieve, C., Goldenberg, D.D., Kenney, A., Kim, G., McMillan, A., Matthay, K.K., Rowitch, D., and Weiss, W.A. 2006. Inhibition of phosphatidylinositol 3-kinase destabilizes Mycn protein and blocks malignant progression in neuroblastoma. Cancer Res 66: 8139–8146. Cohen, B.D., Baker, D.A., Soderstrom, C., Tkalcevic, G., Rossi, A.M., Miller, P.E., Tengowski, M.W., Wang, F., Gualberto, A., Beebe, J.S., et€ al. 2005. Combination therapy enhances the inhibition of tumor growth with the fully human anti-type 1 insulin-like growth factor receptor monoclonal antibody CP-751,871. Clin Cancer Res 11: 2063–2073. Cornelison, D.D., and Wold, B.J. 1997. Single-cell analysis of regulatory gene expression in quiescent and activated mouse skeletal muscle satellite cells. Dev Biol 191: 270–283. Davicioni, E., Finckenstein, F.G., Shahbazian, V., Buckley, J.D., Triche, T.J., and Anderson, M.J. 2006. Identification of a PAX-FKHR gene expression signature that defines molecular classes and determines the prognosis of alveolar rhabdomyosarcomas. Cancer Res 66: 6936–6946. Davis, R.J., D’Cruz, C.M., Lovell, M.A., Biegel, J.A., and Barr, F.G. 1994. Fusion of PAX7 to FKHR by the variant t(1;13)(p36;q14) translocation in alveolar rhabdomyosarcoma. Cancer Res 54: 2869–2872. De Giovanni, C., Melani, C., Nanni, P., Landuzzi, L., Nicoletti, G., Frabetti, F., Griffoni, C., Colombo, M.P., and Lollini, P.L. 1995. Redundancy of autocrine loops in human rhabdomyosarcoma cells: induction of differentiation by suramin. Br J Cancer 72: 1224–1229. De Pitta, C., Tombolan, L., Albiero, G., Sartori, F., Romualdi, C., Jurman, G., Carli, M., Furlanello, C., Lanfranchi, G., and Rosolen, A. 2006. Gene expression profiling identifies potential relevant genes in alveolar rhabdomyosarcoma pathogenesis and discriminates PAX3FKHR positive and negative tumors. Int J Cancer 118: 2772–2781. DePinto, W., Chu, X.J., Yin, X., Smith, M., Packman, K., Goelzer, P., Lovey, A., Chen, Y., Qian, H., Hamid, R., et€ al. 2006. In vitro and in€ vivo activity of R547: a potent and selective cyclindependent kinase inhibitor currently in phase I clinical trials. Mol Cancer Ther 5: 2644–2658. Dias, P., Kumar, P., Marsden, H.B., Gattamaneni, H.R., Heighway, J., and Kumar, S. 1990. N-myc gene is amplified in alveolar rhabdomyosarcomas (RMS) but not in embryonal RMS. Int J Cancer 45: 593–596. Doranz, B.J., Filion, L.G., Diaz-Mitoma, F., Sitar, D.S., Sahai, J., Baribaud, F., Orsini, M.J., Benovic, J.L., Cameron, W., and Doms, R.W. 2001. Safe use of the CXCR4 inhibitor ALX40-4C in humans. AIDS Res Hum Retroviruses 17: 475–486. Dorman, C.M., and Johnson, S.E. 2000. Activated Raf inhibits myogenesis through a mechanism independent of activator protein 1-mediated myoblast transformation. J Biol Chem 275: 27481–27487. Driman, D., Thorner, P.S., Greenberg, M.L., Chilton-MacNeill, S., and Squire, J. 1994. MYCN gene amplification in rhabdomyosarcoma. Cancer 73: 2231–2237. Druker, B.J. 2003. David A. Karnofsky Award lecture. Imatinib as a paradigm of targeted therapies. J Clin Oncol 21: 239s–245s.
Molecular Therapy for Rhabdomyosarcoma
449
Druker, B.J., Guilhot, F., O’Brien, S.G., Gathmann, I., Kantarjian, H., Gattermann, N., Deininger, M.W., Silver, R.T., Goldman, J.M., Stone, R.M., et€al. 2006. Five-year follow-up of patients receiving imatinib for chronic myeloid leukemia. N Engl J Med 355: 2408–2417. Easton, J.B., Kurmasheva, R.T., and Houghton, P.J. 2006. IRS-1: auditing the effectiveness of mTOR inhibitors. Cancer Cell 9: 153–155. Ebauer, M., Wachtel, M., Niggli, F.K., and Schafer, B.W. 2007. Comparative expression profiling identifies an in€vivo target gene signature with TFAP2B as a mediator of the survival function of PAX3/FKHR. Oncogene 26: 7267–7281. Eder, J.P., Heath, E., Appleman, L., Shapiro, G., Wang, D., Malburg, L., Zhu, A.X., Leader, T., Wolanski, A., and LoRusso, P. 2007. Phase I experience with c-MET inhibitor XL880 administered orally to patients (pts) with solid tumors. ASCO Annual Meeting Proceedings Part I. J Clin Oncol 25: 3526. Efeyan, A., Ortega-Molina, A., Velasco-Miguel, S., Herranz, D., Vassilev, L.T., and Serrano, M. 2007. Induction of p53-dependent senescence by the MDM2 antagonist nutlin-3a in mouse cells of fibroblast origin. Cancer Res 67: 7350–7357. El-Badry, O.M., Minniti, C., Kohn, E.C., Houghton, P.J., Daughaday, W.H., and Helman, L.J. 1990. Insulin-like growth factor II acts as an autocrine growth and motility factor in human rhabdomyosarcoma tumors. Cell Growth Differ 1: 325–331. Engel, N., Thorvaldsen, J.L., and Bartolomei, M.S. 2006. CTCF binding sites promote transcription initiation and prevent DNA methylation on the maternal allele at the imprinted H19/Igf2 locus. Hum Mol Genet 15: 2945–2954. Epstein, J.A., Shapiro, D.N., Cheng, J., Lam, P.Y., and Maas, R.L. 1996. Pax3 modulates expression of the c-Met receptor during limb muscle development. Proc Natl Acad Sci USA 93: 4213–4218. Ezzat, S., Zheng, L., Winer, D., and Asa, S.L. 2006. Targeting N-cadherin through fibroblast growth factor receptor-4: distinct pathogenetic and therapeutic implications. Mol Endocrinol 20: 2965–2975. Feng, Y., Zhu, Z., Xiao, X., Choudhry, V., Barrett, J.C., and Dimitrov, D.S. 2006. Novel human monoclonal antibodies to insulin-like growth factor (IGF)-II that potently inhibit the IGF receptor type I signal transduction function. Mol Cancer Ther 5: 114–120. Ferracini, R., Olivero, M., Di Renzo, M.F., Martano, M., De Giovanni, C., Nanni, P., Basso, G., Scotlandi, K., Lollini, P.L., and Comoglio, P.M. 1996. Retrogenic expression of the MET proto-oncogene correlates with the invasive phenotype of human rhabdomyosarcomas. Oncogene 12: 1697–1705. Ferrara, N. 2002. VEGF and the quest for tumour angiogenesis factors. Nat Rev Cancer 2: 795–803. Flores, E.R. 2007. The roles of p63 in cancer. Cell Cycle 6: 300–304. Flores, E.R., Tsai, K.Y., Crowley, D., Sengupta, S., Yang, A., McKeon, F., and Jacks, T. 2002. p63 and p73 are required for p53-dependent apoptosis in response to DNA damage. Nature 416: 560–564. Flores, E.R., Sengupta, S., Miller, J.B., Newman, J.J., Bronson, R., Crowley, D., Yang, A., McKeon, F., and Jacks, T. 2005. Tumor predisposition in mice mutant for p63 and p73: evidence for broader tumor suppressor functions for the p53 family. Cancer Cell 7: 363–373. Flygare, J., Gustafsson, K., Kimby, E., Christensson, B., and Sander, B. 2005. Cannabinoid receptor ligands mediate growth inhibition and cell death in mantle cell lymphoma. FEBS Lett 579: 6885–6889. Fredericks, W.J., Galili, N., Mukhopadhyay, S., Rovera, G., Bennicelli, J., Barr, F.G., and Rauscher, F.J., 3rd. 1995. The PAX3-FKHR fusion protein created by the t(2;13) translocation in alveolar rhabdomyosarcomas is a more potent transcriptional activator than PAX3. Mol Cell Biol 15: 1522–1535. Fry, D.W., Harvey, P.J., Keller, P.R., Elliott, W.L., Meade, M., Trachet, E., Albassam, M., Zheng, X., Leopold, W.R., Pryer, N.K., et€al. 2004. Specific inhibition of cyclin-dependent kinase 4/6 by PD 0332991 and associated antitumor activity in human tumor xenografts. Mol Cancer Ther 3: 1427–1438. Furman, W.L., Stewart, C.F., Poquette, C.A., Pratt, C.B., Santana, V.M., Zamboni, W.C., Bowman, L.C., Ma, M.K., Hoffer, F.A., Meyer, W.H., et€al. 1999. Direct translation of a protracted irinotecan schedule from a xenograft model to a phase I trial in children. J Clin Oncol 17: 1815–1824.
450
R.T. Kurmasheva et al.
Gable, K.L., Maddux, B.A., Penaranda, C., Zavodovskaya, M., Campbell, M.J., Lobo, M., Robinson, L., Schow, S., Kerner, J.A., Goldfine, I.D., et€al. 2006. Diarylureas are small-molecule inhibitors of insulin-like growth factor I receptor signaling and breast cancer cell growth. Mol Cancer Ther 5: 1079–1086. Galili, N., Davis, R.J., Fredericks, W.J., Mukhopadhyay, S., Rauscher, F.J., 3rd, Emanuel, B.S., Rovera, G., and Barr, F.G. 1993. Fusion of a fork head domain gene to PAX3 in the solid tumour alveolar rhabdomyosarcoma. Nat Genet 5: 230–235. Gallicchio, M.A., Kneen, M., Hall, C., Scott, A.M., and Bach, L.A. 2001. Overexpression of insulin-like growth factor binding protein-6 inhibits rhabdomyosarcoma growth in€vivo. Int J Cancer 94: 645–651. Ganjavi, H., Gee, M., Narendran, A., Freedman, M.H., and Malkin, D. 2005. Adenovirusmediated p53 gene therapy in pediatric soft-tissue sarcoma cell lines: sensitization to cisplatin and doxorubicin. Cancer Gene Ther 12: 397–406. Garcia, A., Rosen, L., Cunningham, C.C., Nemunaitis, J., Li, C., Rulewski, N., Dovholuk, A., Savage, R., Chan, T., Bukowksi, R., and Mekhail T. 2007. Phase 1 study of ARQ 197, a selective inhibitor of the c-Met RTK in patients with metastatic solid tumors reaches recommended phase 2 dose. ASCO Annual Meeting Proceedings Part I. J Clin Oncol 25: 3525. Gee, M.F., Tsuchida, R., Eichler-Jonsson, C., Das, B., Baruchel, S., and Malkin, D. 2005. Vascular endothelial growth factor acts in an autocrine manner in rhabdomyosarcoma cell lines and can be inhibited with all-trans-retinoic acid. Oncogene 24: 8025–8037. Gerber, H.P., Kowalski, J., Sherman, D., Eberhard, D.A., and Ferrara, N. 2000. Complete inhibition of rhabdomyosarcoma xenograft growth and neovascularization requires blockade of both tumor and host vascular endothelial growth factor. Cancer Res 60: 6253–6258. Gerber, A.N., Wilson, C.W., Li, Y.J., and Chuang, P.T. 2007. The hedgehog regulated oncogenes Gli1 and Gli2 block myoblast differentiation by inhibiting MyoD-mediated transcriptional activation. Oncogene 26: 1122–1136. Ginsberg, J.P., Davis, R.J., Bennicelli, J.L., Nauta, L.E., and Barr, F.G. 1998. Up-regulation of MET but not neural cell adhesion molecule expression by the PAX3-FKHR fusion protein in alveolar rhabdomyosarcoma. Cancer Res 58: 3542–3546. Goga, A., Yamng, D., Tward, A.D., Morgan, D.O., and Bishop, J.M. 2007. Inhibition of CDK1 as a potential therapy for tumors over-expressing MYC. Nature Med 13: 820–827. Goldstein, M., Meller, I., Issakov, J., and Orr-Urtreger, A. 2006. Novel genes implicated in embryonal, alveolar, and pleomorphic rhabdomyosarcoma: a cytogenetic and molecular analysis of primary tumors. Neoplasia 8: 332–343. Gordon, A.T., Brinkschmidt, C., Anderson, J., Coleman, N., Dockhorn-Dworniczak, B., PritchardJones, K., and Shipley, J. 2000. A novel and consistent amplicon at 13q31 associated with alveolar rhabdomyosarcoma. Genes Chromosomes Cancer 28: 220–226. Gressner, O., Schilling, T., Lorenz, K., Schulze Schleithoff, E., Koch, A., Schulze-Bergkamen, H., Lena, A.M., Candi, E., Terrinoni, A., Catani, M.V., et€ al. 2005. TAp63alpha induces apoptosis by activating signaling via death receptors and mitochondria. EMBO J 24: 2458–2471. Grob, T.J., Novak, U., Maisse, C., Barcaroli, D., Luthi, A.U., Pirnia, F., Hugli, B., Graber, H.U., De Laurenzi, V., Fey, M.F., et€ al. 2001. Human delta Np73 regulates a dominant negative feedback loop for TAp73 and p53. Cell Death Differ 8: 1213–1223. Guba, M., von Breitenbuch, P., Steinbauer, M., Koehl, G., Flegel, S., Hornung, M., Bruns, C.J., Zuelke, C., Farkas, S., Anthuber, M., et€al. 2002. Rapamycin inhibits primary and metastatic tumor growth by antiangiogenesis: involvement of vascular endothelial growth factor. Nat Med 8: 128–135. Guba, M., Graeb, C., Jauch, K.W., and Geissler, E.K. 2004. Pro- and anti-cancer effects of immunosuppressive agents used in organ transplantation. Transplantation 77: 1777–1782. Guba, M., Koehl, G.E., Neppl, E., Doenecke, A., Steinbauer, M., Schlitt, H.J., Jauch, K.W., and Geissler, E.K. 2005. Dosing of rapamycin is critical to achieve an optimal antiangiogenic effect against cancer. Transpl Int 18: 89–94. Guerreiro, A.S., Boller, D., Doepfner, K.T., and Arcaro, A. 2006. IGF-IR: potential role in antitumor agents. Drug News Perspect 19: 261–272.
Molecular Therapy for Rhabdomyosarcoma
451
Hachitanda, Y., Toyoshima, S., Akazawa, K., and Tsuneyoshi, M. 1998. N-myc gene amplification in rhabdomyosarcoma detected by fluorescence in situ hybridization: its correlation with histologic features. Mod Pathol 11: 1222–1227. Hahn, H., Wojnowski, L., Specht, K., Kappler, R., Calzada-Wack, J., Potter, D., Zimmer, A., Muller, U., Samson, E., Quintanilla-Martinez, L., et€al. 2000. Patched target Igf2 is indispensable for the formation of medulloblastoma and rhabdomyosarcoma. J Biol Chem 275: 28341–28344. Haluska, P., Carboni, J.M., Loegering, D.A., Lee, F.Y., Wittman, M., Saulnier, M.G., Frennesson, D.B., Kalli, K.R., Conover, C.A., Attar, R.M., et€al. 2006. In vitro and in€vivo antitumor effects of the dual insulin-like growth factor-I/insulin receptor inhibitor, BMS-554417. Cancer Res 66: 362–371. Hao, Y., Crenshaw, T., Moulton, T., Newcomb, E., and Tycko, B. 1993. Tumour-suppressor activity of H19 RNA. Nature 365: 764–767. Haruta, T., Uno, T., Kawahara, J., Takano, A., Egawa, K., Sharma, P.M., Olefsky, J.M., and Kobayashi, M. 2000. A rapamycin-sensitive pathway down-regulates insulin signaling via phosphorylation and proteasomal degradation of insulin receptor substrate-1. Mol Endocrinol 14: 783–794. Hernandez-Sanchez, C., Werner, H., Roberts, C.T., Jr., Woo, E.J., Hum, D.W., Rosenthal, S.M., and LeRoith, D. 1997. Differential regulation of insulin-like growth factor-I (IGF-I) receptor gene expression by IGF-I and basic fibroblastic growth factor. J Biol Chem 272: 4663–4670. Hoar, K., Chakravarty, A., Rabino, C., Wysong, D., Bowman, D., Roy, N., and Ecsedy, J.A. 2007. MLN8054, a small-molecule inhibitor of Aurora A, causes spindle pole and chromosome congression defects leading to aneuploidy. Mol Cell Biol 27: 4513–4525. Honda, R., Tanaka, H., and Yasuda, H. 1997. Oncoprotein MDM2 is a ubiquitin ligase E3 for tumor suppressor p53. FEBS Lett 420: 25–27. Hope, K.J., Jin, L., and Dick, J.E. 2003. Human acute myeloid leukemia stem cells. Arch Med Res 34: 507–514. Hope, K.J., Jin, L., and Dick, J.E. 2004. Acute myeloid leukemia originates from a hierarchy of leukemic stem cell classes that differ in self-renewal capacity. Nat Immunol 5: 738–743. Houghton, P.J., Morton, C.L., Tucker, C., Payne, D., Favours, E., Cole, C., Gorlick, R., Kolb, E.A., Zhang, W., Lock, R., et€ al. 2007. The pediatric preclinical testing program: description of models and early testing results. Pediatr Blood Cancer 49:928–940. Houghton, P.J., Morton, C.L., Kolb, E.A., Gorlick, R., Lock, R., Carol, H., Reynolds, C.P., Maris, J.M., Keir, S.T., Billups, C.A., et€ al. 2008. Initial testing (stage 1) of the mTOR inhibitor rapamycin by the pediatric preclinical testing program. Pediatr Blood Cancer 50: 799–805. Huang, S., Shu, L., Dilling, M.B., Easton, J., Harwood, F.C., Ichijo, H., and Houghton, P.J. 2003. Sustained activation of the JNK cascade and rapamycin-induced apoptosis are suppressed by p53/p21(Cip1). Mol Cell 11: 1491–1501. Hudson, C.C., Liu, M., Chiang, G.G., Otterness, D.M., Loomis, D.C., Kaper, F., Giaccia, A.J., and Abraham, R.T. 2002. Regulation of hypoxia-inducible factor 1alpha expression and function by the mammalian target of rapamycin. Mol Cell Biol 22: 7004–7014. Hurwitz, H.I., Fehrenbacher, L., Hainsworth, J.D., Heim, W., Berlin, J., Holmgren, E., Hambleton, J., Novotny, W.F., and Kabbinavar, F. 2005. Bevacizumab in combination with fluorouracil and leucovorin: an active regimen for first-line metastatic colorectal cancer. J Clin Oncol 23: 3502–3508. Huttinger-Kirchhof, N., Cam, H., Griesmann, H., Hofmann, L., Beitzinger, M., and Stiewe, T. 2006. The p53 family inhibitor DeltaNp73 interferes with multiple developmental programs. Cell Death Differ 13: 174–177. Hwa, V., Oh, Y., and Rosenfeld, R.G. 1999. Insulin-like growth factor binding proteins: a proposed superfamily. Acta Paediatr Suppl 88: 37–45. Issa, J.P., Kantarjian, H.M., and Kirkpatrick, P. 2005. Azacitidine. Nat Rev Drug Discov 4: 275–276. Jaboin, J., Wild, J., Hamidi, H., Khanna, C., Kim, C.J., Robey, R., Bates, S.E., and Thiele, C.J. 2002. MS-27-275, an inhibitor of histone deacetylase, has marked in€vitro and in€vivo antitumor activity against pediatric solid tumors. Cancer Res 62: 6108–6115.
452
R.T. Kurmasheva et al.
Jankowski, K., Kucia, M., Wysoczynski, M., Reca, R., Zhao, D., Trzyna, E., Trent, J., Peiper, S., Zembala, M., Ratajczak, J., et€al. 2003. Both hepatocyte growth factor (HGF) and stromalderived factor-1 regulate the metastatic behavior of human rhabdomyosarcoma cells, but only HGF enhances their resistance to radiochemotherapy. Cancer Res 63: 7926–7935. Kalebic, T., Tsokos, M., and Helman, L.J. 1994. In vivo treatment with antibody against IGF-1 receptor suppresses growth of human rhabdomyosarcoma and down-regulates p34cdc2. Cancer Res 54: 5531–5534. Kaleko, M., Rutter, W.J., and Miller, A.D. 1990. Overexpression of the human insulinlike growth factor I receptor promotes ligand-dependent neoplastic transformation. Mol Cell Biol 10: 464–473. Kelly, P.N., Dakic, A., Adams, J.M., Nutt, S.L., and Strasser, A. 2007. Tumor growth need not be driven by rare cancer stem cells. Science 317: 337. Khan, J., Bittner, M.L., Saal, L.H., Teichmann, U., Azorsa, D.O., Gooden, G.C., Pavan, W.J., Trent, J.M., and Meltzer, P.S. 1999. cDNA microarrays detect activation of a myogenic transcription program by the PAX3-FKHR fusion oncogene. Proc Natl Acad Sci USA 96: 13264–13269. Knudsen, E.S., Pazzagli, C., Born, T.L., Bertolaet, B.L., Knudsen, K.E., Arden, K.C., Henry, R.R., and Feramisco, J.R. 1998. Elevated cyclins and cyclin-dependent kinase activity in the rhabdomyosarcoma cell line RD. Cancer Res 58: 2042–2049. Kolb, A.E., Gorlick, R., Houghton, P.J., Morton, C.L., Lock, R., Carol, H., Reynolds, C.P., Maris, J.M., Keir, S.T., Billups, C.A., et€al. 2008. Initial testing (stage 1) of a monoclonal antibody (SCH 717454) against the IGF-1 receptor by the pediatric preclinical testing program. Pediatr Blood Cancer 50: 1190–1197. Komdeur, R., Hoekstra, H.J., Molenaar, W.M., Van Den Berg, E., Zwart, N., Pras, E., PlazaMenacho, I., Hofstra, R.M., and Van Der Graaf, W.T. 2003. Clinicopathologic assessment of postradiation sarcomas: KIT as a potential treatment target. Clin Cancer Res 9: 2926–2932. Kranz, D., and Dobbelstein, M. 2006. Nongenotoxic p53 activation protects cells against S-phasespecific chemotherapy. Cancer Res 66: 10274–10280. Kurmasheva, R.T., Dudkin, L., Billups, C., Debelenko, L.V., Morton, C.L., Houghton, P.J. 2009. The insulin-like growth factor-1 receptor-targeting antibody, CP-751,871, suppresses tumorderived VEGF and synergizes with rapamycin in models of childhood sarcoma. Cancer Res 69: 7662–71. Kurmasheva, R.T., and Houghton, P.J. 2006. IGF-I mediated survival pathways in normal and malignant cells. Biochim Biophys Acta 1766: 1–22. Kurmasheva, R.T., Peterson, C.A., Parham, D.M., Chen, B., McDonald, R.E., and Cooney, C.A. 2005. Upstream CpG island methylation of the PAX3 gene in human rhabdomyosarcomas. Pediatr Blood Cancer 44: 328–337. Kurmasheva, R.T., Harwood, F.C., and Houghton, P.J. 2007. Differential regulation of vascular endothelial growth factor by Akt and mammalian target of rapamycin inhibitors in cell lines derived from childhood solid tumors. Mol Cancer Ther 6: 1620–1628. Kutko, M.C., Glick, R.D., Butler, L.M., Coffey, D.C., Rifkind, R.A., Marks, P.A., Richon, V.M., and LaQuaglia, M.P. 2003. Histone deacetylase inhibitors induce growth suppression and cell death in human rhabdomyosarcoma in€vitro. Clin Cancer Res 9: 5749–5755. Lae, M., Ahn, E.H., Mercado, G.E., Chuai, S., Edgar, M., Pawel, B.R., Olshen, A., Barr, F.G., and Ladanyi, M. 2007. Global gene expression profiling of PAX-FKHR fusion-positive alveolar and PAX-FKHR fusion-negative embryonal rhabdomyosarcomas. J Pathol 212: 143–151. Langenau, D.M., Keefe, M.D., Storer, N.Y., Guyon, J.R., Kutok, J.L., Le, X., Goessling, W., Neuberg, D.S., Kunkel, L.M., and Zon, L.I. 2007. Effects of RAS on the genesis of embryonal rhabdomyosarcoma. Genes Dev 21: 1382–1395. LaRusch, G.A., Jackson, M.W., Dunbar, J.D., Warren, R.S., Donner, D.B., and Mayo, L.D. 2007. Nutlin3 blocks vascular endothelial growth factor induction by preventing the interaction between hypoxia inducible factor 1alpha and Hdm2. Cancer Res 67: 450–454. Lazaro, J.B., Bailey, P.J., and Lassar, A.B. 2002. Cyclin D-cdk4 activity modulates the subnuclear localization and interaction of MEF2 with SRC-family coactivators during skeletal muscle differentiation. Genes Dev 16: 1792–1805.
Molecular Therapy for Rhabdomyosarcoma
453
LeRoith, D., Baserga, R., Helman, L., and Roberts, C.T., Jr. 1995. Insulin-like growth factors and cancer. Ann Intern Med 122: 54–59. Leuschner, I., Langhans, I., Schmitz, R., Harms, D., Mattke, A., and Treuner, J. 2003. p53 and mdm-2 expression in Rhabdomyosarcoma of childhood and adolescence: clinicopathologic study by the Kiel Pediatric Tumor Registry and the German Cooperative Soft Tissue Sarcoma Study. Pediatr Dev Pathol 6: 128–136. Li, L., Zhou, J., James, G., Heller-Harrison, R., Czech, M.P., and Olson, E.N. 1992. FGF inactivates myogenic helix-loop-helix proteins through phosphorylation of a conserved protein kinase C site in their DNA-binding domains. Cell 71: 1181–1194. Lin, R., Connolly, P.J., Lu, Y., Chiu, G., Li, S., Yu, Y., Huang, S., Li, X., Emanuel, S.L., Middleton, S.A., et€al. 2007. Synthesis and evaluation of pyrazolo[3,4-b]pyridine CDK1 inhibitors as antitumor agents. Bioorg Med Chem Lett 17: 4297–4302. Linardic, C.M., Downie, D.L., Qualman, S., Bentley, R.C., and Counter, C.M. 2005. Genetic modeling of human rhabdomyosarcoma. Cancer Res 65: 4490–4495. Logan, I.R., McNeill, H.V., Cook, S., Lu, X., Lunec, J., and Robson, C.N. 2007. Analysis of the MDM2 antagonist nutlin-3 in human prostate cancer cells. Prostate 67: 900–906. Lollini, P.L., De Giovanni, C., Del Re, B., Landuzzi, L., Nicoletti, G., Prodi, G., Scotlandi, K., and Nanni, P. 1989. Myogenic differentiation of human rhabdomyosarcoma cells induced in€vitro by antineoplastic drugs. Cancer Res 49: 3631–3636. Longati, P., Comoglio, P.M., and Bardelli, A. 2001. Receptor tyrosine kinases as therapeutic targets: the model of the MET oncogene. Curr Drug Targets 2: 41–55. Lu, Y.J., Williamson, D., Clark, J., Wang, R., Tiffin, N., Skelton, L., Gordon, T., Williams, R., Allan, B., Jackman, A., et€al. 2001. Comparative expressed sequence hybridization to chromosomes for tumor classification and identification of genomic regions of differential gene expression. Proc Natl Acad Sci USA 98: 9197–9202. Lu, X., Pearson, A., and Lunec, J. 2003. The MYCN oncoprotein as a drug development target. Cancer Lett 197: 125–130. Ma, P.C., Schaefer, E., Christensen, J.G., and Salgia, R. 2005. A selective small molecule c-MET Inhibitor, PHA665752, cooperates with rapamycin. Clin Cancer Res 11: 2312–2319. Manfredi, M.G., Ecsedy, J.A., Meetze, K.A., Balani, S.K., Burenkova, O., Chen, W., Galvin, K.M., Hoar, K.M., Huck, J.J., LeRoy, P.J., et€al. 2007. Antitumor activity of MLN8054, an orally active small-molecule inhibitor of Aurora A kinase. Proc Natl Acad Sci USA 104: 4106–4111. Mao, J., Maye, P., Kogerman, P., Tejedor, F.J., Toftgard, R., Xie, W., Wu, G., and Wu, D. 2002. Regulation of Gli1 transcriptional activity in the nucleus by Dyrk1. J Biol Chem 277: 35156–35161. Maris, J.M., Courtright, J., Houghton, P.J., Morton, C.L., Gorlick, R., Kolb, E.A., Lock, R., Tajbakhsh, M., Reynolds, C.P., Keir, S.T., et€al. 2008a. Initial testing of the VEGFR inhibitor AZD2171 by the pediatric preclinical testing program. Pediatr Blood Cancer 50: 581–587. Maris, J.M., Courtright, J., Houghton, P.J., Morton, C.L., Kolb, E.A., Lock, R.B., Tajbakhsh, M., Reynolds, C.P., Keir, S.T., Wu, J., and Smith, M.A. 2008b. Initial testing (stage 1) of sunitinib by the pediatric preclinical testing program. Pediatr Blood Cancer 51:42–48. Matsuoka, S., Edwards, M.C., Bai, C., Parker, S., Zhang, P., Baldini, A., Harper, J.W., and Elledge, S.J. 1995. p57KIP2, a structurally distinct member of the p21CIP1 Cdk inhibitor family, is a candidate tumor suppressor gene. Genes Dev 9: 650–662. McAllister, S.D., Chan, C., Taft, R.J., Luu, T., Abood, M.E., Moore, D.H., Aldape, K., and Yount, G. 2005. Cannabinoids selectively inhibit proliferation and induce death of cultured human glioblastoma multiforme cells. J Neurooncol 74: 31–40. Melino, G., De Laurenzi, V., and Vousden, K.H. 2002. p73: friend or foe in tumorigenesis. Nat Rev Cancer 2: 605–615. Melino, G., Bernassola, F., Ranalli, M., Yee, K., Zong, W.X., Corazzari, M., Knight, R.A., Green, D.R., Thompson, C., and Vousden, K.H. 2004. p73 Induces apoptosis via PUMA transactivation and Bax mitochondrial translocation. J Biol Chem 279: 8076–8083. Mercer, S.E., Ewton, D.Z., Shah, S., Naqvi, A., and Friedman, E. 2006. Mirk/Dyrk1b mediates cell survival in rhabdomyosarcomas. Cancer Res 66: 5143–5150.
454
R.T. Kurmasheva et al.
Minniti, C.P., Tsokos, M., Newton, W.A., Jr., and Helman, L.J. 1994. Specific expression of insulin-like growth factor-II in rhabdomyosarcoma tumor cells. Am J Clin Pathol 101: 198–203. Misawa, A., Hosoi, H., Arimoto, A., Shikata, T., Akioka, S., Matsumura, T., Houghton, P.J., and Sawada, T. 2000. N-Myc induction stimulated by insulin-like growth factor I through mitogenactivated protein kinase signaling pathway in human neuroblastoma cells. Cancer Res 60: 64–69. Mitsiades, C.S., Mitsiades, N.S., McMullan, C.J., Poulaki, V., Shringarpure, R., Akiyama, M., Hideshima, T., Chauhan, D., Joseph, M., Libermann, T.A., et€al. 2004. Inhibition of the insulin-like growth factor receptor-1 tyrosine kinase activity as a therapeutic strategy for multiple myeloma, other hematologic malignancies, and solid tumors. Cancer Cell 5: 221–230. Mohammadi, M., Froum, S., Hamby, J.M., Schroeder, M.C., Panek, R.L., Lu, G.H., Eliseenkova, A.V., Green, D., Schlessinger, J., and Hubbard, S.R. 1998. Crystal structure of an angiogenesis inhibitor bound to the FGF receptor tyrosine kinase domain. EMBO J 17: 5896–5904. Mora-Bermudez, F., Gerlich, D., and Ellenberg, J. 2007. Maximal chromosome compaction occurs by axial shortening in anaphase and depends on Aurora kinase. Nat Cell Biol 9: 822–831. Morgenstern, D.A., and Anderson, J. 2006. MYCN deregulation as a potential target for novel therapies in rhabdomyosarcoma. Expert Rev Anticancer Ther 6: 217–224. Morison, I.M., Ramsay, J.P., and Spencer, H.G. 2005. A census of mammalian imprinting. Trends Genet 21: 457–465. Muller, M., Schilling, T., Sayan, A.E., Kairat, A., Lorenz, K., Schulze-Bergkamen, H., Oren, M., Koch, A., Tannapfel, A., Stremmel, W., et€al. 2005. TAp73/Delta Np73 influences apoptotic response, chemosensitivity and prognosis in hepatocellular carcinoma. Cell Death Differ 12:1564–1577. Nabarro, S., Himoudi, N., Papanastasiou, A., Gilmour, K., Gibson, S., Sebire, N., Thrasher, A., Blundell, M.P., Hubank, M., Canderan, G., et€al. 2005. Coordinated oncogenic transformation and inhibition of host immune responses by the PAX3-FKHR fusion oncoprotein. J Exp Med 202: 1399–1410. Novitch, B.G., Mulligan, G.J., Jacks, T., and Lassar, A.B. 1996. Skeletal muscle cells lacking the retinoblastoma protein display defects in muscle gene expression and accumulate in S and G2 phases of the cell cycle. J Cell Biol 135: 441–456. O’Reilly, K.E., Rojo, F., She, Q.B., Solit, D., Mills, G.B., Smith, D., Lane, H., Hofmann, F., Hicklin, D.J., Ludwig, D.L., et€al. 2006. mTOR inhibition induces upstream receptor tyrosine kinase signaling and activates Akt. Cancer Res 66: 1500–1508. Pappo, A.S., Lyden, E., Breitfeld, P., Donaldson, S.S., Wiener, E., Parham, D., Crews, K.R., Houghton, P., and Meyer, W.H. 2007. Two consecutive phase II window trials of irinotecan alone or in combination with vincristine for the treatment of metastatic rhabdomyosarcoma: the Children’s Oncology Group. J Clin Oncol 25: 362–369. Pavelic, J., Matijevic, T., and Knezevic, J. 2007. Biological and physiological aspects of action of insulin-like growth factor peptide family. Indian J Med Res 125: 511–522. Pavlakovic, H., Havers, W., and Schweigerer, L. 2001. Multiple angiogenesis stimulators in a single malignancy: implications for anti-angiogenic tumour therapy. Angiogenesis 4: 259–262. Perry, R.L., Parker, M.H., and Rudnicki, M.A. 2001. Activated MEK1 binds the nuclear MyoD transcriptional complex to repress transactivation. Mol Cell 8: 291–301. Pession, A., and Tonelli, R. 2005. The MYCN oncogene as a specific and selective drug target for peripheral and central nervous system tumors. Curr Cancer Drug Targets 5: 273–283. Petricoin, E.F., 3rd, Espina, V., Araujo, R.P., Midura, B., Yeung, C., Wan, X., Eichler, G.S., Johann, D.J., Jr., Qualman, S., Tsokos, M., et€al. 2007. Phosphoprotein pathway mapping: Akt/ mammalian target of rapamycin activation is negatively associated with childhood rhabdomyosarcoma survival. Cancer Res 67: 3431–3440. Propper, D.J., McDonald, A.C., Man, A., Thavasu, P., Balkwill, F., Braybrooke, J.P., Caponigro, F., Graf, P., Dutreix, C., Blackie, R., et€ al. 2001. Phase I and pharmacokinetic study of PKC412, an inhibitor of protein kinase C. J Clin Oncol 19: 1485–1492. Qualman, S.J., and Morotti, R.A. 2002. Risk assignment in pediatric soft-tissue sarcomas: an evolving molecular classification. Curr Oncol Rep 4: 123–130. Radhakrishnan, S.K., and Gartel, A.L. 2006. A novel transcriptional inhibitor induces apoptosis in tumor cells and exhibits antiangiogenic activity. Cancer Res 66: 3264–3270.
