Electrokinetics in Microfluidics
INTERFACE SCIENCE AND TECHNOLOGY Series Editor: ARTHUR HUBBARD In this series: Vol. 1: Clay Surfaces: Fundamentals and Applications Edited by F. Wypych and K.G. Satyanarayana Vol. 2: Electrokinetics in Microfluidics By Dongqing Li
INTERFACE SCIENCE AND TECHNOLOGY - VOLUME 2
Electrokinetics in Microfluidics
Dongqing Li Department of Mechanical & Industrial Engineering University of Toronto Toronto, Ontario, Canada
2004
ELSEVIER ACADEMIC PRESS
Amsterdam - Boston - Heidelberg - London - New York - Oxford - Paris San Diego - San Francisco - Singapore - Sydney - Tokyo
ELSEVIER B.V. Sara Burgerhartstraat 25 P.O. Box 211, 1000 AE Amsterdam The Netherlands
ELSEVIER Inc. 525 B Street, Suite 1900 San Diego, CA 92101-4495 USA
ELSEVIER Ltd The Boulevard, Langford Lane Kidlington, Oxford OX5 1GB UK
ELSEVIER Ltd 84 Theobalds Road London WC1X8RR UK
© 2004 Elsevier Ltd. All rights reserved. This work is protected under copyright by Elsevier Ltd, and the following terms and conditions apply to its use: Photocopying Single photocopies of single chapters may be made for personal use as allowed by national copyright laws. Permission of the Publisher and payment of a fee is required for all other photocopying, including multiple or systematic copying, copying for advertising or promotional purposes, resale, and all forms of document delivery. Special rates are available for educational institutions that wish to make photocopies for non-profit educational classroom use. Permissions may be sought directly from Elsevier's Rights Department in Oxford, UK: phone (+44) 1865 843830, fax (+44) 1865 853333, e-mail:
[email protected]. Requests may also be completed on-line via the Elsevier homepage (http://www.elsevier.com/locate/permissions). In the USA, users may clear permissions and make payments through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA; phone: (+1) (978) 7508400, fax: (+1) (978) 7504744, and in the UK through the Copyright Licensing Agency Rapid Clearance Service (CLARCS), 90 Tottenham Court Road, London W1P 0LP, UK; phone: (+44) 20 7631 5555; fax: (+44) 207631 5500. Other countries may have a local reprographic rights agency for payments. Derivative Works Tables of contents may be reproduced for internal circulation, but permission of the Publisher is required for external resale or distribution of such material. Permission of the Publisher is required for all other derivative works, including compilations and translations. Electronic Storage or Usage Permission of the Publisher is required to store or use electronically any material contained in this work, including any chapter or part of a chapter. Except as outlined above, no part of this work may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without prior written permission of the Publisher. Address permissions requests to: Elsevier's Rights Department, at the fax and e-mail addresses noted above. Notice No responsibility is assumed by the Publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made. First edition 2004 Library of Congress Cataloging in Publication Data A catalog record is available from the Library of Congress. British Library Cataloguing in Publication Data A catalogue record is available from the British Library. ISBN: 0-12-088444-5 ISSN: 1573-4285 (series) The paper used in this publication meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). Printed in The Netherlands.
Preface A lab-on-a-chip device is a microscale laboratory on a credit-card sized glass or plastic chip with a network of microchannels, electrodes, sensors and electronic circuits. Electrodes are placed at strategic locations on the chip. Applying electrical fields along microchannels controls the liquid flow and other operations in the chip. These labs on a chip can duplicate the specialized functions as performed by their room-sized counterparts, such as clinical diagnoses, PCR and electrophoretic separation. The advantages of these labs on a chip include significant reduction in the amounts of samples and reagents, very short reaction and analysis time, high throughput and portability. Generally, a lab-on-a-chip device must perform a number of microfiuidic functions: pumping, mixing, thermal cycling/incubating, dispensing, and separating. Precise manipulation of these microfiuidic processes is key to the operation and performance of labs on a chip. The recent, rapid development of the emerging microfiuidics and lab-on-achip technology brings a strong demand of understanding the interfacial electrokinetic phenomena in microfiuidic processes. This increasing demand results from two aspects. First, essentially all microfiuidic processes in lab-on-achip applications are electrokinetic processes, which are critical to the operation and performance of the lab-chip devices. Secondly, due to the micron scale of the fluidic channels, and the applications of the advanced microfabrication and the micro surface modification technology, the electrokinetic phenomena in microfiuidics become more complicated and have many unique features, such as the synergetic effects of the flow field and the electrical double layer field, and the coupled effects of surface roughness, surface heterogeneity and electrokinetics. The interfacial electrokinetic phenomena in microfiuidics have not been systematically discussed in any published books. Without sufficient knowledge of these phenomena, one cannot systematically design the microfiuidic and labchip devices and cannot control their operations. The objective of this book is to provide a fundamental understanding of the interfacial electrokinetic phenomena in several key microfiuidic processes, and to show how these phenomena can be utilised to control the microfiuidic processes. For this purpose, this book emphasises the theoretical modelling and the numerically simulation of these
vi
Preface
electrokinetic phenomena in micro fluidics. However, experimental studies of the electrokinetic microfluidic processes are also highlighted with sufficient detail. Chapter 1 gives an overview of the correlation between microfluidics and interfacial electrokinetics. Chapter 2 introduces the basic theory of the electrical double layer field that is required to understand the electrokinetic phenomena in the later chapters. Electrokinetic phenomena such as the streaming potential and the electro-viscous effect in pressure-driven flows in microchannels are discussed in Chapter 3. Since most on-chip microfluidic transport processes are driven by applied electrical fields, Chapters 4, 5, 6, 7, 8 and 9 are devoted to electrokinetic transport processes in microchannels under applied electrical fields. Chapter 4 introduces the basics of electroosmotic flows, transient electroosmotic flows, solution displacing processes, and the Joule heating effects in electroosmotic flow. Chapter 5 discusses the effects of surface heterogeneity on the flow structure and solution mixing. Both homogeneous and heterogeneous three-dimensional surface roughness elements in microchannels have a great influence on electrokinetic flow and mass transport, which is covered in Chapter 6. Chapter 7 introduces several experimental techniques for measuring the electroosmotic flow velocity, visualizing the flow field and temperature field in microchannels. Chapter 8 focuses on the on-chip electrokinetic sample dispensing processes. Chapter 9 discusses electrokinetic motion of micron-sized particles in microchannels. Since zeta potential is a key parameter in the studies of electrokinetic phenomena, Chapter 10 reviews two commonly-used techniques for measuring zeta potential. This book is not intended to provide a comprehensive review of all aspects of electrokinetic processes in microfluidics. The purpose is to introduce a fundamental understanding of the interfacial electrokinetic phenomena in microfluidics by studying some key processes. Considering the researchers and the graduate students who are new to this field, this book tries to provide many details of theoretical modelling, numerical simulations, and experimental set-up and procedures. This book reviews many recent research works in my laboratory. I greatly appreciate the contributions of my graduate students in these research works. Here I would like to name a few: David Erickson, David Sinton (now assistant professor at the University of Victoria), Liqing Ren (now assistant professor at the University of Waterloo), Chunzhen Ye, Yandong Hu and Xiangchun Xuan. Together we learn new things and make progress every day. I would also like to thank Professor Charles Chun Yang, Singapore Technological University, for providing me with his papers on transient electroosmotic flow. Finally, I am indebted to my wife, Liping Wang, and my daughter, Daphne, without their love and constant support this book would not have been written. Dongqing Li
Table of Contents Preface
V
Chapter 1 Lab-on-a-chip, microfluidics and interfacial electrokinetics 1 References 5 Chapter 2 Basics of electrical double layer 2-1 2-2
7 8
Introduction to electrical double layer (EDL) Basic electrokinetic phenomena in microfluidics References
28 29
Chapter 3 Pressure-driven flows in microchannels
30
3-1 3-2 3-3 3-4
Chapter 4 4-1 4-2 4-3 4-4 4-5 4-6 4-7 4-8
Pressure-driven electrokinetic flows in slit microchannels Pressure-driven electrokinetic flows in rectangular microchannels Measured electro-viscous effects New understanding of electro-viscous effects References
32 44 63 74 91
Electroosmotic flows in microchannels Electroosmotic flow in a slit microchannel Electroosmotic flow in a cylindrical microchannel Electroosmotic flow in rectangular microchannels Transient electroosmotic flow in cylindrical microchannels AC electroosmotic flows in a rectangular microchannel Electroosmotic flow with one solution displacing another solution Analysis of the displacing process between two electrolyte solutions Joule heating and thermal end effects on electroosmotic flow References
92 94 98 104 119 131 151 167 184 202
Chapter 5 Effects of surface heterogeneity on electrokinetic flow 5-1 5-2 5-3 5-4 5-5 5-6 5-7
Pressure-driven flow in microchannels with stream-wise heterogeneous strips Pressure-driven flow in microchannels with heterogeneous patches Electroosmotic flow in microchannels with continuous variation of zeta potential Electroosmotic flow in microchannels with heterogeneous patches Solution mixing in T-shaped microchannels with heterogeneous patches Heterogeneous surface charge enhanced micro-mixer Analysis of electrokinetic flow in microchannel networks References
204 206 215 238 251 268 288 298 318
viii
Table of Contents
Chapter 6 Effects of surface roughness on electrokinetic flow 6-1 6-2
Electroosmotic transport in a slit microchannel with 3D rough elements Effects of 3D heterogeneous rough elements References
Chapter 7 Experimental studies of electroosmotic flow 7-1 Measurement of the average electroosmotic velocity by a current method 7-2 Measurement of the average electroosmotic velocity by a slope method 7-3 Microfluidic visualization by a laser-induced dye method 7-4 Velocity profiles of electroosmotic flow in microchannels 7-5 Comparison of the current method and the visualization technique 7-6 Flow visualization by a micro-bubble lensing induced photobleaching method 7-7 Joule heating and heat transfer in chips with T-shaped microchannels 7-8 Joule heating effects on electroosmotic flow References
321 323 344 353 354 356 369 376 390 403 411 427 446 460
Chapter 8 Electrokinetic sample dispensing in crossing microchannels 463 8-1 8-2 8-3 8-4 8-5
Analysis of electrokinetic sample dispensing in crossing microchannels Experimental studies of on-chip microfluidic dispensing Dispensing using dynamic loading Effects of spatial gradients of electrical conductivity Controlled on-chip sample injection, pumping and stacking with liquid conductivity differences References
466 484 496 510 526 541
Chapter 9 Electrophoretic motion of particles in microchannels 9-1 Single spherical particle with gravity effects 9-2 Single cylindrical particle without gravity effects 9-3 Spherical particle in a T-shaped microchannel 9-4 Two particles in a rectangular microchannel References
542 546 563 579 599 616
Chapter 10 Microfluidic methods for measuring zeta potential 10-1 Streaming potential method 10-2 Electroosmotic flow method References
617 618 627 640
Subject Index
641
Lab-on-a-chip, Microfluidics and Interfacial Electrokinetics
1
Chapter 1
Lab-on-a-chip, microfluidics and interfacial electrokinetics The microfabrication technology has advanced microelectronics and computer technologies in an amazing speed, making the modern telecommunication and Internet technology possible, and consequently changed the way we work and the way we live profoundly. It allows rapid new technology development and dramatic cost reduction. Scientists and engineers have realized the tremendous advantages of the microfabrication technology and the enormous potential of applying the microfabrication technology to other fields such as mechanical engineering and biomedical engineering. This leads to the recent rapid development of Micro-Electro-Mechanical Systems (MEMS) and Laboratory-on-a-Chip (LOC) devices. A lab-on-a-chip (LOC) is a microscale chemical or biological laboratory built on a thin glass or plastic plate with a net work of microchannels, electrodes, sensors and electronic circuits. The width or the height of a typical microchannel ranges from 20 to 200 um. Applying electrical fields through the electrodes along microchannels controls the liquid flow and other operations in the chip. These labs on a chip can duplicate the specialized functions as their room-sized counterparts, such as clinical diagnoses of bacteria and viruses, and DNA electrophoretic separation. The advantages of these labs on a chip include dramatic reduction in the amount of the samples and reagents, very short reaction and analysis time, high throughput, automation and portability [1-5]. In conventional chemistry and biology laboratories, an experiment is generally carried out as a series of separate operations (such as measuring samples, mixing solutions, and incubating) using separate tools and techniques. Many different instruments are involved from simple devices such as beakers, pipettes, stirring hot plate, centrifuge and incubator to more sophisticated instruments for PCR amplification, electrophoresis and fluorescent microscopy. Generally, the sample preparation prior to measurements is conducted manually and is labour intensive. Due to the relatively large size of the instruments, a large amount of the reagents or samples is required. This results in higher operating cost and longer time for completing the reaction and analysis. Conducting experiments with different samples or reagents require performing the costly and time-consuming separate experiments. These manual and individual experimental procedures may result in more chances for errors.
2
Electrokinetics in Microfluidics
A LOC device generally consists of a number of integrated microfluidic components such as pump, mixer, reactor, dispenser and separator, as illustrated in Figure 1. Therefore, multiple steps of an experiment can be conducted automatically on a single chip. For example, a sample of an unknown, singlestranded DNA solution and a solution containing a known, single-stranded DNA tagged with fluorescent dye are pumped from the reagent loading wells into a mixer by applying electric fields through the related electrodes. The mixed solution will then flow into a reactor where the unknown DNA fragments will react with the dye tagged DNA probe molecules (i.e., hybridization) at a specified temperature. The matched DNA samples will bind with the DNA probe. Following that, the reaction product will be pumped to the dispenser section. Then, by switching on another electrical field, a plug of DNA molecules will be dispensed into a buffer solution and flow into a separation microchannel where they are separated according to the charge to mass ratio by electrophoresis. Finally, when the separated DNA molecules enter the detection section of the microchannel, a laser beam is applied. The dye causes the DNA fragments to give off light when a laser beam is shone on them. The larger the separated fragment, the stronger the fluorescence. The detected light intensities are fed to a computer which sorts through signals from separated fragments to provide a sample analysis. Because of the size of the microchannels, the amount of the liquids involved in such a LOC is of the order of nanoliters, and hence the required amount of the samples and reagents are significantly less than that required in conventional lab experiments. Furthermore, using the microfabrication technology, we can easily make many parallel microchannel systems on a single chip, so that one chip can perform multiple tasks at the same time.
Figure 1.1.
Illustration of microfluidic components in a lab-on-a-chip device.
Lab-on-a-chip, Microfluidics and Interfacial Electrokinetics
3
Currently, improving the technology and reducing the cost in health care is a major driving force for rapid development of LOC technology. The demand to apply LOC technology to genomics and proteomics research, high-throughput screening, drug discovery, point-of-care clinical diagnostic devices has been increasing remarkably over the last decade. There are many examples of the applications of LOC, including micro-total analysis system (u-TAS) [6], microfluidic capillary electrophoretic separation [7,8], electrochromatography [9], PCR amplification [10-14], mixing [15,16], flow cytometry [17], sample injection of proteins for analysis via mass spectrometry [18-20], DNA analysis [21-24], cell manipulation [25], cell separation [26], cell patterning [27,28], fluid handling [29], immunoassay [30-37], enzymatic reactions [38-41], and molecular detection [42]. A recent review of integrated LOC devices can be found elsewhere [43] The most important media in the biomedical analysis and diagnostics are liquids. Common liquids used in LOC devices include whole blood samples, bacterial cell suspensions, protein or antibody solutions and various buffers. Therefore, a key to the functions of the LOC is the quantitatively controlled flow, mass (e.g., sample molecules and particles) transport and heat transfer processes in microchannels. The studies of the transport processes in microchannels are referred to as the microfluidics. Generally, a lab-on-a-chip device must perform the following microfluidic functions: pumping, metering, mixing, flow switching, thermal cycling or incubating, sample dispensing or injection, and separating molecules or particles, etc. Precise manipulation of these microfluidic processes is key to the operation and performance of LOC. Generally, we may classify the transport processes into three categories according to the characteristic dimension, Lc, of the systems: (1) Macroscale systems: Lc > 200 um. (2) Microscale systems: 100 nm < Lc < 200 um. (3) Nanoscale systems: Lc< 100 nm. The characteristics of the transport processes change significantly as the characteristic dimension of the system changes from one category to another category. It should be noted that microchannels have very large surface area to volume ratio. For example, for a microchannel of 100 um in diameter, the surface area-volume ratio is: (2TTRL/7IR2L) = (2/R) = 2x10 (m~ ). Therefore, one can expect significant influence of the liquidchannel wall interface on the microfluidic processes. Because most solid-liquid interfaces have electrostatic charge and consequently an electrical field near the interface, the interfacial electrokinetic phenomena are very important to microfluidic processes. In fact, most of the microfluidic processes on a LOC are electrokinetic processes. For example, electroosmosis is used to generate liquid motion or pump the liquid through microchannels; electrophoresis is used to separate molecules and particles in microchannels. Therefore, interfacial electrokinetic phenomena dominate these microscale transport processes.
4
Electrokinetics in Microfluidics
Because of the complex of the electrokinetic phenomena, the characteristics and controlling parameters of the microfluidic processes vary from system to system and from application to application. Conventional theories of the transport phenomena for macroscopic systems are generally not applicable in microfluidics. At the present, the lack of understanding of the complicated electrokinetic transport phenomena in microchannels makes it difficult to do systematic design and precise operation control of the labs on a chip. It is true that the microfabrication capability is needed to make a LOC device, one must realise that the fundamental understanding of the microfluidic transport processes is essential for the design and for the operation control of LOC devices. Therefore, this book is devoted to provide basic understanding of electrokinetic phenomena in some key microfluidic processes involved in LOC devices.
Lab-on-a-chip, Microfluidics and Interfacial Electrokinetics
5
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33]
J. Knight, Nature, 418 (2002) 474-5. M. Freemantle, Chemical & Engineering News, 77 (1999) 27-36. I. Wickelgren, Popular Science, Nov. 1998, 57-61. G. Sinha, Popular Science, Aug. 1999, 48-52. S. Borman, Chemical & Engineering News, Feb. 1999, 30-31. S.C. Jakeway, A.J. de Mello and E.L. Russell, Fresenius J. Anal. Chem., 366 (2000) 525-39. D.J. Harrison, K. Fluri, K. Seiler, K. Seiler, Z. Fan, C. Effenhauser and A. Manz, Science, 261 (1993) 895-7. S.B. Cheng, C. D. Skinner, J. Taylor, S. Attiya, W. E. Lee, G. Picelli, and D. J. Harrison, Anal. Chem., 73 (2001) 1472-1479. R.D. Oelschuk, L.L. Shultz-Lockyear, Y. Ning and D. J. Harrison, Anal. Chem., 72 (2000) 585-90. M.U. Kopp, A.J. de Mello and A. Manz, Science, 280 (1998) 1046-8. D. Erickson and D. Li, Int. J. Heat Mass Transfer, 45 (2002) 3759-3770. P. Belgrader, M. Okuzumi, F. Pourahmadi, D.A. Borkholder and M. Northrup, Biosensors & Bioelectronics, 14 (2000) 849-852. J. Khandurina, et al., Anal. Chem., 72 (2000) 2995-3000. E.T. Lagally, I. Medintz and R.A. Mathies, Anal. Chem., 73 (2001) 565-570. A.D. Stroock, S.K.W. Dertinger, A. Ajdari, I. Mezic, H.A. Stone and G.M. Whitesides, Science, 295 (2002) 647-651. D. Erickson and D. Li, Langmuir, 18 (2002) 1883-1892. L.L. Sohn, et al., Proceedings of the National Academy of Sciences of the United States of America, 97 (2000) 10687-10690. D. Figeys, S.P. Gygi, G. McKinnon and R. Aebersold, Anal. Chem., 70 (1998) 37283734. Y. Jiang, P.C. Wang, L.E. Locascio and C.S. Lee, Anal. Chem., 73 (2001) 2048-2053. J. Gao, J.D. Xu, L.E. Locascio and C.S. Lee, Anal. Chem., 73 (2001) 2648-2655. B.A. Buchholz, et al., Anal. Chem., 73 (2001) 157-164. Z.H. Fan, et al., Analy. Chem, 71 (1999) 4851-4859. L. Koutny, et al. Anal. Chem, 72 (2000) 3388-3391. G.B. Lee, S.H. Chen, G.R. Huang, W.C. Sung and Y.H. Lin, Sensors and Actuators BChemical, 75 (2001) 142-148. I.K. Glasgow, et al, IEEE Transactions On Biomedical Engineering, 48 (2001) 570-578. J. Yang, Y. Huang, X.B. Wang, F.F. Becker and P.R. Gascoyne, Anal. Chem, 71 (1999) 911-918. D.T. Chiu, et al. Proceedings of the National Academy of Sciences of the United States of America, 97 (2000) 2408-2413. A. Folch, B.H. Jo, O. Hurtado, D.J. Beebe and M. Toner, J. Biomedical Materials Research, 52 (2000) 346-353. D.D. Cunningham, Anal. Chim. Acta, 429 (2001) 1-18. A. Hatch, et al. Nature Biotechnology, 19 (2001) 461-465. E. Eteshola and D. Leckband, Sensors and Actuators B-Chemical, 72 (2001) 129-133. S.B. Cheng, et al. Anal. Chem, 73 (2001) 1472-1479. T.L. Yang, S.Y. Jung, H.B. Mao and P.S. Cremer, Anal. Chem, 73 (2001) 165-169.
6
Electrokinetics in Microfluidics
[34] D.L. Stokes, G.D. Griffin and T. Vo-Dinh, Fresenius J. Anal. Chem., 369 (2001) 295301. [35] K. Sato, M. Tokeshi, T. Odake, H. Kimura, T. Ooi, M. Nakao and T. Kitamori, Anal. Chem., 72 (2000) 1144-7. [36] A. Dodge, K. Fluri, E. Verpoorte and N.F. de Rooij, Anal. Chem., 73 (2001) 3400-9. [37] M. Sharma, A. Saxena, S. Ghosh, J.C. Samantaray and G.P. Talwar, Indian J. Med. Res., 88(1998)409-15. [38] A.G. Hadd, D.E. Raymond, J.W. Halliwell, S.C. Jacobson and J.M. Ramsey, Anal. Chem., 69(1997)3407-3412. [39] D.C. Duffy, H. L. Gillis, J. Lin, N.F. Sheppard and G J . Kellogg, Anal. Chem., 71 (1999) 4669-4678. [40] A.G. Hadd, S.C. Jacobson and J.M. Ramsey, Anal. Chem, 71 (1999) 5206-5212. [41] S.C. Jacobson and J.M. Ramsey, Anal. Chem, 68 (1996) 720-3. [42] B.H. Weigl and P. Yager, Science, 283 (1999) 346-7. [43] D. Erickson and D. Li, Anal. Chem. Acta, 507 (2004) 11-26.
Basics of Electrical Double Layer
7
Chapter 2
Basics of electrical double layer There are many excellent books dealing with electrokinetics and interfaces associated with colloidal particles, such as the books by Hunter [1] and Lyklema [2]. This chapter does not intend to duplicate these references, instead, the objective of this chapter is to provide the basic understanding of the electrical double layer field required in the later chapters dealing with different microfluidic processes. More in-depth discussions of various electrokinetic phenomena will be provided in these chapters whenever appropriate.
8
2-1
Electrokinetics in Microfluidics
INTRODUCTION TO ELECTRICAL DOUBLE LAYER (EDL)
2-1.1 Electrical field in a dielectric medium In this chapter, we will discuss the electrical filed near solid-liquid and liquid-fluid interfaces. Since we will deal with dielectric media in most applications, it is worthwhile to mention some features of dielectric materials. Dielectric materials include plastics, organic liquids, water (including aqueous electrolyte solutions) and gases. Generally, molecules of many dielectric materials are permanently polarized due to their asymmetrical molecular structure. A simple example is HC1. For dielectric materials containing symmetrical molecules or atoms they can also become polarized when exposed to an electrical field, resulting from the relative displacement of orbital electrons, such as the case of He. Therefore, in the presence of an electrical field, molecules of all dielectric materials have a dipole that comprises two equal and opposite charges separated by a distance. The Poisson equation describes the electrical potential in a dielectric medium.
(i) where y/ is the electrical field potential, p is the free charge density and s and s0 are the dielectric constants in the medium and in the vacuum, respectively. The assumption for this form of the Poisson equation is that the dielectric constant s is a constant, independent of position. The Poisson equation is one of the fundamental equations used to evaluate the electrical potential in an electrolyte solution such as NaCl in water. In situations where there is no free charge, for example, in pure organic liquids such as pure oils, p « 0, or in an electrically neutral aqueous solution (i.e., net charge density is zero), the Poisson equation becomes the well-known Laplace equation: (2)
Examples of typical values of dielectric constant are given in the Table 1. 2-1.2 Origin of surface charge Most materials obtain surface electric charge when they are brought into contact with an aqueous solution. The origin of the surface charge includes: (1) Different affinities for ions of different signs to two phases. This may include the following situations: (a) The distribution of anions and cations
9
Basics of Electrical Double Layer
between two immiscible phases such as oil and water, (b) The preferential adsorption of certain type ions from an electrolyte solution on to a solid surface. For example, surfactant ions specifically adsorbed on a surface result in a positively charged surface if the surfactant is a cationic surfactant, and a negatively charged surface if the surfactant is an anionic surfactant, (c) The differential solution of one type ion over the other from a crystal lattice. For example, when a silver iodide crystal (Agl) is placed in water, dissolution takes place until the product of ionic concentration equals the solubility product [Ag+][I~] = 10~16 (mol/L)2. If equal amounts of Ag+ and F ions were to dissolve, then [Ag+] = [F] =10~8 and the surface would be uncharged. However, silver ions dissolve preferentially, leaving a negatively charged surface. (2) Ionization of surface groups. If a surface has acidic groups (e.g., COOH on the surface), the acidic groups dissociation (e.g., COO" on the surface and H+ in the aqueous solution) will result in a negatively charged surface. If a surface contains basic groups (e.g., OH on the surface), the dissociation of the basic groups (e.g., release OFF in the aqueous solution, and leave a positive charge on the surface) will generate a positively charged surface. In both cases, the magnitude of the surface charge depends on the acidic or basic strength of the surface groups and on the pH of the solution. Decreasing pH will reduce the surface charge for a surface containing acidic groups. Increasing pH will reduce the surface charge for a surface containing basic
Table 1 Examples of dielectric constants Liquid or Solid
Dielectric constant
Water
H2O
E = 80.1 at T = 20°C
Water
H2O
8 = 78.5 at T = 25°C
Ethylene glycol
C2H4(OH)2
£ = 40.7 at T = 25°C
Methanol
CH3OH
E = 32.6 at T = 25°C
Dodecane
Cj2H26
s = 2.0 a t T = 25°C
Sodium chloride
NaCl
s = 6.0 atT = 25°C
Quartz
SiO2
£ = 4.5
atT = 25°C
An empirical relationship between water's dielectric constant and temperature is 0.391T + 0.000741T2. For gases at atmospheric pressure, e « 1 and s increases with pressure. For example, = 1.0006 at 1 atm, and = 1.05 at 100 atm.
10
Electrokinetics in Microfluidics
groups. The surface charge can thus be reduced to zero. Most metal oxides can have either positive or negative surface charge depending on the bulk pH. (3) Charged crystal surfaces. When a crystal is broken, different surfaces may have different properties including different surface charges. 2-1.3 Electrical double layer (EDL) When in contact with an aqueous solution, most solid surfaces carry electrostatic charges or an electrical surface potential. Generally, the solution is electrically neutral (having an equal number of positively charged ions and negatively charged ions). However, the electrostatic charges on the solid surface will attract the counterions in the liquid (i.e., an electrolyte solution or a liquid with impurities). Because of the electrostatic attraction, the counterion concentration near the solid surface is higher than that in the bulk liquid far away from the solid surface. The coion concentration near the surface, however, is lower than that in the bulk liquid far away from the solid surface, due to the electrical repulsion. Therefore there is a net charge (excess counterions) in the region close to the surface. This net charge should balance the charge at the solid surface. The charged surface and the layer of liquid containing the balancing charges is called the electrical double layer (EDL) [1,2], as illustrated in Figure 2.1. Immediately next to the charged solid surface, there is a layer of ions that are strongly attracted to the solid surface and are immobile. This layer is called the compact layer, normally about several Angstroms thick. The charge and potential distributions in the compact layer are mainly determined by the geometrical restrictions of ion and molecule size and the short-range interactions between ions, the wall and the adjoining dipoles. From the compact layer to the electrically neutral bulk liquid, the net charge density gradually reduces to zero. Ions in this region are affected less strongly by the electrostatic interaction and are mobile. This region is called the diffuse layer of the EDL. As will be discussed below, an equation called the Poisson-Boltzmann equation is used to describe the ion and potential distributions in the diffuse layer. The thickness of the diffuse layer is dependent on the bulk ionic concentration and electrical properties of the liquid, usually ranging from several nanometers for high ionic concentration solutions up one or two microns for pure water and pure organic liquids (with extremely low ionic concentration). The boundary between the compact layer and the diffuse layer is usually referred to as the shear plane. In conventional fluid mechanics, the liquid velocity at the shear plane is usually set to be zero, and used as a boundary condition. From the above discussion of the compact layer, one can easily understand why the velocity at the shear plane is zero. Generally, The electrical potential at the solid-liquid interface is difficult to
Basics of Electrical Double Layer
11
Figure 2.1a. Illustration of the ionic concentration field in an electrical double layer for a flat surface in contact with an aqueous solution.