Molecular Therapy for Rhabdomyosarcoma
455
Radhakrishnan, S.K., Halasi, M., Bhat, U.G., Kurmasheva, R.T., Houghton, P.J., and Gartel, A.L. 2008. Proapoptotic compound ARC targets Akt and N-myc in neuroblastoma cells. Oncogene 27: 694–699. Rao, S.S., and Kohtz, D.S. 1995. Positive and negative regulation of D-type cyclin expression in skeletal myoblasts by basic fibroblast growth factor and transforming growth factor beta. A role for cyclin D1 in control of myoblast differentiation. J Biol Chem 270: 4093–4100. Rao, S.S., Chu, C., and Kohtz, D.S. 1994. Ectopic expression of cyclin D1 prevents activation of gene transcription by myogenic basic helix-loop-helix regulators. Mol Cell Biol 14: 5259–5267. Rees, H., Williamson, D., Papanastasiou, A., Jina, N., Nabarro, S., Shipley, J., and Anderson, J. 2006. The MET receptor tyrosine kinase contributes to invasive tumour growth in rhabdomyosarcomas. Growth Factors 24: 197–208. Romer, J.T., Kimura, H., Magdaleno, S., Sasai, K., Fuller, C., Baines, H., Connelly, M., Stewart, C.F., Gould, S., Rubin, L.L., et€al. 2004. Suppression of the Shh pathway using a small molecule inhibitor eliminates medulloblastoma in Ptc1(+/−)p53(−/−) mice. Cancer Cell 6: 229–240. Romualdi, C., De Pitta, C., Tombolan, L., Bortoluzzi, S., Sartori, F., Rosolen, A., and Lanfranchi, G. 2006. Defining the gene expression signature of rhabdomyosarcoma by meta-analysis. BMC Genomics 7: 287. Rubin, B.P., Heinrich, M.C., and Corless, C.L. 2007. Gastrointestinal stromal tumour. Lancet 369: 1731–1741. Saab, R., Bills, J.L., Miceli, A.P., Anderson, C.M., Khoury, J.D., Fry, D.W., Navid, F., Houghton, P.J., and Skapek, S.X. 2006. Pharmacologic inhibition of cyclin-dependent kinase 4/6 activity arrests proliferation in myoblasts and rhabdomyosarcoma-derived cells. Mol Cancer Ther 5: 1299–1308. Sachdev, D., and Yee, D. 2006. Inhibitors of insulin-like growth factor signaling: a therapeutic approach for breast cancer. J Mammary Gland Biol Neoplasia 11: 27–39. Sachdev, D., and Yee, D. 2007. Disrupting insulin-like growth factor signaling as a potential cancer therapy. Mol Cancer Ther 6: 1–12. Santos, E.S., Rosenblatt, J.D., and Goodman, M. 2004. Role of farnesyltransferase inhibitors in hematologic malignancies. Expert Rev Anticancer Ther 4: 843–856. Sarfaraz, S., Afaq, F., Adhami, V.M., and Mukhtar, H. 2005. Cannabinoid receptor as a novel target for the treatment of prostate cancer. Cancer Res 65: 1635–1641. Sarfaraz, S., Afaq, F., Adhami, V.M., Malik, A., and Mukhtar, H. 2006. Cannabinoid receptor agonist-induced apoptosis of human prostate cancer cells LNCaP proceeds through sustained activation of ERK1/2 leading to G1 cell cycle arrest. J Biol Chem 281: 39480–39491. Sayan, A.E., Roperch, J.P., Sayan, B.S., Rossi, M., Pinkoski, M.J., Knight, R.A., Willis, A.E., and Melino, G. 2007. Generation of DeltaTAp73 proteins by translation from a putative internal ribosome entry site. Ann N Y Acad Sci 1095: 315–324. Schoenherr, C.J., Levorse, J.M., and Tilghman, S.M. 2003. CTCF maintains differential methylation at the Igf2/H19 locus. Nat Genet 33: 66–69. Scholl, F.A., Betts, D.R., Niggli, F.K., and Schafer, B.W. 2000. Molecular features of a human rhabdomyosarcoma cell line with spontaneous metastatic progression. Br J Cancer 82: 1239–1245. Scotlandi, K., Manara, M.C., Nicoletti, G., Lollini, P.L., Lukas, S., Benini, S., Croci, S., Perdichizzi, S., Zambelli, D., Serra, M., et€al. 2005. Antitumor activity of the insulin-like growth factor-I receptor kinase inhibitor NVP-AEW541 in musculoskeletal tumors. Cancer Res 65: 3868–3876. Scrable, H.J., Witte, D.P., Lampkin, B.C., and Cavenee, W.K. 1987. Chromosomal localization of the human rhabdomyosarcoma locus by mitotic recombination mapping. Nature 329: 645–647. Scrable, H., Witte, D., Shimada, H., Seemayer, T., Sheng, W.W., Soukup, S., Koufos, A., Houghton, P., Lampkin, B., and Cavenee, W. 1989. Molecular differential pathology of rhabdomyosarcoma. Genes Chromosomes Cancer 1: 23–35. Seale, P., and Rudnicki, M.A. 2000. A new look at the origin, function, and “stem-cell” status of muscle satellite cells. Dev Biol 218: 115–124. Seale, P., Sabourin, L.A., Girgis-Gabardo, A., Mansouri, A., Gruss, P., and Rudnicki, M.A. 2000. Pax7 is required for the specification of myogenic satellite cells. Cell 102: 777–786. Secchiero, P., Corallini, F., Gonelli, A., Dell’Eva, R., Vitale, M., Capitani, S., Albini, A., and Zauli, G. 2007. Antiangiogenic activity of the MDM2 antagonist nutlin-3. Circ Res 100: 61–69.
456
R.T. Kurmasheva et al.
Sepp-Lorenzino, L. 1998. Structure and function of the insulin-like growth factor I receptor. Breast Cancer Res Treat 47: 235–253. Shi, Y., Yan, H., Frost, P., Gera, J., and Lichtenstein, A. 2005. Mammalian target of rapamycin inhibitors activate the AKT kinase in multiple myeloma cells by up-regulating the insulin-like growth factor receptor/insulin receptor substrate-1/phosphatidylinositol 3-kinase cascade. Mol Cancer Ther 4: 1533–1540. Siehl, J., and Thiel, E. 2007. C-kit, GIST, and imatinib. Recent Results Cancer Res 176: 145–151. Skapek, S.X., Rhee, J., Spicer, D.B., and Lassar, A.B. 1995. Inhibition of myogenic differentiation in proliferating myoblasts by cyclin D1-dependent kinase. Science 267: 1022–1024. Skapek, S.X., Rhee, J., Kim, P.S., Novitch, B.G., and Lassar, A.B. 1996. Cyclin-mediated inhibition of muscle gene expression via a mechanism that is independent of pRB hyperphosphorylation. Mol Cell Biol 16: 7043–7053. Skaper, S.D., Kee, W.J., Facci, L., Macdonald, G., Doherty, P., and Walsh, F.S. 2000. The FGFR1 inhibitor PD 173074 selectively and potently antagonizes FGF-2 neurotrophic and neurotropic effects. J Neurochem 75: 1520–1527. Sorensen, P.H., Lynch, J.C., Qualman, S.J., Tirabosco, R., Lim, J.F., Maurer, H.M., Bridge, J.A., Crist, W.M., Triche, T.J., and Barr, F.G. 2002. PAX3-FKHR and PAX7-FKHR gene fusions are prognostic indicators in alveolar rhabdomyosarcoma: a report from the children’s oncology group. J Clin Oncol 20: 2672–2679. Steenman, M.J., Rainier, S., Dobry, C.J., Grundy, P., Horon, I.L., and Feinberg, A.P. 1994. Loss of imprinting of IGF2 is linked to reduced expression and abnormal methylation of H19 in Wilms’ tumour. Nat Genet 7: 433–439. Stewart, C.E., and Rotwein, P. 1996. Growth, differentiation, and survival: multiple physiological functions for insulin-like growth factors. Physiol Rev 76: 1005–1026. Stiewe, T., and Putzer, B.M. 2002. Role of p73 in malignancy: tumor suppressor or oncogene? Cell Death Differ 9: 237–245. Stoeltzing, O., Liu, W., Reinmuth, N., Fan, F., Parikh, A.A., Bucana, C.D., Evans, D.B., Semenza, G.L., and Ellis, L.M. 2003. Regulation of hypoxia-inducible factor-1alpha, vascular endothelial growth factor, and angiogenesis by an insulin-like growth factor-I receptor autocrine loop in human pancreatic cancer. Am J Pathol 163: 1001–1011. Stone, R.M., De Angelo, J., Galinsky, I., Estey, E., Klimek, V., Grandin, W., Lebwohl, D., Yap, A., Cohen, P., Fox, E., et€al. 2004. PKC 412 FLT3 inhibitor therapy in AML: results of a phase II trial. Ann Hematol 83 Suppl 1: S89–S90. Stratton, M.R., Fisher, C., Gusterson, B.A., and Cooper, C.S. 1989. Detection of point mutations in N-ras and K-ras genes of human embryonal rhabdomyosarcomas using oligonucleotide probes and the polymerase chain reaction. Cancer Res 49: 6324–6327. Takahashi, Y., Oda, Y., Kawaguchi, K., Tamiya, S., Yamamoto, H., Suita, S., and Tsuneyoshi, M. 2004. Altered expression and molecular abnormalities of cell-cycle-regulatory proteins in rhabdomyosarcoma. Mod Pathol 17: 660–669. Taulli, R., Scuoppo, C., Bersani, F., Accornero, P., Forni, P.E., Miretti, S., Grinza, A., Allegra, P., Schmitt-Ney, M., Crepaldi, T., et€al. 2006. Validation of met as a therapeutic target in alveolar and embryonal rhabdomyosarcoma. Cancer Res 66: 4742–4749. Taylor, A.C., Shu, L., Danks, M.K., Poquette, C.A., Shetty, S., Thayer, M.J., Houghton, P.J., and Harris, L.C. 2000. P53 mutation and MDM2 amplification frequency in pediatric rhabdomyosarcoma tumors and cell lines. Med Pediatr Oncol 35: 96–103. Thimmaiah, K.N., Easton, J., Huang, S., Veverka, K.A., Germain, G.S., Harwood, F.C., and Houghton, P.J. 2003. Insulin-like growth factor I-mediated protection from rapamycin-induced apoptosis is independent of Ras-Erk1-Erk2 and phosphatidylinositol 3¢-kinase-Akt signaling pathways. Cancer Res 63: 364–374. Tiffin, N., Williams, R.D., Shipley, J., and Pritchard-Jones, K. 2003. PAX7 expression in embryonal rhabdomyosarcoma suggests an origin in muscle satellite cells. Br J Cancer 89: 327–332. Tomescu, O., Xia, S.J., Strezlecki, D., Bennicelli, J.L., Ginsberg, J., Pawel, B., and Barr, F.G. 2004. Inducible short-term and stable long-term cell culture systems reveal that the PAX3-FKHR fusion oncoprotein regulates CXCR4, PAX3, and PAX7 expression. Lab Invest 84: 1060–1070.
Molecular Therapy for Rhabdomyosarcoma
457
Tonelli, R., Purgato, S., Camerin, C., Fronza, R., Bologna, F., Alboresi, S., Franzoni, M., Corradini, R., Sforza, S., Faccini, A., et€al. 2005. Anti-gene peptide nucleic acid specifically inhibits MYCN expression in human neuroblastoma cells leading to cell growth inhibition and apoptosis. Mol Cancer Ther 4: 779–786. Tostar, U., Malm, C.J., Meis-Kindblom, J.M., Kindblom, L.G., Toftgard, R., and Unden, A.B. 2006. Deregulation of the hedgehog signalling pathway: a possible role for the PTCH and SUFU genes in human rhabdomyoma and rhabdomyosarcoma development. J Pathol 208: 17–25. Tovar, C., Rosinski, J., Filipovic, Z., Higgins, B., Kolinsky, K., Hilton, H., Zhao, X., Vu, B.T., Qing, W., Packman, K., et€al. 2006. Small-molecule MDM2 antagonists reveal aberrant p53 signaling in cancer: implications for therapy. Proc Natl Acad Sci USA 103: 1888–1893. Tsutsumi, N., Yonemitsu, Y., Shikada, Y., Onimaru, M., Tanii, M., Okano, S., Kaneko, K., Hasegawa, M., Hashizume, M., Maehara, Y., et€ al. 2004. Essential role of PDGFRalphap70S6K signaling in mesenchymal cells during therapeutic and tumor angiogenesis in€vivo: role of PDGFRalpha during angiogenesis. Circ Res 94: 1186–1194. Vassilev, L.T. 2007. MDM2 inhibitors for cancer therapy. Trends Mol Med 13: 23–31. Vogelstein, B., Lane, D., and Levine, A.J. 2000. Surfing the p53 network. Nature 408: 307–310. Wachtel, M., Dettling, M., Koscielniak, E., Stegmaier, S., Treuner, J., Simon-Klingenstein, K., Buhlmann, P., Niggli, F.K., and Schafer, B.W. 2004. Gene expression signatures identify rhabdomyosarcoma subtypes and detect a novel t(2;2)(q35;p23) translocation fusing PAX3 to NCOA1. Cancer Res 64: 5539–5545. Wang, W., Kumar, P., Wang, W., Epstein, J., Helman, L., Moore, J.V., and Kumar, S. 1998. Insulin-like growth factor II and PAX3-FKHR cooperate in the oncogenesis of rhabdomyosarcoma. Cancer Res 58: 4426–4433. Wang, X., Le, P., Liang, C., Chan, J., Kiewlich, D., Miller, T., Harris, D., Sun, L., Rice, A., Vasile, S., et€ al. 2003. Potent and selective inhibitors of the Met [hepatocyte growth factor/scatter factor (HGF/SF) receptor] tyrosine kinase block HGF/SF-induced tumor cell growth and invasion. Mol Cancer Ther 2: 1085–1092. Wang, Y., Hailey, J., Williams, D., Wang, Y., Lipari, P., Malkowski, M., Wang, X., Xie, L., Li, G., Saha, D., et€al. 2005. Inhibition of insulin-like growth factor-I receptor (IGF-IR) signaling and tumor cell growth by a fully human neutralizing anti-IGF-IR antibody. Mol Cancer Ther 4: 1214–1221. Wedge, S.R., Kendrew, J., Hennequin, L.F., Valentine, P.J., Barry, S.T., Brave, S.R., Smith, N.R., James, N.H., Dukes, M., Curwen, J.O., et€al. 2005. AZD2171: a highly potent, orally bioavailable, vascular endothelial growth factor receptor-2 tyrosine kinase inhibitor for the treatment of cancer. Cancer Res 65: 4389–4400. Weinstein, I.B., and Joe, A.K. 2006. Mechanisms of disease: oncogene addiction – a rationale for molecular targeting in cancer therapy. Nat Clin Pract Oncol 3: 448–457. Weksberg, R., Teshima, I., Williams, B.R., Greenberg, C.R., Pueschel, S.M., Chernos, J.E., Fowlow, S.B., Hoyme, E., Anderson, I.J., Whiteman, D.A., et€al. 1993. Molecular characterization of cytogenetic alterations associated with the Beckwith–Wiedemann syndrome (BWS) phenotype refines the localization and suggests the gene for BWS is imprinted. Hum Mol Genet 2: 549–556. Williams, J.A., Guicherit, O.M., Zaharian, B.I., Xu, Y., Chai, L., Wichterle, H., Kon, C., Gatchalian, C., Porter, J.A., Rubin, L.L., et€al. 2003. Identification of a small molecule inhibitor of the hedgehog signaling pathway: effects on basal cell carcinoma-like lesions. Proc Natl Acad Sci USA 100: 4616–4621. Williamson, D., Lu, Y.J., Gordon, T., Sciot, R., Kelsey, A., Fisher, C., Poremba, C., Anderson, J., Pritchard-Jones, K., and Shipley, J. 2005. Relationship between MYCN copy number and expression in rhabdomyosarcomas and correlation with adverse prognosis in the alveolar subtype. J Clin Oncol 23: 880–888. Williamson, D., Selfe, J., Gordon, T., Lu, Y.J., Pritchard-Jones, K., Murai, K., Jones, P., Workman, P., and Shipley, J. 2007. Role for amplification and expression of glypican-5 in rhabdomyosarcoma. Cancer Res 67: 57–65.
458
R.T. Kurmasheva et al.
Wittman, M.D., Balasubramanian, B., Stoffan, K., Velaparthi, U., Liu, P., Krishnanathan, S., Carboni, J., Li, A., Greer, A., Attar, R., et€al. 2007. Novel 1H-(benzimidazol-2-yl)-1H-pyridin-2-one inhibitors of insulin-like growth factor I (IGF-1R) kinase. Bioorg Med Chem Lett 17: 974–977. Yang, Y., Hoeflich, A., Butenandt, O., and Kiess, W. 1996. Opposite regulation of IGF-I and IGF-I receptor mRNA and concomitant changes of GH receptor and IGF-II/M6P receptor mRNA in human IM-9 lymphoblasts. Biochim Biophys Acta 1310: 317–324. Yoo, C.B., and Jones, P.A. 2006. Epigenetic therapy of cancer: past, present and future. Nat Rev Drug Discov 5: 37–50. Yu, Y., Davicioni, E., Triche, T.J., and Merlino, G. 2006. The homeoprotein six1 transcriptionally activates multiple protumorigenic genes but requires ezrin to promote metastasis. Cancer Res 66: 1982–1989. Zaika, A.I., Slade, N., Erster, S.H., Sansome, C., Joseph, T.W., Pearl, M., Chalas, E., and Moll, U.M. 2002. DeltaNp73, a dominant-negative inhibitor of wild-type p53 and TAp73, is upregulated in human tumors. J Exp Med 196: 765–780. Zhan, S., Shapiro, D.N., and Helman, L.J. 1994. Activation of an imprinted allele of the insulinlike growth factor II gene implicated in rhabdomyosarcoma. J Clin Invest 94: 445–448. Zhang, L., and Wang, C. 2003. PAX3-FKHR transformation increases 26S proteasome-dependent degradation of p27Kip1, a potential role for elevated Skp2 expression. J Biol Chem 278: 27–36. Zhang, L., Kashanchi, F., Zhan, Q., Zhan, S., Brady, J.N., Fornace, A.J., Seth, P., and Helman, L.J. 1996. Regulation of insulin-like growth factor II P3 promotor by p53: a potential mechanism for tumorigenesis. Cancer Res 56: 1367–1373. Zhang, L., Zhan, Q., Zhan, S., Kashanchi, F., Fornace, A.J., Jr., Seth, P., and Helman, L.J. 1998a. p53 regulates human insulin-like growth factor II gene expression through active P4 promoter in rhabdomyosarcoma cells. DNA Cell Biol 17: 125–131. Zhang, L., Zhan, S., Navid, F., Li, Q., Choi, Y.H., Kim, M., Seth, P., and Helman, L.J. 1998b. AP-2 may contribute to IGF-II overexpression in rhabdomyosarcoma. Oncogene 17: 1261–1270. Zhang, J.M., Wei, Q., Zhao, X., and Paterson, B.M. 1999. Coupling of the cell cycle and myogenesis through the cyclin D1-dependent interaction of MyoD with cdk4. EMBO J 18: 926–933.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here? Rosanna Ricafort and Richard Gorlick
Introduction/Epidemiology Osteosarcoma at the molecular level is among the most complex pediatric Â�malignancies. Despite its relatively high incidence in the pediatric patient population many aspects of this disease continue to defy understanding. Osteosarcoma is the most common primary malignant bone tumor in children and adolescents (Arndt and Crist 1999). Approximately 400 children and adolescents are diagnosed each year in the United States (Meyers and Gorlick 1997). In the adolescent age range, it is the second most common malignancy, following only lymphoma in incidence. Osteosarcoma continues to be diagnosed solely based upon its histologic appearance. Despite having tremendously variable appearance, characterized descriptively by its histologic subtype, the presence of a malignant spindle cell which produces osteoid is the basis for diagnosing osteosarcoma. Molecular analyses of gene copy number, translocations, and gene expression do not contribute to making the diagnosis. In osteosarcoma numerous genetic abnormalities are present, each of which individually could lead to the development of a malignancy. This redundancy makes it difficult to decipher the relative timing and importance of each of these events. Quite similar to the molecular abnormalities, the epidemiology suggests multiple etiologic factors that may play a role in osteosarcoma development. Some of these factors may have differing levels of importance in different patient populations with the overall contribution of each of these factors unknown. Multiple studies have linked periods of rapid growth to the development of osteosarcoma, and not surprisingly incidence for this disease peaks during the second decade of life. Osteosarcoma has a predilection for the rapidly growing long bones of the extremities, with the distal femur and proximal tibia accounting for approximately 50% of all cases (Meyers and Gorlick 1997). Other factors suggesting R. Gorlickâ•›() Department of Pediatrics, Pediatric Hematology/Oncology, The Children’s Hospital at Montefiore and the Albert Einstein College of Medicine, 3415 Bainbridge Avenue, Rosenthal 3rd Floor, Bronx, NY 10467, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_21, © Springer Science+Business Media, LLC 2010
459
460
R. Ricafort and R. Gorlick
an association with growth include its increased incidence in taller individuals in the majority of reported studies, as well as its earlier peak incidence in females concordant with their earlier growth spurt. Perhaps among the more frequently cited data in making this assertion, is the high incidence of osteosarcoma among large breed dogs and its almost complete absence among smaller dogs. More recent studies have suggested that the incidence of osteosarcoma does not vary within canine breeds suggesting the determining factor may be their underlying genetics rather than height as a phenotype. More recent explanations of canine breed height variability based on a finite number of polymorphisms in the insulin-like growth factor axis may give further credence to that view. Independent of growth, accelerated bone turnover likely plays a role in the etiology of osteosarcoma, given its association with disorders of bone metabolism (i.e., Paget’s disease of the bone, and fibrous dysplasia) (Meyers and Gorlick 1997; Gelberg et€ al. 1997). Other clear predisposing factors include exposure to nonspecific DNA damaging agents such as ionizing radiation. Radiation exposure clearly increases the incidence of osteosarcoma, but the vast majority of patients diagnosed have no known exposures and are too young to permit the latency period that typically occurs between the exposure and tumor development. The mechanism by which these etiologic factors predispose to osteosarcoma remains unclear. Multiple molecular alterations resulting in inactivation of tumor suppressor genes and overexpression of oncogenes are observed with considerable variability in every osteosarcoma. With the multitude of genetic lesions in osteosarcoma, it is difficult to discern which of these events is fundamental in tumorigenesis. Despite the many advances in the surgical and medical management of osteosarcoma in the past three decades, outcome of this disease more recently has not changed, with longterm survival rates remaining 70% in patients with localized disease; and the outcome for those with advanced or relapsed disease is poor (Chou et€al. 2005b). Defining the pathogenesis of osteosarcoma likely will have an impact on the prognosis and therapy of this tumor through defining relevant treatment approaches. The following review will explore the molecular alterations of osteosarcoma in attempts to elucidate its pathogenesis, substantiate pathways as therapeutic targets, and contemplate current strategies and future challenges in the treatment of this tumor.
Molecular Features of Tumorigenesis: Alterations in Osteosarcoma Osteosarcoma Cell of Origin Tumors may be defined by both the cells they originate from as well as the genetic events that lead to tumor formation. Identifying the type of cell in which osteosarcoma arises is critical to understand the molecular mechanisms that drive its growth and differentiation. Given its characteristic production of osteoid, traditionally
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
461
osteosarcoma is thought to be derived from osteoblasts, though evidence suggests that the cell of origin of osteosarcoma may be a more primitive precursor (Gibbs et€al. 2005). Osteosarcoma can have varying histologic appearance with descriptions, including osteoblastic, chondroblastic, fibroblastic, telangiectatic, and small cell subtypes (Bertoni et al. 1989). Many osteosarcomas contain multiple histologic subtypes. These histologic subtypes, with limited clinical relevance, may represent an osteosarcoma’s capacity to differentiate into multiple lineages suggesting a more pluripotent cellular potential than is associated with a mature osteoblast. The formation of the various mesenchymal derivatives from a stem cell is analogous to hematopoiesis. Bone, cartilage, fibrous stroma, muscle, tendon, and adipose are all derived from a pluripotent mesenchymal stem cell. Similar to oncogenesis in the context of the hematopoietic system, sarcomas may in theory differentiate or dedifferentiate during the tumor formation further complicating defining the cell of origin. Several other recent murine findings have also implicated mesenchymal stem cells as a possible progenitor (Tolar et€al. 2007); however, all of these data are far from definitive. Whatever the cell of origin in osteosarcoma, its pathogenesis involves complex molecular disruptions in proliferation and cell cycle control, apoptosis, and DNA damage response, along with disruptions in differentiation. All sarcomas can be divided into those that have recurrent characteristic chromosomal translocations such as the translocations of chromosome 11 and 22 in Ewing sarcoma and PAX gene translocations in alveolar rhabdomyosarcoma. The translocation-associated sarcomas generally have simple karyotypes. Osteosarcoma is characteristic of those sarcomas which do not have a consistent, recurrent chromosomal translocation. Osteosarcoma, typical of this group, has a complex karyotype with multiple alterations in various genetic pathways, including p53. Heterogeneity in osteosarcoma is manifest in many manners, including less robust clustering in comprehensive gene expression profiles.
Cancer Predisposition Syndromes Cancer predisposition syndromes associated with a high incidence of osteosarcoma provide clues as to potential mechanisms for its tumorigenesis. The cancer predisposition syndromes associated with osteosarcoma are summarized in Table╛1. Among those with sporadic osteosarcoma few will be found to be associated with an un-suspected cancer predisposition syndrome or their related germline genetic alteration. The incidence of osteosarcoma is high in those with defined genetic predispositions due to several familial cancer syndromes as well as it is one of the most common tumors associated with other malignancies. Patients with hereditary retinoblastoma have a heterozygous germline retinoblastoma (RB) gene mutation and develop retinoblastoma with 90% penetrance (Aue et€al. 1998). The cumulative incidence of a second malignancy by 50 years after the initial retinoblastoma diagnosis is 51%, with approximately half of these tumors being osteosarcoma
462
R. Ricafort and R. Gorlick
Table€1╅ Hereditary cancer syndromes that predispose to osteosarcoma Hereditary cancer syndrome Retinoblastoma
Chromosome location 13q14.2
Gene RB1
Li-Fraumeni
17p13.1
P53
Paget’s disease
18q21-q22 5q31 5q35 18q24.3
LOH18CR1 SQSTM1 MAPK8 RTS (ReQL4)
8p12-p11.2
WRN (RecQL2)
15q26.1
15q26.1
Rothmund Thomson syndrome Werner syndrome
Bloom syndrome
Function Cell cycle regulation DNA damage response IL-1/TNF signaling RANK signaling DNA helicase
DNA helicase Exonuclease activity DNA helicase
% of malignancies that are OS 50% 10% Not applicable
30%
<10%
<10%
(Wong et€ al. 1997; Draper et€ al. 1986). Both osteosarcomas and retinoblastomas occurring in patients with hereditary retinoblastoma have somatic loss of the normal RB allele and resultant complete absence of a functional RB protein (Hansen et€al. 1985). Individuals with Li Fraumeni syndrome have germline mutations in p53 predisposing them to multiple malignancies, 10% of which are osteosarcoma (Li et€al. 1988). Most tumors in these patients have loss of the normal p53 allele (Malkin et€al. 1990). Familial cancer syndromes caused by mutations in RECQ DNA helicases include Rothmund-Thomson syndrome, Werner’s syndrome, and Bloom syndrome. RECQ DNA helicases are conserved proteins that separate the complementary strands of DNA duplexes; the three RECQ DNA helicases involved in cancer predisposition syndromes are RECQL4 in Rothmund-Thomson syndrome, RECQL2 in Werner syndrome, and RECQL3 in Bloom syndrome; they have related but not identical functions in maintaining genomic integrity, which may account for their differential tendency for developing osteosarcoma (Wang et€ al. 2003; Goto et€al. 1996; Ellis et€al. 1995). Patients with Rothmund-Thomson syndrome have a 30% incidence of osteosarcoma, compared to less than 10% of patients with either Werner syndrome or Bloom syndrome (Wang et€al. 2001; Wang et€al. 2003; Goto et€al. 1996; Ellis et€al. 1995). Unlike in Li Fraumeni and hereditary retinoblastoma, in which the p53 and RB regulated pathways are also frequent targets for somatic mutations in sporadic osteosarcoma unrelated to cancer predisposition syndromes, mutations in the RECQ DNA helicases giving rise to osteosarcoma seem to be restricted to their specific familial cancer syndromes (Nishijo et€ al. 2004). Nevertheless, these cancer predisposition syndromes, summarized in Table€1, offer clues as to genetic instability and complex molecular alterations involved in the pathogenesis of osteosarcoma. The genetic pathways with among the most
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
463
compelling� data for their involvement in osteosarcoma tumorigenesis include p53 and RB based upon the hereditary predisposition syndromes, data from murine models and their frequent alteration in sporadic tumors.
Viral Targeting of Tumor Antigens Simian Virus 40 (SV40) is a monkey polyomavirus that has been shown to be oncogenic when injected into animal models, inducing various tumors including sarcomas and osteosarcoma (Carbone et€al. 1996). SV40 large T-antigen protein affects cellular processes via regulation of p53-mediated transcriptional activation and apoptotic pathways (Mendoza et€al. 1998) as well as cell cycle regulation by the RB tumor suppressor pathway (Ahuja et€ al. 2005). In some studies, SV40 DNA sequences have been demonstrated in up to 50% of osteosarcoma tumor samples (Carbone et€al. 1996) suggesting that contamination of poliovirus by SV40 which occurred in North America between 1955 and 1963 and further spontaneous transmission of the virus may be an etiologic factor in osteosarcoma. However, another study failed to find a correlation with integration of SV40 DNA and p53 or RB mutation status (Mendoza et€ al. 1998). The most definitive study demonstrating SV40 is not a major epidemiologic factor in the development of osteosarcoma was a negative serologic study for antibodies in a large cohort of osteosarcoma patients (Carter et€al. 2003).
Murine Models Although genetic alterations in humans, which are associated with osteosarcoma, have a more directly established relevance, murine models offer additional evidence as to genetic alterations and exposures which can predispose to osteosarcoma. Transgenic mice engineered to overexpress SV40 develop osteosarcoma, with the clinical pattern dependent upon the promoter driving expression (Knowles et€al. 1990). This is supportive of the viewpoint that the p53 and RB tumor suppressor genes have a role in osteosarcoma development as the SV40 large T antigen abrogates the functions of both of these pathways. Knock out mice with mutant or absent p53 develop osteosarcoma along with a variety of other tumors. Depending upon the particular p53 alteration, osteosarcoma develops in between 5 and 50% of the mice (Donehower et€ al. 1992; Jacks et€ al. 1994; Liu et€ al. 2000; Olive et€ al. 2004). Knocking out the RB gene is lethal at embryonic stages in mice not permitting an assessment of whether or not these mice are prone to developing osteosarcoma (Khidr and Chen 2006). Transgenic overexpression of several oncogenes can also produce osteosarcoma. Almost 100% of transgenic mice which ubiquitously overexpress Fos develop osteosarcoma (Grigoriadis et€al. 1993; Wang et€al. 1995). Transgenic mice which conditionally overexpress Myc in fibroblasts/osteoblasts
464
R. Ricafort and R. Gorlick
develop osteosarcoma (Jain et€al. 2002). Radiating mice leads to the development of osteosarcoma (Erfle et€ al. 1986) as does chronic exposure to parathyroid �hormone (Vahle et€al. 2002).
Cell Cycle Regulation Molecular alterations in cell cycle regulation are an almost constant feature in the pathogenesis of osteosarcoma. Dysregulation of cell cycle transition leading to malignant proliferation may be the result of many different cellular processes, including inhibition of apoptosis, amplification of DNA replication, loss of DNA repair mechanisms, and loss of checkpoint control. Molecular lesions resulting in RB and p53 pathway inactivation account for the majority of these genetic alterations in osteosarcoma. The inactivation of these pathways is accomplished through any one of a variety of genetic alterations. These alterations typically occur in a nonoverlapping manner so that only one mechanism of RB inactivation and one mechanism of p53 inactivation are present in an osteosarcoma. The RB gene, mapped to 13q14, has been implicated to play a role in these various cellular functions (Classon and Harlow 2002) and genetic lesions in the RB gene itself have been shown to be present in approximately 70% of primary osteosarcoma tumor samples (Wadayama et€ al. 1994; Benassi et€ al. 1999; Hansen et€ al. 1985). Once phosphorylated, RB regulates cell cycle progression from G1 to S phase by interacting with the transcription factor E2F (Classon and Harlow 2002). Abnormalities in either RB or in the various proteins responsible for its phosphorylation can result in loss of its checkpoint function (Nevins 1992). Germline RB mutations, either sporadic or inherited, result in a high incidence of osteosarcoma; half of the second malignancies developing in patients with hereditary retinoblastoma are osteosarcomas (Wong et€al. 1997). Approximately 60% of osteosarcoma tumors will have the loss of heterozygosity at 13q, the site of RB (Wadayama et€ al. 1994; Toguchida et€al. 1988; Feugas et€al. 1996). Gross structural rearrangements of the RB gene are present in approximately 30% of osteosarcoma tumors, but mutations are rare, occurring in less than 10% (Wadayama et€ al. 1994). Cyclin dependent kinase-4 (CDK4) in complex with cyclin D1 (CCND1) is responsible for the phosphorylation of RB, and therefore amplification or overexpression of these genes results in functional inactivation of the RB signaling pathway and subsequent cell cycle progression. Amplification of the 12q13-15 chromosomal region which encompasses the CDK4 locus is observed in approximately 5 to 10% of osteosarcomas, and when measured directly by Southern blotting, CDK4 is amplified in approximately 9% of osteosarcoma tumor samples (Wei et€ al. 1999a). The INK4A gene, localized to 9p21, encodes p16INK4A, a tumor suppressor that inhibits the CDK4-CCND1 complex, thereby halting cell cycle transition. CDK4 amplification seems to occur independently from INK4A alterations, suggesting mutually exclusive molecular aberrations being sufficient for RB pathway disruption (Nielsen et€ al. 1998; Wei et€ al. 1999a). Amplification of the INK4A region is observed in approximately
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
465
5 to 10% of osteosarcomas (Wei et€ al. 1999a, b). Interestingly, by virtue of an � alternative reading frame, INK4A also encodes another tumor suppressor, p19INK4A ARF or p14 , which interacts with MDM2 in the p53 pathway (Quelle et€al. 1995). Tumorigenesis in osteosarcoma, as in many other tumor systems, is affected by disruptions in DNA damage control. The tumor suppressor p53 gene product plays a major role in the cellular response to DNA damage (Bode and Dong 2004). Its alteration may allow cells to proliferate despite genetic errors. Once stabilized in the nucleus, p53 binds DNA and initiates cellular responses through transcriptional activation or suppression of distinct target genes that function to prevent proliferation of damaged cells via damage repair, cell-cycle arrest, or apoptosis (Bode and Dong 2004). By the transactivation of its major target, cyclin dependent kinase inhibitor p21WAF1 (p21), p53 inhibits the activities of cyclin D/CDK4 or CDK6, thereby halting cell cycle progression from the G1 to S phase by keeping RB in a hypophosphorylated state (Mulligan et€al. 1990; El-Diery 1998). Mutations in the p53 gene, localized to 17p13, are the most frequently occurring genetic alteration found in human malignancies (El-Diery 1998; Lonardo et€al. 1997). It is known that patients with Li-Fraumeni syndrome, heralded by heterozygous germline p53 mutations, have a higher incidence of osteosarcoma, as the second most common malignancy in these patients (Siddiqui et€al. 2005). Approximately 3% of sporadically occurring osteosarcoma patients have a germline mutation in p53. A significant proportion of sporadic osteosarcoma samples have some molecular modification abrogating normal p53 function (McIntyre et€ al. 1994). In osteosarcoma tumor samples, alterations in p53 consist of point mutations (occurs in 20 to 30%), gross gene rearrangements (occurs in 10 to 20%), and loss of one 17q allele (occurs in 75 to 80%) (Lonardo et€ al. 1997; Toguchida et€ al. 1998). Alterations in genes encoding cellular proteins that regulate p53 have also been implicated in the pathogenesis of osteosarcoma. MDM2 is a ubiquitin E3 ligase which negatively modulates p53 function by binding the protein, concealing its activation site and facilitating its proteasomal degradation (Oliner et€ al. 1993; Oliner et€ al. 1992; Kubbutat et€al. 1997). Amplification leading to MDM2 overexpression functionally eliminates p53 even in the presence of wild-type protein (Oliner et€al. 1993; Oliner et€al. 1992; Kubbutat et€al.; ╛1997). The human homolog of the MDM2 gene, that maps to chromosome 12q13-14, is amplified in about 10% of osteosarcomas (Oliner et€al. 1993; Ladanyi et€al. 1993a; Kubbutat et€al. 1997; Lonardo et€al. 1997). In most tumor samples, amplification of MDM2 gene is mutually exclusive to p53 overeexpression (considered indicative of a mutant protein with greater stability), confirming the reciprocal relationship of these abnormalities, and suggesting a model of differential pathways for p53 pathway inhibition in osteosarcoma tumorigenesis (Lonardo et€al. 1997). Another negative regulator of p53 is COPS3 (constitutive photomorphogenic homolog subunit 3) of the COP9 proteasome. The COP9 proteasome binds and phosphorylates p53, targeting the tumor suppressor to degradation by the ubiquitin system (Bech-Otschir et€al. 2001). The COPS3 gene maps to chromosome 17p11.2, part of a region amplified in about 25% of cases of osteosarcoma (Van Dartel and Hulsebos 2004). COPS3 was found to be consistently overexpressed after gene amplification (Van Dartel et€al. 2004) and additionally via
466
R. Ricafort and R. Gorlick
alternative mechanisms in some tumors lacking COPS3 gene amplification (Henriksen et€ al. 2003). COPS3 amplification seems to be mutually exclusive to MDM2 amplification or p53 mutations, again suggesting that any of these alterations in the p53 tumor suppressor pathway is sufficient in regards to the contribution of this pathway to tumorigenesis in osteosarcoma (Henriksen et€al. 2003).
Growth Factors/Signal Transduction Pathways Another aspect contributing to abnormal cell proliferation in the pathogenesis of osteosarcoma involves growth factor receptors and autocrine or paracrine stimulation, which leads to the loss of regulation of cell proliferation, motility, and angiogenesis. The higher incidence during the pubertal spurt, in anatomic sites of greater growth, and in taller individuals, suggests that growth factors play an important role in the initiation or potentially pathogenesis of osteosarcoma. Several oncogenes involved in cellular growth and differentiation may contribute to this pathogenesis. The coexpression of these growth factors along with their receptors on osteosarcoma tissue samples and cell lines suggests an autocrine or paracrine mechanism of action may be driving the proliferation of malignant cells in this tumor. Difficulty in understanding osteosarcoma lies in the redundancy with a multitude of growth factor receptors simultaneously overexpressed and potentially dysregulated. As mentioned previously, osteosarcoma has been associated with accelerated bone turnover (Meyers and Gorlick 1997). Therefore, it is not surprising that type-1 insulin-like growth factor (IGF-1), a major regulator of skeletal growth, has been found to be mitogenic in in€ vitro models of osteosarcoma (Pollak et€ al. 1990; Kappel et€al. 1994; Bostedt et€al. 2001; MacEwen et€al. 2004). Survival of osteosarcoma cell lines in€vitro has been shown to be dependent on exogenous IGF-I, and proliferation of these cell lines can be inhibited by blocking signaling through the IGF-I receptor (IGF-1R) via monoclonal antibodies or antisense oligonucleotides (Kappel et€ al. 1994; Haluska et al. 2006). IGF-1R has been found to be abundantly expressed in osteosarcoma cells (Burrow et€ al. 1998). IGF-1R has emerged as the predominant receptor in cancer biology that contributes not only to cellular proliferation, but perhaps also more importantly to malignant transformation and protection from apoptosis (Baserga et€al. 2003). Overexpression of IGF-I and its receptor may therefore play a role in the pathogenesis of osteosarcoma by contributing to the oncogenic transformation, and providing a proliferative and survival advantage to malignant cells (Sekyi-Otu et€al. 1995). Of the growth factor circuits involved in the pathogenesis of osteosarcoma, the contribution of IGF-I/ IGF-IR to the malignant potential of this tumor has been identified to be of major importance (Scotlandi et€al. 2005). Overexpression of the c-MET proto-oncogene tyrosine kinase receptor (MET) and its ligand, hepatocyte growth factor (HGF) or scatter factor, in osteosarcoma cell lines suggests a role for the MET in the metastatic phenotype of osteosarcoma. Binding of HGF or scatter factor to the MET/HGF receptor stimulates both cell
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
467
proliferation and motility, features associated with the malignant potential of tumor cells (Scotlandi et€al. 1996; Rong et€al. 1993). The predominant expression of the MET/HGF receptor on epithelial cells, along with its ligand being produced primarily by cells of mesenchymal origin, suggests a paracrine or autocrine signaling system in governing stromal-epithelial interactions (Scotlandi et€al. 1996; Ferracini et€al. 1995). In a study of 17 osteosarcoma tumor samples, 60% expressed the HGF receptor at high levels (Scotlandi et€al. 1996). When primary and metastatic tumors from the same patient are compared, metastatic tumors have higher MET expression than the primary tumors, again indicating its role in potentiating the metastatic phenotype of osteosarcoma (Scotlandi et€al. 1996; Ferracini et€al. 1995; Oda et€al. 2000). Several MET inhibitors are undergoing clinical trials, and both selective and multitargeted tyrosine kinase inhibitors are being developed. In addition, fully human antibodies targeting the ligand, and MET decoys are also being developed. Initial clinical trials with the small molecule MET inhibitor, ARQ197, suggest that this agent is well tolerated. Platelet derived growth factor (PDGF) is another potent mitogen for cells of mesenchymal origin (McGary et€al. 2002). Coexpression of PDGF and its receptors has been observed in various human solid tumors and correlates with inferior prognosis and metastasis in some (Kawai et€al. 1997; Henriksen et€al. 1993; Sulzbacher et€al. 2000, 2001), thereby suggesting an autocrine or paracrine mechanism driving cellular proliferation and differentiation, chemotaxis, and survival in these tumors. The role of PDGF/PDGF-R in osteosarcoma has been explored. In one immunohistochemical study, PDGF-AA and its receptor, PDGFR-a, was found in 34 and 27% of osteosarcoma tumor samples, respectively, with coexpression of both proteins common (Sulzbacher 2000). Another analysis demonstrated PDGF-A and PDGFR-a expression in 80% of osteosarcoma tumor samples, with coexpression found in 76% of the patients; the expression of either growth factor or receptor, or both, correlated significantly with a shorter event-free survival (Kubo et€al. 2002). Other tyrosine kinase receptors have been implicated in the pathogenesis of osteosarcoma. The ERBB2 proto-oncogene (also called HER2/neu), located on 17q12, encodes a protein structurally homologous to the epidermal growth factor receptor, although its actual ligand has not yet been identified. The importance of expression of ERBB family of tyrosine kinase receptors for tumor growth or clinical outcome has been demonstrated in a variety of human cancers, most notably in invasive breast cancers and non-small cell lung cancer (Wen et€ al. 2007). In these carcinomas, activation of intracellular mitogenic signal transduction pathways is implicated in contributing to tumorigenesis (Wen et€al. 2007). The data in osteosarcoma are somewhat controversial, with some investigators reporting overexpression, associated with either favorable or unfavorable prognosis, and other investigators reporting lack of overexpression of ERBB2 (Zhou et€al. 2003; Akatsuka et€al. 2002; Gorlick et€al. 1999; Hughes et€al. 2006; Scotlandi et€al. 2005; Somers et€al. 2005; Thomas et€al. 2002; Maitra et€al. 2001; Kilpatrick et€al. 2001; Anninga et€al. 2004; Scotlandi et al. 2004). In a study of 26 osteosarcoma tumor samples, 42% expressed ERBB2; the expression of ERBB2 in this cohort was associated with a poor �prognosis (Onda et€al. 1996).
468
R. Ricafort and R. Gorlick
Conversely, another study found HER2 overexpression�associated�with increased survival and less metastases (Akatsuka et€al. 2002). A more recent article, utilizing immunohistochemistry and tissue microarray data, failed to find significant overexpression of the HER2 protein or gene amplification at the ERBB2 locus (Somers et€ al. 2005). Various explanations may account for the discrepancies between studies, including the methodology and interpretation of immunohistochemical staining, the antibody clones utilized, and the cross-reactivity with other membrane antigens, along with inherent biologic variability among osteosarcoma samples tested in various institutions representing different geographical areas. Table€2 summarizes key studies investigating the expression of ERBB2 and its clinicopathologic correlates. ERBB4 is another member of the epidermal growth factor receptor family that has been implicated in the pathogenesis of osteosarcoma. Hughes et€al. evaluated the expression pattern of ERBB family members in primary osteosarcoma tumor samples, and found cell surface expression of EGFR and ERBB2, along with nuclear expression of the p80 isoform of ERBB4 (Hughes et€al. 2004). In a followup study, the authors demonstrated constitutive phosphorylation of these ERBB family of receptor tyrosine kinases in osteosarcoma cells, and suggested their activated role in tumorigenesis (Hughes et€al. 2006).
Signal Transduction Identification of growth factors that may play a role in the pathogenesis of osteosarcoma leads one to consider the downstream signal transduction pathways resulting from activation of their receptors. The majority of signaling by the aforementioned receptors are transduced by three inter-related parallel pathways; RAS/RAF/Â� mitogen activated protein kinases (MAPK), phosphatidylinositol 3¢-kinase/Akt (PI3K/AKT), and mammalian target of rapamycin (mTOR). Each of these pathways is comprised of a series of kinases, where signal is transmitted through phosphorylation of amino acids, tyrosine, serine or threonine. Transmission of signal in this manner permits it to occur rapidly as no RNA or protein synthesis is required. This transmission also permits marked signal amplification. RAS/RAF/MAPK – The MAPK pathways play pivotal roles in cell proliferation, differentiation and survival. The closely related MAPK pathways are regulated through a series of phosphorylation steps in a three component module: MAPKs are activated by MAPK kinases (MAPKKs) on dual residues of threonine and tyrosine, and MAPKKs are in turn phosphorylated by MAPK kinase kinases (MAPKKKs) on dual residues of serine/threonine. This unusual feature ensures the tight control on the signals transduced by MAPK pathways in terms of specificity, intensity, and duration. The ERK1 and ERK2 MAPKs (ERK1/2), often activated by growth factors, are widely considered as prosurvival and oncogenic. The constitutiveÂ� active form of the MAPK/ERK kinase (MAPKK of ERK, MEK1/2) is sufficientÂ�to transform murine fibroblast 3T3 cells (Mansour et€al. 1994; Cowley et€al. 1994).
Table€2â•… Overview of clinicopathological studies of erb-B2 (HER-2/neu) expression in osteosarcoma Protein overexpression by IHC Pattern/scoring Reference Samples Antibody system %â•›+â•›samples Gene amplification Onda et€al. 26 PT CB-11 Mem/qualitative 42% (11/26) Southern Blot (1996) CBE1 (cDNA probe) – no amplification or rearrangement in IHCâ•›+â•›samples PCR No activating mutation FISH (Her-2/neu 44% PT 11/25) Zhou et€al. 25 PT Ab-3 Mem and/or Cy probe) 58% PulM (7/12) (2003) 12 PulM + >/=25% 6/7 IHC + *majority Cy+; positive stain 2/5 IHC − rarely Mem+ 5B5 Mem 42% (20/47) PT 47 PT Gorlick et€al. Herceptest +>25% 50% (3/6) PulM (1999) 6 PulM 76% (10/13) Rec 13 Rec 62% (51/81) (either 81 PT, all L CB-11 Mem and/or Cy Akatsuka Mem et€al. (2002) + >30% positive and/or Cy) stain 21 PT; 13 L, A0485 Mem 0% FISH: Ratio of Maitra et€al. (2001) 8M HER-2/neu oncogene and ch17 centromere >/=2.0â•›=â•›amplified in none (0/21) NA
(continued)
Poor HR (pâ•›<â•›0.03), decreased EFS (localized at dx:47 vs 79% 5 years, pâ•›=â•›0.05) Increased EFS (72 vs 46% 5 years, pâ•›=â•›0.03)
Shorter overall DFS (pâ•›<â•›0.04); no correlation with HR
Clinical correlation IHCâ•›+â•›of primary tumor associated with: Decreased DFS (pâ•›<â•›0.01), Early pulmonary mets (pâ•›<â•›0.05), Poor HR (pâ•›<â•›0.01)
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here? 469
40 PT 1 Rec parosteal OS
25 PT 9 PulM 84 PT, all L
15 PT 18 Res
Kilpatrick et€al. (2001)
Thomas et€al. (2002) Scotlandi et€al. (2005)
Anninga et€al. (2004)
Mem + if 2+ stain
Mem and/or Cy + if >/=25% positive stain
3B5 CB-11 Herceptest
Herceptest
Mem and/or Cy
CB-11 monoclonal antibody, and Oncor polyclonal) A0485
Mem and/or Cy
Protein overexpression by IHC Pattern/scoring Antibody system
0% Mem (47% Cy) 32% (27/84; 78% concordance between the 2 antibodies) 28% by Herceptest (12/43) 3% (1/27)
0% Mem (Cyâ•›+â•›in 98% by CB-11, 83% by Oncor)
%â•›+â•›samples
NA
IHCâ•›+â•›associated with inferior EFS (pâ•›=â•›0.03)
FISH: HER2/neu probe 13 IHCâ•›+â•›tested – none showed amplification FISH: no HER-2 gene amplification in the IHCâ•›+â•›case RT-PCR: no overexpression of HER2 mRNA
NA
Cy IHCâ•›+â•›did not correlate with HR, metastasis, or EFS
RT-PCR: 0/19 OS
Gene amplification
Clinical correlation IHCâ•›+â•›of primary tumor associated with:
PT primary tumor; PulM pulmonary metastases; L localized; M metastatic; Res resection; Rec recurrent; Mem membrane staining; Cy cytoplasmic staining; IHC immunohistochemistry; EFS event-free survival; HR histologic response to preoperative chemotherapy; NA not applicable
Samples
Reference
Table€2╅ (continued)
470 R. Ricafort and R. Gorlick
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
471
Activating mutations are frequently found in BRAF (a MAPKKK), or its upstream activator, RAS, in colorectal cancer and melanoma (Davies et€ al. 2002). The ERK1/2 MAPK is among the most common dysregulated pathways identified in tumors, as indicated by activating mutations frequently occurring in RAS or BRAF in human cancer, particularly in colon cancer and malignant melanoma (Davies et€al. 2002). Some evidence suggests that the ERK MAPK pathway may mediate differentiation in PC12 cells and osteoblasts (Cowley et€al. 1994; Salasznyk et€al. 2004). To date, little information is available regarding the role of this pathway in osteosarcoma. Several sources of data that suggest this pathway may be constitutively active in osteosarcoma (Kubo et€al. 2002). PI3K/AKT and mTOR – The PI3K/AKT and mTOR signal transduction pathways play an important role in transducing a variety of receptor signals. mTOR, perhaps most importantly, serves as one of the intracellular pathways through which IGF signaling is transduced (Thimmaiah et€al. 2003). Signal transduction via PDGF receptors occurs predominantly through the PI3K/AKT pathway (Thomas et€ al. 1997). Ezrin, which will be discussed further subsequently, mediates its growth and survival signal via both PI3K/AKT and mTOR but not MAPK (Krishnan et€al. 2006). These signaling pathways have a variety of components and downstream targets, a discussion of which is beyond the scope of this chapter. In the context of osteosarcoma, little is known about these pathways although several publications demonstrate activation of these pathways with a variety of ligands and with associated biological affects. Analogous to the MAPK pathway, both PI3K/ AKT and mTOR have been suggested to be involved in the differentiation of mesenchymal stem cells, in this case toward the adipogenic lineage (Yu et€al. 2007). The pathways have been suggested as being regulators of osteoblast growth and apoptosis. The mTOR pathway plays a major role in cell growth and differentiation in a variety of tumors, including osteosarcoma (Bjornsti and Houghton 2004; Easton and Houghton 2006). PI3K/AKT has been implicated in osteosarcoma’s metastatic process (Fukaya et€al. 2005) as well as chemotherapy sensitivity.
Other Oncogenes The contribution of alterations of other cellular proto-oncogenes to the molecular pathogenesis of osteosarcoma has been investigated. The MYC proto-oncogene, localized to 8q24, functions in the regulation of cell proliferation. c-Myc has the potential to homo and heterodimerize. When complexed with MAX, it activates the transcription of a large number of genes with many involved in the regulation of proliferation, cell growth, inhibition of differentiation and apoptosis. It has been found to be overexpressed in a subset of osteosarcoma tumor samples (Ladanyi et€al. 1993a, b; Pompetti et€al. 1996; Ikeda et€al. 1989; Gamberi et€al. 1998). The MYC gene is amplified in 44% of osteosarcomas (Kochevar et€al. 1990; Ladanyi et€al. 1993b; Gamberi et€al. 1998; Stock et€al. 2000). c-MYC expression has been suggested to be related with metastasis (Gamberi et€al. 1998).
472
R. Ricafort and R. Gorlick
Fos was initially identified as being related to osteosarcoma through recognition that mice inoculated with FBJ murine virus developed these tumors (Finkel et€al. 1966; Van et€al. 1983; Miller et€al. 1984). The viral protein which induces osteosarcoma is v-fos, which led to the identification of the structurally related c-Fos protein as an oncogene (Curran et€al. 1983). Fos heterodimerizes with Jun to regulate the AP1 transcription factor. The AP1 transcription factor is known to be regulated by Vitamin D, transforming growth factor-bTGF-b and parathyroid hormone (PTH) – all known to be involved in the regulation of bone growth, indicating a possible basis for its association with osteosarcoma development. Overexpression of FOS oncoprotein is found in the majority of osteosarcomas, and potentially associated with metastasis or recurrence (Ruther et al. 1989; Wu et€ al. 1990; Pompetti et€al. 1996; Franchi et€al. 1998; Gamberi et€al. 1998; David et al. 2005).
Genetic Complexity In addition to the known loss of tumor suppressor genes and gain of oncogenes, osteosarcoma has tremendous chromosomal complexity with numerous other genes affected by gross genetic changes. Osteosarcoma DNA ploidy ranges from diploid to hexaploid with the numbers of each chromosome quite variable. As one evidence of the genetic complexity, spectral karyotyping reveals an average of 39 chromosomal rearrangements per tumor (Bayani et€al. 2003). Some of the genetic alterations appear to be random but recurrent regions of chromosomal gain and loss are present. In many of these cases, the target of the genetic amplification and loss are known and in many cases they are not. The most common genetic losses and regions with losses of heterozygosity involve chromosomes 2, 3, 6, 9, 10, 13, 17, and 18 with the RB gene located on chromosome 13 and P53 located on chromosome 17. Loss of heterozygosity on chromosome 18 is also seen in Paget’s disease and chromosome 3 in Brachmann-de Lange Syndrome, a bone dysmorphology syndrome (Nellissery et€ al. 1998). The most common regions of amplification include regions within chromosomes 1, 5, 6, 8, 12, 16, and 17 with MYC located on chromosome 8 and MDM2 on chromosome 12. Candidate genes for some of the amplified regions include PRDM16, CDC5L, NFKBIE, IFNG, MGRN1, PMP22, MYCD, TOP3A, MAPK7, and COPS3 (Man et€al. 2004).
Immortalization Immortalization heralded by telomere length stabilization has been explored as contributing to the pathogenesis of osteosarcoma. Telomeres are nucleoprotein structures that cap chromosome ends and serve to potentiate the replicative process in cells by compensating for the loss of genetic material secondary to the inability to replicate end DNA. Maintaining telomere length is regarded as a fundamental
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
473
necessity for overcoming cellular senescence in tumor cells. Some osteosarcoma tumor models attribute the complex unbalanced karyotypes that characterize osteosarcoma to telomere dysfunction. Although this model has not been disproven, the patterns of genetic rearrangements observed in osteosarcoma are not typical of tumors undergoing chromosomal fusion-bridge-breakage cycles secondary to telomere length erosion. Maintaining telomere length is accomplished by most cancers via activation of an enzyme, telomerase (TERT), a ribonuclease protein. In contrast, TERT activity has been shown in a minority of osteosarcoma tumors, and does not seem to be essential to tumorigenesis. When present in osteosarcoma tumor samples, TERT activity has been shown to be low in comparison to HeLa cell positive controls (Aue et€ al. 1998). The clinicopathological correlation of telomerase has varied: an inverse relationship was reported between TERT activity and pulmonary metastases in patients treated with chemotherapy (Sangiorgi et€ al. 2001), while another study found decreased survival associated with TERT expression in primary osteosarcoma tumor samples (Sanders et€ al. 2004). Alternative mechanisms of telomere lengthening (ALT), a recombination-based method characterized by heterogeneous and elongated telomeres, has been the mechanism of maintaining telomere length that is seen more often in osteosarcoma. About 60% of osteosarcoma tumor samples demonstrated elongated telomeres with ALT (Ulaner et€al. 2004). Aggressiveness of osteosarcoma tumors has been associated with telomere integrity: a subset of osteosarcoma patients that lacked both TERT activity and evidence of ALT demonstrated a more favorable prognosis (Ulaner et€al. 2003).
Bone Differentiation The complex molecular alterations in osteosarcoma may not only involve dysregulation of cellular pathways involved in proliferation and immortalization, but also disruptions in osteoblast differentiation. Osteosarcoma can be regarded as evolving from a primitive pluripotent progenitor cell that retains its proliferative capacity while undergoing partial differentiation, or alternatively, a more differentiated precursor cell that de-differentiates and gains a propensity to proliferate, or loses its nonproliferative commitment. Understanding the molecular mechanisms underlying osteogenic differentiation and discovering genetic alterations within these pathways in osteosarcomas, contributes to our understanding of the molecular pathogenesis of osteosarcoma. The differentiation status of osteosarcoma tumors is variable (Haydon et€al. 2007). Transforming growth factor-beta (TGF-b) isoforms play an important role in regulating bone formation. An immunohistochemical analysis demonstrated expression of one or more TGF-b isoforms, particularly TGF-b3, in all of the 25 high grade osteosarcoma tumor samples studied (Kloen et€ al. 1997). TGF-b3 was correlated with disease progression, and the pattern of staining in the endothelial and perivascular layers in the tumor stroma suggested an angiogenic activity of this growth factor (Kloen et€ al. 1997). Another study
474
R. Ricafort and R. Gorlick
demonstrated increased expression of TGF-b2 and VEGF correlated with �osteosarcoma grade (Jung et€al. 2005). Bone morphogenetic proteins (BMPs), part of the TGF-b superfamily, are multifunctional cytokines that regulate bone and skeletal development. BMPs are involved in the differentiation of mesenchymal cells to cells of osteoblastic lineage, and in the differentiation of immature osteoblasts into mature osteoblasts (Wei et€ al. 1999b; Ohta et€ al. 1992). In vitro data suggest that BMPs stimulate the growth of osteosarcoma cells (Ohta et€al. 1992). Numerous BMPs and/or their receptors are highly expressed in osteosarcoma tumors, implicating their importance in autocrine or paracrine growth stimulation of osteosarcoma (Wei et€al. 1999b). Overexpression of BMP Receptor II has been associated with poor prognosis and metastatic potential in osteosarcoma (Wei et€al. 1999b; Weiss et€al. 2006; Yoshikawa et€al. 1988). BMPs have been demonstrated to induce the expression of c-fos mRNA in animal models of osteoblast formation, suggesting that the sustained expression of c-fos has a role in the pathogenesis of osteosarcoma (Ohta et€al. 1992). The process of osteoblast differentiation is driven by wingless (WNT) signaling within the mesenchymal stem cell. WNTs, encompassing the wingless and int genes, encode a family of highly conserved, secreted ligands that play an important role in osteoblast development and tumorigenesis (Hoang et€ al. 2004b). WNT binds to the cell surface receptor frizzled (FRZ) and the coreceptor lipoprotein related proteins 5 and 6 (LRP-5/6), leading to the stabilization and accumulation of b-catenin that in turn elicits a variety of effects, including the induction of differentiation and proliferation (Hoang et€ al. 2004a, b). Aberrant activation of WNT signaling is associated with many human cancers (Polakis 2000). The role of Wnt pathway signaling in the pathogenesis of osteosarcoma purports that canonical Wnt signaling functions in maintaining an undifferentiated, proliferating progenitor population (Boland et€al. 2004). Cytoplasmic or nuclear accumulation of b-catenin, owing to b-catenin deregulation by WNT pathway signaling, has been found in a majority of osteosarcomas (Haydon et€al. 2002a), and has been correlated with its metastatic potential (Iwaya et€ al. 2003). Expression of LRP5 seen in 50% of osteosarcomas is associated with a less-differentiated phenotype and metastasis (Hoang et€al. 2004a, b). Overexpression of dickkopf-3, DKK-3, a WNT agonist, and inhibition of LRP5 by using a dominant-negative LRP5 receptor in osteosarcoma cell lines was shown to inhibit tumor cell motility and invasion, providing evidence linking WNT pathway signaling to osteosarcoma tumorigenesis (Hoang et€ al. 2004a). DKK-1, a secreted inhibitor of canonical WNT signaling, inhibits osteogenesis, and has been found to be elevated in pediatric osteosarcoma samples when compared to unaffected individuals (Lee et€al. 2007). The WNT signaling pathway, along with other signaling pathways that are involved in osteoblast differentiation, can be induced, perhaps in parallel, by PTH (Kulkarni et€al. 2005; Carpio et€al. 2001). PTH plays a central role in the regulation of bone and mineral metabolism, and has been implicated in the pathogenesis of osteosarcoma. In response to parathyroid hormone-related protein (PTHrP), osteoclastic resorption of bone is stimulated, and as a result, growth factors such as TGF-b and IGF-1 are released from the surrounding extracellular matrix, further
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
475
stimulating PTHrP secretion and additional tumor growth. This positive-feedback loop culminates in osteosarcoma tumor growth and metastasis (O’Keefe and Guise 2003; Yang et€al. 2007a).
Metastasis Chemokine stromal cell-derived factor 1 (SDF-1) is a cytokine-like protein expressed on the surface of vascular endothelial cells that, through binding to its chemokine receptor, CXCR4, plays a role in cytoskeleton rearrangement, adhesion to endothelial cells, and chemotaxis (Bleul et€ al. 1996; Gupta et€ al. 1998). Involvement of the CXCR4/SDF-1 pathway has been implicated in the metastatic potential of many cancers, including rhabdomyosarcoma and lymphoma (Taichman et€al. 2002; Libura et€al. 2002; Corcione et€al. 2000). CXR4 mRNA has been shown to be expressed in osteosarcomas and associated with the presence of metastases at the time of diagnosis (Laverdiere et€al. 2005). In vitro assays showed that migration of osteosarcoma cells expressing CXCR4 follows an SDF-1 gradient and that their adhesion to endothelial and bone marrow stromal cells is promoted by SDF-1 treatment (Perissinotto et€al. 2005). This study also provided a rationale for the propensity of osteosarcoma to metastasize to the lung, where SDF-1 concentration is high, and demonstrated the prevention of pulmonary metastasis in a murine model by the administration of a CXCR4 inhibitor (T134 peptide), suggesting molecular strategies inhibiting this axis as a therapeutic target (Perissinotto et€al. 2005). Ezrin is a membrane-cytoskeleton linker protein that allows direct cellular interactions with the microenvironment, facilitating signal transduction through growth factor receptors and adhesion molecules, thereby regulating cell migration and metastasis (Khanna et€al. 2004). The role of ezrin in potentiating the metastatic potential of osteosarcoma has been investigated. In an orthotopic model of murine osteosarcoma, ezrin expression was threefold higher in the more aggressive K7M2 cell line and correlated with the metastatic potential of the aggressive osteosarcoma when compared to the less aggressive K12 cell line (Khanna et€al. 2000, 2001). In follow-up experiments, ezrin expression was found to provide an early survival advantage for pulmonary metastatic osteosarcoma, mediated partly by MAPK but not AKT (Khanna et€ al. 2004). A significant correlation between high ezrin expression and poor outcome has been shown, with shorter disease-free survival and higher risk of metastatic relapse, thereby corroborating the animal model data (Khanna et€al. 2004).
Angiogenesis In addition to cell proliferation and motility, another feature contributing to the �metastatic potential of tumors is neovascularization. Angiogenesis elicits �proliferation and migration of endothelial cells to allow the formation of new
476
R. Ricafort and R. Gorlick
� capillaries (Kaya et€al. 2000). Vascular endothelial growth factor, VEGF, is a �peptide that acts as a mitogen for endothelial cells, directing new vessel formation and vascular maintenance. The role of neoangiogenesis in the pathogenesis of osteosarcoma has been explored via expression analysis of VEGF in tumor samples. In one such study, VEGF expression was seen by immunostaining in 62% of 17 primary osteosarcoma tumor samples; VEGF positivity defined by expression in greater than 30% of tumor cells, correlated with a poor prognosis amid a significantly higher metastatic rate (Kaya et€al. 2000). This association as a negative predictive factor persists following neoadjuvant chemotherapy, as another study illustrated the presence of VEGF expression in greater than 25% of tumor cells at the time of definitive surgery was associated with a poor outcome (Charity et€al. 2006). This finding supports the concept that the degree of necrosis at the time of definitive surgery reflects the innate biology of osteosarcoma rather than treatment specific factors (Davis 1994, Gorlick and Meyers 2003a). Recent research involving VEGFR1-positive hematopoietic bone marrow progenitor cells initiating a premetastatic niche (Kaplan et€al. 2005) exposes the possible role of VEGF in fostering the metastatic potential of osteosarcoma tumors via expression on osteoprogenitor cells.
Drug Resistance The utility of adjuvant and neoadjuvant chemotherapy for osteosarcoma has long been established (Meyers and Gorlick 1997; Carli 2003); however, its efficacy can be hindered by the presence of intrinsic and/or acquired drug resistance. Possible mechanisms of drug resistance in osteosarcoma include alterations in the expression of p-glycoprotein(P-GP), multidrug resistance protein expression, topoisomerase II, glutathione S-transferases, DNA repair enzymes, drug metabolism (inactivation), and reduced intracellular delivery (Gorlick et€ al. 2003). Of these mechanisms, P-GP expression has been the most extensively studied in osteosarcoma. P-GP is a transmembrane ATP-dependent efflux pump protein encoded by the multidrug resistance (MDR1) gene, which is responsible for the efflux of numerous chemotherapeutic agents, including doxorubicin (Hanahan and Weinberg 2000). Immunohistochemistry or RT-PCR quantification studies have explored P-GP expression in osteosarcoma, and demonstrated overexpression of P-GP in 23 to 45%, with decreased survival reported for patients with P-GP-positive tumors (Baldini et€ al. 1995; Chan et€ al. 1997; Hornicek et€ al. 2000; Yamamoto et€al. 2000; Park et€al. 2001; Serra et€al. 2003). A meta-analysis also proposes that P-GP is associated with an increased risk of disease progression (Pakos and Ioannidis 2003). Though the literature suggests that P-GP may be a marker of drug resistance and aggressiveness in osteosarcoma (Baldini et€ al. 1995; Chan et€ al. 1997; Hornicek et€al. 2000; Yamamoto et€al. 2000; Park et€al. 2001; Serra et€al. 2003), a prospective clinicopathological national intergroup study concluded that there was no correlation between P-GP expression and percentage of osteosarcoma tumor necrosis after induction chemotherapy or event-free survival in localized osteosarcoma (Schwartz et€ al. 2000, 2007). Overexpression of MDR1, the gene
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
477
encoding P-GP, also has been explored in regards to its role in the progression of osteosarcoma and its prognostic relevance. Though a small pilot study showed a trend toward a worse outcome in patients exhibiting high levels of MDR1 expression (Wunder et€al. 1993), the larger, prospective investigation did not delineate any correlation between MDR1 expression and disease progression in patients with osteosarcoma, with patients with either very low or very high levels of MDR1 fairing poorly (Wunder et€al. 2000). High dose methotrexate (MTX) with leucovorin rescue is a major component of current protocols for the treatment of osteosarcoma (Meyers and Gorlick 1997; Rosen et€al. 1982; Meyers et€al. 2005). High-dose MTX is vastly more effective than conventional dose methotrexate in the treatment of osteosarcoma – a finding that is not observed in other malignancies (Gorlick et€al. 1996), implying a mechanism of intrinsic MTX resistance within osteosarcoma tumor cells (Guo et€ al. 1999). MTX is a potent inhibitor of dihydrofolate reductase, a key enzyme for intracellular folate metabolism. In experimental systems, resistance to methotrexate can occur through a variety of mechanisms, including impaired intracellular transport of the drug via the reduced folate carrier, upregulation of dihydrofolate reductase, and diminished intracellular retention secondary to polyglutamylation (Betino 1993; Banerjee et al. 1995; Yang et al. 2003). Studies have demonstrated that impairments of drug influx secondary to decreased expression and mutations in the reduced folate carrier gene is the major basis of intrinsic resistance: 65% of osteosarcoma tumor samples were found to have decreased reduced folate carrier expression at the time of initial biopsy (Guo et€al. 1999). In contrast, dihydrofolate reductase overexpression was seen relatively infrequently at initial biopsy – in only 10% of osteosarcoma tumor samples, as opposed to in 62% of the tumors at the time of definitive surgery or relapse, suggesting this means as the major mechanism of acquired MTX resistance in osteosarcoma (Guo et€al. 1999).
Rationale for Targeting Specific Pathways Tumorigenesis in osteosarcoma, as in all human cancers, is a multistep process reflecting genetic alterations that drive the progressive transformation of normal osteoprogenitor cells into highly malignant osteosarcoma. The molecular pathways presented thus far represent the molecular hallmarks of cancer development set forth by Hanahan and Weinberg (2000) – self-sufficiency in growth signals, insensitivity to antigrowth signals, evasion of apoptosis, strong replicative potential, sustained angiogenesis, and tissue invasion and metastasis. The multiplicity of genetic alterations described in osteosarcoma, and the redundancy of these pathways makes it difficult to discern which, if any, of these pathways is central in tumor initiation and progression, and therefore which pathways to target for anticancer drug development. Similar to many pediatric malignancies, the lack of a premalignant lesion and even a clearly defined cell of origin preclude direct investigations of the multiple steps in osteosarcomas pathogenesis. Although tumors that arise in patients with Li Fraumeni syndrome and hereditary retinoblastoma have
478
R. Ricafort and R. Gorlick
defined inciting molecular aberrations in the p53 and RB pathways, and though these pathways encompass the majority of genetic alterations in osteosarcoma patients, they are not known to be the primary events in sporadic osteosarcoma. A summary of the genetic alterations are presented in Table€ 3 and a schema in Fig.€ 1. Given the need for improved therapeutic options for patients with highgrade osteosarcoma, targeted therapies based on molecular knowledge of critical genetic events may offer attractive adjuncts to our current strategies for osteosarcoma. Factors central in the pathogenesis of osteosarcoma are likely to be prognostic and/or relevant therapeutic targets. In the context of other malignancies, therapeutically targeting signaling pathways that the tumor is addicted to, have generally met with the most success. Defining these addictions in osteosarcoma has been problematic with the most central alterations being loss of the p53 and RB tumor suppressor genes. Correcting these alterations, which would require a gain of function, has been difficult to achieve therapeutically. Most clinical trials have been conducted empirically focusing on conventional cyototoxic chemotherapy. Prioritizing clinical trials of chemotherapy directed to specific pathways will be the challenge for the future. Given the redundancy of the genetic alterations in osteosarcoma, clinical success may ultimately require inhibition of multiple pathways simultaneously. Current and future comprehensive expression profiling studies may further shed light as to the genes central in the molecular pathogenesis of osteosarcoma and provide new insight in guiding targeted therapeutic strategies.