Figure 2.1b Illustration of an electrical double layer potential field for a flat surface in contact with an aqueous solution, K = [ 2 z2 e2 nx I er so fa T ]m.
12
Electrokinetics in Microfluidics
measure directly. The electrical potential at the shear plane, however, can be measured experimentally [1,2], this potential is called the zeta potential, g, and is considered as an approximation of the surface potential in most electrokinetic models. 2-1-4 Boltzmann distribution In thermodynamics, Boltzmann's distribution refers to Boltzmann's probability law. It states that the probability of an isolated system taking a thermodynamic equilibrium state with an energy W is proportional to exp(-W/kbT), where T is the absolute temperature (K) and kb is the Boltzmann constant (J/K). Applying the Boltzmann probability to the ionic number concentration (number per unit volume) distribution near a charged surface in a symmetric (i.e., valence ratio z:z) electrolyte solution, we have
Here n+ and n are the ionic number concentrations for the cations and anions at a given location in the liquid, y/ is the electrical field potential (V) at a given position in the liquid, z is the absolute value of the ionic valence, e is the fundamental charge of an electron, « J = n^ = n^, nx is the ionic number concentration in the bulk solution (infinite away from the charged surface). In fact, the Boltzmann distribution can be derived more rigorously from equilibrium thermodynamics. Consider a charged surface and the surrounding electrolyte solution in an equilibrium state. At equilibrium the electrochemical potential of the ions must be constant everywhere, i.e.,
In other words, the electrical force and the diffusion force on the ion must balance each other.
For a one-dimensional flat surface system, the above equation can be rewritten as:
Basics of Electrical Double Layer
13
As the chemical potential (per ion) is given by:
where nx is the number of ions of type i (i.e., n+ or n ) per unit volume, we have
If we integrate this equation from a point in the bulk solution to a point in the EDL, i.e., using the following boundary conditions: In the bulk solution:
where nf is the bulk ionic number density per unit volume of type / ion. This means that at positions infinitely far away from the charged solid surface the bulk solution is electrically neutral or has zero net charge. In the EDL region:
we will obtain the Boltzmann equation:
The above equations give the so-called Boltzman distribution of the ions near a charged surface. In the above derivation, the implied assumptions are: (l)The system is in equilibrium, ions have no macroscopic motion (i.e., no convection and diffusion). (2)The system is subject only to a homogeneous surface electrical double layer field as characterized by the potential i//. For a microscopically heterogeneous surface, the electrical double layer field varies from point to point, Eq. (9) may not be valid. See Chapter 5 for more general treatments.
14
Electrokinetics in Microfluidics
(3) The charged surface is in contact with an infinitely large liquid medium. At positions infinitely away from the charged surface, the potential y/ is zero and the ionic concentration is nx (i.e., the bulk solution boundary condition). Although the majority of the analyses of electrokinetic processes is based on the Boltzman distribution, it should be aware that the conditions listed above may not be satisfied in many situations. For example, in the cases of liquid flow, the system is not in equilibrium. The equilibrium condition, Eq. (4) is not valid in principle. Therefore the applicability of the Boltzman distribution, Eq. (9), is questionable. However, the ion distribution near the solid-liquid interface will not significantly deviate from the Boltzman distribution unless high speed flows are involved (e.g., Re > 10), as will be discussed later (Chapter 5). In the cases of very low speed flows, we can still use the Boltzmann distribution as a good approximation, as proved below. Consider a differential volume element in a flowing electrolyte solution. At a steady state without chemical reaction, the conservation of ion species requires the divergence of the mass fluxes to be zero.
where the jj is the flux of the ith species. Using the Nernst-Planck equation, we have
where the right hand side terms are the flux due to bulk convection, the flux due to the concentration difference (the diffusion process) and the flux due to the migration in an electrical field. Combining Eq. (10) and Eq. (11) yields:
If the flow of the solution is very slow (as it is in most microfluidic applications), the first term can be neglected. If we consider Dt is a constant, Eq. (12) reduces to the following:
Basics of Electrical Double Layer
15
Under the same boundary conditions, the Boltzmann distribution can be easily derived from Eq. (13). Therefore, if the Boltzmann distribution is valid in lowspeed flow situations, we can use the Poisson-Boltzmann equation (see the next section) to describe the electrical double layer field in these situations. This implies that the electrical double layer field is independent of the flow field. For systems under an externally applied electrical field, one may question the validity of the Boltzmann distribution as well. Consider a typical zeta potential value of lOOmV. If the EDL thickness is of the order of lOnm, the EDL field strength can be estimated as 100 mV/lOnm or 1 x 107 V/m. If the applied electrical field strength is not extremely high, for example, less than lx 105 V/m (or 1 x 103 V/cm), its influence on the EDL field and hence on the ionic distribution is negligible in comparison with the EDL field strength. We can therefore use the Boltzmann distribution and hence the Poisson-Boltzmann equation to describe the EDL field in the studies of flow systems with an applied electrical field such as electroosmotic flow. As another caution, if a liquid is confined in a submicron-sized capillary, the electrical double layer fields of the opposite solid surfaces may overlap. The potential and the net charge density in the middle plane are not zero. Obviously the boundary condition in the bulk solution will be different from what was described above. Since the boundary condition is different, the solution to the differential equation, Eq. (8) will be different, in turn, Eq. (9) is no longer valid in this case [3].
2-1.5 Theoretical model and analysis of EDL According to the theory of electrostatics, the relationship between the electrical potential y/ and the local net charge density per unit volume pe at any point in the solution is described by the Poisson equation: (i) where s is the dielectric constant of the solution. Assuming the equilibrium Boltzmann distribution equation is applicable, the number concentration of the type-i ion in a symmetric electrolyte solution is of the form (9)
16
Electrokinetics in Microfluidics
where niao and z7- are the bulk ionic concentration and the valence of type-i ions, respectively, e is the charge of a proton, Kb is the Boltzmann constant, and T is the absolute temperature. The net volume charge density pe is proportional to the concentration difference between symmetric cations and anions, via
For arbitrary electrolyte solutions that may contain asymmetric ions, the net volume charge density pe is
Obviously, when z, = z = constant, i.e., symmetrical ions, Eq. (15) reduces to Eq. (14). Substituting Eq. (14) into the Poisson equation, Eq. (1), leads to the wellknown Poisson-Boltzmann (P-B) equation. (16) Apparently, the combination of Eq. (1) and Eq. (15) will give a different (more general) form of the P-B equation.
By defining the Debye-Huckel parameter dimensional electrical potential
and the non-
the Poisson-Boltzmann equation, Eq.
(16), can be re-written as:
Generally, solving this equation with appropriate boundary conditions, the electrical potential distribution y/ of the EDL can be obtained, and the local charge density distribution pe can then be determined from Eq. (14).
Basics of Electrical Double Layer
It
should
be
17
noted that the Debye-Huckel parameter is independent of the solid surface properties and is determined by the liquid properties (such as the electrolyte's valence and the bulk ionic concentration) only. \lk is normally referred to as the characteristic thickness of EDL and is a function of the electrolyte concentration. Values of \lk range, for example, from 9.6 nm at 10~3 M to 304.0 nm at 10"6 M for a KC1 solution. When the ionic concentration is 10~6 M, the solution is considered practically the pure water. The thickness of the diffuse layer usually is about three to five times of Ilk, and hence may be larger than one micron for pure water and pure organic liquids. Let's see an example of how to calculate 1/k. Consider pure water at T = 298K and use the following parameters: a = 78.5, so = 8.85 x 10~12 C2/Nm2, e = 1.602 x 10~19 C, kb = 1.381 x 10~23 J/K, and Na = 6.022 x 1023 /mol. Note that ttoo is the bulk ionic number concentration and is expressed in terms of the molarity M (mole/liter) by:
Put all the above values into
and we have
here Mis the molarity of a symmetrical (z:z) electrolyte. For example, z =1, we have
18
Electrokinetics in Microfluidics
As seen from the above table, when the bulk ionic concentration increases, more counterions are attracted to the region close to the charged solid surface to neutralize the surface charge. Consequently, the double layer thickness is reduced, and it seems that the EDL is "compressed". 2-1.6 EDL field near a flat surface Consider two flat surfaces separated by a distance 2a in an aqueous solution, as illustrated in the Figure 2.2 below. The EDL filed is onedimensional. If we set the origin of the X-axis in the middle plane between the two plats, the P-B-equation can be re-written as: (19)
If the distance a is much larger than the EDL thickness, the boundary conditions are: (i.e., infinitely away from the surface) (i.e., at the surface or shear plane)
Figure 2.2.
Illustration of a system that consists of two identical flat surfaces.
Basics of Electrical Double Layer
19
If the electrical potential is small compared to the thermal energy of the ions, i.e., (zey/\< A^r|) so that the right-hand side term in Eq. (19) can be approximated by the first terms in a Taylor series. This transforms Eq. (19) to (20) In literature, this treatment is called the Debye-Huckle linear approximation. At 25°C, this linear assumption requires that y/ < 25 mV. The solution of the above equation can be easily obtained as: sinh(£X)
(21)
For a flat surface in contact with an infinitely large aqueous solution, if we set the origin of the coordinate system on the wall, as illustrated in Figure 2.3 below, the P-B equation in dimensional form is given by (22)
with the boundary conditions: (i.e., at the surface or shear plane) (i.e., infinitely away from the surface) Eq. (22) has the following solution, if the Debye-Huckle linear approximation is used, (23) It should be noted that distance EDL thickness
(24)
is often refereed to as the electrokinetic distance, a measure of the distance relative to the EDL thickness.
20
Electrokinetics in Microfluidics
Figure 2.3. The electrical double layer field near a flat surface in an infinitely large liquid.
For the flat surface system, it is possible to obtain an exact analytical solution to the P-B equation without using the linear assumption (see Hunter's book[l]): (25)
(26)
In this equation, k is the Debye-Huckle parameter. The boundary conditions are the following: At the surface or the shear p l a n e , ; and at points in the bulk liquid infinitely away from the surface, 0. A comparison of the EDL potential predicted by the exact solution (Eq.(26)) and the linear solution (Eq. (23)) is given in Figure 2.4. Once is known, the ionic concentration distribution and the net charge density distribution p (x) can be determined by using Eq. (9) and Eq.(14), respectively. Figure 2.5 illustrates the ionic concentration distributions predicted by using the exact solution (Eq.(26)) and the linear solution (Eq. (23)).
Basics of Electrical Double Layer
21
Figure 2.4. Comparison of EDL potential fields predicted by the linear solution and the non-linear (exact) solution.
Figure 2.5. Ion distribution in diffuse layer: Accumulation of positive ions and expulsion of negative ions in the EDL region of a negatively charged surface.
22
Electrokinetics in Microfluidics
As seen from the above figure, the ions in the diffuse layer are not all of the same sign. At a given position from the charged surface, the concentration difference between the positive ions and the negative ions determines the local net charge density. The charge on the wall surface is balanced by (1) an accumulation of charges of opposite sign (i.e., counterions) and (2) a deficit of charges of the same sign (co-ions) compared to their concentrations in the bulk liquid. The charge per unit area on the surface is called the surface charge density, denoted by a 0 (C/m2). a0 must be balanced by the charge in the adjacent solution: (27)
From the Poisson equation,
we have
Integrating the P-B equation
yields (translate the result in dimensional form):
the surface charge density is related to the surface potential via the following equation:
Basics of Electrical Double Layer
23
(28)
2-1.7 EDL field around a spherical surface For the ID EDL field around a spherical particle of radius a, the P-B equation can be written as: (29) The boundary conditions are: (At the sphere surface or the shear plane) (Infinitely away from the sphere surface) Using the Debye-Huckle linear approximation, it can be shown that (30) The charge on the particle surface must balance the charge in the EDL so that
Use the Poisson equation
and the linearized P-B equation
we have
24
Electrokinetics in Microfluidics
(31)
(32) Using the above equations the potential on the surface of the sphere can be written as (33)
The first term represents the potential on the surface due to the charge on the spherical particle itself. The second term is the potential due to the atmosphere of charge of opposite sign, i.e., it is the potential due to a spherical shell of charge -Q and radius {a + 1/k). If we can measure the zeta potential, £, experimentally, the electrokinetic charge on the particle is given by
2-1.8 EDL field around a cylinder If we consider cylindrical symmetry (i.e., neglecting the end effects), the P-B equation for a cylinder of radius a in a symmetrical electrolyte solution is given by, in cylindrical coordinates,
Basics of Electrical Double Layer
25
where R = kr. Under the linear assumption, (zet//) < kt,T, the solution is: (36)
(37)
(38)
(39)
2-1.9 Dependence of surface charge and zeta potential on ion concentration and pH The effects of the bulk ionic concentration on the EDL potential and particularly the surface potential (or approximately the zeta potential) can be analyzed qualitatively as follows. When we derived the Boltzmann distribution, we showed the following equilibrium condition:
This equation can also be re-arranged into (40)
26
Electrokinetics in Microfluidics
Figure 2.6. Illustration of the surface charge density dependence on pH and the bulk ionic concentration.
Figure 2.7. Illustration of the zeta potential dependence on pH and the bulk ionic concentration.
Basics of Electrical Double Layer
Consider Kb = 1.381x10 1, we have
23
27
J/K, e = 1.602 xlO" 19 C, T = 298K, and assume z,•.=
This equation implies that a ten-fold increase in the ion concentration, nh will have a relatively small change in the potential. Note that if we choose z, = - 1 , the sign in the above equation will change, indicating that the effect of positive and negative ions on the potential change is opposite. Generally, for a given solid surface, the surface charge and zeta potential are functions of the bulk ionic concentration and valence as well as pH, as shown in Figures 2.6 and 2.7. Clearly, for a given electrolyte solution, the surface charge and zeta potential can be changed from positive to negative by varying the pH value of the solution.
28
2-2
Electrokinetics in Microfluidics
BASIC ELECTROKINETIC PHENOMENA IN MICROFLUIDICS
Electroosmosis Consider a stationary solid surface in contact with an aqueous solution. When an electric field is applied, the excess counterions in the diffuse layer of the EDL will move under the applied electrical force. This is called the electroosmosis. As the ions move, they drag the surrounding liquid molecules to move with them due to the viscous effect, resulting in a bulk liquid motion. Such a liquid motion is called the electroosmotic flow. For example, by applying an electric field along a microchannel, we can electroosmotically "pump" liquids to flow through the microchannel. Electrophoresis Consider a (solid, liquid or gas) particle in a bulk liquid phase. When an electric field is applied to the bulk liquid, because the particle surface has electrostatic charge, the particle can be induced to move (relative to the stationary or moving liquid) under the applied electrical field. Such a particle motion is called the electrophoresis. Streaming potential In absence of an applied electric field, when a liquid is forced to flow through a capillary or microchannel under an applied hydrostatic pressure difference, the counterions in the diffuse layer (mobile part) of the EDL are carried towards the downstream end, resulting in an electrical current in the pressure-driven flow direction. This current is called the streaming current. Corresponding to this streaming current, there is an electrokinetic potential called the streaming potential. This flow induced streaming potential is a potential difference that builds up along a capillary or microchannel. This streaming potential acts to drive the counterions in the diffuse layer of the EDL to move in the direction opposite to the streaming current, i.e., opposite to the pressure-driven flow direction. The action of the streaming potential will generate an electrical current called the conduction current. At a steady state, the streaming current will be balanced by the conduction current, and hence the net current in the capillary/microchannel is zero.
Basics of Electrical Double Layer
29
REFERENCES [1] [2] [3]
R.J. Hunter, "Zeta Potential in Colloid Science: Principles and Applications", Academic Press, London, 1988. J. Lyklema, "Fundamentals of Interface and Colloid Science", Vol. I and II, Academic Press, 1995. W. Qu and D. Li, J. Colloid Interface Sci., 224 (2000) 397-407.
30
Electrokinetics in Microfluidics
Chapter 3
Electro-viscous effects on pressure-driven liquid flow in microchannels Just as rapid advances in microelectronics have revolutionized computers, appliances, communication systems and many other devices, microfluidic technologies will revolutionize many aspects of applied sciences and engineering, such as heat exchangers, pumps, gas absorbers, solvent extractors, fuel processors, on-chip biomedical and biochemical analysis instruments. These lightweight, compact and high-performance micro-systems will have many important applications in transportation, buildings, military, environmental restoration, space exploration, environmental management, biochemical and other industrial chemical processing. Fundamental understanding of liquid flow in microchannels is critical to the design and process control of various microfluidic and lab-on-chip devices used in chemical analysis and biomedical diagnostics (e.g., miniaturized total chemical analysis system). However, many phenomena of liquid flow in microchannels, such as unusually high flow resistance, cannot be explained by the conventional theories of fluid mechanics. These are largely due to the significant influences of interfacial electrokinetic phenomena such as electroviscous effects at the micron scale. Although interfacial electrokinetic phenomena such as electro-osmosis, electrophoresis and electro-viscous effects are well known to colloidal and interfacial sciences for many years, the effects of these phenomena on transport phenomena (such as liquid flow and mixing in fine capillaries) generally are less well understood. Partially this is because in colloidal sciences the electrokinetic phenomena usually are studied for closed systems, while in the studies of micro transport phenomena the systems are open systems and involve complicated boundary conditions. In various microfluidic processes, a desired amount of a liquid is forced to flow through microchannels from one location to another. Depending on the specific structures of the microfluidic devices, the shape of the cross-section of the microchannels varies. Two typical cases are the slit microchannels (formed between two parallel plates) and the trapezoidal (may be approximated as rectangular) microchannels made by microfabrication processes. This chapter
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
31
will show how to model and evaluate the interfacial electrokinetic effects on pressure-driven liquid flow through these microchannels. The concept of streaming potential, streaming current, conduction current, electro-viscous effects will be introduced. Some existing experimental evidences of the electroviscous effects will also be presented.
32
3-1
Electrokinetics in Microfluidics
PRESSURE-DRIVEN ELECTROKINETIC FLOW IN SLIT MICROCHANNELS
When a liquid is forced through a microchannel under an applied hydrostatic pressure, the counterions in the diffuse layer (mobile part) of the EDL are carried towards the downstream end, resulting in an electrical current in the pressure-driven flow direction. This current is called the streaming current. Corresponding to this streaming current, there is an electrokinetic potential called the streaming potential. This flow induced streaming potential is a potential difference that builds up along a microchannel. This streaming potential acts to drive the counterions in the diffuse layer of the EDL to move in the direction opposite to the streaming current, i.e., opposite to the pressuredriven flow direction. The action of the streaming potential will generate an electrical current called the conduction current, as illustrated in Figure 3.1. It is obvious that when ions move in a liquid, they will pull the liquid molecules to move with them. Therefore, the conduction current will produce a liquid flow in the opposite direction to the pressure driven flow. The overall result is a reduced flow rate in the pressure drop direction. Under the same conditions, if the reduced flow rate is compared with the flow rate predicted by the conventional fluid mechanics theory without considering the presence of the EDL, it seems that the liquid would have a higher viscosity. This is usually referred to as the electro-viscous effect [1,2]. In this section, we consider liquid flow through a microchannel with a slitshaped cross-section [3,4], such as a channel formed between two parallel plates, as illustrated in Figure 2. We assume the length, L, and the width, W, of the slit channel are much larger than the height, H = 2a, of the channel, so that both the electrical double layer (EDL) field and the flow field can be considered as one dimensional (i.e., with variation in the channel height direction only).
Figure 3.1.
Illustration of the flow-induced electrokinetic field in a microchannel.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
Figure 3.2. A slit microchannel. a « W, and a « smaller than the channel's width and length.
33
L, i.e., the channel's height is much
3-1.1 The electrical double layer field Essentially all liquid flows in microchannels are low speed flow. As discussed in Chapter 2,we will assume that the flow field has no appreciable effects on the ion concentration field. The ion distribution is given by the Boltzmann equation. The EDL field is therefore independent of the flow field. Consider a liquid containing simple symmetric ions, i.e., the valences of the ions are the same, z = z+ = z~. The bulk ionic concentration is nx. For the slitmicrochannel, we consider only the EDL fields near the top and the bottom channel surfaces. The one-dimensional EDL field for a flat surface is described by the following form of the Poisson-Boltzmann equation:
(i) The local net charge density in the liquid is given by (2)
Non-dimensionalizing Eqs. (1) and (2) via (3)
34
Electrokinetics in Microfluidics
we obtain the non-dimensional form of the Poisson-Boltzmann equation as:
(4) (5) where
a
n
d
i
s
the Debye-Huckel parameter
and 1/k is the characteristic thickness of the EDL.
is the
electrokinetic separation distance or the ratio of the half channel's height to the EDL thickness. Therefore, the parameter K can be understood as the relative channel's height with respect to the EDL thickness. For example, K = 10 means that the half channel height a is 10 times of the EDL thickness 1/k. If the electrical potential is small as compared to the thermal energy of the ions, i.e., so that the sinh function in Eq. (4) can be approximated by This transforms Eq. (4) to
(6) This treatment is usually called the Debye-Huckel linear approximation [1,3,4]. The solution of Eq. (6) can be obtained easily. As illustrated in Figure 3.2, if the separation distance between the two plates is sufficiently larger so that the EDL fields near the two plates are not overlapped, the appropriate boundary conditions for the EDL fields are: At the centre of the slit channel: At the solid surfaces: With these boundary conditions Eq. (6) can be solved and the solution is given by: (?)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
35
Eq.(7) allows us to plot the non-dimensional EDL potential field in the slit channel. Due to the symmetry, we plot only the non-dimensional EDL potential field from one surface to the centre of the channel, as shown in Figure 3.3. In this figure, the zeta potential is assumed to be 50 mV. For a given electrolyte, a large K implies either a large separation distance between the two plates or a small EDL thickness. If the separation distance {2a) is given, increase in the bulk ionic concentration nx will increase the value of Debye-Huckel parameter k (recall k = (2naoz2e21ssokbT)112), the double layer thickness (1/k) is reduced (or the double layer is "compressed"), and the electrokinetic separation distance is increased. It can be seen from this figure that as K increases, the double layer field (the region) exists only in the region close to the channel wall. For example, appreciable EDL potential exists only in a region less than a few percent of the channel cross-section area for K = 80. However, for dilute solutions such as pure water (infinite dilute solution), the value of the electrokinetic separation distance K = a * k is small, and hence the EDL filed (the region) may affect significant portion of the flow channel, as shown in Figure 3.3.
Figure 3.3. Non-dimensional electrical double layer potential distribution near the channel wall (£= -50 mV). Center of the channel: X= 0; channel wall: X= 1.0.
36
Electrokinetics in Microfluidics
It should be pointed out that Figure 3.3 is the results of using the DebyeHuckel linear approximation, Eq.(7), and that the linear approximation is valid for small surface potential situations (i.e., y/<25mV). In the cases of large surface potentials, it has been found that, in comparison with the exact solution of the Poisson-Boltzmann equation, the linear solution predicts slightly lower values of the potential in the region near the wall. After a small distance from the wall, the difference between the linear solution and the exact solution diminishes. Discussions of the difference between these two solutions can be found elsewhere [1,3,4]. 3-1.2 Flow field Consider a one-dimensional fully-developed, steady-state, laminar flow through the slit microchannel as shown in Figure 3.2. The forces acting on an element of the liquid include the pressure force, the viscous force and the electric body force generated by the flow-induced electrokinetic field (i.e., the streaming potential). The equation of motion is the Z-directional momentum Equation.
(8)
where Vz is the flow velocity in the Z direction; is the electrokinetic potential gradient in the Z direction, and is the electrical body force; Pz = is the pressure gradient in the Z direction. Non-dimensionalizing Eq. (8) with parameters defined in Eq.(3) and and replacing by Eq. (4), we obtain
(9) where the two non-dimensional numbers are given by: (10) Integrating Eq. (9) twice and employing the appropriate boundary conditions, we obtain the non-dimensional velocity distribution in the slit microchannel as follows:
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
37
(ID
From the definitions of G/ and G2, one can clearly see that the 1st term in Eq. (12) represents the contribution of the applied pressure gradient, and the 2nd term represents the EDL's contribution to the velocity. If there is no EDL effects
between two parallel plates.
3-1.3 Electrokinetic field — streaming potential As seen from Eq. (12), the velocity distribution can be calculated only if the streaming potential Es is known. As mentioned before, in the absence of an applied electric field, when a liquid is forced through a channel under a hydrostatic pressure difference, the excess counter-ions in the diffuse layer of the EDL are carried by the liquid to flow to the downstream, forming an electrical current. This current due to the transport of charges by the liquid flow is called the streaming current and is given by (13)
Non-dimensionalizing the velocity Vz and the net charge density p, we obtain the non-dimensional streaming current as (14) Substituting p by using Eq. (4), the non-dimensional streaming current becomes
38
Electrokinetics in Microfluidics
Using the boundary conditions:
one can easily see that the first term on right hand side of Eq. (15) becomes zero. Therefore the streaming current reduces to (16) By using Eqs. (6), (7) and (9), we can show that the non-dimensional streaming current is given by (17) where (18) The flow-induced streaming potential will drive the counter-ions in the diffuse layer of the EDL to move back in the opposite direction to the pressuredriven flow, and produce an electrical conduction current. The conduction current in the microchannel is given by (19) where Xb is the liquid electrical conductivity; Es /L is the streaming potential gradient; Ac is the cross-section area of the channel. Nondimensionalizing Eq.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
(19) by using L =LIa, Es =ES /g0, conduction current is given by
and Ac = Ac Ia2,
39
the nondimensional
(20)
When the flow in the microchannel reaches a steady state, there will be no net electrical current in the flow, i.e. Ic + Is = 0. Using Ic from Eq. (20) and Is from Eq. (17), we obtain the streaming potential from the steady state condition, Ic + Is = 0-
(21)
where the non-dimensional factor (73 = VQn^zeLjq^X^ . In the classical theory of electrokinetic flow, the effect of EDL on the liquid flow is not considered. The streaming potential is related to the zeta potential and liquid properties through the following equation [1,2]. (22)
where AP is the pressure drop cross the flow path. Eq. (21) can be rearranged in a form similar to Eq. (22) by translating it in terms of dimensional parameters, i.e., (23)
where the correction factor 0 to the classical equation, Eq. (22), is given by (24)
40
Electrokinetics in Microfluidics
If the EDL effect on liquid flow is not considered, f5\ = 1 (see Eq. (17) and Eq. (18)), and the 2nd term in the denominator of Eq.(24) is zero, hence 0=1, and Eq. (23) becomes the classical equation, Eq.( 22). Generally, the zeta potential, the pressure gradient and the liquid properties (such as the bulk ionic concentration, the dielectric constant and the bulk conductivity) can be measured [1,2]. By the above analyses, we can calculate the streaming potential, Eq.(21), and the velocity field, Eq.(12), in the slit microchannel. To obtain an estimation of the parameters such as Gt, G2, G3 and K, let us, for example, consider fully developed laminar flow of an infinitely diluted (i.e., pure water, ««, = 6.022x1020 m"3) aqueous 1:1 electrolyte (e.g. KC1) solution through a slit microchannel. The separation distance is 25 um and the channel is 1 cm long. At room temperature, the physical and electrical properties of the liquid are s = 80, Xb = 1.2639xlO"7 (1/fim), \i = 0.90xl0"3 (kg/ms). A pressure difference of 4 atm and an arbitrarily chosen reference velocity Vo = lm/s are considered. With these, a set of values G7 = 5.009, G2= 7.95xlO"5, G3 = 1.6xlO8, and K = 40.8 is obtained for a fixed value of C, = 50mV. Figure 3.4 shows the non-dimensional velocity distribution of pure water in such a slit silicon microchannel Clearly, by comparing the case of no EDL field (i.e., Q= 0), we see that the EDL effect reduces the velocity of the liquid appreciably.