Current Strategies The standard approach to the treatment of osteosarcoma employs multimodality therapy with surgery in addition to systemic neoadjuvant and adjuvant chemotherapy most standardly comprises high-dose methotrexate, cisplatin, and doxorubicin. Despite improvements in supportive care, surgical techniques, and aggressive chemotherapy approaches for osteosarcoma, 30 to 40% of all patients with high-grade osteosarcomas die of metastatic or recurrent disease (Kempf-Bielack et€ al. 2005; Bacci et al. 2000). Therapeutic options for these high-risk patients are therefore needed. The conventional cytotoxic agents currently utilized in the treatment of osteosarcoma were discovered more than 20 years ago. These agents are characterized by their narrow therapeutic index and thus limited by associated toxicities. It appears that the doses of conventional cytotoxic agents have been maximally applied in osteosarcoma (Chou and Gorlick 2006); Accordingly, emerging therapeutic strategies have shifted from dose intensification of current cytotoxics to the development of novel therapies (Casas-Ganem and Healey 2005; Anderson and Pearson 2006). Radiation therapy has not been widely utilized in the upfront treatment of osteosarcoma as conventional teaching dictates that it is not sensitive to radiotherapy at doses that would allow the safe administration of systemic combination chemotherapy. However, radiation therapy is becoming a potential option for patients whose primary tumors are in sites where appropriate surgery is impossible, and
Location 13q14
9p21
12q13-15
17p13
12q13-14
17p11
Gene RB1
INK4A
CDK4
P53
MDM2
COPS3
Amplification
LOH – 75% Mutations – 20–30% Structural changes – 10–20% Amplification
Amplification
Genetic alteration LOH – 60% Structural changes – 30% Mutations – <10% Deletion
Table€3╅ Overview of genetic alterations in osteosarcoma
Ligase than functionally inhibits p53 Phosphorylates p53; functionally inhibits p53
Tumor suppressor; encodes for p16INK4A inhibits CDK4-CCDN1 and p19ink4a inhibits MDM2 Phosphorylates of RB; functionally inactivates RB Tumor suppressor; DNA damage control
Function Tumor suppressor; cell cycle control
25%
(continued)
Insufficient
Good
Good 40–60%
Advanced presentation
Good
5–10%
10%
Good
Negative prognostic factor
Prognostic significance Negative prognostic factor
5–10%
Estimated frequency (%) of alteration in OS 70%
Strength of data regarding relative contribution to osteosarcoma tumorigenesis Good
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here? 479
8q24
14q21-31
7q31
15q25-26
7p22
17q12
2q34
5p15.33
MYC
FOS
MET
IGF-1R
PDGF-AA/ PDGFALPHA ERBB2/HER2
ERBB4
TERT
b2: 1q41b3:14q24
BMP2: 20p12 BMP3: 14p22 BMP4: 14q22-q23 BMP6: 6p12.1 BMP7: 20q13
TGF-BETA3
BMP
ALT
Location
Gene
Table€3╅ (continued)
Amplification
Amplification
Amplification
Genetic alteration
TGF-beta superfamily of cytokines; regulates skeletal development
Oncogene; cell proliferation; motility Mitogenic stimulus; protection from apoptosis Cell proliferation, differentiation, chemotaxis Intracellular mitogenic stimulus Intracellular mitogenic stimulus Maintenance of telomere length Maintenance of telomere length Autocrine growth factor
Oncogene
Oncogene
Function
Up to 94%
Up to 100%
60%
42%
Advanced presentation Negative prognostic factor
Insufficient
Insufficient
Insufficient
Insufficient
Insufficient
Controversial
Poor Negative prognostic factor
34%/27% up to 80%
Insufficient
Good
Insufficient
Good
Advanced presentation Advanced presentation Advanced presentation
Prognostic significance
40–50%
60%
61%
7–11%
Estimated frequency (%) of alteration in OS
Strength of data regarding relative contribution to osteosarcoma tumorigenesis
480 R. Ricafort and R. Gorlick
11q12–13
2q21
6q25.3
6p12
7q21
21q22.3
5q11.2-13.2
17p11
WNT LRP5
CXCR4
EZRIN
VEGF
MDR1
RFC
DHFR
MAPK7
Encodes p-gp; maintains Amplification; genomic integrity rearrangement Deletion or mutation Transport protein for methotrexate Dihydrofolate reductase; Amplification or mutation target of methotrexate
Angiogenesis
Osteoblast differentiation Co-receptor for Wnt; induction of differentiation and proliferation Chemokine receptor; binds SDF-1 Cell migration
10% (initial bx) 65% (post chemo)
35% (p-gp overexpression in 23–45%) 65%
62%
80%
Cell lines
50%
Advanced presentation Advanced presentation Negative prognostic factor Negative prognostic factor
Advanced presentation
Insufficient
Insufficient
Insufficient
Controversial
Poor
Good
Insufficient
Insufficient Insufficient
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here? 481
482
R. Ricafort and R. Gorlick
ERB2 ERB4
IGF / IGF-1R MET/HGF
PTH-R
PDGF-AA / PDGF-α
Pluripotent Mesenchymal Stem Cell
DIFFERENTIATION MDM2
? Precursor Cell
EARLY OSTEOSARCOMA
p14
ARF
COPS3 p16INK4A CDK4-CCND1 p21 p53 c-fos TERT ALT c-myc Rb mTOR wnt MAPK-RAS-AKT pathway
TGF-β BMP
Drug Resistance MDR1
ANGIOGENESIS VEG-F
METASTASES Ezrin SDF-1/CXCR-4
Osteoblast
ADVANCED OSTEOSARCOMA
Fig.€1╅ Potential mechanisms of pathogenesis in osteosarcoma
for palliation (DeLaney 2005a, 2005b). In addition to traditional external beam irradiation, bone-seeking Â�radiopharmaceuticals have also been employed in the treatment of osteosarcoma to provide therapeutic irradiation to osteoblastic bone metastasis. The most widely studied is samarium (Sm)-153 ethylene diamine tetramethylene phosphonate (153Sm-EDTMP). This compound has been shown to have high bone specificity and may be effectively used in the palliative setting (Anderson et€al. 2002). The synergistic use of gemcitabine as a radiation sensitizer after the Sm-153 dose may increase its radiobiological effectiveness (Lawrence et€al. 1999) and offer improved palliation, though a pilot study failed to demonstrate a durable response with this combination (Anderson et€al. 2005). The biologic agent muramyl tripeptide phosphatidylethanolamine [MTP-PE] is a synthetic lipophilic analog of a component of the cell wall of Bacillus CalmetteGuerin, capable of stimulating macrophage cytotoxicity against osteosarcoma in animal models and sporadic canine osteosarcoma (MacEwen et€al. 1989; Kleinerman 1995), ultimately leading to the development of a phase III clinical trial in pediatrics. However, its development, and hence its role in osteosarcoma therapy, has been hampered by the difficult to interpret initial results of the Children’s Oncology Group phase III clinical trial in pediatrics in which addition of MTP-PE trended toward having a favorable impact on event-free survival only when combined with the addition of ifosfamide to the standard 3-drug chemotherapy regimen (Meyers et€ al. 2005). This interaction is not evident in later survival curves which more clearly demonstrate an advantage with the inclusion of MTP-PE but further development and use will depend on additional supplies of the drug becoming clinically
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
483
available. Other standard agents considered for use in osteosarcoma include the bisphosphonates. Via their direct inhibitory effects on osteoclast-mediated bone resorption, effects on osteoblast activity, and possible antitumor properties, bisphosphonates may have a potential role in the prevention of bone tumor metastases and primary tumor progression (Diel et€al. 2000). In vitro osteosarcoma models have demonstrated the ability of bisphosphonates to inhibit tumor growth and to disrupt tumor cell interactions with the surrounding microenvironment (Ashton et€al. 2005; Cheng et€al. 2004; Ory et€al. 2005). A recent Phase II study conducted at Memorial Sloan-Kettering Cancer Center assessing the role of pamidronate in the treatment of osteosarcoma along with standard chemotherapy has closed to accrual; early indications are that bisphosphonates can be safely given in the context of osteosarcoma therapy. However, whether the addition of bisphosphonates to classical threedrug chemotherapy affords survival benefit remains to be proven in a larger subsequent study. Another avenue of research has been the development of novel delivery mechanisms to deliver focused high doses of agents to sites of disease, while avoiding systemic toxicity. Innovations in drug formulation may improve the delivery efficiency of current cytotoxic agents and improve the therapeutic index. Liposomally encapsulated doxorubicin has been shown to be an effective mode of delivery of doxorubicin, and has been shown to be safe in a limited number of studies (Chidiac et€al. 2000; Judson et€al. 2001; Muggia 2001; Skubitz 2003). Its activity in relapsed sarcomas suggests liposomal doxorubicin that may be able to overcome tumor resistance mechanisms (Alberts et€al. 2004). Aerosolized granulocyte macrophage stimulating factor is being tested in a Children’s Oncology Group Phase II trial for relapsed osteosarcoma patients (Anderson 1999). An aerosolized liposomal formulation of cisplatin (Sustained release Lipid Inhalation Targeting [SLIT™] cisplatin) recently has been tested in a Phase Ib/ IIa setting for osteosarcoma with promising early results (Chou et€al. 2007). Though targeting nonspecific therapies to sites of disease is one alternative in developing novel strategies against osteosarcoma, targeted therapies based on molecular knowledge of critical genetic events may offer new insight in guiding treatment. Identifying the genes and signal transduction pathways responsible for the malignant behavior in osteosarcoma may lead to the development of agents with better tumor specificity and therapeutic efficacy. Hence, there has been intense focus on the efficacy of pathway specific agents in osteosarcoma. Pathway selective agents, in the form of small molecule inhibitors and monoclonal antibodies, can function to antagonize ligand/receptor interactions, inhibit the anabolic activity of key tyrosine kinases, and intercept intracellular downstream functions. Though the majority of molecular alterations involve inactivation of the retinoblastoma or p53 pathways, attempts to target these genes have not been successful. Targeted therapies aimed at inhibiting growth factor and signal transduction pathways involved in osteosarcoma tumorigenesis have been more recently explored. Heat shock protein 90 (Hsp90) inhibition has been shown to simultaneously disrupt multiple signal transduction pathways associated with various cancer models, including osteosarcoma (Bagatell et al. 2005). However, a Phase I trial of a novel Hsp90 inhibitor, 17-N-Allylamino-17-demethoxygeldanamycin (17-AAG), failed to
484
R. Ricafort and R. Gorlick
show benefit in this tumor (Bagatell et al. 2007). PDGF-AA and its receptor, PDGF-a, have been shown to be coexpressed on osteosarcoma tumor cells, correlating with an inferior prognosis (Kawai et€al. 1997; Henriksen et€al. 1993; Sulzbacher et€al. 2001; Kubo et€ al. 2002), making this loop an attractive target for novel therapeutic approaches in osteosarcoma. Imatinib mesylate (also known at ST1571, Gleevec), is a tyrosine kinase inhibitor developed as an inhibitor of the ABL protein tyrosine kinase (Druker et€al. 1996), and in Â�addition also has been noted to inhibit the PDGF receptor (Buchdunger et€ al. 1995, 2000). A Children’s Oncology Group Phase II study of imatinib in children with refractory or relapsed solid tumors, including osteosarcoma, demonstrated that this tyrosine kinase inhibitor had no activity against osteosarcoma as a single agent at conventional doses (Bond et€al. 2008). Several investigators have demonstrated that overexpression of ERBB2, has prognostic significance in osteosarcoma and that it may play a role in maintaining the malignant phenotype in osteosarcoma, although these data are controversial (Zhou et€al. 2003; Akatsuka et€al. 2002; Gorlick et€al. 1999; Hughes et€al. 2006; Scotlandi et€al. 2004, 2005; Somers et€al. 2005; Thomas et€al. 2002). Preclinical studies of trastuzumab, the humanized monoclonal antibody targeting the HER2 receptor, have suggested its ability not only to antagonize the function of growthsignaling properties of the HER2 system, but also to recruit immune effectors against the target (Weiner and Adams 2000), and to augment chemotherapy-induced cytotoxicity (Hancock et€ al. 1991; Baselga et€ al. 1998). Promising early clinical data of trastuzumab in HER2 overexpressing breast cancer (reviewed by Hudis 2007) have led to a Phase II study of trastuzumab in patients with metastatic osteosarcoma whose tumors overexpress HER2. The results of this study, which closed in October 2007, are eagerly awaited. Small molecule inhibitors of ERBB2 (Lapatinib) or pan-ERBB inhibitors (ARRY334543) are approved or in clinical development, and may have utility in treating osteosarcoma. The VEGF pathway is another growth factor signaling system recognized as a potential target of therapy in pediatric osteosarcoma, as studies have shown that VEGF expression correlates with metastasis and poor outcome (Charity et€al. 2006; DuBois and Demetri 2007; Kaya et€al. 2000; Kaya et€al. 2002; Lee et€al. 1999; Lee et€al. 2007a, b). A number of VEGF pathway inhibitors are available now and are being tested in the preclinical arena. Encouraging results for AZD2171, a specific VEGF-receptor inhibitor, demonstrated growth inhibition in solid tumor xenograft models including osteosarcoma (Maris et€al. 2008). Trials of bevacizumab (Avastin) an antibody that binds VEGF, added to conventional chemotherapy, are planned. Preclinical studies have suggested that the mTOR pathway may be a potentially effective target in osteosarcoma (Viella-Bach 1999; Wan et€al. 2005; Houghton et€al. 2007b). Rapamycin is widely available and several new rapamycin analogs (generally with less toxicity than rapamycin itself) have completed early stage clinical trials in sarcomas. A large phase III trial using the rapamycin analog (AP23573) is underway as a multi-institutional consortium study for the treatment of various sarcomas. Dysregulated insulin-like growth factor signaling is emerging as an important molecular target for cancer therapeutics development. Blockade of IGF-1 has been investigated using the sustained release sandostatin (OncoLar), which can reduce
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
485
IGF-1 levels. A Phase I study of OncoLar with and without tamoxifen in relapsed osteosarcoma patients demonstrated a sustained 40 to 50% decrease in IGF-1 levels; however, there was no measurable tumor response (Mansky et€al. 2002). The level of suppression of IGF-1 achieved in this trial may have been insufficient to produce a biologically relevant effect. Furthermore, the presence of redundant signal �transduction pathways along with autocrine and paracrine secretion of growth �factors may have compensated for the suppression of serum IGF-1 levels, rendering this therapy ineffective (Benini et€al. 1999). Alternatively, targeting the receptor itself may result in improved inhibition of the IGF-1 signal transduction pathway and thus lead to better cytotoxic effects. In preclinical cancer models, antibody mediated down regulation of IGF-1R significantly inhibits the growth of neoplastic cells (Kappel et€ al. 1994; Sachdev et€al. 2003; Hailey et€al. 2002; Quelle et€al. 1995; Scotlandi et€al. 2005), and induces regressions alone (Kolb et€al. 2008) or when combined with cytotoxic agents (Khanna et al. 2002; Cohen et€ al. 2005; Goetsch et€ al. 2005). Many new IGF-1R antagonistic monoclonal antibodies are being introduced into Phase I and II studies for sarcomas and specifically in osteosarcoma. As an example, a phase II trial of R1507, a recombinant monoclonal antibody to IGF-1R, which includes a stratum for patients with osteosarcoma has already been initiated by the Sarcoma Alliance for Research through Collaboration (SARC). Multiple competing trials of IGF-1R antibodies are being performed simultaneously with slightly differing eligibility and response criteria using different monoclonal antibodies. As each of the antibodies has different epitope specificity, antibody subclass and toxicity, relative efficacy between these agents may ultimately need to be clarified. Further testing of these biological agents alone and in combination will need to be done in order to fully assess their role in osteosarcoma treatment. Particularly attractive are combinations with rapamycin or rapamycin analogs as IGF-1R signals via mTOR with combined inhibition potentially being more efficacious. Other targets of potential interest include the receptor activator for nuclear factor (RANK)/RANK-Ligand (RANKL)/Osteoprotegrin (OPG) system. The RANK/RANKL)/OPG system has been shown to be important in maintaining normal bone metabolism (Itoh et€al. 2002; Wuyts et€al. 2001). Recent studies suggest that this system may play a role in osteosarcoma pathogenesis (Miyamoto et€al. 2002; Wittrant et€al. 2006) and thus interference with this system represents another potential therapeutic strategy. RANKL antagonists (OPG, anti-RANKL antibodies and a soluble nonsignaling form of RANK, [RANK-Fc]) that bind ligand and prevent its interaction with the RANK receptor, have been used successfully in various animal models of bone metastases (Akatsu et€al. 1998).
Targeting Bone Differentiation Pathways Given the indication that the complex molecular alterations in osteosarcoma may involve disruptions in osteoblast differentiation, agents promoting bone differentiation and/or circumventing differentiation defects may play a role in the therapy of
486
R. Ricafort and R. Gorlick
osteosarcoma. In vitro studies suggest that targeting the WNT signaling pathway via immunodepletion of DKK1 or GSK3b inhibitors, may prevent inappropriate differentiation in osteosarcoma cell lines (Lee et€al. 2007a, b). In addition, Â�strategies aimed at disrupting the PTHrP-based stimulation of growth factors such as TGF-b and IGF-1, leading to differentiation and tumor growth, may be exploited in Â�osteosarcoma therapy (O’Keefe and Guise 2003; Yang et€ al. 2007a). Promoting terminal differentiation with nuclear receptors may represent another means of therapy in osteosarcoma. In vitro data suggest the treatment with PPARg agonists or 9-cis retinoic acid that was able to reduce proliferation rate and cell viability and induce differentiation (Haydon et€al. 2002a, b).
The Pediatric Preclinical Testing Program With the bewildering number of agents under development, there needs to be a focused effort to systematically and rigorously evaluate their relevance to osteosarcoma therapy. The Pediatric Preclinical Testing Program (PPTP) is one such effort underway to facilitate the introduction of new, active agents into clinical trials for all childhood cancers. The number of children who are eligible for Phase I clinical trials are few, underscoring the importance of identifying potentially effective agents quickly and moving these agents into clinical trials in a timely manner. With a consortium of laboratories in the United States and abroad, the PPTP is able to quickly screen a large number of agents using in€ vitro and in€ vivo models. Preclinical testing potentially may predict the activity of new agents in patients with childhood cancers rationally designed clinical trials utilizing new agents can be formulated and completed quickly. The PPTP has evaluated several standard and novel agents, including cyclophosphamide, vincristine, topotecan, SCH717454 (anti-IGF-1R antibody) (Kolb et€ al. 2008), AZD2171 (a specific inhibitor of VEGF-receptor 2) (Maris et€al. 2007), AZD6244 (MEK1/2 inhibitor), bortezomib (proteosome inhibitor) (Houghton et€al. 2007a, b, 2008), ABT-263 (a Bcl-2 inhibitor) (Lock et€al. 2008), MLN8237 (aurora A kinase inhibitor), rapamycin (Houghton et€ al. 2007a, b), vorinostat (histone deacetylase inhibitor), lapitinib (EGFR and ERBB2 inhibitor), and sunitinib (Houghton et€al. 2007a, b, 2008). Thus far, of all the novel agents tested; SCH717454, AZD2171, and rapamycin have been among the most effective for the treatment of osteosarcoma. These data may help prioritize further development of these agents in pediatric clinical trials. The success of the PPTP will not be measured by the rapidity of the screening process but rather by the rapid identification of an active agent in osteosarcoma clinical trials. Other appealing model systems exist, most notably in domestic dogs which develop osteosarcoma spontaneously. The osteosarcomas in canines are quite similar to those that occur in humans and have the advantage of arising spontaneously, and therefore of being heterogeneous. The high incidence of osteosarcoma in dogs and the more accelerated time course for the progression of their disease facilitates the conduct of clinical trials in dogs analogous to those in humans.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
487
Linkages have increasingly been developed between the veterinary oncology consortiums and the oncology cooperative groups. Most notably, the National Cancer Institute and the Children’s Oncology Group are involved in comparative oncology programs in which osteosarcoma is a major interest.
Summary Osteosarcoma is a genetically complex cancer making the identification of �pathways which can be therapeutically targeted extremely difficult. Recent focus has been an increasing reliance on the use of preclinical model systems. Although selection of agents to test in these models remains empiric, the increased throughput possible in these systems may facilitate the identification of active chemotherapy. Perhaps recognition of effective therapies ultimately may inform the identification of pathways that are truly central in osteosarcomas pathogenesis or at a minimum which are worthy of further analysis.
References Ahuja D, Sáenz-Robles MT, Pipas JM. 2005. SV40 large T antigen targets multiple cellular pathways to elicit cellular transformation. Oncogene. 24:7729–45. Akatsu T, Murakami T, Nishikawa M, Ono K, Shinomiya N, Tsuda E, Mochizuki S, Yamaguchi K, Kinosaki M, Higashio K, Yamamoto M, Motoyoshi K, Nagata N. 1998. Osteoclastogenesis inhibitory factor suppresses osteoclast survival by interfering in the interaction of stromal cells with osteoclast. Biochem Biophys Res Commun 250:229–34. Akatsuka T, Wada T, Kokai Y, Kawaguchi S, Isu K, Yamashiro K, et€al. 2002. ErbB2 expression is correlated with increased survival of patients with osteosarcoma. Cancer 94:1397–404. Alberts DS, Muggia FM, Carmichael J, Winer EP, Jahanzeb M, Venook AP, et€al. 2004. Efficacy and safety of liposomal anthracyclines in phase I/II clinical trials. Semin Oncol 31:53–90. Anderson PM, Markovic SN, Sloan JA, Clawson ML, Wylam M, Arndt CA, et€al. 1999. Aerosol granulocyte macrophage-colony stimulating factor: a low toxicity, lung-specific biological therapy in patients with lung metastases. Clin Cancer Res 5:2316–23. Anderson PM, Pearson M. 2006. Novel therapeutic approaches in pediatric and young adult sarcomas. Curr Oncol Rep 8:310–5. Anderson PM, Wiseman GA, Dispenzieri A, Arndt CA, Hartmann LC, Smithson WA, et€al. 2002. High-dose samarium-153 ethylene diamine tetramethylene phosphonate: low toxicity of skeletal irradiation in patients with osteosarcoma and bone metastases. J Clin Oncol 20:189–96. Anderson PM, Wiseman GA, Erlandson L, Rodriguez V, et€al. 2005. Gemcitabine radiosensitization after high-dose samarium for osteblastic osteosarcoma. Cin Cancer Res 11:6895–6900. Anninga JK, van der Vijver MJ, Cleton-Jansen AM, Kristel PMP, Taminiau AHM, Nooij M, Egeler RM, Hogendoorn PCW. 2004. Overexpression of the HER-2 oncogene does not play a role in high-grade osteosarcomas. Eur J Cancer 40:963–970. Arndt CA, Crist WM. 1999. Common musculoskeletal tumors of childhood and adolescence. N Engl J Med 341:342–52. Ashton JA, Farese JP, Milner RJ, Lee-Ambrose LM, van Gilder JM. 2005. Investigation of the effect of pamidronate disodium on the in€vitro viability of osteosarcoma cells from dogs. Am J Vet Res 66:885–91.
488
R. Ricafort and R. Gorlick
Aue G, Muralidhar B, Schwartz HS, Butler MG. 1998. Telomerase activity in skeletal sarcomas. Ann Surg Oncol 5:627–634. Bacci G, Ferrari S, Bertoni F, Ruggieri P, Picci P, Longhi A, et€al. 2000. Long-term outcome for patients with nonmetastatic osteosarcoma of the extremity treated at the istituto ortopedico rizzoli according to the istituto ortopedico rizzoli/osteosarcoma-2 protocol: an updated report. J Clin Oncol 18:4016–27. Bagatell R, Beliakoff J, David CL, Marron MT, Whitesell L. 2005. Hsp90 inhibitors deplete key antiapoptotic proteins in pediatric solid tumor cells and demonstrate synergistic anticancer activity with cisplatin. Int J Cancer 113:179–88. Bagatell R, Gore L, Egorin MJ, Ho R, Heller G, Boucher N, et€al. 2007. Phase I pharmacokinetic and pharmacodynamic study of 17-N-allylamino-17-demethoxygeldanamycin in pediatric patients with recurrent or refractory solid tumors: a pediatric oncology experimental therapeutics investigators consortium study. Clin Cancer Res 13:1783–8. Baldini N, Scotlandi K, Barbanti-Brodano G, et€al. 1995. Expression of p-glycoprotein in highgrade osteosarcomas in relation to clinical outcome. N Engl J Med 333:1380–1385. Banerjee D, Ercikan-Abali E, Waltham M, Schnieders B, Hochhauser D, Li WW, et€ al. 1995. Molecular mechanisms of resistance to antifolates, a review. Acta Biochim Pol. 42:457–64. Baselga J, Norton L, Albanell J, Young-Mee K, Medelsohn J. 1998. Recombinant humanized antiHER-2 antibody (HERceptin) enhances the antitumor activity of paclitaxel and doxorubicin against HER2/neu overexpressing human breast cancer xenografts. Cancer Res 58:2825–2831. Baserga R, Peruzzi F, Reiss K. 2003. The IGF-1 receptor in cancer biology. Int J Cancer 107:873–877. Bayani J, Zielenska M, Pandita A, Al-Romaih K, Karaskova J, Harrison K et€al. 2003. Spectral karyotyping identifies recurrent complex rearrangements of chromosomes 8, 17, and 20 in osteosarcomas. Genes Chromosomes Cancer 36:7–16. Bech-Otschir, Kraft R, Huan X, Henklein P, Kapelari B, Pollmann C, Dubiel W. 2001. COP9 signalosome-specific phosphorylation targets p53 to degradation by the ubiquitin system. EMBO J 20:1630–1639. Benassi MS, Molendini L, Gamberi G, Ragazzini P, Sollazzo MR, Merli M, et€al. 1999. Alteration of pRb/p16/cdk4 regulation in human osteosarcoma. Int J Cancer. 84:489–93. Benini S, Baldini N, Manara MC, Chano T, Serra M, Rizzi S, et€al. 1999. Redundancy of autocrine loops in human osteosarcoma cells. Int J Cancer. 80:581–8. Bertoni F, Present D, Bacchini P, Pignatti G, Picci P, Campanacci M. 1989. The Istituto Rizzoli experience with small cell osteosarcoma. Cancer 64:2591–9. Betino JR. 1993. Karnofsky memorial lecture. Ode to methotrexate. J Clin Oncol 11:5–14. Bjornsti MA, Houghton PJ. 2004. The TOR pathway: a target for cancer therapy. Nat Rev Cancer 4:335–48. Bleul CC, Fuhlbrigge RC, Cassnovas JM, Aiuti A, Springer TA. 1996. A highly efficacious lymphocyte cheomattractant, stromal cell-derived factor 1 (SDF-1). J Exp Med 184:1101–1109. Bode AM, Dong Z. 2004. Post-translational modification of p53 in tumorigenesis. Nat Rev Cancer. 4:793–805. Boland GM, Perkins G, Hall DJ, Tuan RS. 2004. Wnt3a promotes proliferation and suppresses osteogenic differentiation of adult human mesenchymal stem cells. J Cell Biochem 93:1210–1230. Bond M, Bernstein ML, Pappo A, Schultz KR, Krailo M, Blaney SM, Adamson PC. 2008. A phase II study of imatinib mesylate in children with refractory or relapsed solid tumors: a Children’s Oncology Group study. Pediatr Blood Cancer 50:254–258. Bostedt KT, Schmid C, Ghirlanda-Keller C, Ollie R, Winterhalter KH, Zapf J. 2001. Insulin-like growth factor (IGF) I down-regulates type 1 IGF receptor (IGF 1R) and reduces the IGF I response in A549 non-small-cell lung cancer and Saos-2/B-10 osteoblastic osteosarcoma cells. Exp Cell Res 271:368–377. Buchdunger E, Cioffi CL, Law N, Stover D, Ohno-Jones S, Druker BJ, Lydon NB. 2000. Abl protein-tyrosine kinase inhibitor ST1571 inhibits in€vitro signal transduction mediated by c-kit and platelet-derived growth factor receptors. J Pharmacol Exp Ther 295:139–145. Buchdunger E, Zimmermann J, Mett H, Meyer T, Müller M, Druker BJ, Lydon NB. 1995. Selective inhibition of the platelet-derived growth factor signal transduction pathway by a
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
489
protein-tyrosine kinase inhibitor of the 2-phenylaminopyrimidine class. Proc Natl Acad Sci USA. 92:2558–2562. Burrow S, Andrulis IL, Pollak M, Bell RS. 1998. Expression of insulin-like growth factor receptorÂ�, IGF-1, and IGF-2 in primary and metastatic osteosarcoma. J Surg Oncol 69:21–7. Carbone M, Rizzo P, Procopio A, Giuliano M, Pass HI, Gebhardt MC, Mangham C, Hansen M, Malkin DF, Bushart G, Pompetti F, Picci P, Levine AS, Bergsagel JD, Garcea RL. 1996. SV40like sequences in human bone tumors. Oncogene 13:527–535. Carli M, Passone E, Perilongo G, Bisogno G. 2003. Ifosfamide in pediatric solid tumors. Oncology 65 Suppl 2:99–104. Carpio L, Gladu J, Goltzman D, Rabbani SA. 2001. Induction of osteoblast differentiation indexes by PTHrP in MG-63 cells involves multiple signaling pathways. Am J Phsyiol Endocrinol Metab. 281:E489–E499. Carter JJ, Madeleine MM, Wipf GC, Garcea RL, Pipkin PA, Minor PD, Galloway DA. 2003. Lack of serologic evidence for prevalent simian virus 40 infection in humans. J Natl Cancer Inst. 95:1522–30. Casas-Ganem J, Healey JH. 2005. Advances that are changing the diagnosis and treatment of malignant bone tumors. Curr Opin Rheumatol. 17:79–85. Chan HS, Grogan TM, Haddad G, DeBoer G, Ling V. 1997. P-glycoprotein expression: critical determinant in the response to osteosarcoma chemotherapy. J Natl Cancer Inst 89:1706–1715. Charity RM, Foukas AF, Deshmukh NS, Grimer RJ. 2006. Vascular endothelial growth factor expression in osteosarcoma. Clin Orthop Relat Res 448:193–8. Cheng YY, Huang L, Lee KM, Li K, Kumta SM. 2004. Alendronate regulates cell invasion and MMP-2 secretion in human osteosarcoma cell lines. Pediatr Blood Cancer. 42:410–5. Chidiac T, Budd GT, Pelley R, Sandstrom K, McLain D, Elson P, et€ al. 2000. Phase II trial of liposomal doxorubicin (Doxil) in advanced soft tissue sarcomas. Invest New Drugs. 18:253–9. Chou A, Merola P, Sowers R, Yang R, Kolb E, Healey J, et€al. 2005a. Analysis of Aberrant signal transduction pathways in osteosarcoma cell lines. Proc Am Assoc Cancer Res 46: Abstract #4551, 1075. Chou AJ, Merola PR, Wexler LH, Gorlick RG, Vyas YM, Healey JH, et€al. 2005b. Treatment of osteosarcoma at first recurrence after contemporary therapy: the Memorial Sloan-Kettering Cancer Center experience. Cancer 104:2214–21. Chou AJ, Bell MD, Mackinson C, Gupta R, Meyers PA, Gorlick R. 2007. Phase Ib/IIa study of sustained release lipid inhalation targeting cisplatin by inhalation in the treatment of patients with relapsed/progressive osteosarcoma metastatic to the lung. J Clin Oncol ASCO Annual Meeting Proceedings Part I. 25:18S; Abstract #9525. Chou AJ, Gorlick R. 2006. Chemotherapy resistance in osteosarcoma: current challenges and future directions. Expert Rev Anticancer Ther. 6:1075–85. Classon M, Harlow E. 2002. The retinoblastoma tumour suppressor in development and cancer. Nat Rev Cancer 2:910–7. Cohen BD, Baker DA, Soderstrom C, Tkalcevic G, Rossi AM, Miller PE, Tengowski MW, Wang F, Gualberto A, Beebe JS, Moyer JD. 2005. Combination therapy enhances the inhibition of tumor growth with the fully human anti-type 1 insulin-like growth factor receptor monoclonal antibody CP-751,871. Clin Cancer Res 11:2063–2073. Corcione A, Ottonello L, Tortolina G, Facchetti P, Airoldi I, Guglielmino R, Dadati P, Truini M, Sozzani S, Dallegri F, Pistoia V. 2000. Stomal cell-derived factor-1 as a chemoattractant for follicular center lymphoma B cells. J Natl Cancer Inst 92:628–635. Cowley S, Paterson H, Kemp P, Marshall CJ. 1994. Activation of MAP kinase kinase is necessary and sufficient for PC12 differentiation and for transformation of NIH 3T3 cells. Cell. 77:841–52 Curran T, MacConnell WP, van Straaten F, Verma IM. 1983. Structure of the FBJ murine osteosarcoma virus genome: molecular cloning of its associated helper virus and the cellular homolog of the v-fos gene from mouse and human cells. Mol Cell Biol. 3:914–21 David JP, Mehic D, Bakiri L, Schilling AF, Mandic V, Priemel M, et€al. 2005. Essential role of RSK2 in c-Fos-dependent osteosarcoma development. J Clin Invest 115:664–72.
490
R. Ricafort and R. Gorlick
Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, Teague J, et€al. 2002. Mutations of the BRAF gene in human cancer. Nature. 417:949–54. Davis AM, Bell RS, Goodwin PJ. 1994. Prognostic factors in osteosarcoma: a critical review. J Clin Oncol. 12:423–31. DeLaney TF, Park L, Goldberg SI, Hug EB, Liebsch NJ, Munzenrider JE, et€ al. 2005a. Radiotherapy for local control of osteosarcoma. Int J Radiat Oncol Biol Phys. 61:492–8. DeLaney TF, Trofimov AV, Engelsman M, Suit HD. 2005b. Advanced-technology radiation therapy in the management of bone and soft tissue sarcomas. Cancer Control 12:27–35. Diel IJ, Solomayer EF, Bastert G. 2000. Bisphosphonates and the prevention of metastasis: first evidences from preclinical and clinical studies. Cancer 88:3080–3088. Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA Jr, Butel JS, Bradley A. 1992. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 356:215–21. Draper GJ, Sanders BM, Kingston JE. 1986. Second primary neoplasms in patients with retinoblastoma. Br J Cancer 53:661–671. Druker BJ, Tamura S, Buchdunger E, Ohno S, et€al. 1996. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat Med 2:561–566. DuBois S, Demetri G. 2007. Markers of angiogenesis and clinical features in patients with sarcoma. Cancer. 109:813–9. Easton JB, Houghton PJ. 2006. mTOR and cancer therapy. Oncogene 25:6436–46. El-Diery WS. 1998. Regulation of p53 downstream genes. Sem Cancer Biol 8:345–57. Ellis NA, Groden J, Ye TZ, Straughen J, Lennon DJ, Ciocci S, et€al. 1995. The Bloom’s syndrome gene product is homologous to RecQ helicases. Cell 83:655–666. Erfle V, Schmidt J, Strauss GP, Hehlmann R, Luz A. 1986. Activation and biological properties of endogenous retroviruses in radiation osteosarcomagenesis. Leuk Res 10:905–13. Ferracini R, Renzo MFD, Scotlandi K, et€ al. 1995. The Met/HGF receptor is overexpressed in human osteosarcomas and is activated by either a paracrine or autocrine circuit. Oncogene 10:739–49. Feugas O, Guriec N, Babin-Boilletot A, Marcellin L, Simon P, Babin S et€ al. 1996. Loss of heterozygosity of the RB gene is a poor prognostic factor in patients with osteosarcoma. J Clin Oncol 14:467–472. Finkel MP, Biskis BO, Jinkins PB. 1966. Virus induction of osteosarcomas in mice. Science 151:698–701. Franchi A, Calzolari A, Zampi G. 1998. Immunohistochemical detection of c-fos and c-jun expression in osseous and cartilaginous tumours of the skeleton. Virchows Arch 432:515–519. Fukaya Y, Ishiguro N, Senga T, Ichigotani Y, Sohara Y, Tsutsui M, Shioura T, Iwamoto T, Hamaguchi M. 2005. A role for PI3K-Akt signaling in pulmonary metastatic nodule formation of the osteosarcoma cell line, LM8. Oncol Rep.14:847–52. Gamberi G, Benassi MS, Bohling T, Ragazzini P, Molendini L, Solazzo MR, et€al. 1998. C-myc and c-fos in human osteosarcoma: prognostic value of mRNA and protein expression. Oncology 55:556–563. Gelberg KH, Fitzgerald EF, Hwang S, Dubrow R. 1997. Growth and development and other risk factors for osteosarcoma in children and young adults Int J Epidemiol. 26:272–278. Gibbs CP, Kukekov VG, Reith JD, Tchigrinova O, Suslov ON, Scott EW, et€al. 2005. Stem-like cells in bone sarcomas: implications for tumorigenesis. Neoplasia. 7:967–76. Goetsch L, Gonzalez A, Leger O, et€al. 2005. A recombinant humanized anti-insulin-like growth factor receptor type I antibody (h7C10) enhances the antitumor activity of vinorelbine and anti-epidermal growth factor receptor therapy against human cancer xenografts. Int J Cancer 113:316–328 Gorlick R, Goker E, Trippett T, Waltham M, Banerjee D, Bertino JR. 1996. Intrinsic and acquired resistance to methotrexate in acute leukemia. N Engl J Med 11:5–14. Gorlick R, Huvos AG, Heller G, Aledo A, Beardsley GP, Healey JH, et€al. 1999. Expression of HER2/erbB-2 correlates with survival in osteosarcoma. J Clin Oncol. 17:2781–8.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
491
Gorlick R, Meyers PA. 2003a. Osteosarcoma necrosis following chemotherapy: innate biology versus treatment-specific. J Pediatr Hematol Oncol 25:840–1. Gorlick R, Anderson P, Andrulis I, Arndt C, et€ al. 2003b. Biology of childhood osteogenic Â�sarcoma and potential targets for therapeutic development: meeting summary. Clin Cancer Res. 9:5442–5453. Goto M, Miller RW, Ishikawa Y, Sugano H. 1996. Excess of rare cancers in Werner syndrome (adult progeria). Cancer Epidemiol Biomarkers Prev 5:239–246. Grigoriadis AE, Schellander K, Wang ZW, Wagner ER. 1993. Osteoblasts are target cells for transformation in c-fos transgenic mice. J Cell Biol 122:685–701. Guo W, Healey JH, Meyers PA, Ladanyi M, Huvos AG, Bertino JR, et€al. 1999. Mechanisms of methotrexate resistance in osteosarcoma. Clin Cancer Res. 5:621–627. Gupta SK, Lysko PG, Pillarisetti K, Ohlstein E, Stadel JM. 1998. Chemokine receptor in human endothelial cells. Functional expression of CXCR4 and its transctiptional regulation by inflammatory cytokines. J Biol Chem. 273:4282–4287. Hailey J, Maxwell E, Koukouras K, Bishop WR, Patcher JA, Wang Y. 2002. Neutralizing antiinsulin-like growth factor receptor 1 antibodies inhibit receptor function and induce receptor degradation in tumor cells. Mol Cancer Ther. 1:1349–1353. Haluska P, Carboni JM, Loegering DA, Lee FY, Wittman M, Saulnier MG, et€al. 2006. In vitro and in€vivo antitumor effects of the dual insulin-like growth factor-I/insulin receptor inhibitor, BMS-554417. Cancer Res. 66:362–371. Hanahan D and Weinberg RA. 2000. The hallmarks of cancer. Cell 100:57–70. Hancock MC, Langton BC, Chan T, Toy P, Monahan JJ, Mischak RP, Shawver LK. 1991. A monoclonal antibody against the c-erbB-2 signaling enhances the cytotoxity of cis-diamminedichloroplatinum against human breast and ovarian tumor cell lines. Cancer Res 51:4575–4580. Hansen MF, Koufos A, Gallie BL, Phillips RA, Fodstad O, Brogger A, et€al. 1985. Osteosarcoma and retinoblastoma: a shared chromosomal mechanism revealing recessive predisposition. Proc Natl Acad Sci 82:6216–6220. Haydon RX, Deyrup A, Ishikawa A, Heck R, et€al. 2002a. Cytoplasmic and/or nuclear accumulation of the b-catenin protein is a frequent event in human osteosarcoma. Int J Cancer 102:338–342. Haydon RC, Zhou L, Feng T, Breyer B, Cheng H, Jian W, et€al. 2002b. Nuclear receptor agonists as potential differentiation therapy agents for human osteosarcoma. Clin Cancer Res 8:1288–1294. Haydon RC, Luu HH, He TC. 2007. Osteosarcoma and osteoblastic differentiation: a new perspective on oncogenesis. Clin Orthop Relat Res 454:237–246. Henriksen R, Funa K, Wilander E, Backstrom T, Ridderheim M, and Oberg K. 1993. Expression and prognostic significance of platelet-derived growth factor and its receoptors in epithelial ovarian neoplasms. Cancer Res. 53:4550–4554. Henriksen J, Aagesen TH, Maelsandsmo GM, Lothe RA, Myklebost O, Forus A. 2003. Amplification and overexpression of COPS3 in osteosarcomas potentially target TP53 for proteasome-mediated degradation. Oncogene 22:5358–5361. Hoang BH, Kubo T, Healey JH, Sowers R, Mazza BA, Yang R, Huvos AG, Meyers PA, Gorlick R. 2004a. Expression of LDL receptor-related protein (LRP5) as a novel marker for disease progression in high-grade osteosarcoma. Int J Cancer 109:106–111. Hoang BH, Kubo T, Healey JH, Yang R, Nathan SS, Kolb EA, Mazza BA, Meyers PA, Gorlick R. 2004b. Dickkopf 3 inhibits invasion and motility of Saos-2 osteosarcoma cells by modulating the Wnt-b-catenin pathway. Cancer Res. 64:2734–2739. Hornicek FJ, Gebhardt MC, Wolfe MW, Kharrazi FD, Takashita H, Parekh SG, et€al. 2000. P-glycoprotein levels predict poor outcome in patients with osteosarcoma. Clin Orthop Relat Res. 373:11–17. Houghton PJ, Maris JM, Friedman HS, Keir ST, Lock RB, Carol H, et€ al. 2007a. Pediatric Preclinical Testing Program (PPTP) evaluation of the multi-targeted kinase inhibitor Sunitinib. In: American Association for Cancer Research Annual Meeting: Proceedings. Abstract #527. Houghton PJ, Morton CL, Kolb EA, Gorlick R, Lock R, Carol H, Reynolds CP, Maris JM, Keir ST, Billups CA, Smith MA. 2007b. Initial testing (stage 1) of the mTOR inhibitor rapamycin by the pediatric preclinical testing program. Pediatr Blood Cancer 50(4):799–805.