Figure 3.4. Comparison of the non-dimensional velocity distribution of pure water in a 25 |im slit microchannel between £= -50 mV and £= 0 cases.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
41
3-1.4 Volume flow rate and the apparent viscosity The volume flow rate through the parallel plates can be obtained by integrating the velocity distribution over the cross sectional area of the microchannel, (25)
Using Eq. (12) the volume flow rate in non-dimensional form can be written as: (26) For steady flow under an applied pressure gradient, the volume flow rate is given by Eq. (26). The 1st term in Eq.(26) is the volume flow rate without the EDL effect. The other two terms clearly reflect the EDL effect on the flow rate and contribute to reduce (the negative sign) the net flow in the pressure drop direction. This reduced flow rate seems to suggest that the liquid have a higher viscosity. If we define an apparent viscosity /ua, the classical Poiseuille volume flow rate (without considering the EDL effect) for flow between two parallel plates separated by a distance "2a" is then given by
(27) Nondimensionalizing Eq. (27), we have the volume flow rate: (28) Equalizing Eq. (26) with Eq. (28), i.e., Q = Qp, we obtain the ratio of the apparent viscosity to the true liquid viscosity: (29)
42
Electrokinetics in Microfluidics
Obviously, if there is no EDL or electrokinetic effect, the 2nd and the 3rd terms in the denominator of Eq.( 29) will be zero, and hence \xa = fi. Because the 2nd and the 3rd terms in the denominator of Eq.(29) make the denominator smaller than the numerator, it follows that \ia > /J. This is usually referred to as the electro-viscous effect. From Eq.(29), such an electro-viscous effect depends, in addition to the pressure gradient or the flow rate via the parameter Gj, on the ionic properties of the liquid via the Debye-Huckel parameter k, the channel's height, a, via the electrokinetic separation distance K = a * k, and the zeta potential g. As explained above, the flow-induced streaming potential drives the counterions to move in the direction opposite to the pressure drop. These moving ions drag the surrounding liquid molecules with them. This generates an opposite flow to the pressure-driven flow and hence a reduced flow rate in the pressuredrop direction. According to the classical Poiseuille flow equation, Eq.(27), this reduced flow rate seems to suggest that the liquid have a higher viscosity. As an example, the ratio of the apparent viscosity to the bulk viscosity, / V A is plotted by using Eq.(29) as a function of non-dimensional electrokinetic separation
3 I
£>
2
1
-
1
1
1
1
4
6
8
10
\
V 0
2
Electrokinetic Separation Distance K
Figure 3.5. Variation of the ratio of apparent viscosity to the bulk viscosity with the electrokinetic separation distance Kat £= -50 mV.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
43
distance K in Figure 3.5. It is observed that for C, = -50mV, the ratio \ij\i is approximately 2.75 when K = 2, and then decreases as K increases, approaching a constant value equal to one for very large values of K. For lower values of zeta potential, the trend is the same except the value of the ratio, [ij\x, is lower. Generally, the higher the zeta potential C,, the higher the ratio nJn.
44
3-2
Electrokinetics in Microfluidics
PRESSURE-DRIVEN ELECTROKINETIC FLOWS IN RECTANGULAR MICROCHANNELS
For most lab-on-a-chip devices, the cross-section of the microchannels made by micromachining technology is close to a rectangular (trapezoidal, more exactly) shape. As the EDL field depends on the geometry of the solid-liquid interfaces, the EDL field will be two-dimensional in a rectangular microchannel. In such a situation, the two-dimensional Poisson-Boltzmann (P-B) equation is required to describe the electrical potential distribution in the rectangular channel; and the corner of the channel may have particular contribution to the EDL field, subsequently to the fluid flow field. [5,6] 3-2.1 EDL field in a rectangular microchannel In order to consider the electro-viscous effects on liquid flow in rectangular microchannels, we must evaluate the body force generated by the flow induced electrokinetic field in the equation of motion. To do so, we must evaluate the distributions of the electrical potential and the net charge density of the EDL. Consider a rectangular microchannel of width 2W, height 2H, and length L as illustrated in Figure 6.
Figure 3.6.
A rectangular microchannel (height 2H, width 2W).
In this case, the two-dimensional Poisson equation provides the relationship between the electrical potential y/ and the net charge density per unit volume pe at any point in the solution,
(30)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
45
where e is the dielectric constant of the solution. Assuming the equilibrium Boltzmann distribution equation is applicable, the number concentration of the type-i ion in a symmetric electrolyte solution is of the form (31) where «7CO and z7- are the bulk concentration and the valence of type-i ions, respectively, e is the charge of a proton, Kb is the Boltzmann constant, and T is the absolute temperature. The net volume charge density pe is proportional to the concentration difference between symmetric cations and anions, via. (32)
Substituting Eq.(32) into the Poisson equation, Eq. (30), leads to the well-known Poisson-Boltzmann equation in a two-dimensional form. (33)
By defining the Debye-Huckel parameter
and the hydraulic
diameter of the rectangular microchannel Df, = /TTLL and introducing the dimensionless groups: Y = -£-, Z = -^-, K = KD/,, and vF = :f^-, the above equation can be non-dimensionlized as (34)
Due to symmetry, Eq. (34) is subjected to the following boundary conditions in a quarter of the rectangular cross section (35a)
46
Electrokinetics in Microfluidics
(35b)
where g, defined byg = f^,
is the non-dimensional zeta potential at the channel
wall (here g is the zeta potential). For small values of y/ (Debye-Huckel approximation, which physically means that the electrical potential is small in comparison with the thermal energy of ions, i.e. zey/ | < K^T), the Poisson-Boltzmann equation can be linearized as
(36) By using the separation of variable method, the solution to the linearized P-B equation, Eq. (36), can be obtained. Therefore, the electrical potential distribution in the rectangular microchannel is of the form
(37)
For large values of \\i, the linear approximation is no longer valid. The EDL field has to be determined by solving Eq. (33) or (34). In order to solve this non-linear, two-dimensional, differential equation, a numerical finite-difference scheme may be introduced to derive this differential equation into the discrete, algebraic equations by integrating the governing differential equation over a control volume surrounding a typical grid point. The non-linear source term of Eq.(34) is linearized as (38)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
47
where the subscript (n+1) and n represent the (n+l)th and the nth iterative value, respectively. The derived discrete, algebraic equations can be solved by using the Gauss-Seidel iterative procedure. The solution of the linearized P-B equation with the same boundary conditions may be chosen as the first guess value for the iterative calculation. The under-relaxation technique can be employed to make this iterative process converge fast. The details of how to obtain the numerical solutions of Eq. (34) can be found elsewhere [6,7]. After the electrical potential distribution inside the rectangular microchannel is computed, the local net charge density can be obtained from Eq.(32) as (39) Figure 3.7 shows a comparison of the EDL field in a rectangular microchannel predicted by the linear solution and by the complete numerical solution. In the calculation, the liquid is a dilute aqueous 1:1 electrolyte solution (concentration is lxlO"6 M) at 18°C. The rectangular microchannel has a crosssection 30 x 20 urn with a zeta potential of-75 mV. Because of the symmetry, the EDL field is plotted only in one quarter of the microchannel. With such a relatively high zeta potential, the linear approximation is obviously not good. It can be seen that there is a very steep decrease in the potential in the case of the complete solution, while the linear solution predicts a more gradual decay of the potential. The linear solution predicts a thicker layer from the wall that has an appreciable non-zero EDL field. 3-2.2 Flow field in a rectangular microchannel Consider the case of a forced, two-dimensional, laminar flow through a rectangular microchannel as illustrated in Figure 6. The equation of motion for an incompressible liquid is given by (40)
In this equation, pt and Hf are the density and viscosity of the liquid, respectively. For a steady-state, fully-developed flow, the components of velocity V satisfy u = u(y,z) and v = w = 0 in terms of Cartesian coordinates. dV
- -
-
Thus both the time term -r- and the inertia term (F-V)F vanish. Also, the
48
Electrokinetics in Microfluidics
hydraulic pressure P is a function of xonly and the pressure gradient ~
is
constant. If the gravity effect is negligible, the body force F is only caused by the action of an induced electrical field Ex (see the explanation of the electrokinetic potential) on the net charge density p e (Y, Z) in the electrical double layer region, i.e. Fx = Expe. With these considerations, the Eq. (40) is reduced to
(41)
Defining the reference Reynolds number
PfDhU
Re 0 =—
and non-
dimensionalizing the Eq. (41) via the following dimensionless parameters (42a)
(where U is a reference velocity, Po is a reference pressure, and go is a reference electrical potential), one can obtain the non-dimensionlized equation of motion
<43)
Substituting p e (Y, Z)by Eq. (39) and defining a new dimensionless number G\ =
^-^-, the equation of motion may, therefore, be written as PfU (44)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
49
Figure 3.7. The absolute value of the non-dimensional EDL potential distribution in one quarter of a rectangular microchannel, with q = -75 mV: (a) linear solution; and (b) complete numerical solution.
50
Electrokinetics in Microfluidics
The boundary conditions that apply for the velocity u are (45a) (45b) Here, Eq. (45a) is the symmetric condition and Eq. (45b) is the non-slip condition at the walls of the microchannel. By using the Green's function formulation, the solution of Eq. (44) subjecting to the above boundary conditions is
(46)
Here G(Y,Z,t Y , Z ,r)is the Green's function that may be found by using the separation of variables method [8]. The expression for G(Y, Z, t Y , Z ,r)is
(47)
Substituting Eq. (47) into Eq. (46) and rearranging it, one can obtain the nondimensional fluid velocity profile in the microchannel as follows:
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
51
If there is no electrostatic interaction, the second term on the right hand side of the above equation vanishes. The fluid velocity reduces to
which is the well-known Poiseuille flow velocity profile through the rectangular channel. Using Eqs. (48) and (49), the mean velocity with and without the consideration of the effects of the EDL may be written, respectively, as
52
Electrokinetics in Microfluidics
Thus the non-dimensional volumetric flow rate through the rectangular microchannel, defined by Qy = ^~L , is given by Qv = Uave
(52)
Correspondingly, in absence of the electrical double layer, the nondimensional volumetric flow rate is expressed as Qov = Uoave
(53)
In order to calculate the fluid velocity distribution, the analytical solution Eq. (48) for the velocity is used to obtain the "exact" solution" which, in practice, usually means an error of 0.01 % or less. As seen from Eq. (48), the velocity distribution is expressed by two infinite series. Therefore, usually a very large number of terms in series are needed to achieve this error criteria. To reduce computation time, the Aitken's procedure [8] may be employed for accelerating series. 3-2.3 Electrokinetic field in a rectangular microchannel As seen from Eq. (48), the local and the mean velocity can be calculated only when the non-dimensional induced electrokinetic field strength or the electrokinetic potential, Ex, is known. Similar to the previous slit microchannel case, when a liquid is forced through a microchannel under a hydrostatic pressure difference, the ions in the mobile part of the EDL are carried towards the downstream end. This causes an electrical current, called the streaming current, to flow in the direction of the liquid flow. Corresponding to this streaming current, there is an electric field with an electrokinetic potential called the streaming potential. This field generates a current, called the conduction current, to flow in the opposite direction. When the conduction current is equal to the streaming current, a steady state is reached. Usually the net electrical current, / , flowing in the axial direction of the microchannel, is the algebraic summation of the electrical convection current (i.e. steaming current)/5and the electrical conduction current Ic. In a steady-state situation, this net electrical current should be zero. I = Is+Ic=0
(53)
Due to symmetry of the rectangular microchannel, the electrical streaming current can be calculated by:
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
53
(54) The electrical conduction current in the microchannel generally consists of two parts: one is due to the conductance of the bulk liquid; the other is due to the surface conductance of EDL. Generally, the surface conduction is the excess conduction tangential to a charged surface, and originates from the excess counterions1 concentrations in the EDL region near the solid-liquid interface [1,2,9]. Surface conduction takes place mostly through the compact layer of the EDL. Particularly in the cases of low bulk ionic concentrations (below 10~4 M) and small microchannels, the surface conduction may have a significant contribution to the total conduction current. Usually, the surface conductivity, Xs, is considered as the conductivity of a sheet of material of negligible thickness, with a unit ohnT'nT1. Specific surface conductivity values are of the order 1(T9 ~ 10~8 for water in glass capillaries. (55)
where Ibc and Isc are the bulk and surface electrical conduction currents, respectively; Xt is the total electrical conductivity, and can be calculated by XP Xt = Xfr + - ^ [1,2,9]. Here Ps and Ac are the wetting parameter and the crosssection area of the channel, respectively. Xb is the bulk conductivity of the solution and Xs is the surface conductivity that can be determined experimentally [9]. Substituting Eq. (39) for p e (Y,Z)into Eq. (54) and employing Eq.(53) and Eq. (55), the non-dimensional induced field strength can be expressed as
(56)
Here the nondimensional number The substitution of w(Y,Z) from Eq. (48) into Eq. (56) finally gives the non-dimensional flow-induced electrokinetic field strength as
54
Electrokinetics in Microfluidics
Consider a fully developed, laminar flow of a diluted aqueous 1:1 electrolyte (e.g., KC1) solution through a rectangular microchannel with a height of 20 urn, width of 30 urn and length of lcm. At a typical room temperature T = 298 K, the physical and electrical properties of the liquid are e = 80, nx = 6.023xlO 19 (m~ 3 ), nf = 0.90x10~ 3 (kg/ms), and
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
55
concentration of the aqueous solution. This may be understood as follows: If the ionic concentration is higher, which implies a larger Debye-Huckel length, i.e., a smaller EDL thickness, the effect of the EDL is less. Therefore fewer ions are carried to the downstream with the flow and hence a weaker streaming current and streaming potential are produced. It should be noted that the present model modified the simple proportionality relationship between the zeta potential and the streaming potential in the classical electrokinetic theory (as shown in Eq.(22)) by considering the EDL effects on the liquid flow. 3-2.4 Electrokinetic effects on the velocity field As seen from Eq.(48), the velocity field in a rectangular microchannel depends on the EDL field. Since the bulk ionic concentration and the shape of the channel's cross-section will affect the EDL field, these factors will in turn influence the velocity field. In this section, we wish to examine these effects.
Figure 3.8. Variation of non-dimensional streaming potential with non-dimensional pressure difference for different zeta potential.
56
Electrokinetics in Microfluidics
First, let us consider a fully developed, laminar flow of a diluted aqueous 1:1 electrolyte (e.g. KC1) solution through a rectangular microchannel with a height of 20^m, width of 30um and length of lcm. At a typical room temperature 7 = 298K, the physical and electrical properties of the liquid are £ = 80, ««,= 6.023xlO 19 (m" 3 ), /uf = 0.90x 10"3(kg/ms), and £ = - 7 5 m V . The total electrical conductivity Xt value was taken from the experimental result [10]. The computation of non-dimensional velocity distribution according to Eq.(48) is carried out for a fixed external pressure difference. In Figures 3.9 (a) and (b), the distribution of non-dimensional velocity is plotted for the channel of 20 x 30um with and without consideration of the EDL effects. As seen in Figure 3.9(a), the EDL field exhibits significant effects on the flow pattern. The maximum velocity in the center of the channel is lower when the EDL field effects are considered. The flow velocity near the channel wall approaches zero due to the action of the EDL field and the streaming potential. Moreover, the flow around the channel corner greatly deviates from the classical Poisseuille flow pattern as shown in Figure 3.9(b). It should point out that the surface conduction current plays an important role in the total electrical conduction current for a dilute solution flow in microchannels. It makes a significant contribution to the streaming potential and therefore to the flow field. This is clearly demonstrated in Figure 3.9(c), which is the non-dimensional velocity distribution with the same hydrodynamic and electrokinetic conditions but without consideration of the surface conduction current. One may overestimate the electrokinetic effects on the microchannel flow if the surface conduction current is not included. Figures 3.10(a)-(d) show the distribution of non-dimensional velocity as a function of the geometric ratio of height to width with the same hydraulic diameter as a channel of 20 x 30(xm. For a fixed hydraulic diameter, a small geometric ratio represents a smaller channel height but a larger channel width. As shown in Figure 3.10, it is obvious that the channel shape has a significant influence on the flow across the microchannel because of the EDL effects. The general pattern is the smaller the channel size, such as in Figure 3.10(a), the stronger the EDL effects and the larger portion of the flow field is affected. The explanation for this is that as the channel size decreases, the EDL thickness becomes relatively larger and hence the EDL effects exhibit stronger. In Figures 3.11 (a) and (b), the distribution of non-dimensional velocity is plotted for the channel of 20 x 30um for two different bulk ionic concentrations. It is generally known that the zeta potential has a dependence on the bulk ionic concentration. It was found [10] that the zeta potential changes from 100 to 200 mV while the ionic concentration varies from 10~4 to 10~6 M (mol/dm3) for the P-type silicon microchannels. This experimentally determined correlation
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
57
Figure 3.9. Non-dimensional velocity distribution in one quarter of a rectangular microchannel (geometric ratio of height to width = 2/3). (a) With EDL effects (non-dimensional electrokinetic diameter K=24.7) (b) Without EDL effects, and (c) With EDL effects (nondimensional electrokinetic diameter K=24.7) but without considering the surface conduction.
58
Electrokinetics in Microfluidics
between zeta potential and ionic concentration is used in the calculation here. As seen, the EDL exhibits significantly stronger effects on the flow pattern for the dilute solution in Figure 3.11 (b) than that in the higher concentration in Figure 3.1 l(a). In Figure 3.12, the non-dimensional volumetric flow rate is plotted as a function of non-dimensional pressure difference for different concentrations of the aqueous solution and zeta potentials. As expected, the flow rate is reduced because of the electrokinetic effects. Basically Figure 3.12 shows that the flow rate exhibits the same EDL dependence as the streaming potential does. 3-2.5 Electro-viscous effects It is apparent from the above analysis that the presence of an EDL exerts an electrical force on the ions in the liquid, and hence has a profound influence on the flow behaviour. As discussed above, the streaming potential is established when ions in the liquid are carried to the downstream. The streaming potential in turn will produce a backward ion flow (the conduction current). It is easy to understand that when ions move in a liquid, they will pull the liquid molecules to move with them. Thus the streaming potential will produce a liquid flow in the direction opposite to the pressure-driven flow. The net effect is a reduced flow rate in the forward direction. The liquid thus appears to exhibit an enhanced viscosity if its flow rate is compared with that in the absence of the EDL effects (here the viscosity independent of electrolyte concentration is assumed [1]). As we have already shown, the non-dimensional flow rate through the microchannel with and without the consideration of the EDL effects are given by Eq (52) and Eq. (53), respectively. Equalizing Eq. (52) with Eq. (53), i.e. Qv - Qov, and using expressions for uave and uOave in Eq. (50) and Eq. (51), one may obtain the ratio of the apparent viscosity to the bulk viscosity as follows: (58)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
59
Figure 3.10. Non-dimensional velocity distribution as a function of geometric ratio of height to width in the presence of EDL (non-dimensional electrokinetic diameter K=24.7). (a) height/width=l/8, (b) height/width=l/4, (c) height/width=l/2, and (d) height/width=l/l.
60
Electrokinetics in Microfluidics
and
Since the non-dimensional pressure gradient is negative and both Ci and C2 are greater than zero, it is easy to show that this ratio is greater than 1, which is the electro-viscous effects. Using Eq. (58), the ratio of the apparent viscosity to the bulk viscosity, \iat I fif, is plotted as function of the non-dimensional electrokinetic diameter for different values of the zeta potential of the solid surface in Figure 3.13. It is seen from Figure 3.13 that /uaf I Hf is strongly dependent on the strength of the EDL effects. This indicates that the higher the zeta potential, the bigger the ratio naf I [if. In addition, of particular interest is the prediction of a maximum in jj.af I /Uf with respect to the non-dimensional electrokinetic diameter.
Figure 3.11. Non-dimensional velocity distribution in one quarter of a rectangular microchannel (ratio of height to width = 2/3).
-5,
and (b)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
61
Figure 3.12 Variation of non-dimensional volume flow rate with non-dimensional pressure difference for different concentrations and zeta potentials.
Figure 3.13. Variation of ratio of apparent viscosity to bulk viscosity with nondimensional electrokinetic diameter for different zeta potential.
62
Electrokinetics in Microfluidics
The similar pattern has been reported by Rice and Whitehead [3] and Levine et al [4]. There is a critical value for the non-dimensional electrokinetic diameter K corresponding to the strongest EDL effect, at which the largest occur in reduction in the flow and hence the maximum ratio of naflnf rectangular microchannels. When K increases from this critical value, which implies either a larger hydraulic diameter or a thinner EDL thickness (i.e. a higher electrolyte concentration), the EDL effects become weaker. Thus a smaller reduction in the flow and hence a lower ratio of naf I \ij can be observed. On the other hand, for the case that K is less than this critical value, this means either a smaller hydraulic diameter or a larger EDL thickness (i.e. a lower electrolyte concentration). However, as discussed in the preceding theory, the EDL effects on the fluid flow in the microchannels is considered by an additional body force in the equation of motion. By definition this additional body force is proportional to both the streaming potential and the net charge density, which directly relate to the liquid velocity and the electrolyte ionic concentration. If K becomes very small, we may have either extremely small microchannels (which allow very little flow) or a relatively diluted electrolyte. If the flow is very slow, the streaming potential is very small. If the ionic concentration in the solution is very low, the net charge density in the EDL is also very small. Therefore, the EDL effect or the electro-viscous effect becomes weaker and hence the smaller ratio of fxaj-1Hf is shown in Figure 3.13. In conclusion, there is no monotonic relation among the EDL effects on the fluid flow in rectangular microchannels and the channel size and the electrolyte ionic concentration, because the streaming potential, the ionic net charge density, and the liquid velocity depend upon a large number of basic parameters.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
3-3
63
MEASURED ELECTRO-VISCOUS EFFECTS
The theoretical analyses in the previous sections reveal significant electroviscous effects on liquid flow in microchannels. A key question must be answered: Is the electro-viscous effect as large as predicted by these theoretical models? It is desirable to examine if the electro-viscous effect can be measured and to compare the measured effect with the theoretical model prediction. Recently, some limited but direct experimental studies have provided qualitative verification to these models [11]. If the observed non-conventional flow behavior is due to the interfacial electrokinetic or the electro-viscous effects, it must depend on the ionic concentrations and the ionic valence of the testing liquids. For this purpose, pure water, aqueous KC1 solutions of two different concentrations, an A1C13 solution and a LiCl solution were chosen as the testing liquids in the experimental study. To avoid the complication due to irregular cross-section geometry and the surface roughness of the flow channels (micromachined channels on glass or silicon plates generally have a trapezoidal cross-section and relatively rough surface [12]), microchannels formed by two parallel, smooth silicon plates were used in the experimental studies. Figure 3.14 shows the experimental system used to study electrokinetic effect on the flow characteristics of a liquid flowing through a microchannel. This system consists of a flow loop, a test section including a slit microchannel, instruments for measuring flow and electrokinetic parameters, and a computer data acquisition system.
Figure 3.14. Schematic of the experimental system used to measure the flow and the electrokinetic parameters in microchannels.
64
Electrokinetics in Microfluidics
De-ionized ultra filtered water (DIUF) (Fisher Scientific) and aqueous KC1 solutions of two different concentrations, an A1C13 solution and a LiCl solution were used as the testing liquids. The concentrations of the KC1 solutions are 10"4 M (kmol/m3) and 10~2 M (kmol/m3) respectively. The concentration of both the AICI3 and the LiCl solutions is 10~4 M (kmol/m3). In the experiments, the testing liquid was pumped from a liquid reservoir to the flow loop by a high precision displacement pump (Ruska Instruments, Model: 2248-WII) which has a flow rate range of 2.5-560 cm3/hr and can generate a pressure up to 4000 psi (27.6 MPa). A 0.1 um filter was installed in the flow loop between the outlet of the pump and the inlet of the test section. The liquid was forced to flow through this submicron filter before entering the test section to avoid any particles or bubbles from flowing through the test section and blocking the microchannel. In order to minimize the environmental electrical interference on the measurement of the electrokenitic parameters, such as the streaming potential across the microchannel and the bulk conductivity of the liquid, the whole flow loop is made of plastic tubes and plastic valves. The silicon plates (30 mm in length, 14 mm in width and 1 mm in thickness) were used to form the microchannels in this study. The surface roughness of these plates is approximately 20 run. To form a microchannel, two strips of a thin plastic shim (Small Parts Inc.) were used as the spacer and put between a pair of silicon plates in the length direction along the sides of the plates, so that a flow passage of 5 mm width was formed. Then, a specially designed clapper was used to fix the relative position of the plates and the thin film spacers. Finally, epoxy resin was applied to bond the silicon plates together and to seal all the openings except the inlet and outlet of the microchannel. Three microchannels were made in this way and tested in this study. The channels have identical width and length which are 5 mm and 30 mm, respectively, and different heights. By choosing different shim thickness, the heights of the three microchannels are 14.1 um, 28.2 um and 40.5 um, respectively. The width and length of the microchannel can be accurately measured by using a precision gauge. The height of the microchannels was first directly measured by a microscope (Leica MS5 Stereomicroscope)—computer image analysis system with a resolution of 0.8 um. Then the channel height was calibrated by an indirect method that involves the flow of a high ionic concentration solution through the microchannel. For a high concentration electrolyte solution, the EDL thickness is very small (about a few nanometers) and the electrokinetic effect on the flow is negligible. The liquid flow in such a case is basically a Poiseuille laminar flow. Therefore, the channel height can be determined from the measured pressure drop and flow rate by using the Poiseuille flow equation. The channel heights determined in this way were used in this study. It was found
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
65
that the difference between the measured channel heights from these two methods was less than 0.5 um. A microchannel was placed in a two-part symmetrical Plexiglas assembly to form a test section, as shown in Figure 3.14. The epoxy resin was applied to bond the microchannel and the assembly together to avoid leaking. It was found that the height of the microchannel might be altered if the pressure of the liquid in the microchannel was too high. This is because neither the clapper nor the epoxy resin can stand very high pressure and deformation may happen. This limited the experiments to a small Reynolds number range especially for smaller microchannels. Two pairs of sumps were machined in the assembly and were used for the pressure drop and streaming potential measurement. A diaphragm type differential pressure transducer (Validyne Engineering Corp., Model: DPI5) with ±0.5% FS accuracy was connected to one pair of sumps to measure the pressure drop along the microchannel. The pressure transducer was calibrated by using a standard deadweight pressure source before used in the experiments. The flowinduced electrokinetic potential, the streaming potential, was measured by a pair of Ag/AgCl electrodes (Sensortechnik Meinsberg GmbH) and an electrometer (Keithley Instruments Inc., Model 6517). The volume flow rate of water flowing through the microchannels was measured by the weighting method as described below. The liquid exiting the test section was accumulated in a glass beaker whose weight was measured before. A stopwatch was employed to measure the time spent for the accumulation. Then, an electronic balance (Mettler Instrument AG, model: BB240) with an accuracy of 0.001 gram was used to measure the weight of the accumulated liquid. Usually one to two grams of the liquid was collected over approximately 20 to 30 minutes depending on the channel size and the flow rate. Evaporation effect was examined and found to be negligible. The total volume of the liquid was determined by dividing the weight by the liquid's density. The flow rate was then obtained by dividing the total volume of the accumulated liquid by the time. The accuracy of the flow rate measurement was estimated to be ± 2%. In an experiment, the pump was set to maintain a constant flow rate. The readings of the pressure drop along the microchannel were monitored and recorded. The flow was considered to have reached a steady state when the readings of the pressure drop did not change any more. At such a steady state, the flow rate, the pressure drop, the streaming potential and the bulk liquid conductivity were measured. The data quoted here are for steady state flow. For a given channel and a given testing liquid, the measurements for all the parameters were repeated at least twice for the same flow rate and in the same flow direction. The flow direction was then switched by adjusting the control
66
Electrokinetics in Microfluidics
valves, and the measurements were conducted for the same flow rate (i.e., the same pump setting) after the steady state was reached. Again, the measurements were repeated for at least twice. After the measurements for both flow directions were completed, the pump was set to a different flow rate and the measurements described above were repeated for the same microchannel, and so on. When changing a different testing liquid, the flow loop and the test section were flushed thoroughly by DIUF water and then the testing liquid for several hours to remove all the ions and/or other possible contamination left from the previous test. Because the electrical conductivity of the liquid is very sensitive to an even very small change in the ionic concentration, an online electrical conductivity sensor (Model:CR 7300, Metter-Toledo Process Analytical) was used as a monitor. The flushing process was continued until a steady liquid conductivity reading was achieved and that value was the same as the standard value for that liquid. Considering a rectangular microchannel of width 2W, height 2//and length L, as illustrated in Figure 3.6. In the entrance region, the liquid flow was not fully developed laminar flow. The entrance region length is given by [13]:
where Z)/, =
4HW
is the hydraulic diameter of the rectangular channel, um is the
mean velocity. One can easily estimate that the entrance length for a liquid flow at Re = 10 in a microchannel of 20 um in height is 8 um. In this region, the pressure drop is calculated by
where kin is the friction coefficient given by [13]: (61)
At the exit of the flow, the cross-section is suddenly increased greatly as the liquid leaves the slit microchannel and enters a big channel (about 10 mm in diameter). The liquid flow may become turbulent. The pressure loss at the exit is estimated by
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
67
(62) The net pressure drop without the losses at the entrance and at the exit is then: (63) The pressure drop reported in this paper is the net pressure drop. Using the measured flow rate, Q, and the measured pressure drop, the pressure gradient and the Reynolds number can be calculated as follows: (64)
(65) where APnet is determined from Eq.(63), L is the length of the channel (30 mm in this case), Dh is the hydraulic diameter of the channel, and Ac is the crosssection area of the channel. For the microchannels of three different heights, the experimentally determined pressure gradient and the Reynolds numbers were plotted in Figures 3.15, 3.16 and 3.17. Generally, for a given liquid, there are six data points for one and approximately the same Reynolds number. All data were obtained at 21°C. For KC1 solution at the high concentration 10~2 M (i.e., kmol/m3), the EDL thickness is approximately several nanometers. The EDL effect on flow can be neglected as the EDL thickness is so small in comparison with the channel height. The flow in the microchannels is considered as the conventional Poiseuille laminar flow. However, the EDL thickness is approximately 100 nm for the 10~4 M solution and 1 um for the DIUF water, respectively. As seen clearly from all these figures, for the lower concentration (10 4 M) solution and the pure water, the measured pressure gradients at the same Reynolds number (i.e., the same flow rate) are significantly higher (up to 20%) than that for the high concentration solution where there is no EDL effect on flow. However, the difference in dP/dx between the pure water and the 10~2 M solution diminishes as the channel height increases. Furthermore, for all the three microchannels, there are clear differences between the dP/dx ~ Re relationships for the 10"4 M solution and that for the pure water. That is, at the same Reynolds number, the dP/dx for the pure water is always higher than that of the 10~4 M solution. These
68
Electrokinetics in Microfluidics
experimental results show a strong dependence of the flow characteristics in microchannels on the channel size and on the ionic concentration of the liquid. KC1 solutions have symmetrical 1:1 ions (the valence is 1). However, the ions in the A1C13 solution are not symmetrical and the valence of Al+ ions is 3. Comparing the 10~4 M KC1 solution with the 10~4 M AICI3 solution, we see from Figures 3.15-3.17 that the measured pressure gradient for the 10~4 M AICI3 solution is always lower than that for the 10~4 M KC1 solution under the same Re number. In fact, the dP/dx~Re relationship of the 10~4 M A1C13 solution is essentially the same as that of 10~2 M KC1 solution. This clearly shows that the flow characteristics in microchannels depend not only on the ionic concentration but also on the valence and possibly the symmetry of the ions.
Figure 3.15 Experimentally determined pressure gradient and Reynolds number for a microchannel of a height 14.1 p.m.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
69
Figure 3.16. Experimentally determined pressure gradient and Reynolds number for a microchannel of a height 28.2 jim.
As seen from Figures 3.15-17, the 10 4 M LiCl solution has a very unusual behavior. The measured pressure gradient of the 10~4 M LiCl solution is always significantly higher than that of pure water under the same Re number, independent of the channel size. Considering LiCl solution has symmetrical 1:1 ions, one would expect that its flow characteristics are very similar to that of the KC1 solution. There is no obvious explanation to the observed behavior. As discussed in the introduction, because of the presence of the EDL, the pressure-driven flow induces an electrokinetic potential, the streaming potential. The streaming potential in turn will generate a conduction electrical current and hence a liquid flow opposite to the pressure-driven flow. The net result is a reduced flow rate in the pressure-driven flow direction under a given pressure drop or a higher-pressure drop for a given flow rate, in comparison with the conventional Poiseuille flow model. This is the electro-viscous effect. Using the conventional Poiseuille laminar flow equation, the measured pressure gradient and the flow rate, we can calculate the apparent viscosity u.a. Figures 3.18 (a), (b) and (c) show the ratio of the apparent viscosity to the true viscosity \ij\x as a
70
Electrokinetics in Microfluidics
Figure 3.17. Experimentally determined pressure gradient and Reynolds number for a microchannel of a height 40.5 urn.
function of the half electrokinetic channel height kH. Since kH = H/(l/k) where 1/k is the characteristic thickness of the EDL, kH can be understood as the channel height relative to the EDL thickness. Within the limited experimental data range, the electro-viscous effect can produce an apparent viscosity from several percent up to 18% higher than the true viscosity of the liquid, depending on the electrokinetic channel height. These results agree with the theoretical model prediction qualitatively. For these liquid-solid systems, the zeta potential and the surface conductance were evaluated from the measured streaming potential and conductivity data and were summarized in Table 1. As seen from Table 1, for a given solution, the surface conductivity increases slightly with the channel height. This phenomenon was reported by others as well [1,14,15].
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
71
Figure 3.18. Apparent viscosity ratio vs. non-dimensional electrokinetic channel height for (a) DIUF water, (b) 10~4 M aqueous KC1 solution and 10^ M LiCl aqueous solution, (c) 10~4 M aqueous AICI3 aqueous solution.
72
Electrokinetics in Microfluidics
Table 1 The zeta potential and the surface conductance for the three liquid-silicon microchannel systems.
Comparing the DIUF water and 10 4 M KC1 solution, the water-silicon surface system has a much higher zeta potential (£DIUF water = -245mV and £ , 0 - W , = - 1 0 7 m V > <"* a nm and -T\in-4\ivrimi°r
=
mUCh
^ ^
EDL fldd
(ilD.UFwater@2lOC=
305
^0-5 nm). The EDL field is stronger in the DIUF
water system than in the 10~4 M KC1 solution system. Correspondingly, the net charge density in the EDL region of the DIUF water is significantly higher than that of the 10~4 M KC1 solution. Furthermore, the bulk conductivity of the 10~4 M KC1 solution is about one order of magnitude higher than that of water (measured values: Xb DIUF water @2i°c = 1.053*10~4 S/m and Xb KT4 M KCI @2ioc = 15.025 xlO"4 S/m). The more conductive the liquid, the smaller the streaming potential that can be generated. Therefore, as discussed above, for the 10"4 M KCI solution system, a smaller streaming potential and a smaller net charge density will result in a smaller electrokinetic body force (with an opposite sign to the pressure gradient) in the equation of motion. This is the reason that the electro-viscous effect in the 10"4 M KCI solution system is smaller that in the DIUF water system. The 10~4 M A1C13 solution system has a smaller zeta potential (-60 mV), a smaller EDL thickness (-)- L-4.. 4ir,1,,~,,o« = 12.45 nm) and a much higher bulk conductivity (Xb 10~4M AICB @2i°c = 47.028 *10~4 S/m), in comparison with the DIUF water system and the 10~4 M KCI solution system. That is why the electro-viscous effect on the 10~4 M A1C13 solution system is very small and the
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
73
dP/dx~Re relationship of the 10 4 M A1C13 solution is more close to that of the 1(T2MKC1 solution. For the LiCl solution system, the measured bulk conductivity is Xb 10~4M UCI @2i°c = 11.95 "lO"4 S/m. The characteristic EDL thickness is Trlin-4MT • r i n ?i o r = 30.5 nm. From these properties, one would expect that the 10~4 M-LiCl solution would behave similarly to the 10"4 M KC1 solution. The electrokinetic flow model presented in this paper cannot explain the measured extra high flow resistance. However, it was noted that the radius ratio of Li+ to K+ is 0.68/1.33 « 0.511. It was suspected that this unusual behavior might have some thing to do with the size and the adsorption of Li+ ions in the compact layer of the EDL. In summary, the correlation among the pressure drop and the volume flow rate of DIUF water, 10"4 M and 10"2 M aqueous KC1 solutions, 10"4 M A1C13 aqueous solution and 10~4 M LiCl aqueous solution in silicon microchannels were experimentally studied. The flow resistance (i.e., dP/dx) of the 10~2 M KC1 solution and the 10~4 M AICI3 solution is essentially the same as that predicted by the Poiseuille laminar flow equation. However, up to 20% higher flow resistance was found for pure water and the 10~4 M KC1 solution. 20% to 40% higher flow resistance for the LiCl solution was found in comparison with the prediction of the Poiseuille laminar flow equation. These results show a strong dependence of dP/dx ~ Re relationship on the channel size, the ionic concentration, the ionic valence and the bulk conductivity of the liquids. The apparent viscosity corresponding to these measured dP/dx ~ Re relationships was found to be up to 18% higher than the true viscosity depending on the liquids and the ratio of the channel height to the EDL thickness. The experimental results show an unusual flow resistance behavior of LiCl solution that cannot be understood at the moment. Except the LiCl solution, at the same flow rate, the pure water has the highest streaming potential, the high concentration (10~2 M) KC1 solution has the lowest (essentially zero) streaming potential. According to the electrkinetic flow theory, the higher the streaming potential, the higher the electro-viscous effect on flow. These correspond well with the measured dP/dx ~ Re relationships.
74
Electrokinetics in Microfluidics
3-4 NEW UNDERSTANDING OF ELECTRO-VISCOUS EFFECTS The traditional understanding of the electro-viscous effect is as follows. Generally, a solid surface bears electrical charge when this surface is brought into contact with a polar liquid. The counter-ions in the liquid will be attracted to the wall and the co-ions will be repelled from the wall. As a result, there is a thin layer of liquid with a net charge or excess counter-ions near the solid-liquid interface. The charge on the solid surface and this charged liquid layer is called electrical double layer (EDL). When a liquid is forced to flow through a microchannel under an applied hydrostatic pressure, the net charges in the EDL are carried toward the downstream end, resulting in an electrical current in the pressure-driven flow direction. This current is called the streaming current. Corresponding to this streaming current, there is an electrokinetic potential called the streaming potential. This flow-induced streaming potential is a potential difference that builds up along the microchannel. This streaming potential acts to drive the net charges in the EDL to move in the direction opposite to the streaming current, i.e., opposite to the pressure-driven flow direction. When ions move in a liquid, they will pull the liquid molecules to move with them. Therefore, this motion of the net charges due to the presence of the induced streaming potential will produce a liquid flow in the opposite direction to the pressure driven flow. The overall result is a reduced flow rate in the pressure drop direction. If the reduced flow rate is compared with the flow rate predicted by the conventional fluid mechanics theory without considering the presence of the EDL, it seems that the liquid would have a higher viscosity. This is usually referred to as the electro-viscous effect. The experimental results discussed in the previous section have shown that the electro-viscous effect is significant for dilute solutions flowing through small microchannels. However, the traditional theoretical models, as presented in Sections 3-1 and 3-2 cannot predict the experimentally observed electro-viscous effects. The traditional model involves the application of the Poisson-Boltzmann equation. The reason that the traditional theoretical model fails to predict the experimentally observed electro-viscous effects is the use of the Boltzmann distribution for the ionic concentration field. The derivation of the Boltzmann distribution requires an infinitely large liquid phase so that the electrical potential is zero in the liquid far away from the charged surface and that the ionic concentrations in the region far away from the charged surface are equal to the original bulk ionic concentration. However, due to the presence of the charged solid-liquid interface, there exist an excess of counter-ions and a deficit of coions in the EDL region. It therefore should be expected that a deficit of counterions and a surplus of co-ions in the bulk liquid outside of the electrical double layer. The conventional Boltzmann distribution cannot show this logical expectation. This is because a key underline assumption of the Boltzmann
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
75
distribution is an infinitely large liquid phase. It is understandable that if the size of the system is sufficiently large and/or bulk ionic concentration is sufficiently high, such a deficit and such a surplus are negligible and the Boltzmann distribution is acceptable. For a system with a dilute electrolyte solution in a small microchannel, the excess of counter-ions and the deficit of co-ions in the EDL region will result in the significant changes in the concentration of counter-ions and co-ions in the region outside the EDL. In order to satisfy the conservation condition that the total ion number is constant in the given system, the concentration of counterions in the bulk liquid region is expected to be lower than the original bulk concentration due to the accumulation of the counter-ions in the EDL region, and the concentration of co-ions in the bulk liquid region is expected to be higher than the original bulk concentration due to the co-ion deficit in the EDL region. Consequently, neither the concentration of counter-ions nor the concentration of co-ions at the center of the microchannel is equal to the original bulk concentration. Therefore, the assumption that the ionic concentration in the bulk liquid region is equal to the original bulk concentration, which is used to derive the Boltzmann distribution, is not correct and hence the Boltzmann distribution cannot be applied to such a system. The purpose of this section is to present a new model to explain the electro-viscous effects [16]. In this model, instead of using the Boltzmann distribution, the conservation condition of ion numbers and the Nernst equation are used to find the ionic concentration field in the microchannels. A correct boundary condition at the center of the microchannel is derived and applied to this model. The ionic concentration field, electrical potential field and flow field are obtained by numerically solving this model. Finally, in order to verify this model, the numerical simulations are compared with the experimentally measured electro-viscous effects. 3-4.1 Derivation of Boltzmann Equation In order to see the difference between the traditional treatment and this new model, let's have a brief review of how the Boltzmann equation is derived. In the conventional treatment, the ionic concentration field in the EDL is described by the Boltzmann distribution. Consider a solid surface in contact with an infinitely large electrolyte solution. The ionic concentration filed in the EDL region is determined by the thermodynamic equilibrium condition, i.e., the electrochemical potential of the ions should be constant everywhere, (66) where / indicates the type-/ ions and the electrochemical potential is defined as
76
Pi = f*i + zieW
Electrokinetics in Microfluidics
( 67 )
where y/ is the electrical potential, ^ and z,- are the chemical potential and the valence of the type-/ ions, respectively; and e is the charge of a proton. The chemical potential of the ions can be further expressed as (68) where fij is a constant for the type-z ions, kf, is the Boltzmann constant, T is the absolute temperature of the solution, and M,- is the ionic number concentration of the type-z' ions. Substituting Eqs. (67) and (68) into Eq. (66), we obtain the Nernst equation, (69) In the classical theory, Eq. (69) is integrated from a point in the bulk solution where (70) to a point in the EDL region, nf is the original bulk concentration of the type-z ion. This process leads to the Boltzmann distribution: (71) It should be pointed out that the Boltzmann distribution, Eq. (71), is derived by employing the boundary condition, Eq. (70), which requires the liquid to be infinitely large and this condition cannot be satisfied for dilute solutions in small microchannels. This is because in small microchannels, the space is confined and the concentrations of ions at the center of the microchannel are no longer equal to the original bulk concentration, as discussed above.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
11
3-4.2 A new model of pressure-driven electrokinetic flow in a microchannel Consider a pressure-driven flow of a simple symmetric (e.g., z: z = 1:1) electrolyte solution such as KCl through a slit channel formed by two flat plates as illustrated in Figure 3.19.
Figure 3.19. Schematic diagram of a slit microchannel formed by two parallel plates with a separation distance of 2H.
Electrical potential field According to the theory of electrostatics, the relationship between the electrical potential, y/, and the net charge density per unit volume, pe, at any point in the liquid is described by the Poisson equation, (72) where s is the dielectric constant of the solution and s 0 is the permittivity of vacuum. The net charge density is proportional to the ionic number concentration difference between positive ions and negative ions, which includes the potential determining ions such as H3O+ and OtT ions and the non-potential determining electrolyte ions such as K+ and CF [1], via (73)
78
Electrokinetics in Microfluidics
where n^+ and n^,- are the number densities of K+ and Cl ions and nTT ,,+ K.
Cl
HjU
and «^,,.r_are the number densities of H3O+ and OfT ions, respectively. On
Combining the above two equations, the Poisson equation becomes, (74)
In order to solve Eq.(74) numerically, the following boundary conditions are required, (75a) (75b) where C, is the electrical potential at the solid-liquid surface. It should be noted that in traditional theoretical studies, the contributions of potential determining ions to the net charge density are not considered; i.e.,
and the Boltzmann distribution,
is used to describe the ionic concentration field for the non-potential determining K+ and Cl~ ions with the assumption of where y = H is the position of the channel's middle plane, and rij is the original bulk concentration of the type-z ion. As discussed above, there exist an excess of counter-ions and a deficit of co-ions in the EDL region due to the presence of the surface charge. This will result in a deficit of counter-ions and extra co-ions in the bulk liquid outside the EDL. Consequently, neither the concentration of counter-ions nor that of co-ions
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
79
at the middle plane between these two flat plates is equal to the original bulk ionic concentration. This implies that the boundary condition, Eq. (70), used to derive the Boltzmann distribution is not valid in such a system. Therefore, the Boltzmann distribution cannot be applied to such systems and the ionic number concentrations of ions have to be obtained through the Nernst equation and the ionic number conservation equation. Ionic concentration fields The relationship between the ionic concentration and the electrical double layer potential is given by the Nernst equation, (69) Integrating Eq. (69) from the wall, where (76) to a point in the bulk liquid, we get, (77) where «r _ wa // is the ionic concentration of the ith ionic species at the wall, which can be determined by satisfying the conservation condition of ion number (i.e., the total number of ions is constant in the system), given below, (78a) (78b) (78c) (78d)
80
Electrokinetics in Microfluidics
In the above,
are the original ionic number
concentration of K+, Cl~, H3O+ and OH" ions with a unit of m"3, respectively. Nj stands for the amount of water molecules dissociated into H3O+ and OH" ions and (To represents the surface charge. A method of determining the amount of the dissociated water molecules and the surface charge is described below. When an aqueous solution is in contact with an oxide surface (i.e. silicon surface), either H3O+and OH~ ions will attach on the solid surface depending on the pH value of the solution. For most silicon-electrolyte interfaces, pHz, the pH value at the point of zero charge, is between 1 and 4 [17] and pHz = 3.5 is used in this study for silicon microchannels [17]. Under this condition, if the pH value of the aqueous solution is above 6.1, the solid surfaces will be negatively charged due to the attachment of OH" ions to the oxide surface. The attachment of OH" ions will then affect the balance between H3O+ and OH" ions in the solution. The equilibrium between liquid water and the ions yields a constant known as the dissociation constant of water, Kw, defined as [18]: (79) where [// 3 (9 + ]and[CW~] are the ionic concentrations. If the unit of ionic concentration is chosen as M (mol/L), Kw has a value of 10~14 at 25°C, which means in any aqueous solution at 25°C, the product of [H-ilO+'\&n&[OH~] must equal 10~14. Normally, the ionic concentration of H3O+ and OH" ions are given by [H3O+] = \0~pHand [OH~] = \0pH~U. When the solution is in contact with a solid surface, the concentration of OH" ion will decrease due to the adsorption on the solid surface. In order to reach the equilibrium condition as show in Eq. (79) again, a certain amount of water molecules will dissociate into H3O+ and OH" ions. Thus, the total ionic number concentrations of both H3O+ and OH" ions can be expressed as
where n
H3O+
, and n
OH~
can be calculated by [18]:
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
81
(81a) (81b) where Na is the Avogadro number, and UQ is the surface charge density, which can be evaluated by [19]: (82) where A^is the site density on the oxide surface, \]/Q 1S non-dimensional surface potential, given by: (83) is the non-dimensional Nernst potential given by: (84) S in Eq.(82) is a parameter defined as:
where ApKis the dissociation constant difference [19]. The following material i o
j
properties were used in the calculation. Ns = 5 x 1 0 site Im , pHz = 3.5, and ApK = 10.0. If we define the average ionic number concentrations of H3O+ and Otf~ ions, n ,.+ and n , through the following equations: tlyU
Utl
(86a) (86b)
Eqs. (80a) and (80b) can be rewritten as:
82
Electrokinetics in Microfluidics
(87a) (87b) If we assume that the average ionic concentrations of H3O+ and OH~ ions also obey the equilibrium condition, Eq. (79), the product of ionic number concentration can be described by (88) The amount of water molecules dissociated into H3O+ and OH ions, A^, can be determined by solving Eqs. (87) and (88). Flow field If we consider the flow is steady, two-dimensional, and fully developed, the equation of the flow motion can be written as: (89)
In the above equation, the electrokinetic potential, Ex, can be obtained through the balance between the streaming current and the electrical conduction current at the steady state (i.e., the net electrical current should be zero at the steady state). (90) Because of symmetry, the electrical streaming current, transport of the net charge in the EDL region with the liquid flow, is given by: (91) The conduction current, generated by the flow-induced electrostatic potential (the streaming potential), is given by (92)
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
83
where A is the measured electrical conductivity of liquid. Combining Eqs. (90) (92), the gradient of the streaming potential can be obtained as:
(93) Substitute Eq. (93) for £ x into Eq. (89), the momentum equation becomes:
(94)
The above equation of motion is subjected to the following boundary conditions: (95a) (95b) Computational methods The EDL potential field, the ionic concentration fields and the velocity field in microchannels can be determined by simultaneously solving the above governing equations. The first step is to determine the zeta potential value. It should be realized that the zeta potential of a solid-liquid system is the material property of the system and generally is independent of the size of the microchannel (as long as the channel size/height is sufficiently large and there is no significant EDL field overlap). Therefore the zeta potential determined from a large channel (e.g., 200 um in height) should be the same as that for a small channel (e.g., 20 um in height). When the channel height is significantly (say 100 or 1000 times) larger than the EDL thickness, the accumulation of counterions and the deficit of co-ions in the EDL region have very little effects on the redistribution of these ions in the bulk liquid (outside of EDL) region. In other words, the concentrations of the ions in the bulk liquid region remain the same as the original bulk ionic concentrations. Therefore, the ionic concentration in such a large channel can be predicted by the traditional Boltzmann equation. Thus, the zeta potential can be determined by the following procedures: 1. Assume a zeta potential value 2. Numerically solve the traditional P-B equation over a very large channel.
84
Electrokinetics in Microfluidics
3.
Integrate the ionic concentration distribution over the half of the channel height to find out if it satisfies the ionic number conservation condition. 4. If yes, the assumed zeta potential value is a proper can be used in the next step. Otherwise, assume another zeta potential value and repeat the above procedure until a proper one is found. It should be pointed out that the experiments of pressure-driven flow in microchannels as described in the previous section do not a pH control. In other words, the pH values of the solutions are not known in these experiments. Since the channel wall surface is negatively charged,CT0, the surface charge density is a negative value. In order to satisfy this condition, the nondimensional Nernst potential must be smaller than the nondimensional surface potential (.From Eqs. (82) - (84), we have: (96) This provides a condition for choosing a pH value. For a given pH value, the amount of water molecules dissociated into H3O+ and OFT ions, Nd, can be determined using Eqs. (87) and (88). Finally, the potential distribution and the ionic concentrations can be obtained by numerically solving Eqs. (74) and (77) with the proper boundary conditions, Eq. (75), and the ionic number conservation conditions, Eq. (78). In these computations, a set of wall concentration values for each type of ions involved in the ionic number conservation conditions has to be assumed first. If the obtained ionic concentration distribution cannot satisfy the ionic number conservation conditions (i.e., the total number of ions is constant in the system), a set of new values for the ionic concentration at wall must be assumed and the above procedures are repeated. If any sets of these boundary concentration values cannot meet the requirement of the ionic number conservation conditions, a new pH value must be assumed, and all of the above procedure are repeated until the ionic number conservation condition can be satisfied under a given pH value, a set of assumed wall ionic concentration values, and the determined zeta potential. Once the electrical double layer potential distribution and ionic concentration distributions are obtained, the flow equation can be numerically solved with the boundary conditions and the applied pressure gradient. The above solving procedure is summarized below: 1. Based on the determined zeta potential, assume a pH value according to Eq. (96). 2. Calculate the dissociated water molecules, Nd, using Eqs. (87) and (88).
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
85
3. Assume an ionic concentration value at the wall for each type of ions (i.e., K+, C r , H 3 O + and OH"). 4. Solve the potential equation, Eq.(74), numerically with Eq.(77) and the boundary conditions specified in Eq. (75). 5. Calculate the ionic concentration using Eq. (77) 6. Check the ionic number conservation conditions, Eq. (78). If these conditions are satisfied, go to the next step. Otherwise, repeat step 3 to 6 and perform the following two procedures. 1) If a correct ionic concentration at the wall for each type of ions can be found, go to the next step. 2) If a correct ionic concentration at the wall for each type of ions cannot be found, repeat step 1 to 6 until these concentrations are found and the ionic number conservation conditions are satisfied. 7. Once the potential distribution and ionic concentration distribution are found, solve the flow equation, Eq. (94), numerically with the boundary conditions specified in Eq. (95) and the applied pressure gradient to find the flow field.
Figure 3.20 Comparison of the model predicted ionic concentration fields in DIUF water (no = 1.0 |iM) in a microchannel with a height 14.1 \im between the traditional model and the newly developed model. The zeta potential is - 150 mV.
86
Electrokinetics in Microfluidics
3-4.3 Effects of the EDL field and streaming potential Because of the EDL near the wall, the counter-ions are attracted to the wall and the co-ions are repelled from the wall. In order to satisfy the condition that the total number of ions is constant in a given system, a deficit of counterions and a surplus of co-ions in the liquid outside the EDL are expected. Consequently, there is an excess of counter-ions near the wall in the EDL region and an excess of co-ions outside the EDL. Figure 3.20 shows the ionic concentration distributions in DIUF water in a microchannel with a channel height 14.1 (am. Since the concentration of counter-ions in the bulk liquid region and the original bulk concentrations are very small as compared with that near the wall, the deficit of counter-ions in the bulk liquid region cannot be clearly seen in a full-scale plot. Figure 3.20 shows the concentration distributions in the region with the concentration value close to the original bulk concentration. In this way, the deficit of the ions in the bulk liquid region can be seen clearly. Figure 3.21 shows the comparison of ionic concentration fields of counterions for two channels with a channel height of 14.1 urn and 40.5 um, respectively. As shown in the figure, when the channel height is increased, the ionic concentration in most of the bulk liquid region becomes closer to the original bulk concentration. This can be understood as follows. When the
Figure 3.21 Comparison of the counter-ion concentration fields for two channels with a channel height of 14.1 u.m and 40.5 \im, respectively. The liquid is DIUF water (n o = 1.0 )iM) and the zeta potential is - 150 mV.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
87
channel height is increased, the double layer thickness, which is a constant for a given channel material and a given solution, becomes a smaller portion of the channel height. The bulk liquid region outside the EDL contains a large number of ions, thus the accumulation of counter-ions within the EDL region has a smaller effect on the counter-ion concentration in the bulk liquid region. It can be expected that when the channel height is extremely large, the concentration of counter-ion in the bulk liquid region will be the same as the original bulk concentration as predicted by the Poisson-Boltzmann equation. This has been verified through numerical simulation. Because of the negatively charged channel walls, there are excess counterions (positively charged ions) near the wall and excess co-ions (negatively charged ions) outside the EDL region. Correspondingly, there will be a net positive charge density near the wall and a net negative charge density in the bulk liquid. Figure 3.22 shows the net charge density distribution in DIUF water in a microchannel with a channel height of 14.1 um. The line of zero net charge density is shown in the figure for reference. Figure 3.22 is an enlarged view for the net charge density values close to the zero net charge. It should be noted that there is no net charge outside the EDL in the description of the traditional Boltzmann distribution.
Figure 3.22 The net charge density distribution in DIUF water (n0 = 1.0 |aM) in a silicon microchannel with a channel height 14.1|im and a zeta potential of- 150 mV.
88
Electrokinetics in Microfluidics
When a liquid is forced to flow through a microchannel by a hydrostatic pressure, both the net charges in the EDL region near the wall and the net charges outside EDL will be carried to the downstream, resulting in the streaming current. In the traditional treatment, only the streaming current resulting from the motion of the net charges in the EDL is considered. However, as shown in Figure 3.23, the contribution to the streaming current by the motion of the net charges outside EDL also has to be considered. Although the two streaming currents carry the opposite charges, one must realize that the streaming current resulting from the net charges outside EDL is larger than that resulting from the net charges near the wall. This is because the streaming current depends on the product of the net charge and the local velocity (see Eq. (91)). The local velocity outside the EDL and particularly near the middle plane is much bigger than that in the EDL near the wall for pressure driven flow. Therefore, the streaming current outside the EDL is a determinant one. Thus, in comparison with the streaming current evaluated using the traditional model, the streaming current predicted by using the new model is much bigger. Consequently, the streaming potential, which is dependent on the streaming current, is much bigger than that predicted by using the traditional model. Because the streaming potential always acts to drive the net charges to flow in the opposite direction to the streaming current, i.e., opposite to the pressure driven flow direction, a bigger streaming potential results in a stronger electro-viscous effect. Therefore,
Figure 3.23. Illustration of the net charge distribution in the liquid and the directions of the pressure-driven flow and the conduction flow in a microchannel.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
89
much stronger electro-viscous effects are predicted by using this new model due to the consideration of the net charge outside EDL. Figure 3.24 shows the comparison of the predicted velocity profiles between the model without considering the EDL effects, the traditional EDL model where the Boltzmann distribution is used and the new model. The velocity profile predicted by this new model shows a lower velocity. The velocity profile predicted by the traditional model has little difference from the velocity profile predicted by the model without considering the EDL effects. This indicates that the traditional model cannot predict the experimentally observed electro-viscous effects. The new model presented here can be used to predict the relationship between the pressure gradient and Reynolds number, Re = puH/n ( p is the density of the liquid). As an example, Figure 3.25 shows the comparison between the experimentally determined dP/dx ~ Re relationship for DIUF water flowing through the silicon microchannels and the model predictions. When the EDL effects are not considered, the model predicted Reynolds number is much higher than the experimentally determined Reynolds number under a given pressure gradient value. The traditional flow model predicts a slightly lower Reynolds number than that without considering EDL effects, but far away from the experimental results. The predictions using the new model described here agree very well with the experimental results.