492
R. Ricafort and R. Gorlick
Houghton PJ, Morton CL, Kolb EA, Lock R, Carol H, Reynolds CP, et€al. 2008. Initial testing (stage 1) of the proteasome inhibitor bortezomib by the pediatric preclinical testing program. Pediatr Blood Cancer 50:37–45. Hudis C. 2007. Trastuzumab – mechanism of action and use in clinical practice. N Engl J Med 357: 39–51. Hughes DP, Thomas DG, Giordano TJ, Baker LH, McDonagh KT. 2004. Cell surface expression of epidermal growth factor receptor and her-2 with nuclear expression of her-4 in primary osteosarcoma. Cancer Res. 64:2047–2053. Hughes DP, Thomas DG, Giordano TJ, McDonagh KT, Baker LH. 2006. Essential erbB family phosphorylation in osteosarcoma as a target for CI-1033 inhibition. Pediatr Blood Cancer 46:614–23. Ikeda S, Sumii H, Akiyama K, Watanabe S, Ito S, Inoue H, Takechi H, et€al. 1989. Amplification of both c-myc and c-raf oncogenes in a human osteosarcoma. Jpn J Cancer Res 80:6–9. Itoh Y, Tamai M, Yokogawa K, Nomura M, Moritani S, Suzuki H, et€al. 2002. Involvement of multidrug resistance-associated protein 2 in in€vivo cisplatin resistance of rat hepatoma AH66 cells. Anticancer Res 22:1649–53. Iwaya K, Ogawa H, Kuroda M, Izumi M, Ishida T, Mukai K. 2003. Cytoplasmic and/or nuclear staining of beta-catenin is associated with lung metastasis. Clin Exp Metastasis 20:525–529. Jacks T, Remington L, Williams BO, Schmitt EM, Halachmi S, Bronson RT, Weinberg RA. 1994. Tumor spectrum analysis in p53-mutant mice. Curr Biol. 4:1–7. Jain M, Arvanitis C, Chu K, Dewey W, Leonhardt E, Trinh M, Sundberg CD, Bishop JM, Felsher DW. 2002. Sustained loss of a neoplastic phenotype by brief inactivation of MYC. Science. 297:102–4. Judson I, Radford JA, Harris M, Blay JY, van Hoesel Q, le Cesne A, et€ al. 2001. Randomised phase II trial of pegylated liposomal doxorubicin (DOXIL/CAELYX) versus doxorubicin in the treatment of advanced or metastatic soft tissue sarcoma: a study by the EORTC Soft Tissue and Bone Sarcoma Group. Eur J Cancer 37:870–7. Jung ST, Moon ES, Seo HY, Kim JS, Kim GJ, Kim YK. 2005. Expression and significance of TGF-beta isoform and VEGF in osteosarcoma. Orthopedics 28:755–760. Kaplan RN, Riba RD, Zacharoulis S, Bramley AH, Vincent L, Costa C et€ al. 2005. VEGFR1positive haematopoietic bone marrow progenitors initiate the pre-metastatic niche. Nature 438:820–827. Kappel CC, Velez-Yanguas MC, Hirschfeld S, and Helman LJ. 1994. Human osteosarcoma cell lines are dependent on insulin-like growth factor I for in€vitro growth. Cancer Res 54:2803–2807. Kawai T, Hiroi S, Torikata C. 1997. Expression in lung carcinomas of platelet-derived growth factor and its receptors. Lab Invest 77:431–436. Kaya M, Wada T, Akatsuka T, Kawaguchi S, Nagoya S, Shindoh M, et€al. 2000. Vascular endothelial growth factor expression in untreated osteosarcoma is predictive of pulmonary metastasis and poor prognosis. Clin Cancer Res 6:572–7. Kaya M, Wada T, Kawaguchi S, Nagoya S, Yamashita T, Abe Y, et€al. 2002. Increased pre-therapeutic serum vascular endothelial growth factor in patients with early clinical relapse of osteosarcoma. Br J Cancer 6:864–9. Kempf-Bielack B, Bielack SS, Jurgens H, Branscheid D, Berdel WE, Exner GU, et€ al. 2005. Osteosarcoma relapse after combined modality therapy: an analysis of unselected patients in the Cooperative Osteosarcoma Study Group (COSS). J Clin Oncol 23:559–68. Khanna C, Prehn J, Yeung C, Caylor J, Tsokos M, Helman L. 2000. An orthotopic model of murine osteosarcoma with clonally related variants differing in pulmonary metastatic potential. Clin Exp Metastasis 18:261–271. Khanna C, Khan J, Nguyen P, Prehn J, Caylor J, Yeung C, et€al. 2001. Metastasis-associated differences in gene expression in a murine model of osteosarcoma. Cancer Res 61:3750–3759. Khanna C, Prehn J, Hayden D, Cassaday RD, Caylor J, Jacob S, et€al. 2002. A randomized controlled trial of octreotide pamoate long-acting release and carboplatin versus carboplatin alone in dogs with naturally occurring osteosarcoma: evaluation of insulin-like growth factor Â�suppression and chemotherapy. Clin Cancer Res 8:2406–12.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
493
Khanna C, Wan X, Bose S, et€al. 2004. The membrane-cytoskeleton linker ezrin is necessary for osteosarcoma metastasis. Nat Med 10:182–186. Khidr L, Chen PL. 2006. RB, the conductor that orchestrates life, death and differentiation. Oncogene 25:5210–5219. Kilpatrick SE, Geisinger KR, King TS, Sciarrotta J, Ward WG, Gold SH, Bos GD. 2001. Clinicopathologic analysis of HER-2/neu immunoexpression among various histologic subtypes and grades of osteosarcoma. Mod Pathol 14:1277–1283. Kleinerman ES. 1995. Biologic therapy for osteosarcoma using liposome-encapsulated muramyl tripeptide. Hematol Oncol Clin North Am 9:927–38. Kloen P, Gebhardt MC, Perez-Atayde A, Rosenberg AE, Springfield DS, Gold LI, Makin HJ. 1997. Expression of transforming growth factor-beta isoforms in osteosarcomas: TGF-b3 is related to disease progression. Cancer 80:2230–2239. Kolb EA, Gorlick R, Houghton PJ, Morton CL, Lock R, Carol H, Reynolds CP, Maris JM, Keir ST, Billups CA, Smith MA. 2008. Initial testing (stage 1) of a monoclonal antibody (SCH 717454) against the IGF-1 receptor by the pediatric preclinical testing program. Pediatr Blood Cancer 50(6):1190–1197. Knowles BB, McCarrick J, Fox N, Solter D, Damjanov I. 1990. Osteosarcomas in transgenic mice expressing an alpha-amylase-SV40 T-antigen hybrid gene. Am J Pathol.137:259–62. Kochevar DT, Kochevar J, Garrett L. 1990. Low level amplification of c-sis and c-myc in a spontaneous osteosarcoma model. Cancer Lett. 53:213–22. Krishnan K, Bruce B, Hewitt S, Thomas D, Khanna C, Helman LJ. 2006. Ezrin mediates growth and survival in Ewing’s sarcoma through the AKT/mTOR, but not the MAPK, signaling pathway. Clin Exp Metastasis. 23:227–36. Kubbutat MH, Jones SN, Vousden KH. 1997. Regulation of p53 stability by MDM2. Nature 387:299–303. Kubo T, Gorlick R, Meyers P, Sowers R, et€al. 2002. Expression and potential role of plateletderived growth factor receptor and c-Kit in relation to the antiproliferative effect of ST1571 in osteosarcoma. Proc Am Assoc Cancer Res 43:846. Kulkarni NH, Halladay DL, Miles RR, Gilbert LM, Frolik CA, Galvin RJS, Martin RJ, Gillespie MT, Onyia JE. 2005. Effects of parathyroid hormone on Wnt signaling pathway in bone. J Cell Biochem. 95:1178–1190. Ladanyi M, Cha C, Lewis R, Jhanwar SC, Huvos AG, Healey JH. 1993a. MDM2 gene amplification in metastatic osteosarcoma. Cancer Res 53:16–8. Ladanyi M, Park CK, Lewis R, Jhanwar SC, Healey JH, Huvos AG. 1993b. Sporadic amplification of the MYC gene in human osteosarcomas. Diagn Mol Pathol 2:63–167. Laverdiere C, Kolb EA, Supko JG, Gorlick R, Meyers PA, Maki RG, et€al. 2003. Phase II study of ecteinascidin 743 in heavily pretreated patients with recurrent osteosarcoma. Cancer 98:832–40. 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. 11:2561–7. Lawrence TS, Eishbruch A, McGinn CJ, Fields MT, Shewach DS. 1999. Radiosensitization by gemcitabine. Oncology 13:55–60. Lee J, Hoang B, Chou A, Meyers P, Healey J, Gorlick R. 2007. Expression of vascular endothelial growth factor correlates with chemotherapy response in osteosarcoma patients. Am Assoc Cancer Res Annu Meet: Proc. #3489. 48:830. Lee N, Smolarz AJ, Olson S, David O, Reiser J, Kutner R, et€al. 2007. A potential role for Dkk-1 in the pathogenesis of osteosarcoma predicts novel diagnostic and treatment strategies. Br J Cancer 97:1552–1559. Lee YH, Tokunaga T, Oshika Y, Suto R, Yanagisawa K, Tomisawa M, et€al. 1999. Cell-retained isoforms of vascular endothelial growth factor (VEGF) are correlated with poor prognosis in osteosarcoma. Eur J Cancer 35:1089–93. Li FP, Fraumeni JF Jr, Mulvihill JJ, Blattner WA, Dreyfus MG, Tucker MA, Miller RW. 1988. A cancer family syndrome in twenty-four kindreds. Cancer Res 48:5358–5362.
494
R. Ricafort and R. Gorlick
Libura J, Drukala J, Majka M, Dreyfus MG, Tucker MA, Miller RW. 2002. CXCR4-SDF-1 signalingÂ� is active in rhabdomyosarcoma cells and regulates locomotion, chemotaxis, and adhesion. Blood 100:2597–2606. Liu G, McDonnell TJ, Montes de Oca Luna R, Kapoor M, Mims B, El-Naggar AK, Lozano G. 2000. High metastatic potential in mice inheriting a targeted p53 missense mutation. Proc Natl Acad Sci USA. 97:4174–9. Lock R, Carol H, Houghton PJ, Morton CL, Kolb EA, Gorlick R, Reynolds CP, Maris JM, Keir ST, Wu J, Smith MA. 2008. Initial testing (stage 1) of the BH3 mimetic ABT-263 by the pediatric preclinical testing program. Pediatr Blood Cancer 50(6):1181–9. Lonardo F, Ueda T, Huvos AG, Healey J, Ladanyi M. 1997. p53 and MDM2 alterations in osteosarcomas: correlation with clinicopathologic features and proliferative rate. Cancer 79:1541–47. MacEwen EG, Kurzman ID, Rosenthal RC, Smith BW, Manley PA, Roush JK, Howard PE. 1989. Therapy of osteosarcoma in dogs with intravenous injection of liposome encapsulated muramyl tripeptide. J Natl Cancer Inst 81:935–936. MacEwen EG, Pastor J, Kutzke J, Tsan R, Kurzman ID, Thamm DH, et€al. 2004. IGF-1 receptor contributes to the malignant phenotype in human and canine osteosarcoma. J Cell Biochem 92:77–91. Maitra A, Wanzer D, Weinber AG, Ashfaq R. 2001. Amplification of the HER-2/neu oncogene is uncommon in pediatric osteosarcomas. Cancer 92:677–83. Malkin D, Li FP, Strong LC, Fraumeni JF, Jr, Nelson CE, Kim DH, et€ al. 1990. Germline p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science 250:1233–1238. Man TK, Lu XY, Jaeweon K, Perlaky L, Harris CP, Shah S, Ladanyi M, Gorlick R, Lau CC, Rao PH. 2004. Genome-wide array comparative genomic hybridization analysis reveals distinct amplifications in osteosarcoma. BMC Cancer. 4:45. Mansky PJ, Liewehr DJ, Steinberg SM, Chrousos GP, Avila NA, Long L, et€al. 2002. Treatment of metastatic osteosarcoma with the somatostatin analog OncoLar: significant reduction of insulin-like growth factor-1 serum levels. J Pediatr Hematol Oncol. 24:440–6. Mansour SJ, Matten WT, Hermann AS, Candia JM, Rong S, Fukasawa K, Vande Woude GF, Ahn NG. 1994. Transformation of mammalian cells by constitutively active MAP kinase kinase 1994. Science 265(5174):966–70. Maris JM, Courtright J, Houghton PJ, Morton CL, Gorlick R, Kolb EA, et al. 2008. Initial testing of the VEGFR inhibitor AZD2171 by the pediatric preclinical testing program. Pediatr Blood Cancer 50(3):581–7. McGary EC, Weber K, Mills L, Doucet M, Lewis V, Lev DC, et€al. 2002. Inhibition of plateletderived growth factor-mediated proliferation of osteosarcoma cells by the novel tyrosine kinase inhibitor STI571. Clin Cancer Res 8:3584–91. McIntyre JF, Smith-Sorensen B, Friend SH, Kassel, J, et€al. 1994. Germline mutations of the p53 tumor suppressor gene in children with osteosarcoma. J Clin Oncol. 12:925–30. Mendoza S, Konishi T, Miller CW. 1998. Integration of SV40 in human osteosarcoma DNA. Oncogene 17:2457–2462. Meyers PA, Gorlick R. 1997. Osteosarcoma. Pediatr Clin North Am 44:973–89. Meyers PA, Schwartz CL, Krailo M, Kleinerman ES, Betcher D, Bernstein ML, et€ al. 2005. Osteosarcoma: a randomized, prospective trial of the addition of ifosfamide and/or muramyl tripeptide to cisplatin, doxorubicin, and high-dose methotrexate. J Clin Oncol 23:2004–11. Miller AD, Curran T, Verma IM. 1984. c-fos protein can induce cellular transformation: a novel mechanism of activation of a cellular oncogene. Cell 36:51–60. Miyamoto N, Higuchi Y, Mori K, Ito M, Tsurudome M, Nishio M, et€al. 2002. Human osteosarcoma-derived cell lines produce soluble factor(s) that induces differentiation of blood monocytes to osteoclast-like cells. Int Immunopharmacol. 2:25–38. Muggia FM. 2001. Liposomal encapsulated anthracyclines: new therapeutic horizons. Curr Oncol Rep 3:156–62.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
495
Mulligan LM, Matlashewski GJ, Scrable HJ, Cavenee WK. 1990. Mechanisms of p53 loss in human sarcomas. Proc Natl Acad Sci USA 87:5863–7. Nellissery MJ, Padalecki SS, Brkanac Z, Singer FR, Roodman GD, Unni KK, Leach RJ, Hansen MF. 1998. Evidence for a novel osteosarcoma tumor-suppressor gene in the chromosome 18 region genetically linked with Paget disease of bone. Am J Hum Genet. 63:817–24. Nevins JR. 1992. E2F: a link between the Rb tumor suppressor protein and viral oncoproteins. Science 258:424–429. Nielsen GP, Burns KL, Rosenberg AE, Louis DN. 1998. CDKN2A gene deletions and loss of p16 expression occur in osteosarcomas that lack RB alterations. Am J Pathol 153(1):159–63. Nishijo K, Nakayama T, Aoyama T, Okamoto T, Ishibe T, Yasura K, Shima Y, Shibata KR, Tsuboyama T, Nakamura T, Toguchida J. 2004. Mutation analysis of the RECQL4 gene in sporadic osteosarcomas. Int J Cancer 111:367–372. Oda Y, Naka T, Takeshita M, Iwamoto Y, Tsuneyoshi M. 2000. Comparison of histological changes and changes in nm23 and c-MET expression between primary and metastatic sties in osteosarcoma: a clinicopathologic and immunohistochemical study. Hum Pathol 31:709–16. Ohta S, Hiraki Y, Shigeno C, Suzuki F, Kasai R, Ikeda T, Kohno H, Lee K, Kikuchi H, Konishi J, et€al. 1992. Bone morphogenetic proteins (BMP-2 and BMP-3) induce the late phase expression of the proto-oncogene c-fos in murine osteoblastic MC3T3-E1 cells. FEBS Lett. 314:356–360. O’Keefe RJ, Guise TA. 2003. Molecular mechanisms of bone metastasis and therapeutic implications. Clin Orthop Relat Res 415(Suppl):S100–S104. Oliner JD, Kinzler KW, Meltzer PS, George DL, Vogelstein B. 1992. Amplification of a gene encoding a p53-associated protein in human sarcomas. Nature 358:80–3. Oliner JD, Pietenpol JA, Thiagalingam S, Gyuris J, Kinzler KW, Vogelstein B. 1993. Oncoprotein MDM2 conceals the activation domain of tumor suppressor p53. Nature 362:857–60. Olive KP, Tuveson DA, Ruhe ZC, Yin B, Willis NA, Bronson RT, Crowley D, Jacks T. 2004. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 119:847–60. Onda M, Matsuda S, Higaki S, Iijima T, Fukushima J, Yokokura A, Kojima T, Horiuchi H, Kurokawa T, Yamamoto T. 1996. ErbB-2 expression is correlated with poor prognosis for patients with osteosarcoma. Cancer 77:71–78. Ory B, Heymann MF, Kamijo A, Gouin F, Heymann D, Redini F. 2005. Zoledronic acid suppresses lung metastases and prolongs overall survival of osteosarcoma-bearing mice. Cancer 104:2522–9. Pakos EE, Ioannidis JPA. 2003. The association of P-glycoprotein with response to chemotherapy and clinical outcome in patients with osteosarcoma: A meta-analysis. Cancer 98:581–589. Park YB, Kim HS, Oh JH, Lee SH. 2001. The co-expression of p53 protein and P-glycoprotein is correlated to a poor prognosis in osteosarcoma. Int Orthop 24:307–310. Perissinotto E, Cavalloni G, Leone F, Fonsato V, Mitola S, Grignani G, Surrenti N, Sangiolo D, Bussolino F, Piacibello W, Aglietta M. 2005. Involvement of chemokine receptor 4/stromal cellderived factor 1 system during osteosarcoma tumor progression. Clin Cancer Res. 11:490–497. Polakis P. 2000. Wnt signaling and cancer. Genes Dev. 14:1837–1851. Pollak MN, Polychronakos C, Richard M. 1990. Insulinlike growth factor I: a potent mitogen for human osteogenic sarcoma. J Natl Cancer Inst 82:301–5. Pompetti F, Rizzo P, Simon RM, Freidlin B, Mew DJ, Pass HI, Picci P, Levine AS, Carbone M. 1996. Oncogene alterations in primary, recurrent, and metastatic human bone tumors. J Cell Biochem 63:37–50. Quelle DE, Zindy F, Ashmun RA, Sherr CJ. 1995. Alternative reading frames of the INK4a tumor suppressor gene encode two unrelated proteins capable of inducing cell cycle arrest. Cell 83:993–1000. Rong S, Jeffers M, Resau JH, Tsarfaty I, Oskarsson M, Vande Woude GF. 1993. Met expression and sarcoma tumorigenecity. Cancer Res. 53:5355–60.
496
R. Ricafort and R. Gorlick
Rosen G, Caparros B, Huvos AG, Kosloff C, Nirenberg A, Cacavio A, et€al. 1982. Preoperative chemotherapy for osteogenic sarcoma: selection of postoperative adjuvant chemotherapy based on the response of the primary tumor to preoperative chemotherapy. Cancer 49:1221–30. Ruther U, Komitowski D, Schubert FR, Wagner EF. 1989. C-fos expression induces bone tumors in transgenic mice. Oncogene 4:861–865. Sachdev D, Li SL, Hartell JS, Fujita-Yamaguchi Y, Miller JS, Yee D. 2003. A chimeric humanized single-chain antibody against the type I insulin-like growth factor (IGF) receptor renders breast cancer cells refractory to the mitogenic effects of IGF-I. Cancer Res. 63:627–635. Salasznyk RM, Klees RF, Hughlock MK, Plopper GE. 2004. ERK signaling pathways regulate the osteogenic differentiation of human mesenchymal stem cells on collagen I and vitronectin. Cell Commun Adhes 11:137–53. Sanders RP, Drissi R, Billups CA, Daw NC, Valentine MB, Dome JS. 2004. Telomerase expression predicts unfavorable outcome in osteosarcoma. J Clin Oncol. 22: 3790–7. Sangiorgi L, Gobbi GA, Lucarelli E, Sartorio SM, Mordenti M, Ghedini I, Maini V, Scrimieri F, Reggiani M, Bertoja AZ, Benassi MS, Picci P. 2001. Presence of telomerase activity in different musculoskeletal tumor hitotypes and correlation with aggressiveness. Int J Cancer 95:156–161. Schwartz CL, Gorlick R, Teot LA, Grier HE, Krailo M, Meyers PA. 2000. A prospective study of p-glycoprotein (P-GP) expression in newly diagnosed non-metastatic osteosarcoma: An intergroup study of the Children’s Cancer Group and the Pediatric Oncology Group. Proc Am Soc Clin Oncol. 19:587a. Schwartz CL, Gorlick R, Teot L, Krailo M, Chen Z, Goorin A, Grier HE, Bernstein ML, Meyers P. 2007. Mulitple drug resistance in osteogenic sarcoma: INT0133 from the Children’s Oncology Group. J Clin Oncol. 25:2057–2062. Scotlandi K, Baldini N, Oliviero M, Di Renzo MF, Martano M, Serra M, Manara MC, Comoglio PM, Ferracini R. 1996. Expression of Met/Hepatocyte growth factor receptor gene and malignant behavior of musculoskeletal tumors. Am J Pathol 149:1209–19. Scotlandi K, Manara MC, Hattinger CM, Benini S, Perdichizzi S, Pasello M, et€ al. 2004. Prognostic and therapeutic relevance of HER2 expression in osteosarcoma and Ewing’s sarcoma. Eur J Cancer 41:1349–61. Scotlandi K, Manara MC, Nicoletti G, Lollini PL, Lukas S, Benini S, et€al. 2005. Antitumor activity of the insulin-like growth factor-I receptor kinase inhibitor NVP-AEW541 in musculoskeletal tumors. Cancer Res. 65:3868–76. Sekyi-Otu A, Bell RS, Ohashi C, Pollak M, and Andrulis IL. 1995. Insulin-like growth factor 1 (IGF-1) receptors, IGF-1, and IGF-2 are expressed in primary human sarcomas. Cancer Res. 55:129–134. Serra M, Scotlandi K, Reverter-Branchat G, Ferrari S, Manara MC, Benini S, Incaprera M, Bertoni F, Mercuri M, Briccoli A, Bacci G, Picci P. 2003. Value of P-glycoprotein and clinicopathologic factors as the basis for new treatment strategies in high-grade osteosarcoma of the extremities. J Clin Oncol. 21:536–542. Siddiqui R, Onel K, Facio F, Nafa K, Diaz LR, Kauff N, Huang H, Robson M, Ellis N, Offit K. 2005. The TP53 mutational spectrum and frequency of CHEK2*1100delC in Li-Fraumeni-like kindreds. Fam Cancer 4:177–181. Skubitz KM. 2003. Phase II trial of pegylated-liposomal doxorubicin (Doxil) in sarcoma. Cancer Invest. 21:167–76. Somers GR, Ho M, Zielenska M, Squire JA, Thorner PS. 2005. HER2 amplification and overexpression is not present in pediatric osteosarcoma: a tissue microarray study. Pediatr Dev Pathol. 8:525–32. Stock C, Kager L, Fink FM, Gadner H, Ambros PF. 2000. Chromosomal regions involved in the pathogenesis of osteosarcomas. Genes Chromosomes Cancer 28:329–336. Sulzbacher I, Traxler M, Mosberger I, Lang S, and Chott A. 2000. Platelet-derived growth factorAA and -alpha receptor expression suggests an autocrine and/or paracrine loop in osteosarcoma. Mod Pathol. 13:632–637. Sulzbacher I, Birner P, Trieb K, Muhlbauer M, Lang S, and Chott A. 2001. Platelet-derived growth factor-alpha receptor expression supports the growth of conventional chondrosarcoma and is associated with adverse outcome. Am J Surg Pathol. 25:1520–1527.
Molecularly Targeted Therapy for Osteosarcoma: Where Do We Go from Here?
497
Taichman RS, Cooper C, Keller ET, Penta KJ, Taichman NS, McCauley LK. 2002. Use of the stromal cell-derived factor-1/CXCR4 pathway in prostate cancer metastasis to bone. Cancer Res. 62:1832–1837. Thimmaiah KN, Easton J, Huang S, Veverka KA, Germain GS, Harwood FC, Houghton PJ. 2003. Insulin-like growth factor I-mediated protection from rapamycin-induced apoptosis is independent of Ras-Erk1-Erk2 and phosphatidylinositol 3¢-kinase-Akt signaling pathways. Cancer Res. 63:364–74. Thomas DG, Giordano TJ, Sanders D, Biermann JS, Baker L. 2002. Absence of HER2/neu gene expression in osteosarcoma and skeletal Ewing’s sarcoma. Clin Cancer Res. 8:788–793. Thomas JE, Venugopalan M, Galvin R, Wang Y, Bokoch GM, Vlahos CJ. 1997. Inhibition of MG-63 cell proliferation and PDGF-stimulated cellular processes by inhibitors of phosphatidylinositol 3-kinase. J Cell Biochem. 64:182–95. Toguchida J, Ishizaki K, Sasaki MS, Ikenaga M, Sugimoto M, Kotoura Y et€al. 1988. Chromosomal reorganization for the expression of recessive mutation of retinoblastoma susceptibility gene in the development of osteosarcoma. Caner Res. 48:3939–3943. Tolar J, Nauta AJ, Osborn MJ, Panoskaltsis Mortari A, McElmurry RT, Bell S, et€ al. 2007. Sarcoma derived from cultured mesenchymal stem cells. Stem Cells. 25:371–9. Ulaner GA, Hoffman AR, Otero J, Huang HY, Zhao Z, Mazumdar M, Gorlick R, Meyers P, Healey JH, Ladanyi M. 2004. Divergent patterns of telomere maintenance mechanisms among human sarcomas: sharply contrasting prevalence of the alternative lengthening of telomeres mechanism in Ewing’s sarcomas and osteosarcomas. Genes Chromosomes Cancer 41:155–62. Ulaner GA, Huang HY, Otero J, Zhao Z, Ben-Porat L, Satagopan JM, Gorlick R, Meyers P, Healey JH, Huvos AG, Hoffman AR, Ladanyi M. 2003. Absence of telomere maintenance mechanism as a favorable prognostic factor in patients with osteosarcoma. Cancer Res 63:1759–1763. Vahle JL, Sato M, Long GG, Young JK, Francis PC, Engelhardt JA, Westmore MS, Linda Y, Nold JB. 2002. Skeletal changes in rats given daily subcutaneous injections of recombinant human parathyroid hormone (1-34) for 2 years and relevance to human safety. Toxicol Pathol 30:312–21. Van BC, van SF, Curran T, Muller R, Verma IM. 1983. Analysis of FBJ-MuSV provirus and c-fos (mouse) gene reveals that viral and cellular fos gene products have different carboxy termini. Cell 32:1241–1255. Van Dartel M, Redeker S, Bras J, Kool M, Hulsebos TJM. 2004. Overexpression through amplification of genes in chromosome region 17q11.2-p12 in high-grade osteosarcoma. Cancer Genet Cytogenet. 152:8–14. Van Dartel M, Hulsebos TJM. 2004. Amplification and overexpression of genes in 17p11.2-p12 in osteosarcoma. Cancer Genet Cytogenet. 153:77–80. Vilella-Bach M, Nuzzi P, Fang Y, Chen J. The FKBP12-rapamycin-binding domain is required for FKBP12-rapamycin-associated protein kinase activity and G1 progression. J Biol Chem. 1999 Feb 12;274(7):4266–72. Wadayama B, Toguchida J, Shimizu T, Ishizaki K, Sasaki MS, Kotoura Y et€al. 1994. Mutation spectrum of the retinoblastoma gene in osteosarcomas. Cancer Res. 54:3042–3048. Wan X, Mendoza A, Khanna C, Helman LJ. 2005. Rapamycin inhibits ezrin-mediated metastatic behavior in a murine model of osteosarcoma. Cancer Res. 65:2406–11. Wang LL, Levy ML, Lewis RA, Chintagumpala MM, Lev D, Rogers M, Plon SE. 2001. Clinical manifestations in a cohort of 41 Rothmund-Thomson syndrome patients. Am J Med Genet. 102:11–17. Wang LL, Gannavarapu A, Kozinetz CA, Levy ML, Lewis RA, Chintagumpala MM, RuizMaldanado R, Contreras-Ruiz J, Cunniff C, Erickson RP, Lev D, Rogers M, Zackai EH, Plon SE. 2003. Association between osteosarcoma and deleterious mutations in the RECQL4 gene in Rothmund-Thomson syndrome. J Natl Cancer Inst. 95:669–674. Wang ZQ, Liang J, Schellander K, Wagner EF, Grigoriadis AE. 1995. c-fos-induced osteosarcoma formation in transgenic mice: cooperativity with c-jun and the role of endogenous c-fos. Cancer Res. 55:6244–51.
498
R. Ricafort and R. Gorlick
Wei G, Lonardo F, Ueda T, Kim T, Huvos AG, Healey JH, et€al. 1999a. CDK4 gene amplification in osteosarcoma: reciprocal relationship with INK4A gene alterations and mapping of 12q13 amplicons. Int J Cancer 80:199–204. Wei G, Gorlick R, Ladanyi M, Meyers P, Huvos AG, Bertino JR, Healey JH. 1999b. Expression of bone morphogenetic proteins and receptors in sarcomas. Clin Orthop Relat Res. 365:175–183. Weiner LM, Adams GP. 2000. New approaches to antibody therapy. Oncogene 19:6144–6151. Weiss KR, Cooper GM, Jadlowiec JA, McGough RL, Huard J. 2006. VEGF and BMP expression in mouse osteosarcoma cells. Clin Orthop Relat Res. 450:111–177. Wen YH, Koeppen H, Garcia R, Chiriboga L, Tarlow BD, Peters BA, Eigenbrot C, Yee H, Steiner G, Greco MA. 2007. Epidermal growth factor receptor in osteosarcoma: expression and mutational analysis. Human Pathol 38:1184–91. Wittrant Y, Lamoureux F, Mori K, Riet A, Kamijo A, Heymann D, et€al. 2006. RANKL directly induces bone morphogenetic protein-2 expression in RANK-expressing POS-1 osteosarcoma cells. Int J Oncol. 28:261–9. Wong FL, Boice JD, Abramson DH, Tarone RE, Kleinerman RA, Stoval M, et€al. 1997. Cancer incidence after retinoblastoma: radiation dose and sarcoma risk. JAMA 278:1262–1267. Wu JX, Carpenter PM, Gresens C, Keh R, Niman H, Morris JW, Mercola D. 1990. The protooncogene c-fos is over-expressed in the majority of human osteosarcomas. Oncogene 5:989–1000. Wunder JS, Bell RS, Wold L, Adrulis IL. 1993. Expression of the multidrug resistance gene in osteosarcoma: a pilot study. J Othopaed Res 11:396–403. Wunder JS, Bull SB, Aneliunas V, Lee PD, Davis AM, Beauchamp CP, Conrad EU, Grimer RJ, Healey JH, Rock MJ, Bell RS, Andrulis IL. 2000. MDR1 gene expression and outcome in osteosarcoma: a prospective, multicenter study. J Clin Oncol 18:2685–2694. Wuyts W, Van Wesenbeeck L, Morales-Piga A, Ralston S, Hocking L, Vanhoenacker F, et€ al. 2001. Evaluation of the role of RANK and OPG genes in Paget’s disease of bone. Bone 28:104–7. Yamamoto O, Wda T, Takahashi M, Yamawaki S, Ishii S. 2000. Prognostic value of P-glycoprotein expression in bone and soft-tissue sarcoma. Int J Clin Oncol. 5:164–170. Yang R, Hoang BH, Kubo T, Kawano H, Chou A, Sowers R, et€ al. 2007a. Over-expression of parathyroid hormone type 1 receptor confers an aggressive phenotype in osteosarcoma. Int J Cancer 121(5):943–54. Yang R, Sowers R, Mazza B, Healey JH, Huvos A, Grier H, et€al. 2003. Sequence alterations in the reduced folate carrier are observed in osteosarcoma tumor samples. Clin Cancer Res. 9:837–44. Yang R, Kolb EA, Qin J, Chou A, Sowers R, Hoang B, et€al. 2007b. The folate receptor alpha is frequently overexpressed in osteosarcoma samples and plays a role in the uptake of the physiologic substrate 5-methyltetrahydrofolate. Clin Cancer Res. 13:2557–67. Yoshikawa H, Takaoka K, Masuhara K, Ono K, Sakamoto Y. 1988. Prognostic significance of bone morphogenetic activity in osteosarcoma tissue. Cancer 61:569–573. Yu W, Chen Z, Zhang J, Zhang L, Ke H, Huang L, Peng Y, Zhang X, Li S, Lahn BT, Xiang AP. 2008. Critical role of phosphoinositide 3-kinase cascade in adipogenesis of human mesenchymal stem cells. Mol Cell Biochem 310(1–2):11–8. Zhou H, Randall RL, Brothman AR, Maxwell T, Coffin CM, Goldsby RE. 2003. Her-2/neu expression in osteosarcoma increases risk of lung metastasis and can be associated with gene amplification. J Pediatr Hematol Oncol. 25:27–32.
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments Based on a Better Understanding of Disease Biology Stephen X. Skapek
Introduction Nonrhabdomyosarcoma soft tissue sarcomas (NRSTS) are a large and very Â�heterogeneous group of cancers in children. Although approximately 550 NRSTS diagnoses in the US each year represent only approximately 4% of cases of childhood cancer, NRSTS comprise nearly 60% of all soft tissue sarcomas in this age group [reviewed in Spunt et€ al. (2006)]. The incidence is bimodally distributed, with relatively high incidence in the first year of life and a second peak during later childhood and adolescence. While NRSTS is typically a sporadic disease, some types of NRSTS are associated with cancer susceptibility syndromes, such as the Li–Fraumeni familial cancer syndrome (associated with heritable p53 mutations) and neurofibromatosis type I (associated with heritable mutations in NF1 tumor suppressor genes) (Li et€al. 1988; Malkin et€al. 1990; Sorensen et€al. 1986); with certain environmental exposures (therapeutic ionizing radiation and HHV8, HIV, or EBV virus infection); and with certain chemical carcinogens (Spunt et€al. 2006). Historically, therapy for NRSTS has comprised surgery and radiation for local tumor control and systemic cytotoxic chemotherapy to eradicate disseminated disease. Important prognostic factors include the histological grade of the tumor and the extent of disease. A sarcoma grading system validated for children was developed by the Pediatric Oncology Group (POG) (Parham et€ al. 1995) (Table€1). In many cases, pathological grade is reflected in the histological subtype; in others, grade is determined by morphologic criteria, such as necrosis of more than 15% of the tumor surface or a mitotic count greater than 5 per 10
S.X. Skapekâ•›(*) Pediatric Hematology/Oncology, The University of Chicago Comer Children’s Hospital, 900 E. 57th Street, Chicago, IL 60637, USA e-mail:
[email protected] P.J. Houghton and R.J. Arceci (eds.), Molecularly Targeted Therapy for Childhood Cancer, DOI 10.1007/978-0-387-69062-9_22, © Springer Science+Business Media, LLC 2010
499
500
S.X. Skapek
Table€ 1â•… Pediatric Oncology Group (POG) nonrhabdomyosarcoma soft tissue sarcoma grading system Grade I Myxoid and well-differentiated liposarcoma Deep-seated dermatofibrosarcoma protuberans Well-differentiated or infantile (£4 years old) fibrosarcoma Well-differentiated or infantile (£4 years old) hemangiopericytoma Well-differentiated malignant peripheral nerve sheath tumor Extraskeletal myxoid chondrosarcoma Angiomatoid malignant fibrous histiocytoma Grade II Sarcomas not specifically included in Grades I and III, and in which <15% of the surface area shows necrosis, and the mitotic count is £5/10 high power fields (hpf) using a 40× objective. As secondary criteria, nuclear atypia is not marked and the tumor is not markedly cellular Grade III Pleomorphic or round cell liposarcoma Mesenchymal chondrosarcoma Extraskeletal osteosarcoma Malignant triton tumor Alveolar soft part sarcoma Sarcomas not included in Grade I and with >15% of surface area with necrosis, or with ³5 mitoses/10 hpf using a 40× objective Marked atypia or cellularity are less predictive but may assist in placing tumors in this category Reference: Parham et€al. (1995)
high-power fields. Perhaps the most important parameter for determining��survival is the presence or absence of metastasis: large retrospective studies showed overall 5-year survival to be 89% in children with localized, resected NRSTS (Ferrari et€al. 2005; Spunt et€al. 1999) but less than 20% for those with �metastatic disease (Pappo et€al. 1999). However, overall 5-year survival even in children with localized NRSTS is only approximately 50% unless the primary tumor is initially resected (Ferrari et€ al. 2005; Spunt et€ al. 2002). Among patients with resected tumors, survival is poorer if the tumor is greater than 5€cm in size or is locally invasive (Ferrari et€al. 2005; Spunt et€al. 1999). Systemic chemotherapy for children with NRSTS has typically included alkylating agents and anthracyclines. Standard chemotherapy has failed to improve disease control in most retrospective studies [e.g., Spunt et€ al. (1999)] and in prospective POG trials that (1) compared adjuvant vincristine, actinomycin D, cyclophosphamide, and doxorubicin (VACA) to observation alone (Pratt et€al. 1999) and (2) assessed the addition of dacarbazine to VACA chemotherapy (Pratt et€al. 1998). These particular clinical trials are limited by their inclusion of a variety of NRSTS. There is hope that the next therapeutic breakthrough for high-risk disease will come from better understanding of NRSTS biology and the development of therapeutic agents that specifically block molecular events crucial to the malignant features of the tumor.