Figure 3.24 Comparison of the predicted velocity profiles between the model without considering EDL effects, the traditional EDL model using the Boltzmann distribution and the new model for DIUF water flowing in a silicon microchannel with a height 14.1 um at a pressure gradient of 4.5 M Pa/m.
90
Electrokinetics in Microfluidics
Figure 3.25 Comparison of experimentally determined dP/dx ~ Re relationship with the model predictions for DIUF water for different channel heights: (a) 14.1 |am, (b) 28.2 |xm and (c) 40.5 urn.
Electro-viscous Effects on Pressure-driven Liquid Flow in Microchannels
91
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19]
R. J. Hunter, Zeta Potential in Colloid Science, Principles and Applications, Academic Press, New York, 1981. J. Lyklema, Fundamentals of Interface and Colloid Science, Vol. II, Academic Press, New York, 1995. C. L. Rice and R. Whitehead, J. Phys. Chem., 69 (1965) 4017. S. Levine, J. R. Marriott, G. Neale and N. Epstein, J. Colloid Interface Sci., 52 (1975) 136. C. Yang and D. Li, Colloids Surfaces A, 143 (1998) 339-353. C. Yang and D. Li, J. Colloid Interface Sci, 194 (1997) 95-107. M. Mala, C. Yang and D. Li, Colloids and Surfaces A. 139 (1998) 109-116. J.V. Beck, Heat Conduction Using Green's Functions, Hemisphere Publishing Co, London, 1992. D. Li, "Measurement of Surface Conductance", Encyclopedia of Surface and Colloid Science, A. Hubbard, Editor, Page 3167-3177, Marcel Dekker, New York, 2002. M. Mala, D. Li, C. Werner, H. Jacobasch and Y. Ning, Int. J. Heat Fluid Flow, 18 (1997) 489-496. L. Ren, D. Li and W. Qu, J. Colloid Interface Sci, 233 (2001) 12-22. W. Qu, M. Mala and D. Li, Int. J. Heat Mass Transfer, 43 (2000) 353. C. Werner, H. Korber, R. Zimmermann, S. Dukhin and H.J. Jacobasch, J. Colloid Interface Sci, 208 (1998) 329-346. F.A. Morrison and J.F. Osterle, J. Chem. Phys, 43 (1965) 2111. S. Arulanandam and D. Li, J. Colloid Interface Sci, 225 (2000) 421-428. L. Ren, Ph.D. thesis, University of Toronto, 2004. R. Raiteri, B. Margesin, M. Grattarola, Sensors and Actuators B, 46 (1998) 126 - 132. E. R.Toon, G. L. Ellis, "Foundations of Chemistry", Holt, Rinehart and Winston, Toronto, 1978. T. W. Healy and L. R. White, Adv. Colloid Interface Science, 9 (1978) 303 - 345.
92
Electrokinetics in Microfluidics
Chapter 4
Electroosmotic flows in microchannels As discussed in Chapter 2, most surfaces obtain a surface electrical charge when they are brought into contact with a polar medium. The surface charge, in turn, influences the ion distribution in the polar medium, forming the electrical double layer (EDL). The EDL is the region near the charged surface where counter-ions and co-ions in a polar medium are preferentially distributed, such that the local net charge density is not zero. For instance, if the surface is negatively charged, the EDL in the liquid phase will have net positive charge. If an electrical field is applied tangentially to the EDL, an electrical body force is exerted on the excess counter-ions in the diffuse layer of the EDL. The excess counterions will move under the influence of the applied electrical field, pulling the liquid with them. The liquid movement is carried through to the rest of the liquid in the channel by viscous force, resulting in an electroosmotic flow (EOF). This electrokinetic process was first introduced by Reuss in 1809 [1]. In the literature, many studies have been reported on the steady state electroosmotic flow in microchannels of various shapes such as annulus [2,3], elliptical microchannels [4,5], rectangular microchannels [6], and T- and Yshaped microchannels [7,8,9]. On the unsteady electroosmotic flows, Li et al. studied electroosmotic solution displacing processes [10,11]. Soderman and Jonsson [12] developed a theoretical framework describing the temporal resolution of the electroosmotic flow in both planar and cylindrical geometries under the effect of pulsed electric fields. Using slip velocity approach, Santiago [13] studied the effects of fluid inertial and pressure on transient electroosmotic flows between two parallel plates. Without assumption of the thin EDL thickness, Keh and Tseng [14] obtained an analytical solution for the transient electrokinetic flow in a capillary tube and in a capillary slit. However, these theoretical analyses were based on the Debye-Hiickel linear approximation, which is valid only for low zeta potentials. Using a complete numerical approach, Yang et al. [15] studied the time and space development of electroosmotic flow in a slit microchannel without the linear assumption. Yang et al also developed analytical solutions for electroosmotic flow in cylindrical capillary under AC fields [16]. Electroosmotic flow in microchannels is of critical importance to labs-ona-chip, because most lab-on-a-chip devices use electroosmostic flow, for
Electroosmotic Flows in Microchannels
93
example, to transport liquids and mix different solutions through the microchannel network. In practice, this can be realized by applying an electrical field (i.e., a voltage difference) via two electrodes inserted in two wells (liquid reservoirs) connected by a microchannel filled with an aqueous solution. If the channel wall surface is negatively charged (e.g., a microchannel made in a glass chip), the counterions in the EDL are positive ions and they will move towards the negative electrode, generating an electroosmotic flow in that direction. Transporting liquids in microchannels by applying an electrical field (instead of a pressure difference) is called electroosmotic pumping. Electroosmotic pumping is widely used not only to transport bio- or chemical solutions and samples in lab-on-a-chip devices, but also to pump liquids in various MEMS devices, for example, pumping liquids in micro heat sinks for cooling microchips and laser diode arrays. The electroosmotic pump is truly a microscale pump (i.e., the micron-sized electrodes can be embedded in the microchannels during the microfabrication), has no moving parts, and can control the flow direction and flow rate easily and accurately. Although the phenomenon of electroosmosis has been known for nearly two centuries, its application as a microscale pump has only been studied recently. When designing and operating a conventional liquid pump, we need to know the quantitative relationships among the pressure, liquid properties, flow rate and pump size. Similarly, in order to design an electroosmotic pump and control the flow rate precisely, one must understand the electroosmotic flow in microchannels, i.e., the dependence of the flow rate on the channel size, liquid properties, zeta potential of the channel wall and the applied electrical field strength. The purpose of this chapter is to provide an introduction in this regard by discussing electroosmotic flows in ID and 2D microchannels, electroosmoticdriven solution displacing processes and transient electroosmotic flow processes.
94
4-1
Electrokinetics in Microfluidics
ELECTROOSMOTIC FLOW IN A SLIT MICROCHANNEL
Consider an electroosmotic flow of an aqueous solution in a slit microchannel formed between two parallel plates, as shown in Figure 4.1. Generally, most electroosmotic flows, owing to the inherently low Reynolds number, reach a steady state very quickly and the entrance length tends to be extremely short. For electroosmotic flow in microchannels of 100 ~ 200 urn in hydraulic diameter, the time to reach a steady state is on the order of microseconds and the entrance length is approximately lOum. More discussions on the transient aspects of electroosmotic flow will be given in a later section. For the slit microchannel illustrated in Figure 4.1, as the width of the channel is much larger than the height of the microchannel, w » a, the edge effects are negligible and thus the velocity profile essentially uniform along the x-axis. Therefore, the flow can be assumed as steady, one-dimensional and fully developed and can be described by the momentum equation shown below:
(1) where v2 = vz(y) is the velocity in a z-direction, Ez is the applied electrical field strength (V/m) and p(y) is the net charge density. The second term in Eq. (1) describes the electrical body force. In the above equation we assume no hydrostatic pressure gradient in the microchannel. The electrical double layer field between the two parallel plates is considered as one-dimensional only, i.e., in the y direction, and is given by the Poisson equation: (2)
Therefore, the net charge density can be expressed by using the Poisson equation as: (3)
Electroosmotic Flows in Microchannels
Figure 4.1
95
Schematic of a slit microchannel formed between two parallel plates.
Inserting Eq. (3) into Eq. (1) yields:
(4)
By integrating Eq. (4) twice and applying no-slip boundary condition for flow, and constant ^-potential for the EDL at the shear plane (i.e., at y = ± a, y(y) = C,, vz(y) = 0), the following equation is obtained (after some rearrangement): (5) which describes the flow field in terms of the EDL potential distribution, y/ (y). For a symmetric electrolyte solution, the net charge density is proportional to the concentration difference between the cations and the anions, and is given by the Boltzmann equation as: (6) By substituting Eq. (6) into the Poisson equation, Eq. (2), a nonlinear secondorder differential equation (P-B equation) is obtained:
96
Electrokinetics in Microfluidics
(7) In order to find an analytical solution, Debye-Hiickel linear approximation
(8) where K is the Debye-Hiickel parameter. Integrating Eq. (8) with the boundary condition y/ (y = ±a) = C, yields: (9) Eq. (9) can then be substituted into Eq. (5) to obtain the velocity profile in the slit microchannel: (10) Using Eq. (10) the average velocity, vav, can be obtained by integrating vz(y) over the j-domain and dividing by the channel height: (11) If Ka = CI/(1/K) is large, for example, when the double layer thickness (1/K) is small, it can be shown that sinh(fca) « Ka and cosh(fca) « 1. In this case Eq. (11) is reduced to: (12) Eq.(12) indicates that the electroosmotic flow velocity is linearly proportional to the applied electrical field strength and linearly proportional to the zeta potential. The negative sign indicates the flow direction and has to do with the sign of the £
Electroosmotic Flows in Microchannels
97
potential. If C, potential is negative (i.e., a negatively charged wall surface), the excess counterions in the diffuse layer are positive, therefore the electroosmotic flow in the microchannel is towards the negative electrode.
98
4-2
Electrokinetics in Microfluidics
ELECTROOSMOTIC FLOW IN A CYLINDRICAL MICROCHANNEL
Consider an electroosmotic flow of an aqueous electrolyte solution in a cylindrical microchannel with a circular cross-section, as illustrated in Figure 4.2. We assume the bulk composition of the solution is uniform and the channel wall has uniform and constant charge or zeta potential along the entire channel. The two ends of the microchannel are subject to the atmosphere pressure.
Figure 4.2
Illustration of a cylindrical microchannel with a circular cross-section.
4-2.1 Electrical double layer in a cylindrical microchannel According to the theory of electrostatics, the relationship between the electrical potential, y/(r), and the net charge density per unit volume, pe, at any point in the liquid is described by the Poisson equation, (13) where s is the dielectric constant of the solution and £0 is the permittivity of vacuum. Assuming that the equilibrium Boltzmann distribution is applicable, the ion number concentration per unit volume in an electrolyte solution is of the form (14)
Electroosmotic Flows in Microchannels
99
where njao and z ; are the bulk ionic concentration and the valence of type / ion, respectively, e is the charge of a proton, kb is Boltzmann constant and T is the temperature. The net volumetric charge density pe is proportional to the concentration difference between cations and anions, given by: (1 5 ) For a symmetric electrolyte solution such as KC1 (z: z = 1:1) solution, the above equation becomes: (16) For a non-symmetric electrolyte solution such as LaCl3 (z : z ^ 1:1), it takes the following form: (17) Substituting the equation of the net charge density into the Poission equation, Eq. (13), and introducing the dimensionless variables
where d is the diameter of the microchannel, the non-dimensional PossionBoltzmann equation can be written as follows, for symmetric electrolyte solution such as KC1 solution: (18) for non-symmetric electrolyte solution such as LaCl3 solution: (19)
100
Electrokinetics in Microfluidics
Because of the symmetry of the EDL field in the cylindrical capillary, Eq. (18) and Eq. (19) are subjected to the following non-dimensional boundary conditions: (20a)
(20b) where £ is the zeta potential, a measurable electrical potential at the shear plane. 4-2.2 Electro-osmotic flow field in a cylindrical microchannel The motion of the aqueous electrolyte solution is governed by the NavierStokes equations and the continuity equation, i.e. (21) (22) where u is the velocity vector, fj. is the viscosity, p is the density of the fluid, pe is the local net charge density and E is the electrical field strength applied to the capillary. If we assume that the flow is one dimensional, steady state and fully developed, then the velocity components are described by: (23a) (23b) (23c) With these conditions, the inertial term on the left hand side of Eq.(21) drops out and the equation of motion is reduced to: (24)
Electroosmotic Flows in Microchannels
101
We consider that the pressures at the two ends of the microchannel are the same (e.g. atmospheric pressure). For a simple electro-osmotic flow in an open channel, there is no pressure gradient along the capillary. Therefore, AP term in the above equation is zero. However, in the case of electro-osmotic flow with one solution replacing another solution, or in the case of electroosmotic flow in a channel with variable surface charge, the driving force of the electro-osmotic flow, the net charge density and the applied electrical field strength, are different in different sections of the microchannel. This would imply different velocity fields and different flow rates in different sections. For incompressible liquids, however, the continuity condition requires a constant volume flow rate through the channel. Therefore, an induced pressure gradient along the channel is required to satisfy the continuity condition. These cases will be discussed in later sections. Substituting the Eq. (16) and Eq. (17) for the net charge density into Eq. (24) and introducing the following non-dimensional variables,
where D is the diffusion coefficient, the non-dimensional equation of motion can be obtained, for symmetric electrolyte solution such as KC1 solution: (25) for non-symmetric electrolyte solution such as LaCl3 solution: (26) Eq. (25) and Eq. (26) are subject to the no-slip and symmetric boundary conditions: (27a)
(27b)
102
Electrokinetics in Microfluidics
The above equations may not have an analytical solution. Therefore, to obtain information about the electro-osmotic flow field, we solve the above equations numerically. Figure 4.3 show the EDL field and EOF velocity field in a cross-section of the cylindrical microchannel. In these Figures, the electrolyte is KCL, the concentration is lxlO"6 M, the zeta potential is chosen as -200 mV, the applied electrical field strength is 30 kV/m, and the diameter of the microchannel is 100 um. If the surface potential or zeta potential is sufficiently low, we may use the Debye-Huckel approximation to simplify the Poisson-Boltzmann equation such as Eq.(18). The Debye-Huckel approximation assumes that the value of y/ is small ( i.e., y/< 25mV) so that the following approximation can be made: (28) For the linearized Eq.(18), the following exact solutions for the local net charge density distribution and the velocity profile can be obtained analytically: (29)
(30) where Io is the zero-order modified Bessel function of the first kind; s is the dielectric constant of the liquid, the subscript DH indicates the Debye-Huckel approximation. Again, Eq.(30) shows that the electro-osmotic flow velocity is a linear function of the applied electrical field strength and zeta potential.
Electroosmotic Flows in Microchannels
103
Figure 4.3 The EDL field and electroosmotic flow velocity field of a 10 6 M KC1 solution (pure water) in a cross-section of a cylindrical microchannel of 100 um in diameter with C, = -200 mV. The applied field strength is 300 V/cm.
104
Electrokinetics in Microfluidics
4-3 ELECTROOSMOTIC FLOW IN RECTANGULAR MICROCHANNELS In electroosmotic flow, since the liquid motion is initiated by the electrical body force (the driving force) acting on the ions in the diffuse layer of the EDL, electroosmotic flow depends not only on the applied electrical field but also on the local net charge density in the liquid. The local net charge density will depend on the EDL filed and hence depend on the microchannel's cross-section geometry. Earlier studies of EDL and electroosmotic flows were limited to systems with simple geometries, such as cylindrical capillaries with circular cross-sections or a slit-type channel formed by two parallel plates. However, the channels in modern microfluidic devices and MEMS are made by micromachining technologies. The cross-section of these channels is close to a rectangular shape. In such a situation, the EDL field is two-dimensional and will affect the two-dimensional flow field in the rectangular microchannel. In order to understand the characteristics of electroosmotic flow in rectangular microchannels, and to control electroosmotic pumping as a means of transporting liquids in microstructures, we must understand the characteristics of electroosmotic flow in rectangular microchannels. In this section we examine the numerical solutions of the 2D Poisson—Boltzmann equation and the 2Dmomentum equation for electroosmotic flows in rectangular microchannels [6]. The EDL field, the flow field and the volumetric flow rate will be studied as functions of the zeta potential, the liquid properties, the channel geometry and the applied electrical field.
Figure 4.4. Illustration of a rectangular microchannel. The shaded region indicates the computational domain.
Electroosmotic Flows in Microchannels
105
4-3.1 Electrical double layer field in a rectangular microchannel Consider a rectangular microchannel of width 2 W, height 2H and length L, as illustrated in Figure 4.4. Because of the symmetry in the potential and the flow fields, the solution domain can be reduced to a quarter section of the channel (as shown by the shaded area in Figure 4.4). We are interested in the potential and velocity fields within the diffuse layer (not including the immobile layer) of the EDL. Therefore, the exterior surfaces of the solution domain coincide with the shear plane. Before considering the effect of an applied voltage potential along a microchannel, we first solve the EDL field to find the net charge density in the EDL region. The 2D EDL field in the rectangular crosssection of the microchannel can be described by Poisson equation: (31) Assuming that the Boltzmann distribution applies, the equilibrium Boltzmann distribution equation can be used to describe the ion concentration as follows:
where «, is the ionic number concentration of the /th species, z, is the valence of type-r ions, nx is the ionic number concentration in the bulk solution, e is the fundamental electric charge, T is the absolute temperature and kb is Boltzmann's constant. The net charge density can then be expressed in terms of the Boltzmann distribution, which for a symmetric electrolyte is given by: (33)
Substituting the above expression into Poisson's equation, the 2D Poisson-Boltzmann equation is obtained: (34)
Along the planes of symmetry, the following symmetry boundary conditions apply:
106
Electrokinetics in Microfluidics
(35a)
(35b) Along the surfaces of the solution domain, the potential is the zeta potential: (35c) (35d) The Poisson-Boltzmann equation and the boundary conditions can be transformed into non-dimensional equations by introducing the following dimensionless variables: (36a) (36b) (36c) where Dh is the hydraulic diameter, for the rectangular channel shown in Figure 4.4, given by: (37) Substituting the above dimensionless variables into Eq. (31), the nondimensional form of the Poisson-Boltzmann equation is obtained: (38) where K, the Debye-Huckle parameter, is defined as follows:
Electroosmotic Flows in Microchannels
107
and 1/K is the characteristic thickness of the EDL. The non-dimensional parameter xDh is a measure of the relative channel diameter, compared to the EDL thickness. tcDh is often referred to as the electrokinetic diameter. The corresponding non-dimensional boundary conditions are given by: (40a) (40b) (40c) (40d) Eq. (38) is a non-linear two-dimensional partial differential equation that must be solved numerically, subject to the boundary conditions given in Eq. (40). Taylor series expansion can be used to linearize the non-linear source term on the right-hand side of Eq. (38) as follows: (41) where i//*' is the value for y* obtained in the previous iteration. A numerical scheme can then used to discretize, the governing differential equation, and the resulting system of algebraic equations may be solved using the Gauss-Seidel iterative technique, with successive over-relaxation employed to improve the convergence time [17,18]. Since the electrical potential field in the EDL varies greatly within a small distance of the channel walls, variable grid spacing should be employed to ensure that as the surface is approached, the grid spacing is refined enough to capture the sharp gradients. Once the electrical potential distribution \f/*(y*, z*) has been found, the net charge density at any point in the channel can be found by using Eq. (33).
108
Electrokinetics in Microfluidics
4-3.2 Electroosmotic flow field in a rectangular microchannel The general equation of motion for laminar conditions in a liquid with constant density and viscosity is given by: (42)
If we consider that the flow is steady, two-dimensional, and fully developed, then the velocity components are described by: (43a) (43b) (43c) In addition, if there is no pressure gradient, the first term on the right hand side of Eq. (42) drops off. The only force in such a simple electroosmotic flow is from the externally applied electrical field. In a later chapter, we will discuss the electroosmotic flow induced pressure gradient only when the solution properties or the surface charge (zeta potential) of the channel wall changes along the flow direction. Under the above conditions, the general equation of motion can be reduced to a balance between the viscous or shear stresses in the fluid and the externally imposed electrical field force: (44)
where Fx is the electrical force per unit volume of the liquid, which is related to the electric field strength Ex and the local net charge density as follows: (45) Substituting Eq. (45) into Eq. (44) and replacing pe by Eq.(33) results in: (46)
The following boundary conditions apply along the planes of symmetry:
Electroosmotic Flows in Microchannels
109
(47a)
(47b) Along the surface of shear (the surface of the solution domain), the velocity boundary conditions are given by: (47c) (47d) The equation of motion can be non-dimensionalized using the following additional transformations: (48a)
(48b) where U is a reference velocity and L is the distance between the two electrodes. Substituting the above dimensionless variables into Eq. (46), the nondimensionalized equation of motion becomes: (49) M is a new dimensionless group, which is a ratio of the electrical force to the frictional force per unit volume, given by:
The corresponding non-dimensional boundary conditions are given by:
110
Electrokinetics in Microfluidics
(51a) (51b) (51c) (5 Id) Eq. (49) is a 2D partial differential equation that can be solved numerically, using a finite difference scheme and Gauss-Seidel iteration, once a solution for has been obtained. From the dimensionless equations and boundary conditions, it can be concluded that y/* and u* are dependent on the following variables: (52a)
(52b) The performance of an electroosmotic pumping system can be characterized by using the following two summary parameters: the volumetric flow rate and the average velocity. The volumetric flow rate through the entire channel cross-section is given by: (53) and the average velocity can then be determined using: (54)
4-3.3 EDL and flow fields The above equations and boundary conditions for the EDL and the electroosmotic velocity fields in a rectangular microchannel were solved numerically [6]. The physical properties of an aqueous KC1 solution at a
Electroosmotic Flows in Microchannels
in
concentration of 1x10 6 M were used as the properties of the fluid, i.e., £ = 80 and /j. = 0.90 x 10~3 kg/(m-s) [19]. An arbitrary reference velocity of U = 1 mm/s was used to non-dimensionalize the velocity. Using experimental results [20], zeta potential values changes from 100 to 200mV, corresponding to three concentrations of the KC1 solution, l x l O 6 , lxlO" 5 and lxl0~ 4 M. In the calculation, the hydraulic diameter of the channel varied from 12 to 250 um, while the aspect ratio varied from 1:4 to 1:1, and the applied voltage difference ranged from 10 V to 10 kV. The EDL potential distribution in the diffuse double layer region is shown in Figure 4.5. The nondimensional EDL potential profile across a quarter section of the rectangular channel exhibits characteristic behavior. The potential field drops off sharply very close to the wall. The region where the net charge density is not zero is limited to a small region close to the channel surface. Figure 4.6 shows the non-dimensional electroosmotic velocity field for an applied potential difference of 1 kV/cm. The velocity field exhibits a profile similar to plug flow, however, in electroosmotic flow, the velocity increases rapidly from zero at the wall (shear plane) to a maximum velocity near the wall, and then gradually drops off to a slightly lower constant velocity that is maintained through most of the channel. This unique profile can be attributed to the fact that the externally
Figure 4.5. Non-dimensional electric double layer potential profile in a quarter section of a rectangular microchannel with KDR — 79, C,* = 8 and HIW= 2/3.
112
Electrokinetics in Microfluidics
Figure 4.6. Non-dimensional velocity field in a quarter section of a rectangular microchannel with KDh = 79, C* = 8, HIW= 2/3, E* = 5000 and M = 2.22.
imposed electrical field is driving the flow. In the region very close to the wall, the mobile part of the EDL region, the larger electrical field force exerts a greater driving force on the fluid because of the presence of the net charge in the EDL region. The flow in this region may be viewed as the 'active' flow. By contrast, the flow in the center part of the channel may be considered as 'passive' flow, caused by the viscous drag force. Consequently, the flow in the center of the channel has a lower velocity in comparison with the 'active' flow near the wall. Because the thickness of the EDL field is characterized by the Debye-Huckel length, the Debye-Huckel length can also be used to characterize the location of the maximum velocity in the electroosmotic flow profile. 4-3.4 Effect of hydraulic diameter Variation of Dh affects the following nondimensional parameters: the electrokinetic diameter, and the strength of the viscous forces in the ratio of
Electroosmotic Flows in Microchannels
113
electrical to viscous forces. The volumetric flow rate increased with approximately D\ as seen in Figure 4.7a. This is expected, since the crosssectional area of the channel also increases proportionate to D\. When larger pumping flow rates are desired, larger diameter channels would seem to be a better choice. However, there is no corresponding increase in the average velocity with increased hydraulic diameter. In addition, for closed systems, gains made by larger diameter channels may be offset by increases in the back-pressure. Changing the size of the channel affects both the EDL and the velocity profile, and the effects can be seen in Figures 4.7b and 4.7c. Note that in Figure 4.7c, only a very small fraction of the velocity profiles near the channel wall is shown, in order to see the differences clearly.
Figure 4.7. (a) Variation of volumetric flow rate with hydraulic diameter for three different combinations of concentration and zeta potential, with H/W= 2/3, and Ex = 1 kV/cm.
114
Electrokinetics in Microfluidics
Figure 4.7 (b) Non-dimensional electric double layer potential profile and (c) non-dimensional velocity profile along the width of a microchannel at ylH = 0 for two different channel sizes. In these figures, with C* = 8, HIW= 2/3 and Ex = 1 kV/cm. zlW= 1.0 represents the channel wall, and z/W=0 represents the center of the channel.
Electroosmotic Flows in Microchannels
115
4-3.5 Effect of aspect ratio With a rectangular microchannel not only the hydraulic diameter but also the channel shape will influence the velocity profile. This is because of the impact of the channel geometry on the EDL. Figure 4.8 shows the relationship between the aspect ratio (H/W) and the volumetric flow rate for a fixed hydraulic diameter. As expected, when the channel was rotated by 90 degrees, the same volumetric flow rates were achieved, demonstrating that the orientation of the channel is not significant as long as the electric field is applied tangentially to the channel. As the ratio of H:Wapproaches 1:1 (for a square channel), the flow rate decreases. This is because of the larger role that corner effects have on the development of the EDL and the velocity profile in square channels. 4-3.6 Effect of concentration Figure 4.9 illustrate the variation in volumetric flow rate with electrolyte (ionic) concentration. From Eqs. (39) and (50), it is clear that increasing the
Figure 4.8. Variation of volumetric flow rate with aspect ratio for three different combinations of concentration and zeta potential, with A, = 24 um and Ex = 1 kV/cm. In this case, z/W=1.0 represents the channel wall, and z/W= 0 represents the center of the channel.
116
Electrokinetics in Microfluidics
concentration results in an increase in K as well as M. Increasing the ionic strength or concentration results in the potential field falling off to zero more rapidly with distance. This effect is also known as "compression of the doublelayer" [21,22]. The ionic concentration effect on the velocity or the flow rate can be understood as follows. Since ionic concentration or the ionic strength influences the zeta potential, as the ionic strength or concentration is increased, the zeta potential decreases in value. As the zeta potential decreases, so does the electroosmotic flow velocity and the volumetric flow rate. 4-3.7 Effect of electric field strength Electro-osmotic flow, as explained earlier, is the result of a tangentially applied electric field on a channel with the presence of an EDL. As assumed in the model development, higher order terms were neglected in the equations of
Figure 4.9. Variation of volumetric flow rate with concentration for three different combinations of concentration and zeta potential, with A, =24 (xm, H/W=2/3, and Ex =1 kV/cm.
Electroosmotic Flows in Microchannels
117
motion, implying a linear relationship between voltage and velocity. Thus, as expected, the results indicate a linear variation in volumetric flow rate with the applied voltage along a rectangular microchannel, with all other parameters held constant. Experimentally it has been observed that for fields up to 3xlO7 V/m the velocity is always directly proportional to the applied field [23]. Electroosmotic flow differs from conventional laminar flow in that the driving force is not a pressure gradient but an externally applied electric field. The "effective pressure" or the "equivalent pressure" required to produce the same electroosmotic flow rate can be calculated using the Poseuille flow equation. Figure 4.10 shows the variation in Re with EJL and the corresponding pressure gradient (dP/dx) required to produce the same flow rate. From the figure it is clear that the pressure gradients required to drive the flow through microchannels may be prohibitively large, in comparison to the externally applied electrical field required. However, in applying large electrical fields, there are other considerations that must be accounted for. The three main considerations are, power consumption of the system, safety considerations, and the potential temperature increase caused by the heat generated with the applied electrical field. High power consumption is an overall drain on the system and could make implementing electroosmotic flow (EOF) systems unfeasible in many situations. Fortunately, because of the high resistance, the power requirements are usually quite low, for example, less than 1 mW for 10 kV/cm applied potential [24]. Safety issues were mentioned in a recent experimental study, where it was shown that the potential hazards are minimized with the low stored energy in most power supplies [20]. Finally, the increase in temperature due to Joule heating during electroosmosis may be estimated as follows. When an electrical field is applied along the channel, Joule heating occurs. An energy balance in the channel results in the following equation to describe the temperature increase over a time At:
(55) The above expression shows that AT is proportional to the applied voltage field squared. For a typical ionic concentration C = 1xlO~6 M, X = 149.79xl(T4 m2 Q/mol (the resistivity of KC1 solution), Cv =1.0 kcal /(kg °C) (the specific heat capacity) and p = 1000 kg/m3 (the liquid density), AT is related to E and A7 as follows: (56)
118
Electrokinetics in Microfluidics
If the electric field is applied for At = 10 s, then a temperature increase of AT = 5°C occurs at E =3.7xlO5 V/m. At a higher concentration, such as C = lx 1(T4 M, the upper limit for E increases to 1.4xlO8 V/m. Below these limits, for given concentrations, Joule heating will not affect the fluid properties appreciably.
Figure 4.10. Applied electric field strength and the equivalent pressure gradient versus Re, with A, = 24 urn, HIW = 2/3, KDH = 79, C* = 8, C =lx 10' 6 M.
Electroosmotic Flows in Microchannels
4-4
119
TRANSIENT ELECTROOSMOTIC FLOW IN CYLINDRICAL MICROCHANNELS
Most studies of electroosmotic flows deal with steady state problems. However, the transient aspects of electroosmotic flow are also important in various applications such as mixing solutions and dielectrophoresis in lab-on-achip devices. This section will review an analysis of transient electroosmotic flow in a cylindrical microcapillary [16]. 4-4.1 Transient electroosmotic flow Consider an electroosmotic flow in a cylindrical capillary of radius a. The liquid in the microcapillary is assumed to be an incompressible and Newtonian liquid of density, p , and viscosity, n with a symmetric mono-valence electrolyte. The capillary wall is uniformly charged with a zeta potential, £ . When an external electric field, E(t) is applied along the axis of the capillary, the liquid starts to move as result of the interaction between the net charge density in the electric double layer (EDL) and the applied electric field. The motion of liquid through the cylindrical microcapillary is governed by the Navier-Stokes equation (57) where u(r,t) is the transient velocity field. pe{r) is the local volumetric net charge density of the electrolyte due to the presence of the EDL. If the Boltzmann distribution is considered valid, pe (r) can be expressed as (58) where e is the elementary charge, nx is the ionic concentration in the bulk liquid (i.e., far from the charged surfaces), kb is the Boltzmann constant, T is the absolute temperature, and y/(r) is the electric potential of the EDL. Introducing the following dimensionless parameters, (59)
120
Electrokinetics in Microfluidics
we can nondimensionlize Eq. (57) and Eq. (58) respectively as (60) and (61) Here us is the steady state electroosmotic velocity and used here as a reference velocity,
where sr is the dielectric constant of the electrolyte and e 0 is the permittivity of vacuum. The reference velocity us can also be regarded as the "slip" velocity on the shear plane of the EDL [25]. Obviously, us is proportional to the magnitude of the external field, E and the zeta potential, C, . Substituting Eq. (61) into Eq. (60), we obtain (63)
is the Debye-Huckel parameter, and 1 / K denotes the characteristic thickness of the EDL. Eq. (63) is subject to the following initial and boundary conditions (64a) (64b) Equation (63) is a 2nd order inhomogenous diffusion equation. Yang et al [16] obtained an analytical solution to it by using the Green's function approach.
Electroosmotic Flows in Microchannels
121
(65)
i are the zero-order and the 1s order Bessel functions, respectively, and Xn are the positive roots of the zero-order Bessel function J 0 (A n ) = 0. Integrating Eq. (65) along the radius of the cylinder, we obtain the non-dimensional mean velocity
(66)
Sinusoidally alternating electric field Consider the application of a sinusoidally alternating electric field with an angle frequency co, (67) Substituting Eq. (67) into Eq. (65), we can show next
(68)
where / is the unit imaginary number. "REAL" denotes the real part of the solution. Here a new parameter f3 is defined as [26] (69)
122
Electrokinetics in Microfluidics
P represents the aspect ratio of the capillary radius a to the Stokes penetration depth 8S, defined as [26] (70)
where / = — is the frequency of the applied electric field. Accordingly, the In mean velocity is given by
(71)
Step-change electric field In this case, the external electrical field is applied and remains constant from the time t = 0 (i.e., the electrical field follows step function change) E(t) = E0H(t)
(72)
where
H«-8)
=\ [0
(73) (t<8)
is the well-known Heaviside step function. Substituting Eq. (72) into Eq. (65) and Eq. (66) gives (74)
and
Electroosmotic Flows in Microchannels
123
(75) Square wave electric field Let's consider a more general situation where the electric field is switched on at the time t = 0 and remains constant until it is switched off at the time t = 8, the electric field is then described by (76) Substituting Eq. (76) to Eq. (65) and Eq. (66) yields: (77)
and (78) The total volumetric net flow quantity from when the electric field is switched on to the termination of the entire flow due to frictional stress can be determined from (79) It is obvious that the total net flow quantity is proportional to the duration of time 8 (here 8 =
pa
8 ), during which the electric field is applied.
4-4.2 Electric double layer (EDL) field in the cylindrical capillary In order to calculate the dynamic electroosmotic flow velocity by using the derived expressions, we must know the electrical double layer field in the capillary. As clearly shown in Eq. (1), the driving force for the electroosmotic flow is the electrical body force term that is determined by both the applied electrical field strength and the net charge distribution in the EDL. Rigorous mathematical modeling of ion distribution in the EDL region under alternating
124
Electrokinetics in Microfluidics
electric fields should take into account of unsteady effects. Generally, the time scale related to electro-migration in the EDL is of the order of 10~8~10~7 sec [27]. This value is at least two orders of magnitude smaller than the characteristic time associated with the evolution of the electroosmotic flow, which is of the order of 10 5~ 10 3 sec [28]. If the frequency of the applied external electric field is not very high, the transient effect of the EDL "relaxation" can be safely neglected, and the ion distribution in the EDL may still be assumed to obey the equilibrium Boltzmann distribution. Consequently, the EDL potential distribution, y/(r) can be described by Poisson-Boltzmann equation that, in polar coordinates, takes the following non-dimensional form: (80) Eq. (80) is subject to the following boundary conditions (81)
However, no exact analytical solution to Eq. (80) is available because of the nonlinearity. However, the hyperbolic sine function sinhT may be approximated as
(82)
Such a treatment was successfully used by Philip and Wooding [29] in their study concerning the electrical potential distribution outside a charged cylindrical particle immersed in an electrolyte. They found that the obtained potential profile differs only slightly from that obtained by using numerical integration of the complete P-B equation.
Electroosmotic Flows in Microchannels
125
Based on Eq. (82), the entire cylindrical region can be considered as two hypothetical concentric regions (L and H) so that at their junctions (i.e., R = R ), we have XF = 1 as illustrated in Figure 4.11 [16]. The corresponding EDL fields can be described as the following: Low potential region L (0
) (83)
High potential region H {R
These two equations are subject to the coupling boundary conditions (85a)
Figure 4.11. Divided EDL regions in a cross-section of the cylindrical capillary for solving the Poisson-Boltzmann equation with high zeta potentials.
126
Electrokinetics in Microfluidics
(85b) For the case of Ka>\, solutions of Eq. (83) and Eq. (84) can be expressed as Eq.(86) and Eq.(87), respectively. (86) and
(87)
where (88a)
(88b) Io is the first kind, zero-order modified Bessel function. C is the integration constant, given by
(89)
Electroosmotic Flows in Microchannels
127
Figure 4.12 Comparison of the predicted dimensionless EDL potential, y/(r)/£ 0 versus dimensionless radius, rla, obtained from the Debye-Hiickel linear approximation, the analytical solution, and the numerical integration of the complete Poisson-Boltzmann equation in a cylindrical capillary for two cases: (1) Ka = 10, and (2) KO = 2 5 , with a fixed dimensionless zeta potential, ^ = 8.
4-4.3 Behaviors of the transient electroosmotic flow To predict the behavior of the transient electroosmotic flow by using the above-described equations, specific material properties must be chosen. For simplicity, let's consider an aqueous NaCl solution. At room temperature of T = 298K, the electrolyte properties are dielectric constant sr =80, viscosity li = 0.90 x 10"3 kg Ism, and density p = 998 kg/m3. Figure 4.12 shows the EDL potential profiles in a cylindrical capillary for two different electrokinetic diameters, Ka with a fixed non-dimensional zeta potential, VFJ. For comparison, it also includes the results obtained from the Debye-Hiickel linear approximation and the complete numerical method. To exclude the effect of zeta potential on the magnitude of the EDL potential, the EDL potential is normalized by an arbitrary reference potential, C,o. It is shown in Figure 4.12 that the Debye-Huckel linear assumption provides a good approximation for a larger electrokinetic diameter. However, the linearized
128
Figure 4.13.
Electrokinetics in Microfluidics
Dimensionless transient velocity, u(r)/usQ versus dimensionless radius, r / a . T h e
reference velocity is defined as us0 =-^^-E0 H
—. e
electrokinetic diameters (a) KQ = 10 and (b) Ka -100.
Time evolution of the velocity for two
Electroosmotic Flows in Microchannels
129
solution deviates significantly from the complete Poisson-Boltzmann equation for a smaller electrokinetic diameter, i.e., a thicker EDL corresponding to a dilute electrolyte or a smaller capillary. Furthermore, it is noted from Figure 4.12 that the difference between the analytical solution (Eq.(86) and Eq.(87)) and the numerical solution of the complete Poisson-Boltzmann equation is essentially identical. Time evolution of the electroosmotic flow field in cylindrical capillaries under a constant electric field is shown in Figures 4.13a and 4.13b. Here the reference velocity is chosen as us0 =—r—^-EQ— so that the zeta potential e H effect can be shown. As seen from these figures, upon the application of an electric field, the flow is activated in a region adjacent to the channel wall and the velocity increases rapidly from zero at the wall to a maximum within the EDL where the driving force is present. As the time goes, the flow gradually extends to the rest portion of the channel due to the viscous effect. When the flow in entire channel becomes steady, the flow velocity across the channel outside the EDL region remains essentially a constant value, resembling a plug flow. Furthermore, closely examining Eq.(74), we can find the time-dependent component in the velocity profile is given by the exponential function, exp(-A^Z)- Following an approach used for dynamic phenomena such as vibration and transient heat conduction (lumped heat capacity), we can assume the characteristic time is defined by the case that the exponential function decays to exp (-1). Therefore, the characteristic time scale for the electroosmotic flow to reach its steady state can be estimated from Eq. (74) by choosing
(90) where Xx = 2.405 determined form the zero-order Bessel function, J0(Xn) = 0 . Eq.(90) shows that the characteristic time, t * is proportional to the square of the channel radius, a1. For a capillary of a = 10 fim in radius, we can estimate the characteristic time for the electroosmotic flow to reach a steady state is t* = (pa2)/(/u X2)=\9.2/JS . The corresponding eigenfrequency of the system is f*
=\/t*
= JUAJ2 / p a 2
=
52.25KHz.
In Figure 4.14, the time dependent mean velocity is plotted for two-cycle periods of the AC electric field. The flow can follow the sinusoidally oscillation of the electric field, but there exists a phase lag. The phase lag is larger as the
130
Electrokinetics in Microfluidics
frequency goes higher. For the same reason as the maximum transient velocity, the magnitude of the maximum mean velocity is also dependent on the frequency. (91)
Figure 4.14.
Dimensionless mean velocity versus time with a fixed electrokinetic diameter,
ica = 32.57 and the zeta potential, x¥s = 4 . Reference velocity is chosen as us = - ^ ^ - £ 0 £ o . The system eigenfrequency is / * = 52.25KHz.
Electroosmotic Flows in Microchannels
131
4-5 AC ELECTROOSMOTIC FLOWS IN A RECTANGULAR MICROCHANNEL Electroosmotic flow induced by unsteady applied electric fields or electroosmotic flow under an alternating electrical (AC) field has its unique features and applications. Oddy et. al [30] proposed and experimentally demonstrated a series of schemes for enhanced species mixing in microfluidic devices using AC electric fields. In addition Oddy et. al. [30] also presented an analytical flow field model, based on a surface slip condition approach, for an axially applied (i.e. along the flow axis) AC electric field in an infinitely wide microchannel. Comprehensive models for such a slit channel have also been presented by Dutta and Beskok [31], who developed an analytical model for an applied sinusoidal electric field, and Soderman and Jonsson [12], who examined the transient flow field caused by a series of different pulse designs. As an alternative to traditional DC electroosmosis, a series of novel techniques have been developed to generate bulk flow using AC fields. For example Green et. al. [32] experimentally observed peak flow velocities on the order of hundreds of micrometers per second near a set of parallel electrodes subject to two AC fields, 180° out of phase with each other. The effect was subsequently modeled using a linear double layer analysis by Gonzalez et. al. [33]. Using a similar principal, both Brown et. al. [34] and Studer et. al. [35] presented microfluidic devices that incorporated arrays of non-uniformly sized embedded electrodes which, when subject to an AC field, were able to generate a bulk fluid motion. Also using embedded electrodes, Selvaganapathy et. al. [36] demonstrated what they termed fr-EOF or bubble free electroosmotic flow in which a creative periodic waveform was used to yield a net bulk flow while electrolytic bubble formation was theoretically eliminated. AC fields in microfluidic structures have also proven to be promising for dielectrophoretic separation of particles on Lab-OnChip devices [37] or as a method to enhance electroosmotic flow through ion exchange membranes [38]. This section reviews a combined theoretical and numerical approach that was taken to investigate the time periodic electroosmotic flow in a rectangular microchannel [39]. An analytical solution to the velocity field, based on a Green's function formulation, is presented for a linearized Poisson-Boltzmann double layer model and an applied sinusoidal electric field, while more complex cases (including different waveforms) are studied numerically using the BLOCS (Bio-Lab-On-a-Chip Simulation), a finite element simulation software developed by the Microfluidics Laboratory in the University of Toronto. The rectangular microchannel is more representative of the channel geometry encountered in actual microfluidic systems. The model and the analysis methods allow us to demonstrate the effects of, among other things, non-uniform ^-potential distribution and channel aspect ratio on the AC electroosmotic flow field.
132
Electrokinetics in Microfluidics
Figure 4.15 illustrates the time periodic electroosmotic flow in a straight, rectangular microchannel and the analytical domain. For generality the model and the solution will be developed to consider that the upper and lower surfaces can have a different ^-potential from that of the side walls (denoted by C,\ and £2, respectively). Assuming a fully developed flow field and considering the geometric symmetry, the analytical domain can be reduced to the upper left-hand quadrant of the channel cross section as shown in Figure 4.15.
Figure 4.15 Illustration of AC electroosmotic flow in a rectangular microchannel and the analytical domain.
Electroosmotic Flows in Microchannels
133
As explained previously, when a liquid comes into contact with a solid, the formation of an interfacial charge causes a rearrangement of the local free ions in the liquid and produces a thin region of non-zero net charge density near the interface, referred to as the electrical double layer (EDL). The application of an external electric field then results in a net body force on the free ions within the EDL inducing a bulk fluid motion called electroosmotic flow. For pure electroosmotic flows (i.e. absent of any pressure gradients) of incompressible liquids, the Navier-Stokes equations may take the following form, (92) where v is the flow velocity, / is time, pf is the fluid density, pe is the net charge density, fi is the fluid viscosity and E(a> t) is a general time-periodic function with a frequency co = 27tfand describes the applied electric field strength. As was observed experimentally by Oddy et. al. [30] at very high applied electric fields strengths (greater than 100 V/mm) the flow system may become unstable (chaotic) and thus this may be considered the upper limit on the applicability of Eq. (92). The electric double layer distribution, i//, in a rectangular microchannel can be described by the Poisson equation: (93) where e is the dielectric constant of the fluid medium. In the absence of a significant convective or electrophoretic disturbance to the double layer, the net charge density field can be described by a Boltzmann distribution, which takes the following form (assuming a symmetric electrolyte): (94) where z is the valence, e the charge of an electron, nm the bulk electrolyte concentration, &A the Boltzmann constant and T the temperature. Combining Eq. (93) and Eq. (94) and introducing the non-dimensional double layer potential, f = zey//kbT, and non-dimensional double layer thickness K = DhK (where Dh is the hydraulic diameter of the channel, Dh = 4lxly/2(lx+ly), and K is the Deybe-Hiickel parameter) yields the non-linear Poisson-Boltzmann distribution equation:
134
Electrokinetics in Microfluidics
(95) where the ~ signifies that the spatial variables in the gradient operator have been non-dimensionalized with respect to Dh (i.e. X = x/Dh, Y = y/Dh). Eq. (95) is subject to the symmetry conditions along the channel center axes, i.e.,
and the appropriate conditions at the channel walls,
Generally, for multidimensional space and complex geometry Eq. (95) must be solved numerically. In the interest of developing an analytical solution for the situation of interest here, Eq. (95) will be linearized using the DeybeHuckel approximation, yielding, (96) The Deybe-Hiickel approximation is considered valid only when zeQ
(97)
Electroosmotic Flows in Microchannels
135
where Xn and ]xn are the eigenvalues given by
and
respectively. Introducing the non-dimensional time,
and the non-dimensional frequency,
scaling the flow velocity by the electroosmotic slip velocity,
where Ez is a constant equivalent to the strength of the applied electric field, and combining Eq. (92) with Eq. (93) and Eq. (95) yield the non-dimensional flow equation, (98) where F is a general periodic function of unit magnitude such that E(f20) = EZF(Q9). Along the channel axes Eq. (98) is subject to a symmetry boundary conditions,
while no-slip conditions are applied along the solid channel walls,
136
Electrokinetics in Microfluidics
In order to obtain an analytical solution to Eq. (98), the Deybe-Hiickel approximation is implemented and F(Q0) is chosen to be a sinusoid function, resulting in the following form of the equation, (99) Using a Green's Function formulation, an analytical solution to Eq. (99), subject to the homogeneous boundary conditions discussed above, can be obtained. (100) where G(X,Y,9\X',Y',T) is the Green Function. For the finite domain used here the Green's function can be found from the eigenvectors of the homogeneous problem [40]:
(101)
where A/ and /um are the same eigenvalues as those for Eq. (97). Substituting Eq. (101) and Eq. (97) into Eq. (100) and solving for V(X,Y,9) yields:
(102)
where fi,™ and f2in are given by:
Electroosmotic Flows in Microchannels
137
Figure 4.16: Steady state time periodic electroosmotic velocity profiles in a square microchannel at (a) Q = 30 and (b) Q = 625 (equivalent to an applied electric field frequency of (a) 500hz and (b) 10 kHz in a 100u,m by 100(J.m square channel).
138
Electrokinetics in Microfluidics
(103a)
(103b)
Eq. (102) along with Eq. (103a) and Eq. (103b) represent the full solution to the transient flow problem. The equation can be somewhat simplified in cases where the quasi - steady state time periodic solution (i.e. after the influence of the initial conditions has dissipated) is of interest. As these initial effects are represented by the exponential term in Eq. (102), the quasi-steady state time periodic solution has the form shown below,
(104)
In the above analytical description of the uniaxial electroosmotic flow in a rectangular microchannel, the governing parameter is Q which represents the ratio of the diffusion time scale (tdiff = pjDh2//j) to the period of the applied electric field (tE = 1/co). Figure 4.16 compares the time-periodic velocity profiles (as computed from Eq. (104)) in the upper left hand quadrant of a square channel for two cases: (a) Q = 31 and (b) Q = 625. These two Q. values correspond to frequencies of 500hz and 10khz in a lOOum square channel or equivalently a 100am and a 450(im square channel at 500hz. To illustrate the essential features of the velocity profile a relatively large double layer thickness has been used, K = 3x106 m"1 (corresponding to a bulk ionic concentration nx = 10"6 M), and a uniform surface potential of £ = -25 mV was selected (within the bounds imposed by the Deybe-Hiickel linearization). For a discussion on the effects of
Electroosmotic Flows in Microchannels
139
Figure 4.17: Transient and steady state time periodic velocity of (a) channel midpoint and (b) double layer maxima at Q = 30. The dashed line represents transient solution, Eq. (102), the solid line represents steady state solution, Eq. (104), and the dotted line represents the scaled magnitude of applied electric field.
140
Electrokinetics in Microfluidics
double layer thickness the reader is referred to Dutta and Beskok [31]. From Figure 4.16, it is apparent that the application of the electrical body force results in a rapid acceleration of the fluid within the double layer. In the case where the diffusion time scale is much greater than the oscillation period (high Q, Figure 4.16(b)) there is insufficient time for fluid momentum to diffuse far into the bulk flow and thus while the fluid within the double layer oscillates rapidly the bulk fluid remains almost stationary. At Q = 31 there is more time for momentum diffusion from the double layer, however the bulk fluid still lags behind the flow in the double layer (this out of phase behavior will be discussed shortly). Extrapolating from these results, when Q < 1, such that momentum diffusion is faster than the period of oscillation, the plug type velocity profile characteristic of steady state electroosmotic flow would be expected at all times. Another interesting feature of the velocity profiles shown in Figure 4.16 is the local velocity maximum observed near the corner (most clearly visible in the Q = 625 case at t = 7t/2co and t = 7t/(o). The intersection of the two walls results in a region of double layer overlap and thus an increased net charge density over that region in the double layer. This peak in the net charge density corresponds to a larger electrical body force and hence a local maximum electroosmotic flow velocity, it also increases the ratio of the electrical body force to the viscous retardation allowing it to respond more rapidly to changes in the applied electric field. The finite time required for momentum diffusion will inevitably result in some degree of phase shift between the applied electric field and the flow response in the channel. From Figure 4.16, however, it is apparent that within the limit of Q > 1, this phase shift is significantly different in the double layer region than in the bulk flow. Figure 4.17 (Q = 30) and Figure 4.18 (D = 625) illustrate this steady state phase shift at two points (a) the channel mid point, showing the characteristic of the bulk liquid motion and (b) at the velocity maximum near the channel corner, illustrating the characteristic of the flow velocity in the double layer. Also shown in Figures 4.17 and 4.18 is the fully transient flow solution given by Eq. (102) (the dashed line). From Figure 4.17 it is apparent that the response of the fluid within the double layer is essentially immediate, however, the bulk liquid lags behind the applied field by a phase shift of the order of 7i/4. Additionally while the velocity in the double layer reaches its steady state oscillation almost immediately, the bulk flow requires a period before the transient effects are dissipated. In Eq. (104) the out-of-phase cosine term (Q cos(Q6 )) is proportionally scaled by Q, thus as expected when Q is increased in Figure 4.18, the phase shift for both the double layer and bulk flow velocities is increased as is the number of cycles required to reach the steady state. Although the magnitude of the velocity at the channel midpoint is significantly decreased, to approximately 1% of that for the Q. - 30 case,
Electroosmotic Flows in Microchannels
141
Figure 4.18: Transient and steady state velocity of (a) channel midpoint and (b) double layer maxima at Q = 625. The dashed line represents transient solution, Eq. (102); the solid line represents steady state solution, Eq. (104); and the dotted line represents the scaled magnitude of the applied electric field.
142
Electrokinetics in Microfluidics
it is interesting to note the net positive velocity at the channel midpoint within the transient period before decaying into the steady state behavior. This is a result of the initial positive impulse given to the system when the electric field is first applied and is reflected by the exponential term in Eq. (102). The transient oscillations were observed to decay at an exponential rate, as expected from this transient term in Eq.(102). Similar to the out-of-phase cosine term, this exponential term is also proportionally scaled by the non-dimensional frequency, suggesting that the effect of the initial impulse becomes more significant with increasing Q. Referring back to Figure 4.17a this effect can also be observed at the peak of the first oscillation, the transient solution velocity (dashed line) is slightly higher than that the time periodic (solid line) velocity. However since Q is much lower in this case the decay of this initial impulse is significantly faster. 4-5.1 Effect of channel aspect ratio While the square geometry is of significant theoretical interest, in most practical cases microchannels are significantly wider than deep (as a consequence of the micromachining techniques). Therefore it is of interest to examine the influence of the channel aspect ratio on the velocity profile. Figure 4.19 shows the time periodic velocity profile, from Eq. (104), in the upper right quadrant, at t = n/co for an aspect ratio (width:depth) of 2:1 and 4:1 at the same oscillation frequencies as those shown in Figure 4.16. As can be seen within the corner region, the velocity profiles are very similar to those presented in Figure 4.16. However at lower values of X (i.e., nearer the middle of the channel) the influence of the active flow in the double layer region (near the channel side wall) is reduced and the magnitude of the velocity is decreased, particularly prevalent in the low frequency case. Extending this trend to the infinitely wide slit channel case where Dh,sm = 2Dh,sqUare and thus QsUt = 4Qsquare^ it can be concluded that the flow response in a slit channel will be comparable to that in an equivalent square channel at four times the applied frequency or with double the channel dimensions. 4-5.2 Influence of the magnitude and distribution of the zeta potential As discussed above an inherent assumption in the analytical model, Eq.(102), is the electrical double layer field linerization via the Deybe-Huckel approximation. This approximation is applicable to cases where |Q<25mV, and is somewhat restrictive since most practical applications of electroomostic flow occur outside of this range. Thus it is desirable to examine both the influence of the ^-potential on the flow field and the inherent error in the linearized model by comparing it with a numerical solution to the exact, non-linear governing
Electroosmotic Flows in Microchannels
143
Figure 4.19. Steady state time periodic electroomostic velocity profiles at t = TZ/CO in a microchannel with (a) 2:1 aspect ratio and Q = 30, (b) 4:1 aspect ratio and Q = 30, (c) 2:1 aspect ratio and Q = 625 and (d) 4:1 aspect ratio and Q = 625.
144
Electrokinetics in Microfluidics
equations, Eq. (95) and Eq. (98). The set of these equations was solved by using the BLOCS (Bio-Lab-On-a-Chip Simulation) finite element simulation software (For details on the numerical technique the reader is referred to references [41~43]). The numerical simulation results of the time periodic velocity profiles along the Y= 0 edge are compared with that of the linearized solution (Eq.(102)) in Figure 4.20 for the (a) £ = -25mV and (b) C, = -lOOmV for both the Q = 31 and Q. = 625 cases. By comparing these figures it is apparent that increasing the ^"-potential increases the magnitude of the velocity, as expected, however it has little effect on the flow structure, as both cases yield essentially identical velocity profiles. In each case it is also apparent that the linearized analytical solution tends to slightly underestimate the actual flow velocity. An examination of the cause of this error revealed that the largest contributor was the sinhCP) « T approximation in the body force term of the Stokes equation, Eq. (99), whereas the error induced by the linear double layer equation, Eq. (96), had a much smaller effect on the computed flow field. However, the magnitude of the error for all cases is less than 2%, suggesting the linearized analytical solution is sufficiently accurate for the flow field in most practical cases. Many microfluidic channels have non-uniform surface properties either as a result of the fabrication process (for example micro-devices fabricated in PDMS devices are commonly bonded to a glass microscope slide to enclose the channel) or by heterogeneous adsorption of transported species/contaminants. It is of interest to examine how such a non-uniform ^-potential distribution will affect the flow field. Figure 4.21 shows the time periodic velocity profiles at t = n/2co, as computed using the analytical solution, for the Q. = 31 and Q. = 625 cases with a neutral surface {C, = OmV) (a and b) and a positive surface charge {C, = +25mV) (c and d) along the Y = Ly/2 surface, while the ^-potential along the X = LJ2 surface was maintained at C, = -25mV. Along the zero charged Y = Ly/2 surface there is a zero net charge density in the nearby liquid (and thus no net body force applied to the fluid continua), the liquid remains stationary. This has a significant influence on the bulk flow field in the region near this surface but the effect tended to diminish as the distance from this wall increased. Consequently, the maximum velocity occurs along the Y = 0 edge as opposed to the corner as seen for the homogeneous case. Similarly for the positive surface charge case the negative net charge density in the double layer induced a net flow opposite to that along the negatively charged surface, resulting in zero net volume flow rate at all times. For both cases it is interesting to note that while the surface heterogeneity does significantly affect the bulk velocity profile, the velocity profile near the X = LJ2 surface {C, = -25mV) only exhibits a significant change in the corner region.