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
501
NRSTS Biology Two Fundamental Groups of NRSTS There are many histological subtypes of NRSTS in children; synovial sarcoma and malignant peripheral nerve sheath tumor (MPNST) are the most common [e.g., Ferrari et€al. (2005)]. The clinical behavior of most NRSTS depends on the histological grade and is similar in children and adults. Infantile fibrosarcoma and infantile hemangiopericytoma (a very rare childhood NRSTS) are exceptions to this rule. They are defined as occurring in children younger than 4 years and are considered low grade tumors despite their histological appearance, due to their relatively good outcome (Parham et€al. 1995). Despite the daunting complexity of the numerous histological subtypes, childhood NRSTS can be divided into two fundamentally different groups: tumors with complex genomic abnormalities (e.g., aneuploidy, chromosomal breaks and translocations, and end-to-end chromosome fusions) and tumors with specific chromosomal rearrangements but relatively few cytogenetically evident genomic abnormalities [reviewed in Helman and Meltzer (2003)]. The mechanisms underlying generalized genomic instability have not been precisely defined, but some insight has been gleaned from mouse models. For example, p53 plays a critical role in preserving genomic integrity. Sarcomas frequently develop in mice with inactivated p53 or Arf (which activates p53) genes (Donehower et€al. 1992; Jacks et€al. 1994a; Kamijo et€al. 1999). In mice lacking the Ink4a/Arf locus, loss of one allele of the gene encoding DNA ligase IV, an enzyme involved in nonhomologous DNA end joining, causes a wide range of soft tissue sarcomas with chromosomal translocations, amplifications, and deletions similar to those observed in human NRSTS (Sharpless et€al. 2001). When bred into a line of mice with short telomeres because of telomerase deficiency, mice that lack functional p53 and Atm, which integrates DNA-damage sensing and repair mechanisms (Bakkenist and Kastan 2004), develop lymphomas with complex karyotypic abnormalities, some of which influence genes known to be involved in human cancer (Maser et€al. 2007). Similar mechanisms may underlie the development of NRSTS with complex karyotypes, but specific studies are needed to identify potential therapeutic targets. Development of NRSTS in the presence vs. absence of global genomic derangement is likely to be fundamentally different. Most of these sarcomas have specific chromosomal defects initially identified by routine karyotypic analysis. In many cases, specific defects are now detectable by sensitive molecular techniques [such as fluorescence in situ hybridization (FISH)] even when the karyotype appears grossly normal. Many of the translocations generate novel fusion proteins (often transcriptional regulators) that presumably drive sarcomagenesis. The striking correlation of specific chromosomal rearrangements with histological subtype suggests their key role in driving the biology of particular sarcomas. Nonetheless, a single genetic event is not sufficient to cause malignancy. Therefore, even in cases of NRSTS with
502
S.X. Skapek
specific chromosomal translocations, additional genetic or epigenetic events are required to confer all of the “hallmarks” of cancer (Hanahan and Weinberg 2000).
What Is the Cellular Origin of NRSTS? NRSTS, like other sarcomas, are derived from mesenchymal cells. In some cases, for example, MPNST arising from a preexisting benign, plexiform neurofibroma, the cellular origin of the sarcoma is easy to explain. In most cases, however, the exact cell of origin is not known. In principle, different histological subtypes of NRSTS may originate from different types of mesenchymal cells. As in lymphoblastic leukemia, the relative degree of cellular differentiation within a class of tumors may indicate transformation at an earlier or later stage of development. It is also plausible that a common mesenchymal progenitor cell may give rise to different types of NRSTS. Just as a single transcription factor (MyoD) can determine the skeletal muscle differentiation of mesenchymal progenitors, NRSTS histology could be driven by abnormal transcription factors, such as EWS(or TLS)-CHOP, EWS-ATF1, and EWS-WT1, generated by tumor-restricted balanced chromosomal translocations. These fusion proteins are expressed exclusively in myxoid liposarcoma, clear cell sarcoma of soft parts, and desmoplastic small round cell tumor [reviewed in Skapek and Chui (2000), Spunt et€al. (2006)].
Coupling Molecular Biology to Targeted Therapy for Childhood NRSTS Although there are many histological subtypes of NRSTS, current treatment strategies are largely independent of subtype and are instead based on pathological stage and extent of disease. If the molecular basis of individual NRSTS subtypes is clarified, it may be possible to direct subtype-specific therapy at essential molecular targets. The following sections review how understanding of the biology of the more common childhood NRSTS subtypes has led or may lead to targeted therapies. They also provide examples of less common tumors for which targeted therapies have already been identified.
Synovial Sarcoma and the SYT-SSX Gene Fusions Synovial sarcoma is the most common NRSTS in children, accounting for �approximately 27 to 30% of cases (Ferrari et€al. 2005; Parham et€al. 1995). Although it �commonly arises near joints, the cell of origin and its specific relationship to the
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
503
joint synovium is not clear. Histologically, synovial sarcoma is composed either entirely of spindle-shaped mesenchymal cells or an admixture of spindle and Â�epithelial cells; these subtypes are termed monophasic and biphasic synovial sarcoma, respectively (Spunt et€al. 2006). As in most types of NRSTS, surgical resection with or without ionizing radiation is the standard of care for localized disease. Important insight into the biology of synovial sarcoma came from the identification of nonrandom translocations between chromosomes X and 18 [t(X;18)(p11.2; q11.2)] in some cases (Turc-Carel et€ al. 1987). Cloning and sequencing of the translocation breakpoints showed that the translocation fuses most of the SYT gene (encoding all but the last eight amino acids) on chromosome 18 to DNA encoding the carboxyl end of either SSX1 or SSX2 [(Clark et€al. 1994; Ladanyi 2001)(review)]. Alternative translocations generating fusion proteins with similar motifs have occasionally been reported [e.g., Storlazzi et€al. (2003)]. Molecular diagnostic studies have established that the SYT-SSX1 or SYT-SSX2 transcript (detected by reversetranscriptase polymerase chain reaction) is highly specific for synovial sarcoma and is present in almost all cases. The presence of SYT-SSX1 vs. SYT-SSX2 is correlated with histological subtype: the SYT-SSX1 fusion is found in both monophasic and biphasic tumors, in both histologic “phases” (Birdsall et€al. 1999; Kawai et€al. 1998). In contrast, SYT-SSX2 is found mostly in monophasic tumors. This finding implies that SYT-SSX2 may influence the expression of genes required for mesenchymal-to-epithelial cell transition. While some studies show that the SYT-SSX1 fusion portends a somewhat poorer outcome (Kawai et€ al. 1998; Nilsson et€ al. 1999), this conclusion is debated (Guillou et€al. 2004). The molecular biology of the SYT-SSX1 and SYT-SSX2 fusion proteins is not fully understood [reviewed in Ladanyi (2001)]. Functional motifs in the SYT protein that are preserved in the fusion include interaction domains for (1) the transcriptional coactivator p300; (2) Brg1 and BRM (components of the SWI/SNF chromatin remodeling machinery); and (3) a QPGY motif acting as a transactivation domain. Normal SSX proteins include an amino-terminal region similar to a Kruppelassociated box (KRAB) transcriptional repression domain (which is not present in the fusion protein) and a carboxy-terminal repressor domain (RD) that mediates colocalization with Polycomb group proteins (Pgps), which also remodel chromatin. The predominantly nuclear localization of SYT-SSX and the aforementioned interaction motifs imply that SYT-SSX acts as a transcriptional regulator. Unlike most other fusion transcription factors (see below), it has no sequence-specific DNA binding motif but likely influences transcription by interacting with other proteins. In this regard, SYT-SSX has been shown to repress Candidate of Metastasis 1 (COM1) (Ishida et€al. 2007) and E-cadherin (Saito et€al. 2006); to colocalize with b-catenin in the nucleus (Pretto et€al. 2006); to stabilize cyclin D1 (Xie et€al. 2002); and to induce insulin-like growth factor-2 (Igf2) (Sun et€al. 2006). Gene expression profiling and analyses of candidate proteins have also been used to identify molecular mechanisms operating downstream of SYT-SSX fusion (Allander et€ al. 2002; Lee et€ al. 2003; Nagayama et€ al. 2002). Gene expression clustering observed in synovial sarcoma cases provides insight into both its Â�possible neuroectodermal lineage and its occasional epithelial differentiation. Protein studies
504
S.X. Skapek
confirm a number of candidate gene products that could contribute to tumor biology: synovial sarcomas frequently express detectable levels of the insulin-like growth factor1 receptor 1 (IGF-1R) (Xie et€al. 1999); the epidermal growth factor receptor (EGFR) (Nielsen et€al. 2003); the antiapoptotic protein Bcl2 (Nielsen et€al. 2003); HER2/neu (Nielsen et€al. 2003; Nuciforo et€al. 2003); and platelet-derived growth factor receptor (PDGFR)-b (Tamborini et€ al. 2004). However, there are conflicting data on the frequency of c-KIT expression (Nielsen et€al. 2003; Smithey et€ al. 2002; Tamborini et€ al. 2004). More compelling evidence that a particular protein may drive tumor biology can come from genetic studies. For example, an activating mutation in the gene encoding PDGFR-a, which was found in one in a series of 12 cases, implies a causative role (Lopez-Guerrero et€al. 2005). Similarly, frequent deletion of CDKN2A, which encodes the Cdk4/6-specific inhibitor p16Ink4a, suggests that Cdk4/6 inhibition may impede synovial sarcoma development or progression (Subramaniam et€al. 2006). The importance of some of the aforementioned candidate therapeutic targets has been validated in experimental models. The SYT-SSX protein associates with chromatin remodeling proteins to repress gene expression; the histone deacetylase inhibitor FK228 (whose net effect can block gene repression) inhibits synovial sarcoma cell growth in€vitro and in xenograft models (Ito et€al. 2005). The antiapoptotic protein Bcl-2 is expressed in many synovial sarcomas; the Bcl-2-directed antisense oligonucleotide G3139 enhances doxorubicin-mediated apoptosis in two synovial sarcoma cell lines (Joyner et€al. 2006). SYT-SSX1 can induce insulin-like growth factor (IGF)-2; blocking this signaling pathway promotes apoptosis and impedes tumor formation in cells expressing the fusion protein (Sun et€al. 2006). Members of the Capecchi laboratory recently reported the generation of a mouse in which the conditional expression of human SYT-SSX2 causes a synovial sarcomalike tumor (Haldar et€ al. 2007); this elegant model should be valuable in further clarifying the genetic and biochemical events that cooperate with SYT-SSX expression to drive synovial sarcomagenesis and in testing novel therapeutic strategies.
Malignant Peripheral Nerve Sheath Tumor and NF1 Mutation MPNST is the second most common NRSTS in children [e.g., Ferrari et€al. (2005)]. It can occur in children with neurofibromatosis type 1 or sporadically in children with no known predisposition. In either case, the tumor often arises near major peripheral nerves, nerve roots, or the brachial plexus or from a preexisting plexiform neurofibroma [reviewed in Webber and Parham (1996)]. Although MPNST is commonly accepted as arising from Schwann cells, it must be emphasized that both plexiform neurofibromas and MPNST also contain fibroblasts, perineurial cells, and mast cells that can contribute to their biology (discussed more below). Surgical resection with or without ionizing radiation is central to the effective treatment. The use of intensive ifosfamide and doxorubicin-based therapy has not greatly influenced survival.
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
505
Cytogenetic and molecular genetic analyses have not provided major insight into MPNST biology, instead revealing a variety of complex genomic abnormalities involving chromosome arms 7p, 9p, 17q, and 22q (Skapek and Chui 2000). However, the close link of MPNST with plexiform neurofibromas and neurofibromatosis type 1 strongly implies that loss of the NF1 tumor suppressor gene plays a role. Children who are born with only one normal NF1 allele are at increased risk of certain brain tumors, pheochromocytoma, myeloid leukemia, and a variety of other disorders, including cutaneous and plexiform neurofibromas [reviewed in Cichowski and Jacks (2001)]. A relatively small subset of the latter can degenerate into MPNST. NF1, residing at human chromosome 17q, encodes neurofibromin, which acts as a GTPase activating protein (GAP) (Cichowski and Jacks 2001). In response to signaling from a receptor tyrosine kinase, such as EGFR, the RAS protein is activated by binding to GTP and transduces a variety of signals that contribute to tumor biology. RAS signaling is “turned off” as its enzymatic activity converts bound GTP to GDP. As a GAP, neurofibromin facilitates this transition to curb RAS signaling; while RAS-independent effects of neurofibromin may exist, their importance to neurofibromin’s tumor suppression is not clear. While NF1 haploinsufficiency can alter the biology of perineurial cells and fibroblasts within a neurofibroma, there is good evidence that the normal NF1 allele is lost in the Schwann cells within these tumors. Additional genetic or epigenetic changes, such as loss of p53, p16INK4a, and p14ARF, appear to foster the transition from neurofibroma to MPNST. Importantly, key aspects of MPNST biology have been verified by using elegant mouse models. First, mere haploinsufficiency of mouse Nf1 predisposes mice to pheochromocytoma and myeloid leukemia but not to peripheral neurofibroma (Jacks et€al. 1994b). Because Nf1 knockout mice are not viable, members of the Jacks laboratory used mouse chimeras possessing both wild-type and Nf1−/− cells to show that the loss of the second Nf1 allele is essential for neurofibroma development (Cichowski et€al. 1999). Second, the combined loss of the second Nf1 allele and the p53 tumor suppressor causes MPNST and other malignant tumors found in children with neurofibromatosis type 1 (Cichowski et€al. 1999; Vogel et€al. 1999). Third, by proving that neurofibromas occur in mice when Nf1 is specifically inactivated in Schwann cells, members of the Parada laboratory provided solid evidence that neurofibromas originate from these cells (Zhu et€al. 2002). Interestingly, they also showed that Nf1 haploinsufficiency in non-Schwann cells greatly enhances the development of peripheral neurofibromas, underlining both the complexity of heterotypic cell–cell interactions in this disease and pointing out the potential for therapeutic avenues distinct from those directed at Schwann cells. One important purpose of better defining the molecular biology of neurofibroma and MPNST formation is the development of rational therapy. Some important progress has been made. One obvious strategy is to block farnesyl protein transferase, which promotes RAS-dependent signaling by altering its subcellular localization (reviewed in Dilworth et€ al. (2006)). A number of so-called farnesyl transferase inhibitors (FTIs), such as lonafarnib, tipifarnib, BMS-214662, FTI-277, L-744,832, and BMS-186511, have been tested in preclinical models with some
506
S.X. Skapek
success. However, because N- and K-RAS can also be activated by geranylgeranyl transferases, FTIs alone may not be sufficient. On the basis of EGFR expression and signaling activity in Schwann cells and MPNSTs from Nf1-deficient mice (DeClue et€al. 2000; Li et€al. 2002), it was recently shown that the genetic loss of EGFR activity dramatically improves survival of Nf1+/−, p53+/− mice (Ling et€ al. 2005). In mouse and human MPNST cell lines, the EGFR inhibitor AG1478 blocks in€ vitro accumulation of MPNST cells (DeClue et€ al. 2000; Li et€ al. 2002). PD184352, an inhibitor of MEK (which is activated downstream of EGFR and RAS), promotes apoptosis of MPNST-derived cells in€vitro (Mattingly et€al. 2006). The combination of erlotinib with oncolytic herpes simplex virus infection inhibits MPNST growth in a xenograft model as well (Mahller et€ al. 2007). Finally, the observation that mTOR is also deregulated in the absence of NF1 implies that rapamycin and related compounds may have selective activity against plexiform neurofibroma and MPNST (Johannessen et€al. 2005).
Infantile Fibrosarcoma and the ETV6-NTRK3 Fusion Protein Infantile fibrosarcomas are composed of fibroblast-like cells and nearly always occur in children less than 4 years old (Spunt et€ al. 2006). Although their histopathological features suggest aggressive behavior (frequent mitotic figures and areas of necrosis), as does fibrosarcoma histology in adults, metastasis is rare and the prognosis is usually good. Insight into this favorable biology initially came from the observation that these tumors have a recurrent translocation involving chromosomes 12 and 15 (Knezevich et€al. 1998). This translocation generates a fusion protein containing the dimerization domain of the ETV6 transcription factor (also known as TEL) and the protein tyrosine kinase domain of the neurotrophin-3 receptor, NTRK3 (also known as TRKC) (Knezevich et€al. 1998). Unlike fusion proteins that show marked specificity for a single tumor, ETV6-NTRK3 has been described in congenital mesoblastic nephroma (Rubin et€ al. 1998), secretory breast cancer (Tognon et€ al. 2002), and acute myelogenous leukemia (Eguchi et€al. 1999) as well as infantile fibrosarcoma, although it is not detected in adult-type fibrosarcoma. The finding that ectopic expression of ETV6-NTRK3 in NIH 3T3 cells induces a morphological change to a “transformed” phenotype and allows growth in soft agar suggests that the expression of ETV6-NTRK3 is integral to infantile fibrosarcoma biology (Wai et€ al. 2000). Hence, elucidation of the function of ETV6-NTRK3 may suggest how its activity can be blocked as a therapeutic strategy. Studies of ETV6-NTRK3 biology have provided tremendous insight into the mechanisms by which it can drive sarcomagenesis. As expected from sequencing data, the fusion protein acts as a kinase capable of auto- or transphosphorylation (Wai et€al. 2000). Its kinase activity and transformation potential in 3T3 cells are blocked by the deletion of the helix-loop-helix dimerization motif in the ETV6
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
507
domain and by point mutations disrupting the kinase activity of the NTRK3 domain. Expression of this fusion protein activates both the RAS/ERK and the PI3K/AKT signaling pathways, leading to induction of cyclin D1 in cultured cells (Tognon et€al. 2001). Pharmacological inhibition of either pathway blocks cellular transformation and cyclin D1 induction. IGF-IR suppresses apoptosis induced by anchorage-independent growth conditions (known as anoikis) in fibroblasts expressing ETV6-NTRK3 (Martin et€ al. 2006). Interestingly, however, it is not needed for cyclin D1 induction. The adaptor protein c-Src fosters an interaction between the fusion protein and the insulin receptor substrate 1 (IRS1); c-Src kinase activity is needed for AKT activation by ETV6-NTRK3 (Jin et€al. 2007). Finally, ETV6-NTRK3 physically interacts with and phosphorylates the transforming growth factor (TGF) b type II receptor, thereby blocking TGFB-dependent signaling (Jin 2005). As TGFBs have well-described antimitogenic properties [reviewed in Bierie and Moses (2006)], this interaction demonstrates a novel mechanism by which a constitutively activated fusion protein kinase may drive tumor formation. It is not yet known whether selective inhibition of ETV6-NTRK3 kinase activity or any of these cooperating pathways will be clinically useful.
Malignant Fibrous Histiocytoma The biology of malignant fibrous histiocytoma (MFH) remains enigmatic, and even its nomenclature is in flux. The 2002 WHO Classification of soft tissue tumors conceded that what had previously been called pleomorphic MFH is not a distinct entity (Fletcher 2006). This conclusion is consistent with recent gene expression array patterns in MFHs; they do not form tight clusters as a single entity, as is the case of synovial sarcoma, for example (Baird et€al. 2005; Lee et€al. 2003; Nagayama et€al. 2002; Nakayama et€al. 2007). Currently, the WHO considers that what was previously called pleomorphic MFH is better classified as undifferentiated pleomorphic sarcoma (Fletcher 2006). Similarly, giant cell MFH is likely to represent a giant cellrich osteosarcoma or other giant cell-containing soft tissue tumors, and most inflammatory MFH would currently be recognized as dedifferentiated liposarcoma. Also, unifying genetic or cytogenetic abnormalities have not been described in MFH, which instead often has complex karyotypic abnormalities similar to those in osteosarcoma; this fact is consistent with the concept that “MFH” may represent a broader array of tumors [reviewed in Fletcher (2006), Skapek and Chui (2000)]. Effective therapy for MFH in children, like therapy for other NRSTS, relies heavily on surgical resection; without resection, survival is poor despite intensive chemotherapy and radiation (Daw et€ al. 2003). Interestingly, three of four tested MFH-derived cell lines express both c-KIT and PDGFRb; treatment with imatinib mesylate slows accumulation of the cells in€ vitro and in xenograft models (Irsan et€ al. 2007). Whether inhibition of these receptor tyrosine kinases will also be a useful therapy is not yet known.
508
S.X. Skapek
Dermatofibrosarcoma Protuberans and the COL1A1-PDGFB Fusion Dermatofibrosarcoma protuberans (DFSP) is a relatively uncommon soft tissue sarcoma that can occur in young adults and children. It is typically a low grade tumor, but higher grade lesions and metastatic disease can occur. Treatment in both children and adults has focused on achieving excision with wide margins, after which the chance of recurrence is relatively low [reviewed in McArthur (2006)]. Higher grade fibrosarcomatous elements (likely indicating gene mutation) can be histologically detected within an otherwise low grade tumor; while this finding can portend a worse prognosis (Abbott et€al. 2006), adequate resection may be protective (Goldblum et€al. 2000). As with synovial sarcoma, initial insight into DFSP biology came from cytogenetic studies showing that a balanced translocation involving chromosomes 17 and 22 is common in DFSP and in giant cell fibroblastoma, which is likely related (McArthur 2006). Molecular characterization of the breakpoint shows that it creates a fusion transcript involving variable regions of collagen type 1a1 and always involving exon 2 of platelet-derived growth factor B (Maire et€al. 2007; McArthur 2006). Expression of the COL1A1-PDGFB fusion transcript is therefore controlled by the COL1A1 promoter but the transcript is processed to generate a normal PDGF-B protein, which is mitogenic for fibroblasts and other mesenchymal cells. Although this lesion is presumed to drive cell proliferation and hence tumorigenesis in an autocrine or paracrine manner, relatively little phosphorylation (activation) of PDGF receptors a or b is evident in protein lysates (McArthur et€al. 2005). Imatinib mesylate, developed as a selective tyrosine kinase inhibitor, is known to block signals stemming from PDGF receptors. Rubin and colleagues reported that this drug significantly reduced the DFSP tumor burden in a single patient with unresectable, metastatic disease (Rubin et€al. 2002). This report led to the successful use of imatinib in a child with unresectable disease (Price et€al. 2005) and a prospective clinical trial in adults (McArthur et€ al. 2005). The latter showed partial or complete responses in 9 of 10 patients enrolled, allowing surgical resection in some cases and long-term disease control. Xenograft studies using a DFSP-derived cell line indicate that imatinib causes apoptosis of the tumor cells but has little effect on cell proliferation (Sjoblom et€al. 2001). While results to date are promising, additional studies are needed to determine whether continuous therapy is needed to eradicate DFSP and to address potential drug resistance (see additional discussion under GIST below).
Gastrointestinal Stromal Tumor and Activation of cKIT and PDGFR Gastrointestinal stromal tumors (GISTs) are relatively uncommon in children. A recent retrospective study identified only seven cases among 276 children with
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
509
NRSTS over a 40-year period at a single institution (Cypriano et€al. 2004). Until recently, it was difficult to distinguish GISTs from leiomyosarcoma of the gastrointestinal tract and therapeutic options were largely limited to surgical resection and radiation. Despite its rarity, GIST in children provides an exceptional example of the clinical application of molecular insight into sarcoma biology. A major breakthrough came from the finding that nearly all GISTs express the cKIT receptor tyrosine kinase (Hirota et€al. 1998); in two smaller series in children, focal or diffuse cKIT expression was detectable in all cases (Cypriano et€al. 2004; Prakash et€al. 2005). This finding alone provides a useful tool to discriminate GIST from leiomyosarcoma. However, it further suggested that GIST arises from interstitial cells of Cajal, which express this receptor. The gene encoding cKIT is now known to be mutated in approximately 78% of GISTs [reviewed in Rubin et€al. (2007)]. In most cases, the mutation alters the amino acid sequence in the juxtamembrane domain of the protein, activating its carboxy-terminal tyrosine kinase domain in the absence of its natural ligand, stem cell factor. The gene encoding the related receptor tyrosine kinase PDGFRA is mutated in a subset of GISTs with normal cKIT, thereby revealing a common theme underlying GIST biology. Activation of cKIT was proved to play a principal role in tumor formation by disruption of the juxtamembrane region of mouse cKit (Sommer et€ al. 2003); survival of cKitV558D mice is significantly decreased and nearly all have gastrointestinal pathology (myenteric plexus hyperplasia in the distal esophagus, stomach, proximal duodenum, and cecum). The morphological appearance and uniform expression of cKit in the neoplastic lesions resemble that in GISTs. It is not clear exactly how constitutively active cKIT drives tumorigenesis. Evidence from cultured cells suggests that it secondarily activates PI3-kinase/AKT prosurvival pathways (Bauer et€al. 2007), but enhanced cell proliferation is also observed in€vivo in mice with mutated cKit (Sommer et€al. 2003). While it is striking that cKit mutation is sufficient to cause tumors in a mouse model, the presence of a wide range of other chromosomal abnormalities in human tumors suggests that cooperating genetic events are important in human tumor progression (Rubin et€al. 2007). As the relationship between cKIT (or PRGFRA) mutations and GIST was becoming clear, imatinib mesylate (STI571) was being developed as a selective inhibitor of the ABL tyrosine kinase (Druker et€al. 1996). Imatinib competes with ATP for binding to the tyrosine kinase domain and is now known to block wild-type and mutant KIT, PDGFRA, and several other receptor tyrosine kinases in addition to ABL (Rubin et€ al. 2007). A striking case report showing essentially complete resolution of tumor cell metabolic activity in an adult with widespread GIST opened the door to a number of clinical trials (Joensuu et€ al. 2001). Numerous phase II and III trials enrolling 27 to more than 400 patients showed objective responses to imatinib in 48 to 71% of patients and tumor control (objective response or stable disease) in 74 to 90% of patients [Rubin et€ al. (2007) and references therein]. When used as an adjuvant, imatinib effectively prevented recurrence after surgical resection (Nilsson et€ al. 2007). The fact that disease recurrence is more common in patients who electively discontinued the drug after 1 year implies that
510
S.X. Skapek
it acts, at least partially, as a cytostatic agent (Blay et€al. 2007). In cultured GIST cells, its proapoptotic effects appear to be due to the induction of phosphorylated histone H2AX (Liu et€al. 2007). While imatinib has been tested in Phase II clinical trials in the Children’s Oncology Group, its effectiveness against childhood GIST is not yet known. Given the efficacy of imatinib in adult patients, attention has now turned to understanding the mechanisms of resistance and defining “second-line” therapies [reviewed in Rubin et€al. (2007), Joensuu (2007)]. Patients whose tumors do not respond to imatinib tend to have activating mutations in exon 9 affecting the extracellular domain of the protein or no detectable KIT mutations. Those whose tumors respond initially but then become resistant usually have acquired new mutations that are likely to limit the ability of imatinib to bind the ATP-binding site. Secondary resistance is highly relevant because GISTs usually recur despite continued therapy [e.g., Blay et€al. (2007)]. Sunitinib, a more broadly acting tyrosine kinase inhibitor, appears to have the activity against imatinib-resistant tumors, including those with exon 9 mutations (Joensuu 2007). In vitro studies of vatalanib (targets multiple tyrosine kinases), N-benzoyl-staurosporine (inhibits PKC), nilotinib (inhibits KIT), dasatinib (inhibits SRC and ABL), and geldanamycin (inhibits HSP-90) in imatinib-resistant GIST have yielded promising results.
Inflammatory Myofibroblastic Tumor and ALK Activation Inflammatory myofibroblastic tumor (IMT) is an uncommon soft tissue tumor composed of fibroblasts and myofibroblasts intermingled with an inflammatory infiltrate consisting mostly of lymphocytes and plasma cells (Webber and Parham 1996). Its classification as a true neoplasm vs. a “pseudotumor” and its status as a single disease vs. a group of tumors were debated until relatively recent studies confirmed that IMTs often have a balanced translocation involving chromosome band 2p23 that disrupts the gene encoding the anaplastic lymphoma kinase (ALK) [reviewed in Dehner (2004)]. IMTs are usually low grade neoplasms treated with surgical resection; their expression of COX-2 provides a rationale for the use of nonsteroidal anti-inflammatory drugs, but the efficacy of this therapy is not established (Applebaum et€al. 2005). More aggressive tumors that metastasize are reported to be generally refractory to cytotoxic chemotherapy [e.g., Dishop et€al. (2003)]. ALK rearrangement was originally reported in anaplastic large cell lymphoma (Morris et€al. 1994). Numerous translocations are now known to generate a fusion protein in which the amino terminus contains a dimerization motif and carboxyl terminus represents the same 563 amino acids containing the ALK kinase domain [reviewed in Duyster et€al. (2001)]. In IMTs, the dimerization motifs include portions of tropomyosin 3, tropomyosin 4, clathrin heavy chain, and ran binding protein 2 (Duyster et€ al. 2001). In all cases, the promoter driving expression of the fusion partner causes increased expression of ALK; hence, immunohistochemical
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
511
detection of ALK supports the diagnosis of IMT, but it may also be observed in other tumors. The oligomerization domain in the fusion protein leads to constitutive activation of ALK, which has potent transforming potential. Downstream effectors of ALK include PLC-g, PI3-K, STAT3, and STAT5. Because of the central role of ALK in IMT, pharmacological inhibition of ALK is an attractive strategy. Chemical biology screens have begun to identify compounds, including WHI-P131 (4-(40-hydroxyphenyl)amino-6,7-dimethoxyquinazoline) and WHI-P154 (4-[(30-bromo40-hydroxyphenyl)amino]-6,7 dimethoxyquinazoline) (Marzec et€ al. 2005), and test their activity in preclinical models. Because NPM-ALK requires HSP90 for normal protein folding, the geldanamycin analog 17-AAG can destabilize NPMALK (Bonvini et€al. 2004). It remains to be seen whether ALK inhibition will be helpful for low grade and malignant IMTs.
Leiomyosarcoma Leiomyosarcoma is an uncommon type of NRSTS in children. In a retrospective series of 276 children with NRSTS at St. Jude Children’s Research Hospital, only 4 (1.4%) were shown to have leiomyosarcoma (Cypriano et€al. 2004). However, it has become the second most common neoplasm in children with HIV infection (Granovsky et€ al. 1998). In either case, it occurs in gastrointestinal sites, liver, spleen, uterus, and other soft tissue sites. As with other NRSTS, survival is relatively poor unless the tumor can be completely removed surgically (Cypriano et€al. 2004). Relatively little is known about leiomyosarcoma biology. The tumor is presumed to arise from smooth muscle cells or their progenitors but rarely arises from benign leiomyomas [reviewed in Sandberg (2005)]. Most biology studies have been performed on tumors in adults. Cytogenetic and molecular genetic analyses (e.g., comparative genomic hybridization) show that these tumors have complex and variable gains and losses involving a large number of chromosomes. This variability has made it difficult to develop a unifying hypothesis for the origin of leiomyosarcoma. Some have taken a “candidate gene” approach and have made some progress by using preclinical models. For example, building on the fact that the Wnt signaling pathway is generally important in a range of cancers and that Cripto-1 (CR-1) is a growth factor that can be induced by Wnt signaling, Strizzi et€al. (2007) tested whether the deregulated expression of CR-1 could cause leiomyosarcoma in transgenic mice. Indeed, 20% of mice expressing this transgene developed uterine leiomyosarcoma as they aged. These investigators subsequently showed by immunostaining that 9 of 13 cases of human leiomyosarcoma expressed CR-1. Using similar logic, members of the Cordon-Cardo laboratory tested whether the activation of the PI3-kinase/AKT pathway in smooth muscle cells is sufficient to cause leiomyosarcoma (Hernando et€ al. 2007). They found that cell type-specific genetic ablation of the PTEN tumor suppressor, which functions to block PI3K/AKT signaling, caused generalized
512
S.X. Skapek
smooth muscle hyperplasia in the gastrointestinal and urinary tracts (but not the uterus) that progressed to leiomyosarcoma in 80% of mice more than 2 months of age. Interestingly, p53 gene mutations were not observed, but Mdm2, which acts as a negative regulator of p53 tumor suppressor function, was highly expressed. Inactivation of this important pathway by Mdm2 supports the potential usefulness of therapeutic compounds aimed at disrupting Mdm2–p53 interactions. These investigators also observed molecular evidence of mTOR activation as a downstream effector of PTEN loss in smooth muscle cells. Survival was significantly prolonged in mice treated with the mTOR inhibitor everolimus, suggesting that mTOR activation plays a causative role. In view of these findings, mTOR and PI3K/AKT are very attractive therapeutic targets for this disease, especially if functional genomics approaches largely pursued in adult leiomyosarcoma [e.g., Segal et€al. (2003)] can be extended to children.
Current Clinical Trials Using Targeted Therapies for Childhood NRSTS Currently, there are relatively few histology-specific clinical trials of targeted therapy for NRSTS in the US. A search of ClinicalTrials.gov revealed the following trials to be open as of June 2007: NCT00029354:â•… A randomized phase II trial of the FTI R115777 (tipifarnib) in children (ages 3 to 25 years) with neurofibromatofsis type I and progressive plexiform neurofibromas; conducted by the National Cancer Institute. NCT00148109:â•… A nonrandomized Phase II trial of cetuximab, an antibody directed against the EGF receptor, in children (16 years and above) and adults; conducted by the University of Michigan Cancer Center in collaboration with Bristol-Myers Squibb and ImClone Systems. NCT00171912:â•… A nonrandomized Phase II trial of imatinib in children (16 years and above) and adults with a variety of cancers, including dermatofibrosarcoma; conducted by Novartis. NCT00187174:â•… A nonrandomized Phase I/II trial of the mTOR inhibitor RAD001C (everolimus) in children (ages 3 to 21 years) with recurrent, refractory sarcoma, including NRSTS; conducted by St. Jude Children’s Research Hospital. NCT00401388:â•… A nonrandomized Phase II trial of the AKT/PKB inhibitor perifosine in children and adults (ages 13 and above) with refractory alveolar soft part sarcoma, chondrosarcoma, or extraskeletal myxoid chondrosarcoma; conducted by the Sarcoma Alliance for Research through Collaboration (SARC) in collaboration with AOI Pharma, Inc. NCT00464620:â•… A nonrandomized Phase II trial of the second-generation tyrosine kinase inhibitor dasatinib in children and adults with any of a variety of unresectable,
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
513
recurrent, or metastatic soft tissue or bone sarcomas; conducted by SARC in collaboration with Bristol-Myers Squibb.
Future Directions and Challenges Great insight into the biology of many types of NRSTS in children and adults has come from cytogenetics, molecular genetics, and molecular biology studies largely using tumor-derived cell lines, human tissue samples, and elegant mouse models of the human diseases. The remarkable initial successes of imatinib against certain previously untreatable NRSTS, such as GIST, suggests great promise for the future of targeted therapies for this class of tumors, which are minimally susceptible to intensive chemotherapy with alkylating agents and anthracyclines. Of course, numerous challenges remain. Perhaps foremost among them are the remarkable diversity of histological subtypes of NRSTS and the relative rarity of any particular subtype in children. The recent opening of a national collaborative trial for NRSTS in children and young adults by the Children’s Oncology Group is hoped to encourage the enrollment of children on research protocols. This trial (ARST0332) tests the usefulness of a risk-based application of standard therapies, and the companion biology study (ARST D9902) will increase the availability of tissue and tumorderived cells to further dissect the molecular mechanisms of NRSTS. The rarity of individual subtypes of NRSTS would necessitate national collaborative studies of any histology-specific targeted therapies. However, histologically dissimilar tumors may be driven by similar molecular processes and may respond to the same therapies; for example, the unrelated GIST and DFSP are both treatable with imatinib mesylate. Therefore, eligibility for trials of targeted therapies may be based on biological rather than histological subtype. It is already becoming clear that the initial successes of targeted therapy for NRSTS are likely to lead us to the next challenge – refractory disease. This challenge must be met by continued molecular biological and genetic studies of refractory tumors; the development of second and subsequent generations of novel agents; better understanding of downstream signaling pathways and their changes in response to targeted therapy; the potential synergism of combined targeted therapies; and better understanding of parallel signaling pathways or genetic or epigenetic events necessary for tumor progression. The increased development and use of elegant, in€vivo models such as those developed for the studies of synovial sarcoma and GIST provide an opportunity to discover cooperating pathways that may serve as additional targets.
References Abbott, J. J., Oliveira, A. M., and Nascimento, A. G. 2006. The prognostic significance of fibrosarcomatous transformation in dermatofibrosarcoma protuberans. Am. J. Surg. Pathol. 30:436–443.