Electroosmotic Flows in Microchannels
145
Figure 4.20. Steady state time periodic electroosmotic velocity profile along the Y= 0 edge for (a) C, = -25mV and (b) £ = -lOOmV. Dashed line represents numerical solution to non-linear Poisson-Boltzmann distribution and solid line represents Green's function solution.
146
Electrokinetics in Microfluidics
Figure 4.21. Steady state electroosmotic velocity profiles at t = JI/2CO in a square microchannel with (a) C,(Y = Ly/2) = OmV and D. = 30, (b) £,(Y = Ly/2) = OmV and Q = 625, (c) C,(Y= Ly/2) = +25mV and Ci = 30, (d) C,(Y = Ly/2) = +25mV and O = 625. In all cases
£(X=LJ/2)=-25mV.
4-5.3 Different excitation waveforms The results presented above are limited to sinusoidal waveforms. However it is of interest to examine how the flow field will respond to different forms of periodic excitation, such as a square (step) or triangular waveforms as defined by Eq. (105) and Eq. (106) respectively. (105)
Electroosmotic Flows in Microchannels
147
Figure 4.22. Steady state time periodic electroosmotic velocity profiles for different applied electric field waveforms along the 7 = 0 edge for (a) 13 = 30 and (b) D. = 625. Solid line represents sin wave, long dashed line represents triangular wave and doted line represents square wave.
148
Electrokinetics in Microfluidics
(106)
Figure 4.22 compares the time periodic velocity profiles, along the 7 = 0 axis, for the three waveforms (sin, square and triangular) at (a) Q = 31 and (b) Q. = 625 in a square channel by numerical solution to the exact, non-linear governing equations, Eq. (95) and Eq. (98). As can be seen in all cases, the square waveform yields higher local velocities owing to the fact that the full strength of the electric field is applied for a longer time. It is apparent from these figures that the particular waveform has a greater effect on the flow field at lower D (i.e. at lower frequencies or smaller channel dimensions). A lower Q corresponds to a smaller diffusion time scale in comparison with the excitation period and thus the velocity in the bulk flow is more representative of the instantaneous magnitude of the applied field. Figure 4.23 shows that the particular waveform also has a significant effect on the transient response of the bulk fluid (again the channel mid point is chosen as a representative point). As can be seen in Figure 4.23a for the Q. = 31 case, a square wave excitation tends to produce higher velocities whereas the triangular wave exhibits slightly smaller bulk velocities when compared with the sinusoidal waveform. As Q is increased, the initial positive impulsive velocity seen earlier in Figure 4.18a is again observed for both additional waveforms as shown in Figure 4.23b. As expected the fluid excited by the square waveform exhibits higher instantaneous velocities, which lead to an increase in the number of cycles required to reach the time periodic quasi-steady state oscillation. In summary, the time periodic electroosmotic velocity profile is governed by a non-dimensional parameter Q=pfDh2co/(i which represents the ratio of the time scale for viscous diffusion to the period of oscillation. In cases of high Q (i.e. high excitation frequency or large channel sizes) the flow is confined to a region near the channel wall while the bulk fluid remains essentially stationary. Both the channel's ^-potential distribution and aspect ratio are shown to have significant effect on the velocity profile. In uniform channels however the C,potential serves only to scale the magnitude of the velocity and does not significantly affect the flow structure. The impulsively started flows (from rest) are also shown to exhibit transient behaviour resulting in a net positive flow
Electroosmotic Flows in Microchannels
149
Figure 4.23. Transient stage velocity at channel midpoint for impulsively started flows using different waveforms at (a) Q = 30 and (b) Q = 625. Solid line represents sin wave, long dashed line represents triangular wave and doted line represents square wave.
150
Electrokinetics in Microfluidics
during the initial cycles for cases of high Q. Additionally the number of cycles required reaching a steady state oscillation increases with Q. At higher D. the excitation waveform (sinusoidal, square or triangular) has little effect on the velocity profile. However as D, approaches unity, viscous diffusion is sufficiently fast to allow the bulk flow to respond to instantaneous changes in the applied electric field, the influence of a particular waveform is more significant.
Electroosmotic Flows in Microchannels
4-6
151
ELECTROOSMOTIC FLOW WITH ONE SOLUTION DISPLACING ANOTHER SOLUTION
Most studies of electroosmotic flow have been focused on the flow of a uniform solution [7,38,44-49]. However, the operation of some lab-on-a-chip devices such as on-a-chip biosensors involves replacing one solution by another solution in a microchannel by electroosmotic pumping. For example, the operation of a fiber-optical nucleic acid biosensor device reported by Bier et al. [50] involves the following procedure: The sensor situated in a continuous flow apparatus is exposed to the target nucleic acid in a hybridization buffer for 60180s, followed by a 120-180s wash with buffer, 45-60s treatment with a solution of fluorochrome, and then 30s wash with buffer. In order to develop such a biosensor on a chip using electroosmotic pumping, we must control the above- mentioned solution replacing processes. In such a solution replacing process, if the two solutions are very different in terms of the ionic type, valence and concentration, the zeta potentials and the EDL fields will be different in the different sections of the microchannel. This will bring more complication to the electroosmotic flow in microchannels. This section will discuss the solution displacing process in a cylindrical capillary under an applied electrical field [10]. Consider the following electroosmotic flow. There are two reservoirs containing two different electrolyte solutions, respectively. The ionic concentration in reservoir 1 is Q . The ionic concentration in reservoir 2 is C2. A cylindrical capillary tube is used to connect the two reservoirs. Initially, the capillary tube is filled with solution 1 that has a concentration C/. Immediately after the two solutions are in contact in the capillary tube, an electrical field is applied along the capillary. The applied electrical field results in an electroosmotic flow in the capillary tube. During the electroosmosis, the solution 2 with a concentration C2 in reservoir 2 gradually displaces the solution 1 in the capillary tube. Because of the difference in chemical potentials between the two solutions, the diffusion takes place across the interface between these two solutions, meanwhile, the convection also occurs due to the movement of the liquid. As a result, the electrolyte concentration and the mixing zone length in capillary changes with time. Therefore, in the displacing process, the capillary can be divided into three sections according to the concentration distribution. The first section is filled with one solution, the second section is the mixing zone and the third section is filled with another solution, as shown in Figure 4.24. During the replacing process, the electrical resistance of the capillary depends on the type of ions and the ionic concentration distribution in the capillary and varies with time. The electrical current, which is generated by the excess ions' motion under the applied electrical field, is axially uniform and varies with time. Consequently, the electrical field strength along the capillary (volts per meter) is different from one section to another section.
152
Electrokinetics in Microfluidics
Figure 4.24 (Top) Illustration of one solution replacing another solution in a capillary during electro-osmotic flow. (Bottom) Schematic diagram illustrating the electrical resistance along the capillary.
4-6.1 Electrical double layer field Consider a cylindrical capillary with a circular cross section. According to the theory of electrostatics, the relationship between the electrical potential, i//(r), and the net charge density per unit volume, pe, at any point in the liquid is described by the Poisson equation, (107) where £ is the dielectric constant of the solution and s0 is the permittivity of vacuum. Assuming that the equilibrium Boltzmann distribution is applicable, the ion number concentration per unit volume in an electrolyte solution is of the form (108) where nix and zt are the bulk ionic concentration and the valence of type i ion, respectively, e is the charge of a proton, kb is Boltzmann constant and T is the
Electroosmotic Flows in Microchannels
153
temperature. The net volumetric charge density pe is proportional to the concentration difference between cations and anions, given by: (109) For a symmetric electrolyte solution such as KCl (z: z = 1:1) solution, the above equation becomes: (110) For a non-symmetric electrolyte solution such as LaCl3 (z: z * 1:1), it takes the following form: (111) Substituting the equation of the net charge density into the Poisson equation, Eq.(107), and introducing the dimensionless variables
where d is the diameter of the capillary tube, the non-dimensional PoissonBoltzmann equation can be written as follows, for KCl solution: (112) For LaCl3 solution: (113)
154
Electrokinetics in Microfluidics
Because of the symmetry of the EDL field in the cylindrical capillary, Eq. (112) and Eq. (113) are subjected to the following non-dimensional boundary conditions:
where £ is the zeta potential. It should be noted that the zeta potential varies with the ionic concentration of the solution, and hence it has different values for different sections. Such a variation of the zeta potential is considered here and the zeta potential values for different concentrations are listed below in the discussion of the numerical simulation. 4-6.2 Electroosmotic flow field As described before, there are three sections in this capillary. As shown in Figure 4.24, the solution on the left side (entering the capillary) has an ionic concentration, C2, and the solution on the right side (leaving the capillary) has an ionic concentration, Cj. If we assume that the flow in these two sections is one dimensional and fully developed, then the equation of motion is given by: (114) We consider that the pressures at the both end of the capillary are the same (e.g. atmospheric pressure). For a simple electroosmotic flow in a capillary with open ends, there is no pressure gradient along the capillary. However, for the case of electroosmotic flow with one solution replacing another solution, the driving force of the electroosmotic flow, the net charge density and the applied electrical field strength are different in different sections. This would imply different velocity fields and different flow rates in different sections. For incompressible liquids, however, the continuity condition requires a constant volume flow rate throughout the channel. Therefore, an induced pressure gradient along the channel is required to satisfy the continuity condition. Substituting the Eq. (110) and Eq. (111) for the net charge density into the Eq.(l 14) and introducing the following non-dimensional variables,
Electroosmotic Flows in Microchannels
155
where D is the diffusion coefficient, the non-dimensional equation of motion can be obtained, for a KC1 solution: (115)
for a LaCl3 solution: (116)
The above two equations are subjected to the no-slip and symmetric boundary conditions:
4-6.3 Concentration field The concentration distribution in the capillary is governed by the mass conservation law, which for this case takes the form: (117) where C is the bulk ionic concentration in capillary tube, uz is the average electroosmosis velocity of the liquid which is determined by the equation of motion. Introducing the following non-dimensional parameters:
where z is the coordinate variable in the flow direction as shown in Figure 4.24. The equation of concentration can be non-dimensionalized as follows:
156
Electrokinetics in Microfluidics
(118)
Since the average velocity is constant axially at the given time, the above equation can be further reduced to: (119)
The above equation is subjected to the following initial and boundary conditions
where Ltotal is the total length of the capillary and Lx is the length of the section with the to-be-replaced electrolyte solution 1 (C;). The capillary tube in this study has the following dimensions: Ltotal=\Q cm and d= lx 10 2 cm. Now we have the complete set of equations, Poisson-Boltzmann equation, motion equation and concentration equation and the matching initial and boundary conditions. In this model, the mixing zone length is defined as follows: The position where C = 99.99%C2 will be the left boundary of the mixing zone. The position where C - 99.99%Cj will be the right boundary of the mixing zone. It is obvious that the length of the mixing zone and the lengths of the other two sections change with time. Consequently, according to the equation of electrical resistance, (120) where Lt is the length of the ith section, Cr is the concentration of the ith section, Xt is the bulk conductivity of the ith section and A is the cross section
Electroosmotic Flows in Microchannels
157
area of the cylindrical capillary. The overall electrical resistance of the liquid in the capillary tube will change with time during the electroosmosis. At a given time, the total resistance is the sum of the resistance of the three sections, as shown in Figure 4.24, given by: (121) where i?j is the electrical resistance of the section with the electrolyte solution 1; R2 is the electrical resistance of the section with the electrolyte solution 2 , and Rmix is the electrical resistance of the mixing zone. When the total resistance and the electrical voltage applied to the capillary are known, the electrical current through the capillary can be determined by: (122) where Vtotal is the total electrical voltage applied to the capillary, which is constant during the replacing process. As the current is axially uniform and the electrical resistance varies from section to section (see Eq. (120)), the electrical field strength for each section will be different and can be determined bv (123) Using the above-defined equations, one can determine the concentration distribution, the electrical field strength for each section and the mixing zone length. First, at a given time, a guess value for concentration profile is chosen. With the guessed concentration values, the zeta potential for each section can be determined by experimental data. The equation of electrical potential, Eq. (112) or Eq. (113), in turn, can be numerically solved to find the EDL potential y/t (r). Once the EDL field is known, the local net charge density pei{r) can be determined. Meanwhile, using the guess value for the concentration profile, the resistance for each section can be determined by Eq. (120), which will be used to determine the total current and the electrical field strength for each section. Then the equation of motion, Eq. (115) or Eq. (116), can be numerical solved with a guess value of the induced pressure gradient for one of the non-mixing zone section. At this time, the continuity condition has to be satisfied by adjusting the pressure gradient. Once the velocity profile is determined, the equation of concentration can be solved to obtain the concentration distribution in capillary. Repeat this iteration procedure until the convergence for concentration field is reached. Finally, with the results obtained in this time step, the above procedure
158
Electrokinetics in Microfluidics
Figure 4.25. Comparison of the experimentally determined current ~ time relationship with the model prediction, (a) a lxlO" 2 M KC1 solution replacing a lxlO" 4 M KC1 solution, and (b) a lxlO~4M KC1 solution replacing a lxl(T 2 M KC1 solution, under an electrical field of 350 V/cm.
Electroosmotic Flows in Microchannels
159
is repeated for the next time step until the solution 2 from the reservoir 2 completely replaces the solution 1 in the capillary tube (i.e., the left boundary of the mixing zone reaches the exit of the capillary tube). In this model, the equation of concentration is a diffusion and convection equation. The finite control volume method is employed to solve this equation and to ensure that the mass conservation is satisfied in the computation domain [51]. Eq. (119) was numerically solved for every control volume using alternative direction implicit method to deal with the unsteady term. In addition, varying-step grid is used to consider the sharp change in electrical potential i// and the velocity u within a small distance from the channel wall. This is achieved by employing the grid-clustering scheme [52]. 4-6.4 Experiments of the solution displacing processes Several such solution displacing experiments were conducted [10]. In the experiments, Polyamide-coated silica capillary tubes of 100 um internal diameter (Polyamide Technologies Inc., Phoenix, AZ) were cut to 10 cm in length and used to connect the electrolyte solutions in reservoir 1 and reservoir 2. In the experiment, reservoir 2 was filled with solution 2. The capillary tube and reservoir 1 were filled with solution 1. All the solutions were prepared by using Deionized ultrafiltered water (Fisher Scientific, Ontario), KC1 (Anachemia Science, Quebec), and LaCl3 (Fisher Scientific Company, New Jersey). Immediately after connecting the two reservoirs by the capillary tube, a voltage difference between the two reservoirs was applied by setting reservoir 1 at a ground potential and reservoir 2 to a higher voltage (HV power source: CZE 1000 R, Spellman, NY) via platinum electrodes. The applied electrical field results in an electro-osmotic flow in the capillary tube. The electrolyte solution from reservoir 2 gradually displaces the electrolyte solution in the capillary tube. As a result, the overall electrical resistance of the liquid in the capillary tube changes, A PGA-DAS 08 data acquisition chip (OMEGA Engineering, Quebec) was used to record the voltage (KV) and current (|aA) as a function of time (second). Once the solution in the capillary tube is completely replaced by the solution from the reservoir 2, the current reaches a constant value. The measured time for the current to reach such a plateau value is the time required for the solution from reservoir 2 to travel through the entire capillary tube. This may provide an indication of the average velocity by (124) where L is the length of the capillary and At is the time required for the electrolyte solution 2 to completely displace the electrolyte solution 1. In the
160
Electrokinetics in Microfluidics
Figure 4.26. Comparison of the experimentally determined current ~ time relationship with the model prediction, (a) a lxlCT2 M KCl solution replacing a lxlO" 4 M KCl solution, and (b) a lxl0" 2 M KCl solution replacing DIUF Water, under an electrical field of350V/cm.
Electroosmotic Flows in Microchannels
161
experimental studies, each measurement was repeated at least five times for a given set of conditions. All experiments were conducted at a room temperature (25°C). 4-6.5 Analysis of the displacing process Because the electrical current depends on the electrical resistance of the solution, which in turn depends on the concentration and the conductivity, when the concentration in the capillary changes, the current will change. The electrical current in the capillary will increase with time as an electrolyte solution with high conductivity (i.e. X w-2 M KCI =14.12 xlO~2 S/m) gradually replaces a solution with low conductivity (i.e. A io-4 M Kci = 15.14 xlO~4 S/m), and vice versa. The increase (or the decrease) in current continues until the capillary is completely filled by the solution with high (or low) conductivity. Thereafter the current reaches a constant value. Using the aforementioned numerical methods, the complete set of non-linear differential equations, governing the diffusion and convection processes of the two solutions in the capillary, were solved to predict the mixing process. As shown in Figures 4.25-4.28, the good agreements were found between the experimentally measured current-time relationships and the numerical model predictions. In this model, the properties of the electrolyte solution in the mixing zone are obtained by using linear interpolation of the properties of the two uniform solutions. Such interpolation may result in the small discrepancies between the experimental results and the numerical predictions as seen in these figures. The change of current depends on the concentration and the conductivity of the solutions in two reservoirs. As seen in Figure 4.26, when the electrolyte solution with high conductivity (i.e. X iO-2M Kci =0.1412 S/m, C, w-2 M KCI = -60 mV) replaces the electrolyte solution with low conductivity (i.e. X DIUF water = 0.7269 xlO~4 S/m, £ DIUF water = -240 mV), the current increases dramatically immediately before the current reaches the constant value as compared to the gradually changing in the case of a 10"2 M KCI solution replacing a 10"4 M KCI solution. This is because the conductivity of DIUF Water is very low as compared to the conductivity of the 10"2M KCI solution. As long as the DIUF Water is still in the capillary, it dominates the resistance and the corresponding current is very low, on the other hand, immediately when the DIUF Water leaves the capillary, the 10"2 M KCI solution dominates the resistance in capillary and the current increases dramatically. The electrical current also depends on the electrical field strength applied to the capillary. The higher the electrical field strength, the higher the current. As shown in Figure 4.27, when the electrical field strength increases (i.e. 350 V/cm in Figure 4.25 and 700 V/cm in Figure 4.26), the increase of current can be seen. In addition, the time required for the electrolyte solution from reservoir 2 to
162
Electrokinetics in Microfluidics
Figure 4.27. Experimentally determined current ~ time relationship for the displacing processes between a lxlO" 2 M KC1 solution and a lxlO" 4 M KC1 solution, (a) under an electrical field of 350V/cm, and (b) under an electrical field of 700 V/cm.
Electroosmotic Flows in Microchannels
163
completely displace the solution 1 in capillary tube depends on the electrical field strength too. When the electrical field strength increases, the required time period is shorter. For the case of high electrical field strength, the required time is about 20.75 seconds and for the case of low electrical field strength, the required time is around 40.2 seconds. From this time period, the average velocity for one solution replaces another solution can be estimated by Eq.(124), for example, the average velocity is 4.819 mm/s for the case of high electrical field strength and 2.488 mm/s for the case with low electrical field strength. The velocity profiles during the replacing process are presented in Figure 4.29 and Figure 4.30. A distorted electroosmotic velocity profile and a similar parabolic velocity profile are founded for different sections. Generally, if the electrolyte solution in capillary is uniform and only the electrical field for driving flow is applied to the capillary, the electrical field strength and the electroosmotic velocity profile are same for different locations. However, during the mixing process, the current is uniform axially and the conductivity is different from location to location due to the different concentration distributions. Consequently, according to the Eq. (123), the electrical field strength is non-uniform along the capillary, which would imply the different volumetric flow rates for each section. It is well known that, for the incompressible fluid, the continuity condition requires the uniform volume flow rates {Qi=Q2) in the fluid flow. This implies that there is an induced pressure gradient in the liquid to ensure the constant flow rate in different sections during the replacing process. For example, in Figure 4.29, the LaCl3 solution on the left-hand side has a higher conductivity {X iO-4 M La ci3 = 40.2 x 10~4 S/m), a lower zeta potential (£ iO-4 M Laci3 = -39.2 mV) and a smaller electrical field strength. Hence the electroosmosis effect is weak and the electroosmotic velocity would be small. In order to achieve the same flow rate as the downstream solution with a higher electroosmotic velocity, a negative pressure gradient is generated to increase the flow. Recall that for this system the pressure at the both ends of the capillary tube is atmospheric pressure, a negative pressure gradient implies that the pressure at the end of the LaCl3 solution section is lower than the atmospheric pressure. Since the electroosmotic effect in this section is so small, the negative pressure gradient becomes the dominant force to generate the flow. That is why the velocity profile in this section is approximately parabolic. For the KC1 solution on the right-hand side of the capillary tube, the conductivity (X 10-4 M Kci =15.14 x 10"4 S/m), the zeta potential (£ io-4 MKCI = -107 mV) and the electrical field strength are relatively higher. The electroosmotic effect and hence the electroosmotic velocity is larger. However, because of the presence of a low pressure at the interface region between the LaCl3 solution and the KCI solution,
164
Electrokinetics in Microfluidics
Figure 4.28. Comparison of the experimentally determined current ~ time relationship with the model prediction, (a) a lxlO"4 M LaCl3 solution replacing a lxlO"4 M KCl solution, and (b) a lxlO"4 M KCl solution replacing a lxl0~4M LaCl3 solution, under an electrical field of 350V/cm.
Electroosmotic Flows in Microchannels
165
Figure 4.29. Simulation of the velocity field (top) and the concentration fields (bottom) in the capillary for a lxlO~4M LaCb solution replacing a lxlO~4M KC1 solution at 36 seconds under an electrical field of 350V/cm
Figure 4.30. Simulation of the velocity field (top) and the concentration fields (bottom) in the capillary for a lxlO"4 M KCl solution replacing a lxlCT4 M LaCU solution at 36 seconds under an electrical field of 350V/cm.
166
Electrokinetics in Microfluidics
a positive pressure gradient exists for the KC1 solution section. This positive pressure gradient tends to reduce the flow in this section and results in a velocity profile as shown in Figure 4.29. These velocity profiles agree with the results presented by Herr et al. [53]. The model developed in this section also reveals the mixing process. In Figures 4.29 and 4.30, the concentration fields for each electrolyte are plotted below the corresponding velocity field at the same time. The mixing zone can be identified according to the concentration distribution in the capillary. The length of the mixing zone increases with time after the two solutions are brought into contact. This is because of the diffusion caused by the concentration difference across the interface.
Electroosmotic Flows in Microchannels
4-7
167
ANALYSIS OF THE DISPLACING PROCESS BETWEEN TWO ELECTROLYTE SOLUTIONS
As discussed in the last section, in an electroosmotic solution displacement process, if the two solutions are very different in terms of the ionic type, valence and concentration, the zeta potential of the microchannel wall and the EDL fields will be different in the different sections of the microchannel. Such a displacement process is more complicated than the simple electroosmotic flow of a uniform solution in microchannels. It is desirable to obtain a better understanding and more information of an electroosmotic solution displacement process, such as the location of the interface between these two solutions, which is very important in operating biochips. For this purpose, theoretical models were developed to predict the interface position [11]. This section will use these models to interpret some experiment data of the current-time relationship in the displacing processes. To model such an electroosmotic flow process, let's look at the electrical current in the capillary during the electroosmsis in detail. The local electric current vector in a microchannel is given by: (125)
where i and u are local electric current density vector and local velocity vector, respectively, y/ is the local electrical potential. The jth ionic species has a valence of z , a diffusion coefficient of Dj and an ionic number concentration of rij. Here, e,&^and T represent the fundamental elementary charge, Boltzmann constant and the system absolute temperature, respectively. When an electric field is applied across the capillary, in the absence of end effects, the current is along the axial direction. The total current is given by (126) where A is the cross-sectional area of the capillary. For fully developed flow, one can neglect the second term of Eq.(126) which defines the current flow due to migration. For large K: a capillary, where the electric double layer is very thin, electroneutrality is assumed to dominate in the channel and the first term of
168
Electrokinetics in Microfluidics
Figure 4.31 Schematic diagram of one solution displacing another solution in a capillary where (a) a sharp interface between the two solutions is assumed, and (b) a mixing zone between the two solutions is considered.
Electroosmotic Flows in Microchannels
169
Eq.(126) defining the current transport due to convection can be neglected. Here a is the capillary radius and K~1 is the Debye double layer thickness. Consequently, for large K: a flows, one can write the electric current as: (127) where E is the electric field strength in x- direction and njm is the ionic number concentration for an electrolyte solution. For a given electrolyte solution, Eq.(127) can be written as (128) where X is the electric conductivity of the electrolyte solution. Eq.(128) is an expression of Ohm's law for electrically neutral dilute solutions or solutions in capillary having large values of K a with negligible axial concentration gradients [54]. We will make use of Eq.(128) to derive the following models to evaluate the interface position in the microchannel with time. 4-7.1 Model I In this model, we assume that there is a sharp interface between the two solutions during the displacement process. The schematic diagram of the solution displacement process is shown in Figure 4.31a. The total length of the capillary is L and the traveling distance of the liquid-liquid interface is x at a given time. In the present configuration, we consider that the electric double layer thickness is very small, and the aspect ratio of the capillary length to diameter is very large. Consequently, we can safely assume the applicability of Eq.(128) to our system. The total resistance of this capillary is assumed to be made up of two resistances, R} and R2, corresponding to the two solutions. That is (129) According to Ohm's law, the electric resistances are given by (130a)
(130b)
170
Electrokinetics in Microfluidics
where X^ is the electric conductivity of the ith electrolyte solution. The current, in turn, can be determined by (131)
From Eq.(131), the traveling distance of the interface can be obtained as (132)
Eq.(132) shows that once the current is known, the traveling distance, x, can be determined since the total applied voltage, the conductivity of the individual solution and the total length of the capillary are constant through the displacement process. However, a detailed analysis [10] shows that there exists a mixing zone between the two solutions, instead of a sharp liquid-liquid interface assumed in Model I. In the mixing zone, the ionic concentration and the electric conductivity vary between the bulk solutions. The size of the mixing zone depends on the diffusion coefficients of the electrolytes in the two solutions, the concentration difference, the speed of the flow and the time. Therefore, in order to make the analysis of the displacement process closer to the real system, the displacement process with the consideration of the mixing zone is analyzed below. 4-7.2 Modelll In this model, we consider a situation where one solution displaces another solution and there exists a mixing zone between the two solutions during the displacement process. The two solutions contain the same electrolyte or different electrolytes, and have different concentrations, Cx and C 2 . The electroosmotic flow is from left to right as shown in Figure 4.31b. The mixing zone, with a length of Lmix, is defined as a region where the left boundary of the mixing zone has a concentration of C2 +0.01%(Cj - C 2 ) and the right boundary of the mixing zone has a concentration of C 2 + 99.99%(C1 - C 2 ). Assuming that x is the position of the left boundary of the mixing zone at a given time, the length of the channel filled with the electrolyte solution 1 is L - x - Lmix. In this model, it
Electroosmotic Flows in Microchannels
171
is assumed that the total resistance is made up of three components corresponding to the three regions and it takes the form of (133) where Rmix is the resistance of the mixing zone. Now assume that the mixing zone is divided into N small sections and the length of each section, Lmix(i), is so small that the conductivity of the liquid, Xmix(i), in this section, can be considered as a constant. Therefore, the total resistance of the mixing zone is a summation of the resistance of each section, Rmix =Y.Rmix(i), where Rmix(i) is the resistance of the /^section of the mixing zone. According to Ohm's law, the resistances can be written as (134a)
(134b)
(134c) The current, in turn, can be given by (135)
Taking the derivative of current with respect to time, we obtain (136) where ux is the cross-sectional average velocity, given by ux = dx/dt. As explained earlier, Lmix(i) is dependent on the concentration distribution in the capillary. The concentration distribution during the displacement process is governed by:
172
Electrokinetics in Microfluidics
(137) From the order of magnitude analysis of the convection and diffusion
\c]
\c]
terms in Eq.(137), [w-J^r and [^]r—n, we found that the order of magnitude of x2\ ix\ the convection term is approximately 10~2, and that of the diffusion term is x~l0~l,D~l0~9). Therefore, the approximately 10"7 (C~\^U~10~3 contribution of the diffusion to the concentration distribution can be neglected. Consequently, the length of the ith section of the mixing zone, Lmix(i), is determined by the initial condition and the convection term. The ith section of the mixing zone has two boundaries, x{f) and x(i + l), the length of the ith section can be calculated by Lmjx(i) = x(i + \)-x(i). Taking a derivative of Z,mfcc(/)with respect to time, we obtain (138) Because of the continuity condition, at a given time, the cross^sectional average velocity is constant through the whole capillary, i.e., ux{i) = ux(i + \). Consequently, Eq. (138) becomes (139) Substituting Eq.(139) into Eq.(136), we obtain (140a)
(140b) The traveling distance, x, can be calculated by
Electroosmotic Flows in Microchannels
173
Figure 4.32. Comparison of the variation of the traveling distance, x, with current in the process of a lOmM KCl solution displacing a O.lmM KCl solution and the variation of L-x with current under (a) an electric field of 35 kV/m, and (b) an electric field of 70 kV/m.