514
S.X. Skapek
Allander, S. V., Illei, P. B., Chen, Y., Antonescu, C. R., Bittner, M., Ladanyi, M., and Meltzer, P. S. 2002. Expression profiling of synovial sarcoma by cDNA microarrays: association of ERBB2, IGFBP2, and ELF3 with epithelial differentiation. Am. J. Pathol. 161:1587–1595. Applebaum, H., Kieran, M. W., Cripe, T. P., Coffin, C. M., Collins, M. H., Kaipainen, A., Laforme, A., and Shamberger, R. C. 2005. The rationale for nonsteroidal anti-inflammatory drug therapy for inflammatory myofibroblastic tumors: a Children’s Oncology Group study. J. Pediatr. Surg. 40:999–1003. Baird, K., Davis, S., Antonescu, C. R., Harper, U. L., Walker, R. L., Chen, Y., Glatfelter, A. A., Duray, P. H., and Meltzer, P. S. 2005. Gene expression profiling of human sarcomas: insights into sarcoma biology. Cancer Res. 65:9226–9235. Bakkenist, C. J. and Kastan, M. B. 2004. Initiating cellular stress responses. Cell. 118:9–17. Bauer, S., Duensing, A., Demetri, G. D., and Fletcher, J. A. 2007. KIT oncogenic signaling mechanisms in imatinib-resistant gastrointestinal stromal tumor: PI3-kinase/AKT is a crucial survival pathway. Oncogene. 26:7560–7568. Bierie, B. and Moses, H. L. 2006. TGFb: the molecular Jekyll and Hyde of cancer. Nat. Rev. Cancer. 6:506–520. Birdsall, S., Osin, P., Lu, Y. -J., Fisher, C., and Shipley, J. 1999. Synovial sarcoma specific translocation associated with both epithelial and spindle components. Int. J. Cancer. 82:605–608. Blay, J. Y., Le, C. A., Ray-Coquard, I., Bui, B., Duffaud, F., Delbaldo, C., Adenis, A., Viens, P., Rios, M., Bompas, E., Cupissol, D., Guillemet, C., Kerbrat, P., Fayette, J., Chabaud, S., Berthaud, P., and Perol, D. 2007. Prospective multicentric randomized phase III study of imatinib in patients with advanced gastrointestinal stromal tumors comparing interruption versus continuation of treatment beyond 1 year: the French Sarcoma Group. J. Clin. Oncol. 25:1107–1113. Bonvini, P., Dalla Rosa, H., Vignes, H., and Rosolen, A. 2004. Ubiquitination and proteasomal degradation of nucleophosmin-anaplastic lymphoma kinase induced by 17-allylaminodemethoxygeldanamycin: role of the co-chaperone carboxyl heat shock protein 70-interacting protein. Cancer Res. 64:3256–3264. Cichowski, K. and Jacks, T. 2001. NF1 tumor suppressor gene function: narrowing the GAP. Cell. 104:593–604. Cichowski, K., Shih, T. S., Schmitt, E., Santiago, S., Reilly, K., McLaughlin, M. E., Bronson, R. T., and Jacks, T. 1999. Mouse models of tumor development in neurofibromatosis type 1. Science. 286:2172–2176. Clark, J., Rocques, P. J., Crew, A. J., Gill, S., Shipley, J., Chan, A. M. L., Gusterson, B. A., and Cooper, C. S. 1994. Identification of novel genes, SYT and SSX, involved in the t(X;18) (p11.2;q11.2) translocation found in human synovial sarcoma. Nat. Genet. 7:502–508. Cypriano, M. S., Jenkins, J. J., Pappo, A. S., Rao, B. N., and Daw, N. C. 2004. Pediatric gastrointestinal stromal tumors and leiomyosarcoma. Cancer. 101:39–50. Daw, N. C., Billups, C. A., Pappo, A. S., Jenkins, J. J., Mahmoud, H. H., Krasin, M. J., and Rao, B. N. 2003. Malignant fibrous histiocytoma and other fibrohistiocytic tumors in pediatric patients: the St. Jude Children’s Research Hospital experience. Cancer. 97:2839–2847. DeClue, J. E., Heffelfinger, S., Benvenuto, G., Ling, B., Li, S., Rui, W., Vass, W. C., Viskochil, D., and Ratner, N. 2000. Epidermal growth factor receptor expression in neurofibromatosis type 1-related tumors and NF1 animal models. J. Clin. Invest. 105:1233–1241. Dehner, L. P. 2004. Inflammatory myofibroblastic tumor: the continued definition of one type of so-called inflammatory pseudotumor. Am. J. Surg. Pathol. 28:1652–1654. Dilworth, J. T., Kraniak, J. M., Wojtkowiak, J. W., Gibbs, R. A., Borch, R. F., Tainsky, M. A., Reiners, Jr. J. J., and Mattingly, R. R. 2006. Molecular targets for emerging anti-tumor therapies for neurofibromatosis type 1. Biochem. Pharmacol. 72:1485–1492. Dishop, M. K., Warner, B. W., Dehner, L. P., Kriss, V. M., Greenwood, M. F., Geil, J. D., and Moscow, J. A. 2003. Successful treatment of inflammatory myofibroblastic tumor with malignant transformation by surgical resection and chemotherapy. J. Pediatr. Hematol. Oncol. 25:153–158. Donehower, L. A., Harvey, M., Slagle, B. L., McArthur, M. J., Montgomery, C. A., Jr., Butel, J. S., and Bradley, A. 1992. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature. 356:215–221.
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
515
Druker, B. J., Tamura, S., Buchdunger, E., Ohno, S., Segal, G. M., Fanning, S., Zimmermann, J., and Lydon, N. B. 1996. Effects of a selective inhibitor of the Abl tyrosine kinase on the growth of Bcr-Abl positive cells. Nat. Med. 2:561–566. Duyster, J., Bai, R. Y., and Morris, S. W. 2001. Translocations involving anaplastic lymphoma kinase (ALK). Oncogene. 20:5623–5637. Eguchi, M., Eguchi-Ishimae, M., Tojo, A., Morishita, K., Suzuki, K., Sato, Y., Kudoh, S., Tanaka, K., Setoyama, M., Nagamura, F., Asano, S., and Kamada, N. 1999. Fusion of ETV6 to neurotrophin-3 receptor TRKC in acute myeloid leukemia with t(12;15)(p13;q25). Blood. 93:1355–1363. Ferrari, A., Casanova, M., Collini, P., Meazza, C., Luksch, R., Massimino, M., Cefalo, G., Terenziani, M., Spreafico, F., Catania, S., Gandola, L., Gronchi, A., Mariani, L., and FossatiBellani, F. 2005. Adult-type soft tissue sarcomas in pediatric-age patients: experience at the Istituto Nazionale Tumori in Milan. J. Clin. Oncol. 23:4021–4030. Fletcher, C. D. 2006. The evolving classification of soft tissue tumours: an update based on the new WHO classification. Histopathology. 48:3–12. Goldblum, J. R., Reith, J. D., and Weiss, S. W. 2000. Sarcomas arising in dermatofibrosarcoma protuberans: a reappraisal of biologic behavior in eighteen cases treated by wide local excision with extended clinical follow up. Am. J. Surg. Pathol. 24:1125–1130. Granovsky, M. O., Mueller, B. U., Nicholson, H. S., Rosenberg, P. S., and Rabkin, C. S. 1998. Cancer in human immunodeficiency virus-infected children: a case series from the Children’s Cancer Group and the National Cancer Institute. J. Clin. Oncol. 16:1729–1735. Guillou, L., Benhattar, J., Bonichon, F., Gallagher, G., Terrier, P., Stauffer, E., Somerhausen, N. S., Michels, J. J., Jundt, G., Vince, D. R., Taylor, S., Genevay, M., Collin, F., Trassard, M., and Coindre, J. M. 2004. Histologic grade, but not SYT-SSX fusion type, is an important prognostic factor in patients with synovial sarcoma: a multicenter, retrospective analysis. J. Clin. Oncol. 22:4040–4050. Haldar, M., Hancock, J. D., Coffin, C. M., Lessnick, S. L., and Capecchi, M. R. 2007. A conditional mouse model of synovial sarcoma: insights into a myogenic origin. Cancer Cell. 11:375–388. Hanahan, D. and Weinberg R. A. 2000. The hallmarks of cancer. Cell. 100:57–70. Helman, L. J. and Meltzer, P. 2003. Mechanisms of sarcoma development. Nat. Rev. Cancer. 3:685–694. Hernando, E., Charytonowicz, E., Dudas, M. E., Menendez, S., Matushansky, I., Mills, J., Socci, N. D., Behrendt, N., Ma, L., Maki, R. G., Pandolfi, P. P., and Cordon-Cardo, C. 2007. The AKT-mTOR pathway plays a critical role in the development of leiomyosarcomas. Nat. Med. 13:748–753. Hirota, S., Isozaki, K., Moriyama, Y., Hashimoto, K., Nishida, T., Ishiguro, S., Kawano, K., Hanada, M., Kurata, A., Takeda, M., Muhammad, T. G., Matsuzawa, Y., Kanakura, Y., Shinomura, Y., and Kitamura, Y. 1998. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science. 279:577–580. Irsan, I., Akisue, T., Hara, H., Fujimoto, T., Imabori, M., Doita, M., Kuroda, R., Fujioka, H., Kawamoto, T., Yamamoto, T., and Kurosaka, M. 2007. Imatinib mesylate inhibits tumorigenicity of malignant fibrous histiocytoma cells in€vivo. Anticancer Res. 27:423–429. Ishida, M., Miyamoto, M., Naitoh, S., Tatsuda, D., Hasegawa, T., Nemoto, T., Yokozeki, H., Nishioka, K., Matsukage, A., Ohki, M., and Ohta, T. 2007. The SYT-SSX fusion protein down-regulates the cell proliferation regulator COM1 in t(x;18) synovial sarcoma. Mol.Cell Biol. 27:1348–1355. Ito, T., Ouchida, M., Morimoto, Y., Yoshida, A., Jitsumori, Y., Ozaki, T., Sonobe, H., Inoue, H., and Shimizu, K. 2005. Significant growth suppression of synovial sarcomas by the histone deacetylase inhibitor FK228 in€vitro and in€vivo. Cancer Lett. 224:311–319. Jacks, T., Remington, L., Williams, B. O., Schmitt, E. M., Halachmi, S., Bronson, R. T., and Weinberg, R. A. 1994a. Tumor spectrum analysis in p53-mutant mice. Curr. Biol. 4:1–7. Jacks, T., Shih, T. S., Schmitt, E. M., Bronson, R. T., Bernards, A., and Weinberg, R. A. 1994b. Tumour predisposition in mice heterozygous for a targeted mutation in Nf1. Nat. Genet. 7:353–361. Jin, W., Kim, B.-C., Tognon, C., Lee, H.-J., Patel, S., Lannon, C. L., Maris, J. M., Triche, T. J., Sorensen, P. H. B., and Kim, S.-J. 2005. The ETV6-NTRK3 chimeric tyrosine kinase suppresses TGF-bsignaling by inactivating the TGF-b type II receptor. Proc. Natl. Acad. Sci. 102:16239–16244.
516
S.X. Skapek
Jin, W., Yun, C., Hobbie, A., Martin, M. J., Sorensen, P. H., and Kim, S. J. 2007. Cellular transformation and activation of the phosphoinositide-3-kinase-Akt cascade by the ETV6-NTRK3 chimeric tyrosine kinase requires c-Src. Cancer Res. 67:3192–3200. Joensuu, H. 2007. Second line therapies for the treatment of gastrointestinal stromal tumor. Curr. Opin. Oncol. 19:353–358. Joensuu, H., Roberts, P. J., Sarlomo-Rikala, M., Andersson, L. C., Tervahartiala, P., Tuveson, D., Silberman, S., Capdeville, R., Dimitrijevic, S., Druker, B., and Demetri, G. D. 2001. Effect of the tyrosine kinase inhibitor STI571 in a patient with a metastatic gastrointestinal stromal tumor. N. Engl. J. Med. 344:1052–1056. Johannessen, C. M., Reczek, E. E., James, M. F., Brems, H., Legius, E., and Cichowski, K. 2005. The NF1 tumor suppressor critically regulates TSC2 and mTOR. Proc. Natl. Acad. Sci. U.S.A. 102:8573–8578. Joyner, D. E., Albritton, K. H., Bastar, J. D., and Randall, R. L. 2006. G3139 antisense oligonucleotide directed against antiapoptotic Bcl-2 enhances doxorubicin cytotoxicity in the FU-SY-1 synovial sarcoma cell line. J. Orthop. Res. 24:474–480. Kamijo, T., Bodner, S., van de Kamp, E., Randle, D. H., and Sherr, C. J. 1999. Tumor spectrum in ARF-deficient mice. Cancer Res. 59:2217–2222. Kawai, A., Woodruff, J., Healey, J. H., Brennan, M. F., Antonescu, C. R., and Ladany, M. 1998. SYT-SSX gene fusion as a determinant of morphology and prognosis in synovial sarcoma. N. Engl. J. Med. 338:153–160. Knezevich, S. R., McFadden, D. E., Tao, W., Lim, J. F., and Sorensen, P. H. B. 1998. A novel ETV6-NTRK3 gene fusion in congenital fibrosarcoma. Nat. Genet. 18:184–187. Ladanyi, M. 2001. Fusions of the SYT and SSX genes in synovial sarcoma. Oncogene. 20:5755–5762. Lee, Y. F., John, M., Edwards, S., Clark, J., Flohr, P., Maillard, K., Edema, M., Baker, L., Mangham, D. C., Grimer, R., Wooster, R., Thomas, J. M., Fisher, C., Judson, I., and Cooper, C. S. 2003. Molecular classification of synovial sarcomas, leiomyosarcomas and malignant fibrous histiocytomas by gene expression profiling. Br. J. Cancer. 88:510–515. Li, F. P., Fraumeni, J. F., Jr., Mulvihill, J. J., Blattner, W. A., Dreyfus, M. G., Tucker, M. A., and Miller, R. W. 1988. A cancer family syndrome in 24 kindreds. Cancer Res. 48:5358–5362. Li, H., Velasco-Miguel, S., Vass, W. C., Parada, L. F., and DeClue, J. E. 2002. Epidermal growth factor receptor signaling pathways are associated with tumorigenesis in the Nf1:p53 mouse tumor model. Cancer Res. 62:4507–4513. Ling, B. C., Wu, J., Miller, S. J., Monk, K. R., Shamekh, R., Rizvi, T. A., DeCourten-Myers, G., Vogel, K. S., DeClue, J. E., and Ratner, N. 2005. Role for the epidermal growth factor receptor in neurofibromatosis-related peripheral nerve tumorigenesis. Cancer Cell. 7:65–75. Liu, Y., Tseng, M., Perdreau, S. A., Rossi, F., Antonescu, C., Besmer, P., Fletcher, J. A., Duensing, S., and Duensing, A. 2007. Histone H2AX is a mediator of gastrointestinal stromal tumor cell apoptosis following treatment with imatinib mesylate. Cancer Res. 67:2685–2692. Lopez-Guerrero, J. A., Navarro, S., Noguera, R., Carda, C., Farinas, S. C., Pellin, A., and Llombart-Bosch, A. 2005. Mutational analysis of the c-KIT AND PDGFRalpha in a series of molecularly well-characterized synovial sarcomas. Diagn. Mol. Pathol. 14:134–139. Mahller, Y. Y., Vaikunth, S. S., Currier, M. A., Miller, S. J., Ripberger, M. C., Hsu, Y. -H., Mehrian-Shai, R., Collins, M. H., Crombleholme, T. M., Ratner, N., and Cripe, T. P. 2007. Oncolytic HSV and erlotinib inhibit tumor growth and angiogenesis in a novel malignant peripheral nerve sheath tumor xenograft model. Mol. Ther. 15:279–286. Maire, G., Fraitag, S., Galmiche, L., Keslair, F., Ebran, N., Terrier-Lacombe, M. J., De, P. Y., and Pedeutour, F. 2007. A clinical, histologic, and molecular study of 9 cases of congenital dermatofibrosarcoma protuberans. Arch. Dermatol. 143:203–210. Malkin, D., Li, F. P., Strong, L. C., Fraumeni, J. F., Jr., Nelson, C. E., Kim, D. H., Kassel, J., Gryka, M. A., Bischoff, F. A., Tainsky, M. A., and Friend, S. H. 1990. Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science. 250:1233–1238. Martin, M. J., Melnyk, N., Pollard, M., Bowden, M., Leong, H., Podor, T. J., Gleave, M., and Sorensen, P. H. 2006. The insulin-like growth factor I receptor is required for Akt activation
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
517
and suppression of anoikis in cells transformed by the ETV6-NTRK3 chimeric tyrosine kinase. Mol. Cell Biol. 26:1754–1769. Marzec, M., Kasprzycka, M., Ptasznik, A., Wlodarski, P., Zhang, Q., Odum, N., and Wasik, M. A. 2005. Inhibition of ALK enzymatic activity in T-cell lymphoma cells induces apoptosis and suppresses proliferation and STAT3 phosphorylation independently of Jak3. Lab. Invest. 85:1544–1554. Maser, R. S., Choudhury, B., Campbell, P. J., Feng, B., Wong, K. K., Protopopov, A., O’Neil, J., Gutierrez, A., Ivanova, E., Perna, I., Lin, E., Mani, V., Jiang, S., McNamara, K., Zaghlul, S., Edkins, S., Stevens, C., Brennan, C., Martin, E. S., Wiedemeyer, R., Kabbarah, O., Nogueira, C., Histen, G., Aster, J., Mansour, M., Duke, V., Foroni, L., Fielding, A. K., Goldstone, A. H., Rowe, J. M., Wang, Y. A., Look, A. T., Stratton, M. R., Chin, L., Futreal, P. A., and DePinho, R. A. 2007. Chromosomally unstable mouse tumours have genomic alterations similar to diverse human cancers. Nature. 447:966–971. Mattingly, R. R., Kraniak, J. M., Dilworth, J. T., Mathieu, P., Bealmear, B., Nowak, J. E., Benjamins, J. A., Tainsky, M. A., and Reiners, J. J., Jr. 2006. The mitogen-activated protein kinase/extracellular signal-regulated kinase kinase inhibitor PD184352 (CI-1040) selectively induces apoptosis in malignant Schwannoma cell lines. J. Pharmacol. Exp. Ther. 316:456–465. McArthur, G. A. 2006. Dermatofibrosarcoma protuberans: a surgical disease with a molecular savior. Curr. Opin. Oncol. 18:341–346. McArthur, G. A., Demetri, G. D., van Oosterom, A., Heinrich, M. C., Debiec-Rychter, M., Corless, C. L., Nikolova, Z., Dimitrijevic, S., and Fletcher, J. A. 2005. Molecular and clinical analysis of locally advanced dermatofibrosarcoma protuberans treated with imatinib: Imatinib Target Exploration Consortium study B2225. J. Clin. Oncol. 23:866–873. Morris, S. W., Kirstein, M. N., Valentine, M. B., Dittmer, K. G., Shapiro, D. N., Saltman, D. L., and Look, A. T. 1994. Fusion of a kinase gene, ALK, to a nucleolar protein gene, NPM, in non-Hodgkin’s lymphoma. Science. 263:1281–1284. Nagayama, S., Katagiri, T., Tsunoda, T., Hosaka, T., Nakashima, Y., Araki, N., Kusuzaki, K., Nakayama, T., Tsuboyama, T., Nakamura, T., Imamura, M., Nakamura, Y., and Toguchida, J. 2002. Genome-wide analysis of gene expression in synovial sarcomas using a cDNA microarray. Cancer Res. 62:5859–5866. Nakayama, R., Nemoto, T., Takahashi, H., Ohta, T., Kawai, A., Seki, K., Yoshida, T., Toyama, Y., Ichikawa, H., and Hasegawa, T. 2007. Gene expression analysis of soft tissue sarcomas: characterization and reclassification of malignant fibrous histiocytoma. Mod. Pathol. 20:749–759. Nielsen, T. O., Hsu, F. D., O’Connell, J. X., Gilks, C. B., Sorensen, P. H., Linn, S., West, R. B., Liu, C. L., Botstein, D., Brown, P. O., and Van de, R. M. 2003. Tissue microarray validation of epidermal growth factor receptor and SALL2 in synovial sarcoma with comparison to tumors of similar histology. Am. J. Pathol. 163:1449–1456. Nilsson, B., Sjolund, K., Meis-Kindblom, J. M., Bumming, P., Nilsson, O., Andersson, J., and Ahlman, H. 2007. Adjuvant imatinib treatmetn improves recurrence-free survival in patients with high-risk gastrointestinal stromal tumors (GIST). Br. J. Cancer. 96:1656–1658. Nilsson, G., Skytting, B., Xie, Y., Brodin, B., Perfekt, R., Mandahl, N., Lundeberg, J., Uhlen, M., and Larsson, O. 1999. The SYT-SSX1 variant of synovial sarcoma is associated with a high rate of tumor cell proliferation and poor clinical outcome. Cancer Res. 59:3180–3184. Nuciforo, P. G., Pellegrini, C., Fasani, R., Maggioni, M., Coggi, G., Parafioriti, A., and Bosari, S. 2003. Molecular and immunohistochemical analysis of HER2/neu oncogene in synovial sarcoma. Hum. Pathol. 34:639–645. Pappo, A. S., Rao, B. N., Jenkins, J. J., Merchant, T., Poquette, C. A., Cain, A., and Pratt, C. B. 1999. Metastatic nonrhabdomyosarcomatous soft-tissue sarcomas in children and adolescents: the St. Jude Children’s Research Hospital experience. Med. Pediatr. Oncol. 33:76–82. Parham, D. M., Webber, B. L., Jenkins, J. J., III, Cantor, A. B., and Maurer, H. M. 1995. Nonrhabdomyosarcomatous soft tissue sarcomas of childhood: formulation of a simplified system for grading. Mod. Pathol. 8:705–710.
518
S.X. Skapek
Prakash, S., Sarran, L., Socci, N., DeMatteo, R. P., Eisenstat, J., Greco, A. M., Maki, R. G., Wexler, L. H., LaQuaglia, M. P., Besmer, P., and Antonescu, C. R. 2005. Gastrointestinal stromal tumors in children and young adults: a clinicopathologic, molecular, and genomic study of 15 cases and review of the literature. J. Pediatr. Hematol. Oncol. 27:179–187. Pratt, C. B., Maurer, H. M., Gieser, P., Salzberg, A., Rao, B. N., Parham, D., Thomas, P. R. M., Marcus, R. B., Cantor, A., Pick, T., Green, D., Neff, J., and Jenkins, J. J. 1998. Treatment of unresectable or metastatic pediatric soft tissue sarcomas with surgery, irradiation, and chemotherapy: a Pediatric Oncology Group study. Med. Pediatr. Oncol. 30:201–209. Pratt, C. B., Pappo, A. S., Gieser, P., Jenkins, J. J., Salzbergdagger, A., Neff, J., Rao, B., Green, D., Thomas, P., Marcus, R., Parham, D., and Maurer, H. 1999. Role of adjuvant chemotherapy in the treatment of surgically resected pediatric nonrhabdomyosarcomatous soft tissue sarcomas: a Pediatric Oncology Group Study. J. Clin. Oncol. 17:1219. Pretto, D., Barco, R., Rivera, J., Neel, N., Gustavson, M. D., and Eid, J. E. 2006. The synovial sarcoma translocation protein SYT-SSX2 recruits beta-catenin to the nucleus and associates with it in an active complex. Oncogene. 25:3661–3669. Price, V. W., Fletcher, J. A., Zielenska, M., Cole, W., Viero, S., Manson, D. E., Stuart, M., and Pappo, A. S. 2005. Imatinib mesylate: an attractive alternative in young children with large, surgically challenging dermatofibrosarcoma protuberans. Pediatr. Blood Cancer. 44:511–515. Rubin, B. P., Chen, C. J., Morgan, T. W., Xiao, S., Grier, H. E., Kozakewich, H. P., Perez-Atayde, A. R., and Fletcher, J. A. 1998. Congenital mesoblastic nephroma t(12;15) is associated with ETV6-NTRK3 gene fusion: cytogenetic and molecular relationship to congenital (infantile) fibrosarcoma. Am. J. Pathol. 153:1451–1458. Rubin, B. P., Heinrich, M. C., and Corless, C. L. 2007. Gastrointestinal stromal tumour. Lancet. 369:1731–1741. Rubin, B. P., Schuetze, S. M., Eary, J. F., Norwood, T. H., Mirza, S., Conrad, E. U., and Bruckner, J. D. 2002. Molecular targeting of platelet-derived growth factor B by imatinib mesylate in a patient with metastatic dermatofibrosarcoma protuberans. J. Clin. Oncol. 20:3586–3591. Saito, T., Nagai, M., and Ladanyi, M. 2006. SYT-SSX1 and SYT-SSX2 interfere with repression of E-cadherin by snail and slug: a potential mechanism for aberrant mesenchymal to epithelial transition in human synovial sarcoma. Cancer Res. 66:6919–6927. Sandberg, A. A. 2005. Updates on the cytogenetics and molecular genetics of bone and soft tissue tumors: leiomyosarcoma. Cancer Genet. Cytogenet. 161:1–19. Segal, N. H., Pavlidis, P., Antonescu, C. R., Maki, R. G., Noble, W. S., DeSantis, D., Woodruff, J. M., Lewis, J. J., Brennan, M. F., Houghton, A. N., and Cordon-Cardo, C. 2003. Classification and subtype prediction of adult soft tissue sarcoma by functional genomics. Am. J. Pathol. 163:691–700. Sharpless, N. E., Ferguson, D. O., O’Hagan, R. C. O., Castrillon, D. H., Lee, C., Farazi, P. A., Alson, S., Fleming, J., Morton, C. C., Frank, K., Chin, L., Alt, F. W., and DePinho, R. A. 2001. Impaired nonhomologous end-joining provokes soft tissue sarcomas harboring chromosomal translocations, amplifications, and deletions. Mol. Cell. 8:1187–1196. Sjoblom, T., Shimizu, A., O’Brien, K. P., Pietras, K., Dal, C. P., Buchdunger, E., Dumanski, J. P., Ostman, A., and Heldin, C. H. 2001. Growth inhibition of dermatofibrosarcoma protuberans tumors by the platelet-derived growth factor receptor antagonist STI571 through induction of apoptosis. Cancer Res. 61:5778–5783. Skapek, S. X. and Chui, C. H. 2000. Cytogenetics and the biological basis of sarcomas. Curr. Opin. Oncol. 12:315–322. Smithey, B. E., Pappo, A. S., and Hill, D. A. 2002. C-kit expression in pediatric solid tumors: a comparative immunohistochemical study. Am. J. Surg. Pathol. 26:486–492. Sommer, G., Agosti, V., Ehlers, I., Rossi, F., Corbacioglu, S., Farkas, J., Moore, M., Manova, K., Antonescu, C. R., and Besmer, P. 2003. Gastrointestinal stromal tumors in a mouse model by targeted mutation of the Kit receptor tyrosine kinase. Proc. Natl. Acad. Sci. U.S.A. 100:6706–6711. Sorensen, S. A., Mulvihill, J. J., and Nielsen, A. 1986. Long-term follow-up of von Recklinghausen neurofibromatosis: survival and malignant neoplasms. N. Engl. J. Med. 314:1010–1015.
Nonrhabdomyosarcoma Soft Tissue Sarcoma in Children: Developing New Treatments
519
Spunt, S. L., Hill, D. A., Motosue, A. M., Billups, C. A., Cain, A. M., Rao, B. N., Pratt, C. B., Merchant, T. E., and Pappo, A. S. 2002. Clinical features and outcome of initially unresected nonmetastatic pediatric nonrhabdomyosarcoma soft tissue sarcoma. J. Clin. Oncol. 20:3225–3235. Spunt, S. L., Poquette, C. A., Hurt, Y. S., Cain, A. M., Rao, B. N., Merchant, T. E., Jenkins, J. J., Santana, V. M., Pratt, C. B., and Pappo, A. S. 1999. Prognostic factors for children and adolescents with surgically resected nonrhabdomyosarcoma soft tissue sarcoma: an analysis of 121 patients treated at St Jude Children’s Research Hospital. J. Clin. Oncol. 17:3697–3705. Spunt, S., Wolden, S., Schofield, D., and Skapek, S. 2006. Non-Rhabdomyosarcoma Soft Tissue Sarcomas. In Pediatric Bone and Soft Tissue Sarcomas, pp. 133–162. New York/Berlin/ Heidelberg: Springer. Storlazzi, C. T., Mertens, F., Mandahl, N., Gisselsson, D., Isaksson, M., Gustafson, P., Domanski, H. A., and Panagopoulos, I. 2003. A novel fusion gene, SS18L1/SSX1, in synovial sarcoma. Genes Chromosomes Cancer. 37:195–200. Strizzi, L., Bianco, C., Hirota, M., Watanabe, K., Mancino, M., Hamada, S., Raafat, A., Lawson, S., Ebert, A. D., D’Antonio, A., Losito, S., Normanno, N., and Salomon, D. S. 2007. Development of leiomyosarcoma of the uterus in MMTV-CR-1 transgenic mice. J. Pathol. 211:36–44. Subramaniam, M. M., Noguera, R., Piqueras, M., Navarro, S., Lopez-Guerrero, J. A., and Llombart-Bosch, A. 2006. p16INK4A (CDKN2A) gene deletion is a frequent genetic event in synovial sarcomas. Am. J. Clin. Pathol. 126:866–874. Sun, Y., Gao, D., Liu, Y., Huang, J., Lessnick, S., and Tanaka, S. 2006. IGF2 is critical for tumorigenesis by synovial sarcoma oncoprotein SYT-SSX1. Oncogene. 25:1042–1052. Tamborini, E., Bonadiman, L., Greco, A., Gronchi, A., Riva, C., Bertulli, R., Casali, P. G., Pierotti, M. A., and Pilotti, S. 2004. Expression of ligand-activated KIT and platelet-derived growth factor receptor beta tyrosine kinase receptors in synovial sarcoma. Clin. Cancer Res. 10:938–943. Tognon, C., Garnett, M., Kenward, E., Kay, R., Morrison, K., and Sorensen, P. H. 2001. The chimeric protein tyrosine kinase ETV6-NTRK3 requires both Ras-Erk1/2 and PI3-kinase-Akt signaling for fibroblast transformation. Cancer Res. 61:8909–8916. Tognon, C., Knezevich, S. R., Huntsman, D., Roskelley, C. D., Melnyk, N., Mathers, J. A., Becker, L., Carneiro, F., MacPherson, N., Horsman, D., Poremba, C., and Sorensen, P. H. 2002. Expression of the ETV6-NTRK3 gene fusion as a primary event in human secretory breast carcinoma. Cancer Cell. 2:367–376. Turc-Carel, C., Dal, C. P., Limon, J., Rao, U., Li, F. P., Corson, J. M., Zimmerman, R., Parry, D. M., Cowan, J. M., and Sandberg, A. A. 1987. Involvement of chromosome X in primary cytogenetic change in human neoplasia: nonrandom translocation in synovial sarcoma. Proc. Natl. Acad. Sci. U.S.A. 84:1981–1985. Vogel, K. S., Klesse, L. J., Velasco-Miguel, S., Meyers, K., Rushing, E. J., and Parada, L. F. 1999. Mouse tumor model for neurofibromatosis type 1. Science. 286:2176–2179. Wai, D. H., Knezevich, S. R., Lucas, T., Jansen, B., Kay, R. J., and Sorensen, P. H. 2000. The ETV6-NTRK3 gene fusion encodes a chimeric protein tyrosine kinase that transforms NIH3T3 cells. Oncogene. 19:906–915. Webber, B. L. and Parham, D. M. 1996. Soft Tissue Tumors Other Than Rhabdomyosarcoma and Peripheral Neuroepithelioma. In Pediatric Neoplasia – Morphology and Biology, pp. 205–257. Little Rock, AR/Philadelphia, NY: Lippincott-Raven Publishers. Xie, Y., Skytting, B., Nilsson, G., Brodin, B., and Larsson, O. 1999. Expression of insulin-like growth factor-1 receptor in synovial sarocma: association with an aggressive phenotype. Cancer Res. 59:3588–3591. Xie, Y., Skytting, B., Nilsson, G., Gasbarri, A., Haslam, K., Bartolazzi, B., Mandahl, N., and Larsson, O. 2002. SYT-SSX is critical for Cyclin D1 expression in synovial sarcoma cells: A gain of function of the t(X;18)(p11.2;q11.2) translocation. Cancer. Res. 62:3861–3867. Zhu, Y., Ghosh, P., Charnay, P., Burns, D. K., and Parada, L. F. 2002. Neurofibromas in NF1: Schwann cell origin and role of tumor microenvironment. Science. 296:920–922.
Index
A Abelson, H.T., 140 Acute lymphoblastic leukemia (ALL). See also Lymphoblastic lymphoma BCR/ABL tyrosine kinase and Ph+ ALL dasatinib and nilotinib, 5–6 imatinib mesylate, 4–5 potential toxicity, 5 SCT, 4 cancer stem cells self-renewing, 12 xenograft mouse model, 13 DNA technology, 3 FLT-3 pathway, MLL event free survival (EFS), 6 juxtamembrane mutations, 7 lestuartinib and midostaurin, 7–8 monoclonal antibodies alemtuzumab, 11–12 epratuzumab, 11 gemtuzumab ozogamicin, 12 rituximab, 10–11 therapeutic, 10 mutated gene, 4 Notch pathway c-Myc, 10 domain, 8–9 GSIs, 9–10 intracellular Notch (ICN), 9 members, 8 Acute megakaryocytic leukemia (AMkL) chromosome 21-localized genes, 112–115 description, 109–110 differential gene expression, 116–117 non-Down syndrome, treatment results, 111 somatic mutations, 116 Acute myeloid leukemia (AML). See also Down syndrome animal models, 74–75
development cellular and molecular targets, 62 cytoreduction, 61 drug resistance mechanisms cyclosporine, 69 P-glycoprotein, 68–69 proteasome inhibition, 70 epigenetic and chromatin remodeling DNA methylation and histone deacetylation, 71 gene expression, 70–71 sequencing, 72 FLI1 homolog ERG translocation, 380 immunotherapy and immunostimulatory gemtuzumab ozogamicin, 72–73 graft vs. host disease, 74 IL-2, 73–74 MDR1 transporter, 73 leukemogenesis mechanisms, 60 MDS and, 157, 163 relevance, 61 risk-adapted therapy, 59 signaling pathways, TK receptors FLT3, 67 FLT-3, 67 mTOR serine/threonine, 68 RAS mutations, 66 toxicity, 74, 75 tyrosine kinases, leukemia cell prolifeÂ� ration and survival CEP-701/lestaurtinib, 63–64 c-KIT mutations, 64–65 cytokines, 62–63 FLT3/ITD mutations, 63 inhibitors, 66 VEGF, 65 Acute promyelocytic leukaemia (APL) alemtuzumab, 101
521
522 Acute promyelocytic leukaemia (APL) (cont.) anthracycline cumulative dose and cardiotoxicity, 96 arsenic trioxide (ATO) induction and consolidation, 97 telomerase, 96 demographic features, 83–84 diagnosis distribution pattern, 86, 88 M3v, 86 reverse-transcription PCR, 88 tools, 87 failure, treatment antileukemic efficacy, 98 extramedullary (EM) relapse, 99–100 GO, 98–99 HSCT, 99 relapse, 97 salvage therapy, 97–98 FLT-3 inhibitors, 100 molecular monitoring MRD detection, 94–95 real-time quantitative RT-PCR assays, 95–96 pathogenesis chimeric protein, 85 PML-RAR and ATRA, 84–85 retinoic acid receptor-a(RARA) gene, 84 RA-sensitive and resistant cells, 100–101 SCT, 96 treatment anthracycline-based, 88–89 consolidation therapy, 92–94 induction therapy, 89–92 maintenance, 94 Adesina, A.M., 272 Akatsuka, T., 469 Al-Ahmari, A., 110 Albritton, K., 341 ALCL. See Anaplastic large cell lymphoma Aldosari, N., 272 ALL. See Acute lymphoblastic leukemia All-trans retinoic acid (ATRA), APL anthracycline-based chemotherapy, 88–89 arsenic trioxide (ATO) and, 96–97 consolidation therapy, 92–94 extramedullary relapses, 99 functions, 85 induction therapy, 89–92 maintenance chemotherapy, 94 PML-RARA fusion gene, 84 synergistic effect, 93 Alveolar RMS (ARMS), 425 AMkL. See Acute megakaryocytic leukemia
Index Ammatuna, E., 95 Anaplastic large cell lymphoma (ALCL) ALK rearrangement, 510 anaplastic lymphoma kinase (ALK) CD30, 197–198 CD52, 198 expression/function, fusion protein, 194–195 fusion partners, 194 inhibitors, 195–196 nucleophosmin (NPM), 193 overexpression, 193 proteasome inhibitors, 197 signaling pathways, 196–197 cDNA microarray analysis, 198 NPM-ALK multiple signaling pathways, 195 subtypes, 192–193 Anaplastic lymphoma kinase (ALK) activation, IMT, 510–511 Anderson, C.P., 356 Angiogenesis inhibition, astrocytomas blockade, VEGF, 245 avb3 integrin inhibitor cilengitide, 246 molecularly targeted therapies, 235 thalidomide, 245–246 VEGF receptor EGFR-mediated signaling, 244 ligand/receptor interactions, 244–245 Anninga, J.K., 470 Antimetabolite therapy, 160 APL. See Acute promyelocytic leukaemia Apoptotic tendency modulation, non-EWS-FLI1 DR4 and DR5, 383, 385 intrinsic and extrinsic pathway, 383 TRAIL receptors function, 385 Arceci, R.J., 59 Arcellana-Panlilio, M.Y., 403 ARMS. See Alveolar RMS Astrocytomas angiogenesis inhibition avb3 integrin inhibitor cilengitide, 246 thalidomide, 245–246 VEGF blockade, 245 VEGFR, 244–245 downstream signaling inhibitors, 240 drug resistance phenotype alkylguanine DNA alkyltransferase inhibition, 247–249 base-excision repair inhibition, 249–250 cell cycle regulatory proteins, 251–252 HDACi, 250–251 HSP and NF-kB, 251
Index epidermal growth factor receptor (EGFR) description, 238 gefitinib, 239 GW572016, 240 molecular signaling profile, 239–240 ZD 1839, 238–239 growth factor receptor and survival signaling angiogenic, 247 cytotoxic drugs, 246–247 high-grade gliomas anaplastic and oligoastrocytoma, 232–233 de€novo, 233 malignant astrocytoma, 233–234 immunological/ligand-based therapies IL13-PE38QQR, 253–254 receptors, 252–253 Tf-CRM107, 253 TP-38, 254 vaccine-based, 252 inhibitors, growth factor receptor, 236 low-grade gliomas juvenile pilocytic, 232 non-pilocytic low-grade, 232 mitogen-activated protein kinase (MAPK) cascade signaling pathway, 242–243 sorafenib, 243 PI3K/Akt pathway inhibitors description and PTEN, 243 glioblastoma multiforme, 244 platelet-derived growth factor (PDGF) receptor expression, 236–237 identification, 236 inhibitors, 237 proliferation, 237–238 STI571, 238 protein kinase C (PKC) inhibition, 242 RAS processing inhibition cell-surface receptors, 240–241 farnesylation, 241 farnesyl transferase inhibitors (FTIs), 241 Athale, U.H., 111 B Babovic-Vuksanovic, D., 340 Badiali, M., 272 Baeza, N., 272 Bagatell, R., 169, 356 Baker, M.J., 251 Bakhshi, S., 270 Banerjee, D., 356
523 Barel, D., 270 Barnard, D.R., 111 Basic fibroblastic growth factor (bFGF), 444 Basso, K., 190 Baudry, D., 403 Bayani, J., 272 Beckwith–Wiedemann syndrome (BWS), 404 Begum, S., 430 Berg, T., 433 Berthold, F., 356 Bevacizumab, 299 Bianco, R., 247 Biondi, A., 83 B lymphocyte-induced maturation protein 1 (BLIMP-1), 186–187 BMPs. See Bone morphogenetic proteins Boehrer, S., 168 Bone differentiation, osteosarcoma BMPs, 474 PTHrP, 474–475 TGF-b isoforms, 473–474 WNT signaling, 474, 486 Bone morphogenetic proteins (BMPs), 474 Bosutinib, 147 Brain tumor polyposis syndrome, 275 Braun, B.S., 123 Bredel, M., 241 Brignole, C., 356 Brown, K.W., 412 Brown, P.A., 31, 34 Burkitt lymphoma (BL) antisense oligodeoxynucleotides, 192 benzodiazepine, Bz-423, 190, 192 description, 188, 190 differential gene signatures, 189 MYC subnetwork, 191 NOTCH1 signaling, 192 1q+ abnormalities, 188 subnetwork, 190 C Cairo, M.S., 177 Calabrese, C., 296 cAMP response element-binding protein (CREB), 407 Cancer predisposition syndromes BWS, 404 medulloblastoma, 270 osteosarcoma hereditary, 461–462 incidence, 461 RECQ DNA helicases, 462 Candidate of metastasis 1 (COM1), 503
524 Carroll, W.L., 3 Castleberry, R., 341 Cell cycle regulation, osteosarcoma INK4A gene, 464–465 malignant proliferation, 464 MDM2 and COPS3 gene amplification, 465–466 p53 gene, 465 Chakraborty, S., 160 Cheung, N.K., 356 Chimeric transcription factors and oncogenes, RMS MYCN amplification and overexpression, 432–433 downregulation, 433 GSK609693, 434 PAX3-FKHR antisense oligonucleotides and siRNA, 430–431 CXCR4, entry fusion inhibitors, 431 downstream consequences, 429 ectopic expression, 429–430 elucidation, pathogenesis, 428 expression, ARMS, 428, 429 specificity, RNAi, 431, 432 TP53 dysregulation, 434–435 Cho, Y.-J., 267 Chronic myelogenous/myeloid leukemia (CML) Bcr-Abl pathophysiology breakpoints, 140 hematopoietic stem cells, 139 kinase cellular proliferation, 140 murine retroviral transplantation, 140–141 structure and signaling function, 141–142 cancer stem cells, 12 chronic phase (CP) accelerated phase (AP), 143 blasts count, 142 description, 139 genetic lesion, 27 imatinib, 64, 145, 388 kinase inhibitors resistance Bcr-Abl domain mutations, 147 bosutinib, 148 dasatinib (See Sprycel™) imatinib therapy, 147 nilotinib (See Tisigna™) second generation tyrosine, response, 148
Index T315I mutation, 148–149 monitoring PCR, 146–147 Philadelphia chromosome, 145–146 pediatric, 142 targeted therapies diuretics, 144–145 imatinib, 144 interferon-a, 143 response criteria, 143–144 transplantation allogeneic stem cell, 149 HLA-matched sibling donor, 149–150 Clifford, S.C., 272 CML. See Chronic myelogenous/myeloid leukemia Coffey, D.C., 356 COM1. See Candidate of metastasis 1 Consolidation therapy, APL daunorubicin and cytarabine, 93 molecular evidence, 92 Corey, S.J., 139 Cortes, J., 139 Cox, C.V., 13 CR-1. See Cripto-1 Creutzig, U., 110 Cripto-1 (CR-1), 511 Cyclin D1 (CDKs), 408 Cytochrome P450 inhibitors, 92, 144 Cytokine-regulated growth pathways, non-EWS-FLI1 IGF-IR, 385–386 imatinib, 388 KIT, 387–388 PDGF, 386–387 Cytotoxic agents, neuroblastoma, 356–357 Czibere, A., 170 D Dahmen, R.P., 272 Dave, S.S., 188, 189 Davicioni, E., 428 De Bortoli, M., 272 de Chadarevian, J.P., 270 Delta-serrate-lag2 (DSL) family, 8 Dermal neurofibromas, 333 Dermatofibrosarcoma protuberans (DFSP), 508 Diamanti, P., 13 Diarrhea, 144 Dickson, P.V., 356
Index Diffuse large B-cell lymphoma (DLBCL) BCL-6 gene CD80, 182 chromatin remodeling, 183 expression deregulation, 181 germinal center-derived DLBCL (GC DLBCL), 180–181 germinal center (GC) phenotype, 180 peptide interference peptides, 183 protein interaction, 182 transcriptional regulation, 181–182 BLIMP-1, 186–187 B-NHL, 178–179 FAB/LMB96 international B-NHL trial, 183–184 genetic subtypes, 179 molecular targeting, 184 NF-kB pathway differential expression, 187 gene expression and transcriptional activation, 184–186 IkB proteins polyubiquitination, 186 rituximab, 186 PMBL children and adolescents, 180 differential expression, 184 gene expression, 180–181 IKK inhibitors, 187–188 Diller, L., 403 Distel, L., 270 DNA methyltransferase (DNMT) inhibitors gene transcription, 164, 166 HDACi, 166 p21Cip1 reexpression, 442 Doepfner, K.T., 67 Dome, J.S., 401 Douer, D., 83 Down syndrome (DS) AAML041, 117 acute myeloid leukemia DS ALL, 112 event-free survival (EFS) rate, 110, 111 GATA1, 115–116 infectious complications, 112 chemotherapy sensitivity anthracycline, 112 GATA1, 115–116 megakaryoblasts, 111 toxicity, 111–112 chromosome 21-localized genes allosteric regulation, 113 cystathionine-b-synthase (CBS), 112–113 GATA1, role of, 114–115 differential gene expression bone marrow stromal cell antigen 2 (BST2), 116–117
525 megakaryocytic blast maturation, 117 microarrays, 116 leukemia, 109 non-Down syndrome AMkL cases, 111 resistance, chemotherapy, 118 treatment outcome, 110 Drug resistance, astrocytomas alkylguanine DNA alkyltransferase inhibition Gliadel wafer implantation, 248–249 mechanism, 247–248 O6-BG, 249 O6-methylguanine, 248 base-excision repair inhibition cell cycle regulatory proteins, 251–252 DNA mismatch deficiency, 249–250 HDACi, 250–251 HSP and NF-kB, 251 INO-1001, 250 DS. See Down syndrome Duchayne, E., 111 Dyspnea, 144–145 E Eberhart, C.G., 272 Ebinger, M., 272 E-cadherin, 503 EGFR. See Epidermal growth factor receptor Ellison, D.W., 272 Ependymoma cancer stem cells and origins gene expression, 295–296 identification, 295 neural progenitor, 296 epidemiology and histopathology, 292 genetics, 294 genomics amplicons, 295 molecular alterations, 294–295 molecular targeted therapies, 297 stem cell niches, 296 therapy conventional chemotherapy, 293–294 surgery and radiation, 292 total tumor resection, 292–293 treatment improvement cancer stem cells, 299 disease-risk stratification system, 298 gene mutations, 298–299 JAGGED ligands binding, 297–298 minimal conventional therapy, 298 subgroups, genetic alterations, 297 Epidermal growth factor receptor (EGFR), 504 Epstein–Barr virus (EBV) apoptotic resistance, 220–221
526 EBV-PTLD molecular targets Epstein–Barr virus (EBV) (cont.) anti-viral medications, 221–222 immunosuppression, 222 monoclonal antibodies, 222–223 T cell adoptive immunotherapy, 223 vaccines, 221 host immune response cellular, 218–220 cytokine modulation and viral IL-10, 220 IgM and IgG titers, 218 infection, human B cell differentiation, 217–218 description, 217 ERBB2 proto-oncogene, 467–468 Erez, A., 272 ETV6-NTRK3 fusion protein, 506–507 Evans, A.E., 356 Ewing’s sarcoma family of tumors (ESFT) CD99MIC2 expression, 374 description, 373 discovery, 374 EWS-FLI1 approaches, small molecule identifiÂ�cation, 382 chromosomal translocations, 376–377 DNA binding, disruption, 383 elimination of, 379 expression, 374–375 single-chain antibodies, 382 small-molecule inhibitors, 380–381 targets of, 382–383 transcription and splicing, 377–379 YK-4-279, 381–382 features, 373–374 models murine, 376 spheroid, 375 xenograft, 375–376 non-EWS-FLI1 apoptotic tendency modulation, 383–385 current clinical trials, 383–384 cytokine-regulated growth pathways, 385–388 molecular targets, 388–390 targeted molecular agents, 380 Ewing’s sarcoma (EWS)-Friend leukemia insertion (FLI1) approaches, small molecule identification, 382 chromosomal translocations, 376–377 DNA binding, disruption, 383 elimination of, 379
Index ESFT molecular targeting, 380 expression, 374–375 RNA splicing mRNA, 377–378 U1C identification, 377 single-chain antibodies, 382 small-molecule inhibitors peptide sequences, 381 protein–protein interactions, 380–381 targets of, 382–383 transcriptional regulation hsRPB7, 377–378 and RNA metabolism, 377 YK-4-279 novel compound, 381–382 RHA region, 381 Extramedullary (EM) relapse, APL, 99–100
F Faderl, S., 5 Farnesyl transferase inhibitors (FTIs) AML, 66 antiproliferative effects, 241 NF1 related tumors, 339 RAS-dependent signaling, 442, 505 Fate-mapping experiments, 269 Feldkamp, M.M., 241 Felix, C.A., 31 Fernandeaz, C.V., 403 Feuerhake, F., 187 FMS-like tyrosine kinase-3 (FLT-3) childhood leukemia, 7 description, 6–7 inhibitors, APL, 100 midostaurin inhibitor, 8 receptor, 6 Fox, E., 356 Frappart, P.O., 273 Friedman, H.S., 249 FTIs. See Farnesyl transferase inhibitors
G Gallicchio, M.A., 439 Gamis, A.S., 110 Garvin, A.J., 412 Gastrointestinal stromal tumors (GISTs) description, 508–509 gene encoding cKIT, 509 phase II and III clinical trials, 509–510 second-line therapies, 510 Geller, J.I., 401
Index Genasense™ BCL-2 mRNA expression, 39–40 refractory/relapsed adult acute leukemia, 39 Gessler, M., 403 Ge, Y., 109 Gibson, B.E.S., 83 Gilbert, J., 251 Gilbertson, R.J., 291 GISTs. See Gastrointestinal stromal tumors Gleevec. See Imatinib Goodrich, L.V., 273 Gore, L., 155 Gorlick, R., 459, 469 Grignani, F., 84 Grimwade, D., 86, 88, 94, 95 Gross, T.G., 215 Grundy, P., 403 Gupta, A., 340 Guran, S., 270 Gutierrez, A., 19 H Hahn, H., 270 Hallahan, A.R., 273 Hamilton, S.R., 270 Hanahan, D., 477 Heat shock proteins (HSP). See also Astrocytomas HSP90, 46–48 anticancer therapeutics, 389 NPM-ALK, 511 Phase I trial, 483 PPTP, 415 HSP 27 and 70, 411 Hedgehog signaling pathway, 441 Hematopoietic stem cell transplantation (HSCT) allogeneic, 98, 123 AML graft vs. host disease, 74 therapy, 59 frontline therapy, 88 modality, 99 molecular disease, 96 PTLD, 215–316 rapamycin, 132 veno-occlusive disease (VOD), 72 Hepatocyte growth factor (HGF) HGF/c-MET, 410–411 ligand binding prevention, 440–441 MET, 439–440 PAX3-FKHR, 440 HER2/neu. See ERBB2 proto-oncogene Hirsch, B., 270
527 Histone deacetylase inhibitors (HDACi) cell cycle arrest, 72 histone deacetylases (HDAC) and, 166 as investigational agent, 356 neuroblastoma, 362–363 p21Cip1, 442 potential role, 100–101 vivo activity, WT, 414 Holcomb, V.B., 273 Ho, R., 356 Hosoi, H., 425 Houghton, P.J., 356, 425 Huang, H., 272 I IGF-2. See Insulin-like growth factor-2 (IGF)-2 IGF1R. See Insulin-like growth factor-1 receptor IGF signaling, WT IGF2/IGFR1, 405 IGFR1, IGFR2 and IGF2, 404–405 and mTOR, 405 IL13-PE38QQR domains, 253–254 resection, 254 Imatinib and chemotherapy, 5 children, 145 CML, 27, 64 description, 143 mesylate Abl, KIT, and PDGFR inhibition, 388 vs. ATP, 509–510 block signals stemming, PDGF receptors, 508 inhibitor, ABL, 484 neuroblastoma, 358 RMS, 426 resistance, 5–6 131 I-metaiodobenzylguanidine (131I-MIBG), 359 Immunomodulatory drugs (IMiDs), 167 IMT. See Inflammatory myofibroblastic tumor Induction therapy, APL ATRA and anthracycline, 89 complications APL differentation syndrome, 91–92 coagulation parameters, 91 headache and pseudotumor cerebri, 92 hemorrhage, 90–91 sequential treatment, 90 Infant ALL amino and carboxyl terminal fragments, 32 anti-apoptotic BCL-2 family members ABT-737, 43–44
528 Infant ALL (cont.) ABT-869, 44 ADVL0816, 42–43 basal expression levels, 42 classes, 39 drug resistance, 43 Genasense™, 39–40 homo-and heterotypic dimers, 40 neuroblastoma, 41 obatoclax, 40–41 silencing, 38 CD33 cell surface antigen, 49 epigenetic strategies, 48 FLT-3 tyrosine kinase inhibition CEP-701, 34–35, 37–38 chemotherapy regimens, 36 LSC, 36–37 microarray studies, 33 mutations, 33–34 plasma inhibitory activity (PIA), 37 signaling, 34 glycogen synthase kinase 3 small molecule inhibitors, 46 transduction, 46–47 HSP90, potential therapeutic target description, 47 inhibition, 47–48 MLL fusion transcripts, 45 mTOR inhibition glucocorticoid resistance, 48–49 serine/threonine kinase, 48 oncoprotein transcriptional deregulation, 33 partner genes, 31 PFWT, MLL partner protein interactions AF4-AF9 complexes, 45 synergistic, 46 treatment options, 31–32 Infantile fibrosarcomas kinase activity, 507 recurrent translocation, 506 Inflammatory myofibroblastic tumor (IMT) ALK rearrangement, 510–511 description, 510 INI1/hSNF5 tumor suppressor, RTs. See also Rhabdoid tumors differentiation cell of origin, 315 nonRT and normal cells, 314 interferon signaling, 314 loss cell characterization, 312–313 G0/G1 arrest, 313 mitotic spindle checkpoint activation, 313
Index mechanisms chromatin remodeling complexes, 310 SWI/SNF complexes, 310–311 Insulin-like growth factor-2 (IGF)-2, 503, 504 Insulin-like growth factor-1 receptor (IGF1R) characteristics, 385 inhibition, ESFT patients, 386 medulloblastomas description, 276–277 gene expression, 277 inhibitors, 277–278 Phase I and II studies, sarcomas, 485 predominant receptor, cancer biology, 466 RMS, 435–436, 438–439 synovial sarcomas, 504 VEGF production, 445 Insulin-like growth factor signaling, RMS CTCF engenders, 438 H19/IGF-2 expression, 437–438 IGFBPs, 439 IGF-2 mRNA, 437 IGF-1R signaling, 435–436, 438 specificity, 438–439 loss of heterzygosity (LOH), 437 International Prognosis Scoring System (IPSS), 158–159 J Jakacki, R., 340 Johnson, R.L., 270 Juvenile myelomonocytic leukemia (JMML), RAS signaling pathways description, 123 GM-CSF and deregulated RAS signaling effector cascades, 125 intracellular phosphoprotein analysis, 126–127 neurofibromin, 124–125 PTPN11 mutations, 125–126 quantitative assays, 126 RAS proteins, 124 JAK2-STAT5, 133–134 molecular genetics, 123–124 mouse models advantages, 128 conditional alleles, 127–128 Cre recombinase protein, 127 PI3K/AKT/mTOR AML, 132–133 rapamycin, 132 PTPN11/SHP-2, 133
Index RAF/MEK/ERK activation, 130–131 CI-1040, MEK inhibitor, 131–132 colony formation, 131 Nf1 mutant, 132 sorafenib, 131 targeted therapeutics activated RAS, inhibition, 128 CAAX protein, 128–129 oncogene addiction, 130 prenylation, 129 Juvenile pilocytic astrocytomas, 232 K Kalpana, G.V., 305 Karp, J.E., 65 Keshelava, N., 356 Kikuchi, K., 425 Kilpatrick, S.E., 470 Knudson’s two-hit hypothesis. See Wilms’ tumor Koch, A., 272 Koesters, R., 403 Kohler, 72 Kruppel-associated box (KRAB) repression domain, 503 transcriptional repressor, 379 Kudo, K., 110 Kurmasheva, R.T., 425 Kushner, B.H., 356 L Lae, M., 428 Lam, C.W., 272 Lam, L.T., 187 Lauchle, J.O., 123 Lee, Y., 273 Leiomyosarcoma, 511–512 Lenalidomide, 167 Leukemia initiating cells (LICs). See Leukemic stem cells Leukemic stem cells (LSCs) self renewal, 12 xenograft model, 36 Lie, S.O., 111 Lim, M.S., 177 Linardic, C.M., 427 Lo-Coco, F., 95 Lode, H.N., 356 Look, A.T., 19 Lorand-Metze, I., 157 Lymphoblastic lymphoma (LLs) ALCL molecular targets, 201 vs. ALL, 199
529 B-precursor (B-LLs) cytogenetic abnormalities, 205 and T-LLs, 199 cytogenetic and molecular changes, 200 description, 198–199 molecular abnormalities, 201–202 TAL1, 202–203 T-ALLs cryptic aberrations, 199–200 gene expression, 202 NOTCH1 mutations, 203–204 T-cells (T-LL) chromosomal translocations, 199 vs. T-ALL, 20, 200–201 M Maitra, A., 469 Major, M.B., 403 Maki, R.G., 341 Malignant fibrous histiocytoma (MFH), 507 Malignant gliomas genetic alterations, 234 molecular feature, 234 oligodendroglial features, 233–234 Malignant peripheral nerve sheath tumors (MPNSTs) description, 504 features, 505 molecular features, 338 and neurofibromas, 505–506 Malignant rhabdoid tumors (MRTs). See Rhabdoid tumors (RTs) Mammalian target of rapamcyin (mTOR) inhibitors, 279 NF1, 337–338 PI3K, 278–279 Marino, S., 273 Maris, J.M., 351, 356 Marks, P.A., 356 Marshall, J.L., 356 Matthay, K.K., 356 Maurer, B.J., 356 McDonald, J.M., 403 McKinnon, P.J., 273 Medulloblastomas cancer predisposition syndromes, 270 cellular origins cancer stem cells, 269, 271 pathogenesis pathways, 271 precursors, 269 classic, 267–268 molecular features CXCR4, 278
530 Medulloblastomas (cont.) EGFR family, 279–280 HDAC inhibitors and retinoic acid, 280–281 IGF1R (See Insulin-like growth factor-1 receptor) NOTCH, 276 PDGFR and TRKC, 280 PI3K/AKT/mTOR, 278–279 SHH (See Sonic hedgehog) WNT/b-catenin, 275–276 molecular pathways and targeted therapies, 271 mouse models, 273 nodular architecture, 268 risk stratification desmoplastic tumors, 268 diagnosis, 268–269 sporadic, mutations, 272 subtype classifications, 267 Methotrexate (MTX) DS patients, 112 human ALL xenografts, 48 osteosarcoma, 477 T-LL, 199 MFH. See Malignant fibrous histiocytoma Michiels, E.M., 272 Miller, B.S., 356 Milstein, 72 Minimal residual disease (MRD) detection, APL fusion transcripts, 95 leucocyte count, 94 relapse risk, 94–95 Mistry, A.R., 84 Mitotic spindle inhibition, neuroblastoma, 362 Mixed lineage leukemia (MLL) BCL2 family members, 38 FLT-3 pathway, 6–8 fusion transcripts, 45 native protein, 32 partner protein interactions, 45–46 rearrangements, 33–34, 36, 38, 42, 71 translocations, 31, 32 Modak, S., 356 Molecular targets, non-EWS-FLI1 ET-743, 388–389 HDACi, 389 HSP, 389 protein synthesis and degradation, 389–390 Morris, E.S., 11 Mossé, Y.P., 351 MPNSTs. See Malignant peripheral nerve sheath tumors MTP-PE. See Muramyl tripeptide phosphatidylethanolamine
Index MTX. See Methotrexate Multidrug resistance (MDR1) gene, 476–477 Multidrug-resistance protein (MRP), WT, 412 Muramyl tripeptide phosphatidylethanolamine (MTP-PE), 482 Myelodysplastic syndromes (MDS) vs. AML, 163 bone marrow angiogenesis, 65 chromatin-based transcriptional therapy DNMT inhibitors, 164–166 epigenetic and chromatin changes, 164 HDACs and HDACi, 166 classification adult and pediatric, 158–159 FAB and WHO, 157 IPSS, 156–157 myelofibrosis/myeloproliferative syndromes, 156 and DS, AML, 110 etiology chemotherapy/toxin exposure, 160 development, 157 genetic abnormalities, 160–162 hematopoietic disorders, 160, 163 predisposing genetic syndromes, 160 5q31 deletions, 163 IMiDs and anti-angiogenic agents, 167 incidence, 156 maturation-directed therapies, 167 SCT adults, 170 pediatric patients, 170 regimens, adults, 169 timing, 169–170 TKIs, 168 Myelosuppression gemtuzumab ozogamicin, 98–99 occurance, 145 N Nabarro, S., 430 Natrajan, R., 403 Neuroblastoma description, 351 locoregional tumors, 353 metastatic tumors common agents, relapse setting, 354 retinoids, 354–355 overall survival probability, high-risk patients, 351–352 relapse setting, high-risk disease angiogenesis inhibitors, 361 cytotoxic agents, 356–357 histone deacetylase inhibitors, 362–363
Index immunotherapy, 359–360 investigational agents, 355–356 mitotic spindle inhibition, 362 retinoids, 360–361 stem cell conundrum, 358–359 targeted delivery, radionucleotides, 359 targeting MYCN, 361–362 tyrosine kinase inhibitors, 357–358 risk classification system chromosomal aberrations, 353 MYCN oncogene amplification, 352 strategies, molecular targets identification, 355 Neurofibromas dermal, 333 and MPNST, 505–506 plexiform, 333–334 Neurofibromatosis type 1 (NF1) age stratification, 344 combination therapy, 345 description, 331 development, targeted treatments completed, ongoing and planned clinical trials, 339–341 phase II trial, TTP, 342–343 refractory cancer trial, 339, 342 diagnosis, 333 infrastructure, clinical trials, 343 molecular features, tumorigenesis angiogenesis, 336–337 growth factor receptors, 337 mTOR, 337–338 RAS pathway, 335–336 rationale, pathways, 339 tumor origin and environment, 336 MPNSTs, 338 pathway structure, 332 pilocytic astrocytomas, 232 preclinical evaluation, agents, 343 RAS activity, 332 target validation, 344 tumor manifestations dermal neurofibromas, 333 JMML, 335 malignant peripheral nerve sheath, 334 optic pathway gliomas, 334–335 PN, 333–334 NHL. See Non-Hodgkin lymphoma Nitta, T., 236 Non-Hodgkin lymphoma (NHL) BL (See Burkitt lymphoma) DLBCL ABC-DLBCL and GC-DLBCL, 183–184 BCL-6 gene, 180–183
531 BLIMP-1, 186–187 CARD11, 184 childhood and adolescent, 178–179 CHOP therapy, 179 gene expression profiles, 179–180 molecular targeting, 184 NF-kB pathway, 184–186 PMBL, 187–188 epratuzumab, 11 LLs (See Lymphoblastic lymphoma) Non-pilocytic low-grade astrocytomas, 232 Nonrhabdomyosarcoma soft tissue sarcomas (NRSTS) adjuvant VACA, 500 cellular origin, 502 clinical trials, 512–513 description, 499 DFSP, 508 fundamental groups chromosomal defects, 501–502 development mechanisms, 501 GISTs, 508–510 IMT, 510–511 infantile fibrosarcomas and ETV6-NTRK3 fusion protein, 506–507 leiomyosarcoma, 511–512 MFH, 507 MPNST and NF1 mutation, 504–506 POG, 499–500 synovial sarcoma, 502–503 SYT-SSX gene fusions, 503–504 NOTCH1 inhibition gastrointestinal toxicity, 23–24 g-secretase complex, 23 pathway description, 22 identification, 21 mutations, 22–23 resistance FBW7 tumor suppressor, 24 PTEN inactivation, 24–26 Nowicki, M., 409 NRSTS. See Nonrhabdomyosarcoma soft tissue sarcomas O Occhipinti, E., 157 Offit, K., 270 Ohgaki, H., 272 Oki, Y., 111 Onda, M., 469 Osenga, K.L., 356
532 Osteosarcoma description, 459 etiologic factors, 459–460 molecular alterations angiogenesis, 475–476 bone differentiation, 473–475, 485–486 cancer predisposition syndromes, 461–463 cell cycle regulation, 464–466 cell of origin, 460–461 drug resistance, 476–477 genetic complexity, 472 growth factors, 466–468 immortalization, 472–473 metastasis, 475 murine models, 463–464 oncogenes, 471–472 signal transduction, 468–471 SV40, 463 murine models, 463–464 pathogenesis genetic alterations, 479–481 mechanisms, 482 pediatric malignancies, 477–478 PPTP, 486–487 signal transduction, 468–471 SV40, 463 treatment strategies cytotoxic agents, 478 HER2 expression, 484 Hsp90 inhibition, 483–484 IGF-1, 484–485 liposomal doxorubicin, 483 MTP-PE, 482 pamidronate role, 483 radiation therapy, 478, 482 RANK/RANKL/OPG system, 485 rapamycin and VEGF pathway, 484 P Pagano, L., 111 Palomero, T., 24 Parathyroid hormone-related protein (PTHrP), 474–475 Parham, D.M., 500 Pasqualucci, L., 187 PDGF. See Platelet-derived growth factor PDGFR-b. See Platelet-derived growth factor receptor-b Pearson, A.D., 270 Pediatric preclinical testing program (PPTP) ABT-751 and Ixabepilone, 414 novel agents, tested, 486 osteosarcoma incidence, 486–487
Index tested agents, 415 xenografts utilization, 415 Perotti, D., 403 Pgps. See Polycomb group proteins Pieters, R., 3 Pietsch, T., 272 Pinkerton, R., 178 Pinthus, J.H., 414 Piva, R., 194 Platelet-derived growth factor (PDGF) and fibroblast growth factor (FGF), 234 isoforms, 236 neutralizing antibody, 236 overexpression, 387 PDGF-C and PDGF-D families, 386–387 potent mitogen, 467 Platelet-derived growth factor receptor (PDGFR) inhibition and inhibitors, 237 kinase activity, 247 PDGF isoforms, 236–237 PDGFRb signaling, 280 pediatric glial neoplasia, 231 signaling, 237–238 Platelet-derived growth factor receptor-b (PDGFR)-b, 504 Plexiform neurofibromas (PN), 333–334 PMBL. See Primary mediastinal B-cell lymphomas Pollack, I.F., 231 Polo, J.M., 183 Polycomb group proteins (Pgps), 503 Pomalidomid, 167 Pomeroy, S.L., 267 Post-transplant lymphoproliferative disorders (PTLD) classification, 215–216 EBV anti-viral medications, 221–222 apoptotic resistance, 220–221 cytotoxic T cells (CTLs), 216 host immune response, 218–220 human, 217–218 immunosuppression, 222 monoclonal antibodies, 222–223 T cell adoptive immunotherapy, 223 treatment, 217 vaccines, 221 occurance, 215 signs and symptoms, 216 PPTP. See Pediatric preclinical testing program Primary mediastinal B-cell lymphomas (PMBL) DLBCL and, 186–188, 205 gene expression, 180–181 prognosis, children and adolescents, 180
Index Pritchard-Jones, K., 403 PTEN Akt activation, 243 signaling, 247 high-grade gliomas, 234 inactivation, NOTCH1 inhibition, 24–26 preservation, 240 T-ALL, 26 transcriptional repression, 23 PTHrP. See Parathyroid hormone-related protein PTLD. See Post-transplant lymphoproliferative disorders R Raetz, E.A., 200 Raffel, C., 272 Rahman, N., 403 Ramburan, A., 403 Ranked list overlap analysis, RMS, 427–428 Rao, A., 110 Rao, G., 273 Rapamycin analogs, 484 medulloblastoma, 279 mTOR pathway, 204 VEGF, 444–445 Ravenel, J., 403 Ravindranath, Y., 109–111 Real-time quantitative RT-PCR (RQ-PCR) assays, APL, 95–96 Reardon, D.A., 238 Reifenberger, J., 272 Reinhardt, D., 111 Retinoids, neuroblastoma, 360–361 Reynolds, C.P., 356 Rhabdoid tumors (RTs) aggressiveness, 323 cell of origin cellular differentiation pathways, 315 INI1/hSNF5 reintroduction, 314 mesenchymal, 315 clinical trials molecular mechanism, 309–310 oxaliplatin, 309 peripheral blood stem cell rescue, 308 cyclin/cdk axis G0–G1 arrest and apoptosis, 318 phosphorylation, 318–319 cyclin D1 mouse model, 318 over-expression, 317–318
533 development, targeted therapy gene expression profile analysis, 321 interferon signaling pathway, 322–323 mitotic spindle checkpoint, 322 diagnosis cells, 306–307 description, 306 molecular genetics, 307 drug action, 323–324 flavopiridol, 321 4HPR apoptosis, 320 description, 319–320 ifosfamide, 308 INI1 loss characterization, 312–313 G0/G1 arrest, induction, 313 interferon signaling, 314 mitotic spindle checkpoint activation, 313 pathology, 306 radiotherapy, 307–308 suppression mechanisms, INI1-mediated core subunits, 311 hypothesis, 316–317 SWI/SNF complex, 310 transcription regulation, 311–312 surgical intervention, 307 Rhabdomyosarcoma (RMS) angiogenesis mTOR signaling, 445 neovascularization, solid tumors, 443 rapamycin, 444–445 VEGF stimulates proliferation, 443–444 ARMS, 425 cell of origin AML, 426 PAX7 and PAX3, 427 chimeric transcription factors and oncogenes MYCN dysregulation, 432–434 PAX3-FKHR, 428–432 TP53 dysregulation, 434–435 description, 425 ERMS, 425–426 genome expression profiling ERMS and ARMS, 427 PAX-FKHR expression, ARMS, 428, 429 ranked list overlap analysis, 427–428 growth factor signaling hedgehog, 441 HGF, 439–441 IGF-1R, 435–439 RAS/MAPK/c-KIT, 441–442 molecular characteristics, 426
534 Rhabdomyosarcoma (RMS) (cont.) myogenic differentiation cyclin D-CDK4 activity, 443 E protein heterodimerization blockage, 442–443 p21Cip1 reexpression, 442 pharmacologic inhibition, CDK1, 446 Ricafort, R., 459 Rivera, M., 403 RMS. See Rhabdomyosarcoma Rosenwald, A., 179, 180 Rossi, M.R., 272 Roy, D., 403 RTs. See Rhabdoid tumors Ruiz-Argüelles, A., 111 Ruud, E., 270 S Safford, S.D., 403 Sato, K., 236 Savasan, S., 117 Saylors, R.L., 272 Schulz, S., 403 Scotlandi, K., 470 g-Secretase inhibitors (GSIs) NOTCH1 activation, 23 T-ALL, 9–10 Shah, N.P., 356 SHH. See Sonic hedgehog Shipp, M.A., 180 Siapati, K.E., 356 Signal transduction pathways, osteosarcoma c-MET overexpression, 466–467 ERBB family genes clinicopathological studies, 469–470 ERBB4, 468 ERBB2 proto-oncogene, 467 IGF-IR, 466 PDGF, 467 PI3K/AKT and mTOR, 471 RAS/RAF/MAPK, 468, 471 Simian virus 40 (SV40), 463 Singh, K.P., 403 Skapek, S.X., 499 Slack, C., 356 Small, D., 59 Small hairpin RNAs (shRNAs), 183 Smith, M.E., 305 Sohara, Y., 356 Solid organ transplantation (SOT), 215 Sonic hedgehog (SHH) medulloblastoma cyclopamine and HhAntag, 274–275
Index pathway, 273 signaling, 273–274 stem cell determination and renewal, 358 Sprycel™, 147–148 Stark-Vance, V., 245 Stem cell transplantation (SCT) imatinib post transplant, 5 myelodysplastic syndromes (See Myelodysplastic syndromes) veno-occlusive disease (VOD), 72 Strizzi, L., 511 Sundberg, T.B., 190 SV40. See Simian virus 40 Synovial sarcoma description, 502–503 and SYT-SSX gene fusions, 503–504 SYT-SSX gene fusions IGF-2 induction, 504 and synovial sarcoma, 503–504 SYT-SSX1 vs. SYT-SSX2, 503 T Takahashi, Y., 434 T-ALL. See T cell acute lymphoblastic leukemia Tallman, M.S., 111 Taub, J.W., 109 Taylor, M.D., 272, 295 T cell acute lymphoblastic leukemia (T-ALL) gene expression, 199–202 mutations, oncogenes, 203–204 NOTCH1 inhibition, 23–24 pathway, 21–23 resistance, 24–26 oncogenic transcription factors chromosomal translocations, 20 protein–protein interactions, 21 PTEN, PI3K-AKT, and mTOR, 26 RAS, 27 surface antigen expression, 19–20 TCR gene rearrangements, 19 tyrosine kinase genes, 26–27 Teitz, T., 356 Tentori, L., 250 Testi, A.M., 83 Tf-CRM107, 253 Thiopurine methyltransferase (TPMT) drug, 160 Thomas, D.A., 5, 11 Thomas, D.G., 470 Thompson, M.C., 272 Tipifarnib time to progression (TTP), 342 Tischkowitz, M.D., 270
Index Tisigna™, 147–148 TKIs. See Tyrosine kinase inhibitors TNF-related apoptosis inducing ligand (TRAIL) receptors, 385 WT, 410 Tomescu, O., 430 Tong, W.M., 273 Toretsky, J.A., 373 TP-38, 255 Tumor suppressor P53 (TP53) dysregulation HDM2-p53 interactions, 434–435 nutlin 3a, 435 TAp73 isoforms production, 434 Turcot’s syndrome, 294 Tyrosine kinase inhibitors (TKIs) CML theraphy, 168 dasatinib and nilotinib, 5–6 neuroblastoma, 357–358 U Üren, A., 373 Uziel, T., 273 V Van Waes, C., 185 Vascular endothelial growth factor (VEGF) AML, 65 angiogenesis, 408–409 astrocytomas blockade, 245 receptor inhibition, 244–245 HDM2 gene, 435 PI3K/Akt pathway, 444 VEGF. See Vascular endothelial growth factor Vincristine, actinomycin D, cyclophosphamide and doxoru bicin (VACA), 500 von Recklinghausen disease. See Neurofibromatosis type 1 (NF1) Vora, A., 11 Vujanic, G., 403 W Wachtel, M., 428 Wang, C., 430 Wang, M., 215 Weinberg, R.A., 477 Wesenberg, F., 270 Wetmore, C., 273 Widemann, B.C., 331, 340 Wilms’ tumor (WT) description, 401
535 molecular pathways apoptosis activators, 409–410 cell cycle, 410 c-KIT, 411 demethylating agents, 411–412 EGFR, 408 HGF/c-MET, 410–411 HSP 27 and 70, 411 IGF signaling, 404–405 MRP, 412 putative genetic loci/genes involvement, 402–403 VGEF role, 408–409 WNT-b-catenin, 406–408 preclinical identification, novel agents anti-EGFR and angiogenesis testing, 414 PPTP, 414–415 in€vitro investigation, 412–413 in€vivo xenograft testing, 413 treatment strategies adjuvant and neoadjuvant chemotherapy, 401 SIOP-2001 and AREN03B2 protocol, 402 WNT-b-catenin, WT cDNA microarray studies, mutated vs. nonmutated WT, 406–407 E-cadherin, 406 inhibitors, 407–408 networking with FGF, Notch, BMP and Hedgehog, 408 WNT-4, 406 Wnt/beta-catenin pathway, medulloblastoma activation and mutations, 275 ZTM000990 and PKF118-310, 275–276 Wong, K.K., 232 WT. See Wilms’ tumor X Xenograft testing, in€vivo, 413 Y Yan, C.T., 273 Yang, Q., 356 Yee, D., 356 Yokoi, A., 414 Yokota, N., 272 Yu, A., 356
536 Z Zeller, B., 110 Zhang, L., 430 Zhan, S., 437 Zhao, J., 403
Index Zhou, H., 356, 469 Zhou, L., 164 Zipursky, A., 110, 118 Zurawel, R.H., 272