174
Electrokinetics in Microfluidics
(141) where Iinitai is the electric current generated at the start of the displacement process when the capillary is filled with the electrolyte solution 1. Hence the initial current can be evaluated by: (142) Substituting Eq.(142) for the initial current into Eq.(141), the traveling distance of the left boundary of the mixing zone is given as: (143) Obviously, Eq.(132) and Eq.(143) are identical. It should be pointed out here that the traveling distance of the interface, Eq.(132), derived in Model I, and the traveling distance of the left boundary of the mixing zone, Eq.(143), derived in Model II are the length of the channel filled with the electrolyte solution 2. Therefore, in terms of predicting the length of the channel filled with the electrolyte solution 2, Model I and Model II are equivalent. This implies that Model I, which approximates the mixing zone as a sharp interface, is sufficient to predict the length of the channel filled with an electrolyte solution 2 during the displacement process. For consistency, the length of the channel filled with the electrolyte solution 2 is referred to as the traveling distance of the interface in the following sections and it is determined by Eq.(132). It should be noted that Eq.(140b) provides a method to determine the cross-sectional average velocity during the displacement process. However, it has been found that the cross-sectional average velocity is highly dependent on the local slope of the current-time relationship during the displacement process. In turn, it strongly depends on the curve fitting to the experimental data of the current-time relationship. A small fluctuation in the experimental data causes a large change in the local slope as determined by the curve-fitting technique. Consequently, such a local fluctuation causes a significant error in the predicted cross-sectional average velocity. Therefore, for accuracy, this work focuses on the investigation of the traveling distance of the interface, except for the special case of displacement between nearly similar solutions.
Electroosmotic Flows in Microchannels
175
Figure 4.33 (a) Current-time relationship in the process of a lOmM KCl displacing a O.lmM KCl solution under an electric field of 35 kV/m and 70 kV/m, respectively, (b) relationship between traveling distance and time in the same displacement process, and (c) relationship between current and traveling distance in the same displacement process.
176
Electrokinetics in Microfluidics
4-7.3 Application of the models to experimental data Electroosmotic solution displacing experiments were conducted by using aqueous KCl and LaCl3 solutions and polyamide-coated silica capillary tubes of lOOum internal diameter. The experimental procedures are the same as described in the previous section. Using the measured current-time relationship, Eq. (132) as derived in the Model I will be employed to predict the position of the interface, x, with time, t, while solution 2 displaces solution 1 from left to right. A careful analysis of Eq. (131) would reveal that the electric current is not dependent on the direction of displacement. In other words, when an interface is located at x, the current is the same whether the interface location has been achieved through solution 2 displacing solution 1 (left to right) or solution 1 displacing solution 2 (right to left). Conversely, by conducting one experiment with solution 2 displacing solution 1 (left to right) and a second experiment with solution 1 displacing solution 2 (again left to right), the current is the same when the interface is at x for the first experiment and at L - x for the second experiment. Figure 4.32 shows the interface location, x, for a lOmM KCl solution displacing a O.lmM KCl solution and the interface location in terms of L - x for a O.lmM KCl solution displacing a lOmM KCl solution. The interface location in Figure 4.32a and Figure 4.32b is for electric field strength of 35 kV/m and 70 kV/m, respectively. As would be expected, the data for each set of experiments coincided with each other. The near perfect data agreement would suggest that the current is the same for an interface location, x, for the first displacement experiment and L-x for the second displacement experiment. This would indicate that the mixing zone is not significant. This needs to be true as, except for the mid-point, the traveling distance in the two sets of experiments is not equal and the mixing length is a function of the traveled distance. Figure 4.33a shows the current-time relationship during the displacement process of a lOmM KCl solution displacing a O.lmM KCl solution under an electric field of 35 kV/m and 70 kV/m, respectively. Figure 4.33b shows the change of the traveling distance of the interface with time and Figure 4.33c shows the variation of the current with the traveling distance of the interface, separately. As shown in Figure 4.33b, the traveling distance increases until it reaches a constant value when the second solution completely displaces the first solution. The variation of the traveling distance with time shows very different behavior from that of the current. When the traveling distance reaches 90% of the total length of the capillary (i.e. x = 0.9L), the corresponding current is still very low (i.e. around 3.8 |aA) as compared with the current when the traveling distance reaches 100% of the total length of the capillary (i.e. around 37.8 uA). This is because the conductivity of the lOmM KCl solution is approximately 100
Electroosmotic Flows in Microchannels
Figure 4.34 (a) Current-time relationship in the process of a 0.1 mM KCl displacing a lOmM KCl solution under an electric field of 35 kV/m and 70 kV/m, respectively, and (b) relationship between traveling distance and time in the same displacement process.
178
Electrokinetics in Microfluidics
times higher than that of a O.lmM KCl solution (see Table 1). Even when the traveling distance is approximately 90% of the total length of the capillary (i.e. only 10% of the total capillary length is filled with the O.lmM KCl solution), the total resistance, given by Rtotal = x/(l1OmM
KC1
-A) + (L- x)/(X0lmM
KC1
• A), is
still dominated by the low conductivity of the O.lmM KCl solution. Consequently, the total resistance is very high and the corresponding current is very low. In addition, Figure 4.33a and 4.33b show that when the electric field strength is increased, the time required for a lOmM KCl displacing O.lmM KCl solution decreases. This is because when the electric field strength is increased, the electroosmotic velocity of the electrolyte solution increases. As a result, the time required for the displacement process decreases. From Figure 4.33c, we can see that when the electric field strength is increased, the electric current increases. This is because the electric resistance, which is independent of the applied electric field strength, is the same as long as the interface travels the same distance during the two replacing processes. Consequently, the current, given by I = (E • L)/Rtotai, increases with the applied electric field strength accordingly at the same traveling distance. Figure 4.34a shows the current-time relationship in the process of a O.lmM KCl solution displacing a lOmM KCl solution under an electric field of 35 kV/m and 70 kV/m, respectively. Figure 4.34b shows the relationship between the traveling distance of the interface and time. As shown in Figure 4.34, when the interface travels a very short distance, the current drops significantly to a very low value. This is because the O.lmM KCl solution has a much lower conductivity than that of the lOmM KCl solution (see Table 1). Once a small amount of the low-conductivity O.lmM KCl solution is pumped into the capillary, the total resistance is dominated by the resistance in this section of the capillary and it increases significantly with the traveling time. Consequently, the current drops sharply. Figure 4.34b shows that the interface moves faster under an electric field of 70 kV/m than that under an electric field Table 1. The measured bulk conductivity of the electrolyte solutions
Electroosmotic Flows in Microchannels
179
Figure 4.35 (a) Current-time relationship in the process of a lOmM KCl displacing DIUF water under an electric field of 35 kV/m, and (b) relationship between traveling distance and time in the same displacement process.
180
Electrokinetics in Microfluidics
of 35 kV/m. This is because as the electric field strength is increased, the electroosmotic velocity of the liquid increases. Although the relationship between current and time shows similar behavior for the two electric field strengths, the slope of the relationship between the traveling distance and time are markedly different for the two electric field strengths. Figure 4.35a shows the current-time relationship during the process of a lOmM KCl solution displacing DIUF water under an electric field strength of 35 kV/m. Figure 4.35b shows the variation of the traveling distance of the interface with time. As shown in Figure 4.35b, the traveling distance increases very quickly at the beginning of the displacement process and then slows down until the interface reaches the end of the capillary. This is because the channel wall in DIUF water has a much higher zeta potential (i.e. around 240 mV) than in the lOmM KCl solution (i.e. around 65 mV). Consequently, the electroosmotic velocity of DIUF water in capillary is higher than that of lOmM KCl solution. At the beginning of the displacement process, most part of the capillary is filled with DIUF water and hence the average velocity of the liquid in the capillary is higher. However, when the interface moves towards the end of the capillary, more and more of the lOmM KCl solution is pumped into the capillary to displace DIUF water. As a result, the section length of the capillary with a lower zeta potential becomes longer, and the average zeta potential of the capillary decreases. Consequently the average electroosmotic velocity decreases. This causes the interface to move slower than that at the beginning of the displacement process. Figure 4.36a shows the current-time relationship in the process of a O.lmM KCl solution displacing a O.lmM LaCl3 solution under an electric field of 35 kV/m and Figure 4.36b shows the change of the traveling distance of the interface with time. As observed in Figure 4.36a, the current decreases fairly linearly during the displacement process. This is because the difference of the conductivity between these two solutions is small (see Table 1), as a result, no significant changes occur in the total resistance and the total current during the displacement process. In addition, the traveling distance almost increases fairly linearly when the second solution gradually displaces the first solution. This is because there is a small difference in the electroosmotic velocity of the two solutions when the two solutions have the same ionic concentration [55], allowing the interface to move at a relatively constant velocity. Figure 4.37a shows the current-time relationship in the process of a O.lmM LaCl3 solution displacing a O.lmM KCl solution under an electric field of 35 kV/m and Figure 4.37b shows the variation of the traveling distance of the interface with time. The traveling distance continuously increases until it reaches a constant value when the first solution is completely displaced by the second solution. The fairly linear change of the traveling distance and the current during
Electroosmotic Flows in Microchannels
181
Figure 4.36 (a) Current-time relationship in the process of a O.lmM KCl solution displacing a O.lmM LaCl3 solution under an electric field of 35 kV/m, and (b) relationship between traveling distance and time in the same displacement process.
182
Electrokinetics in Microfluidics
Figure 4.37 (a) Current-time relationship in the process of a 0.1 mM LaCb solution displacing a O.imM KC1 solution under an electric field of 35 kV/m, and (b) relationship between traveling distance and time in the same displacement process.
Electroosmotic Flows in Microchannels
183
the displacement process, as shown in Figure 4.37, can be understood by recognizing that the electric conductivity of the two solutions are nearly the same as shown in Table 1.
184
Electrokinetics in Microfluidics
4-8 JOULE HEATING AND THERMAL END EFFECTS ON ELECTROOSMOTIC FLOW Generally, electroosmotic flows always involve Joule heating effect, and the Joule heating effect is especially significant when high voltages are applied. This effect produces temperature gradients in both cross-stream and axial directions in the channel. The produced temperature gradients can cause the band spreading of charged analytes and thus a reduction in the electrophoresis separation efficiency [56-59]. In addition, elevated solution temperatures can lead to denaturation of proteins or nucleic acids [60]. The use of glass microfluidic chips greatly mitigates Joule heating effects through rapid heat conduction. However, increasingly popular polymer-based (e.g., PDMS) chips, which have a very low thermal conductivity, can suffer from Joule heating effects significantly. Therefore, it is important to understand the Joule heating effect on the liquid temperature in microfluidic systems. Using a caged-dye based flow visualization method, Sinton and Li [61] found a slight curvature in the electroosmotic velocity profiles obtained at the capillary middle point. Although the presence of the curvature was shown to correlate with expected Joule heating effects, the exact mechanism was not determined. Also, quite a few researchers have carried out theoretical analyses to calculate the liquid temperature. Grushka et al. [56] and Knox [58] assumed a constant electric conductivity of the solution, and obtained a radially parabolic temperature profile. Gobbie et al. [62] and Bello et al. [63] pointed out that assuming a constant electric conductivity resulted in underestimated liquid temperatures because in reality, the electrical conductivity increases with temperature. In these previous analyses, the solution was viewed as a "solid" conductor and the advective effects of electroosmotic flow were neglected, as were the thermal end effects due to the presence of reservoirs. Moreover, they did not consider the dynamic change of electric field intensity when the electric conductance of solutions varied with temperature. We will show in this section that the real liquid temperature is an intermediate value between the two temperatures predicted by the "solid" solution models with and without accounting for the temperature dependence of the electric conductivity. The flow field, electric potential field and temperature field are strongly coupled if the temperature dependence of the electric conductivity is taken into consideration. A number of simulations have been conducted on these coupled fields. Zhao and Liao [64] analyzed the thermal effects on electroosmotic pumping of liquids in a slit channel, but they assumed a constant electric field and imposed isothermal conditions on the channel plates. Tang et al. [65] investigated the Joule heating effect on the steady state electroosmotic flow in a capillary, and obtained a significant radial temperature difference under high electric fields (e.g., over 10 degrees in the capillary with a 100-micron internal
Electroosmotic Flows in Microchannels
185
diameter under a 50 kV/m electric field). However, they also assumed a uniform electric field. In this section, we will discuss the dynamic effect of Joule heating on the temperature, potential and flow fields during capillary electroosmosis [66]. Variations of material properties with respect to temperature are taken into consideration. In this way, the temperature field, the electrical double layer field, the applied electrical field and the flow field are coupled together. In addition, we will consider a capillary with a finite length and hence consider the thermal end effects. The objective here is to present a comprehensive analysis of the effect of Joule heating on the dynamic electroosmotic flow in a whole capillary system (i.e., with the ends). 4-8.1 Mathematical model Figure 4.38 shows the capillary system to be modeled in this section, a capillary joining two reservoirs. The liquid in the reservoirs is assumed to be at the ambient temperature, Tm. Although both reservoirs, particularly the downstream reservoir, receive heat from the liquid flowing through the capillary, this heat contribution is considered negligible. This is because the reservoir volume is on the order of micro-liters, several orders of magnitude lager than the
Figure 4.38. The capillary system to be modeled and the computational domain.
186
Electrokinetics in Microfluidics
electroosmotic flow rate. The capillary is composed of three components: liquid domain with a radius R, fused silica wall with a radius Rw and polyimide coating with a radius Rp. The axial length of the capillary is L. In cylindrical coordinates, it reduces to a 2-D problem. Governing equations Electro-osmotic flow is governed by the continuity equation (144) and the momentum equations subjected to an electrical body force (145) where v is the velocity vector, t the time, p the density of liquid assumed to be incompressible, p the hydrodynamic pressure, /U(T) the temperature dependent viscosity with T the absolute temperature, pe the electric charge density, and E the externally applied electric field. The electric charge density is related to the internal potential field y/ of the electric double layer (EDL), formed by the charge at the capillary internal wall or by the zeta potential C, , via the Poisson equation (146)
where S(T) is the temperature dependent dielectric constant, erey the reference dielectric constant, and £0 the permittivity of the vacuum. For a symmetric electrolyte solution, say KC1, y/ is determined by the Poisson-Boltzmann equation (147)
(148)
Electroosmotic Flows in Microchannels
187
where K:"1 is the characteristic thickness of the EDL with n^ the bulk ionic concentration, z the valence, e the electron charge and K^ the Boltzmann's constant, respectively. The electric field E is calculated from the externally applied electric potential <j> by (149) Since the capillary wall is non-conducting, the conservation of electric current gives (150) where <J(T) = X{T)C is the temperature dependent electric conductivity with C the molar concentration and A(r) the molar conductivity of the solution, respectively. The molar concentration is dependent on the bulk ionic concentration n^'mEq. (148)by nx -NaC with Na the Avogadro'snumber. While the flow field in Eq. (145), EDL potential field in Eq. (147), and applied electric potential field in Eq. (150) are all restricted in the liquid region, the temperature field must be extended to cover the whole computational domain (see Figure 4.38). Within the liquid region, the energy equation in the presence of electroosmotic flow effect is given by (151) where C is the specific heat of the liquid and is assumed to be constant in this work, k(l) the temperature dependent thermal conductivity of the liquid. Here, we have neglected the term associated with viscous dissipation (i.e., the thermal energy converted from the mechanical energy) because it is small compared to the Joule heat. Within the two solid regions, the energy equations become (152)
(153)
188
Electrokinetics in Microfluidics
where the subscripts w and p denote the capillary wall and polyimide coating, respectively. Modeling simplifications It is rather difficult to solve the set of Eqs. (145), (147), (150)-(153) without simplifications. The main problem lies in the simultaneous presence of three separate length scales: the capillary length of millimeters, the capillary radius of micrometers and the EDL thickness of nanometers. A complete solution on all the three length scales would require a prohibitive amount of memory and computational time. Therefore, two methods have been proposed to solve this problem. One is to artificially increase the order of magnitude of the EDL thickness, and a qualitative nature of the flow field can thus be obtained [7,9]. The other way is to apply a slip boundary condition at the wall, and avoid the solution of the EDL field [67,41]. In this section, the second approach is chosen, where the liquid at the capillary internal wall is assumed to slip at a velocity Vwall =neoEz with jj.eo =-S(T)SO£(T)/II(T) the electroosmotic mobility and Ez the local electric field in the axial direction (the axial coordinate z is indicated in Figure 4.38). Another simplification is made with respect to the momentum equations. Since electroosmotic flows are generally limited to small Reynolds numbers, the convection term (i.e., the second term on the left-hand side) in Eq. (145) can be neglected. In addition, it has been shown that the characteristic time tsteady for an electroosmotic flow to reach the steady state is on the order of milliseconds [7,67], which is far less than the characteristic time of thermal diffusion in this system (on the order of seconds) [68]. Hence, we can reasonably omit the transient term in the momentum equations as long as the time step sizes selected in the numerical simulation are greater than tsteady [67]. Finally, the set of non-dimensional equations determining the applied electric potential field, flow field, and temperature field is summarized as
Electroosmotic Flows in Microchannels
189
where the length is scaled by the capillary internal diameter (i.e., 2R in Figure 4.38), O =
/(p with (p the potential applied at the capillary inlet, V = v/Frey with
the
reference
velocity
Vref = srefe0
P = p/pVr2ef,
T = tkrej- pCpR2 and 0 - (T' -T^)krej j(p1arey . The subscript i may be w for the capillary wall orp for the polyimide coating. We now specify the boundary conditions for the problem. For the applied electric field, we impose insulation conditions along the edges of the liquid domain, and non-dimensional values of 1 at the inlet and 0 at the outlet of the capillary. For the flow field, we impose fully developed velocity profiles at both ends of the capillary, a symmetric condition along the axis, and a slip velocity along the charged wall. For the temperature field, we impose isothermal conditions at both ends of the capillary, a symmetric condition along the axis, and a convective boundary condition surrounding the capillary given by (155)
where h is the convective heat transfer coefficient, and the radial coordinate r is indicated in Figure 4.38. Numerical method The equations in (154) were solved by using the finite element method using an in-house written code, using 6-noded quadratic triangle elements for the electric potential, velocity and temperature, and 3-noded linear triangle elements for the pressure. A non-uniform grid is generated with grid refinement along the axial direction in the regions near the capillary inlet and outlet, where field variables were found to vary most strongly. At each time step, properties evaluated with the current temperature field (initially at room temperature) were used to determine in turn the potential field, wall slip velocity and flow field.
190
Electrokinetics in Microfluidics
Next, known values of potential and velocity fields were used to solve for the new temperature field at the current time step. These steps were repeated until the temperature variation between adjacent time steps is less than a tolerance (i.e., attain a steady state) or a prescribed time is exceeded for those cases where steady state is inaccessible. To verify the code, two testing cases for which analytical solutions are available were considered. One is the homogeneous electroosmotic flow in a cylindrical capillary considered by Rice and Whitehead [69], and the other is the 1-D heat conduction across a cylindrical capillary considered by Gobbie and Ivory [62]. The numerical solution results are in good agreement with these analytical formulae. 4-8.2 Joule heating and thermal end effects We assume that the liquid has the same physical properties as water, and all the material properties used in the simulation are summarized in Table 2. The dimensions and properties of the flexible fused silica capillary tubes are provided by Polymicro Technologies, USA. The outer diameter of the capillary is 360 microns with a 20-micron thick polyimide coating (i.e., Rw= 160 microns and ^=180 microns). The capillary internal radius, R, was a parameter in the simulation.
Table 2 Material properties used in the simulation. Too = 298K. Liquid
Capillary Wall
Coating
Density p 103(kgm~3)
1.00
2.15
1.42
Heat capacity Cp (10 3 JKg"'K' r )
4.18
1.00
1.10
1.38+0.0013(T-Too)
0.15
Thermal conductivity 0.61+0.0012(T-Too) KCWm-'K"1) Molar conductivity X (10"3m2Smor')
12.64[l+0.0025(T-Too)]
Dynamic viscosity H (kg m-'s"1)
2.761exp(1713/T)xl0 6
Dielectric constant s 305.7exp(-T/219)
Electroosmotic Flows in Microchannels
191
Figure 4.39. Transient development of temperature field: (a) along the axis, (b) at the middle of the capillary. The insert in (b) is the expanded view of the temperature profile 80s after the voltage is applied. Working parameters include internal radius R=5Q microns, length L=\Q cm, zeta potential £ = -50 mV, molar concentration C=10 mM and the applied electric field 15 KV/m, and the convective heat transfer coefficient h =10 W/(m2K).
192
Electrokinetics in Microfluidics
Temperature field Figure 4.39 shows the transient development of the temperature field. We see that as time goes on, the temperature of the whole capillary is elevated (Figure 4.39a). But, temperature gradients are mainly in the inlet and outlet regions. The inlet region is gradually expanded while the outlet region is shortened. These dynamic changes are due to the advective effect of electroosmotic flow, which continuously pulls the cold solution from the inlet reservoir while pushes the hot solution into the outlet reservoir. The temperature difference between the capillary core and the ambient is very small (Figure 4.39b). It is attributed to the high Biot number of the whole system. The radial temperature gradient, however, might be significant due to the essentially small capillary radius. We will discuss the effects of radial and axial temperature gradients later. External potential field Figure 4.40 shows the transient development of the axial electric field Ez, which has been scaled by q>/L (i.e., the nominal electric field applied to the capillary, Enom). Initially, Ez is uniform because the liquid temperature is uniform. When temperature gradients arise in the inlet and outlet regions, the
Figure 4.40. Transient development of the axial electric field. Working parameters are the same as those in Figure 4.39.
Electroosmotic Flows in Microchannels
193
Figure 4.41. Transient development of electroosmotic velocity: (a) axial distribution, solid lines with hollow symbols represent the bulk liquid velocity along the axis, and dashed lines with full symbols represent the slip velocity along the charged wall (not shown at t=0 s); (b) cross-stream velocity profiles at the middle of the capillary. Working parameters are the same as those in Figure 4.39.
194
Electrokinetics in Microfluidics
axial electric field is altered. The higher the temperature is, the larger the local electric conductivity becomes, resulting in a lower local electric field intensity. As shown in the figure, the electric field is highly non-uniform near the entrance and exit of the capillary. This non-uniformity is the main difference between the current full model and the solid "solution" model we mentioned above, which will be further discussed below. Flow field Figure 4.41 shows the transient development of the electroosmotic velocity. The electroosmotic velocity is gradually increased over time as the temperature rises (Figure 4.41a). Another important finding is the convex velocity profile at the middle of the capillary (Figure 4.41b), resulting from the velocity difference between the electroosmotic wall velocity and the bulk liquid motion. This velocity profile shape is caused by an induced pressure field. Because of the temperature difference along the capillary, the electroosmotic flow velocity is different along the capillary length direction as well. Therefore, an induced pressure filed exists inside the capillary along the flow direction to adjust the axial flow velocity in order to satisfy the continuity equation (i.e., the constant flow rate). Figure 4.42 shows the sine-like pressure field throughout the capillary. In both the inlet and outlet regions, a positive pressure gradient (pressure increases in the flow direction) is induced to decrease the bulk
Figure 4.42. Transient development of the induced pressure field. Working parameters are the same as those in Figure 4.39.
Electroosmotic Flows in Microchannels
195
liquid velocity, while a negative pressure gradient is induced through the rest channel to increase the bulk liquid velocity. Figure 4.43 shows the velocity vector plot 80s after the voltage is applied. We can see a concave velocity profile at both ends of the capillary (Figures. 4.43a and 4.43c), and a nearly flat velocity profile (as we have discussed ahove, it is actually of convex shape) through most portion of the channel (Figure 4.43b). In summary, the Joule heating effect induces axial temperature gradients in the regions near the capillary inlet and outlet, which in turn change the local electric field and thus the local electro-osmotic velocity. These end effects affect the velocity profile through the channel by the conservation of mass requirement. Comparison with "solid" solution models We also calculated the temperature field using the "solid" solution models with and without including the temperature dependence of electric conductivity.
Figure 4.43. Velocity vector plot 80s after the voltage is applied: (a) inlet region; (b) middle region; (c) outlet region. All coordinates are in non-dimensional forms scaled by the capillary internal diameter (100 microns, here). Working parameters are the same as those in Figure 4.39.
196
Electrokinetics in Microfluidics
Figure 4.44 shows the comparison of temperature contours in the whole domain 80s after the voltage is applied. The "solid" solution models always give a symmetric temperature profile due to the absence of advective flow effects. In the present complete model, the electroosmotic flow pushes the high temperature plateau downstream in the capillary. Actually, as the time goes on, the length of the high temperature plateau is gradually shortened, and can disappear depending on the channel length and the extent of Joule heating effect. We can see this trend in Figure 4.39a. Moreover, we see that the Joule heating effect is significantly underestimated if the temperature dependence of the electric conductivity is neglected. When we use the "solid" solution model to consider
Figure 4.44. Temperature contours 80s after the voltage is applied: (a) the "solid" solution model with constant electric conductivity; (b) the "solid" solution model with temperature dependent electric conductivity; (c) the present complete model. The temperature differences between adjacent contour levels are all 2 K except those specially labeled. Working parameters are the same as those in Figure 4.39.
Electroosmotic Flows in Microchannels
197
this temperature dependence, however, the Joule heating effect is overestimated. The results of the model presented here agree qualitatively with the experimental observations from Gobbie et al. [62], who also observed that the model with temperature dependent electric conductivity generally over-predicts the capillary temperature, while the constant-conductivity model under-predicts it. We compared the characteristic time required to reach the steady state for the above three models, as shown in Figure 4.45. For the case with constant electric conductivity, the characteristic time is the shortest. The complete model presented in this section predicts a moderate characteristic time. In the following, we will discuss the effects of the working parameters and capillary size on the electroosmotic flow using the complete model. Figure 4.46 shows the curves of centerline liquid velocity against time, and Figure 4.47 shows the centerline temperature profiles at / =80s. The liquid velocity at the center of the capillary is normalized by the nominal value of the applied electric field intensity, Enom.
Figure 4.45. Transient development of the temperature at the middle point of the capillary along the centerline. Curve A is for the "solid" solution model with constant electric conductivity; curve B is for the "solid" solution model with temperature dependent electric conductivity; curve C is for the present complete model. Working parameters are the same as those in Figure 4.39.
198
Electrokinetics in Microfluidics
Capillary size Here, we consider both the capillary internal diameter and the length. As has been well known, decreasing the capillary internal diameter leads to a less Joule heating (see the rise of liquid velocity in Case B compared with Case A in Figure 4.46). Moreover, the time required to reach the steady state is shortened. It should be pointed out, however, that the faster thermal relaxation is not solely due to the more effective heat dissipation associated with the smaller diameter. The reason is also attributed to the reduction of total Joule heat inside the capillary. This can be qualitatively explained as follows. Let us consider a simple 1-D heat conduction, and assume that the radial temperature difference can be neglected, then the energy conservation inside a capillary gives (156) where T can now be viewed as the average temperature of the liquid. Assuming a constant electric conductivity (we also can use the empirical formula for electric conductivity as given in Table 2 to find a more accurate estimation of the whole-capillary temperature rise), the temperature elevation, AT, of the whole capillary is given by
Figure 4.46. Comparison of liquid velocity at the middle point of the centerline of the capillary center of the mid-capillary when only one working parameter (with respect to the general Case A) is changed for each case, which has been labeled beside the corresponding curves. The velocity is normalized by the nominal electric field intensity. For Case E, Q=-6Q mV is used. All other parameters are the same as those in Figure 4.39.
Electroosmotic Flows in Microchannels
199
(157)
which tells us AT is proportional to the square of the capillary internal radius R. Figure 4.47 includes the cases where the magnitude of EnomR holds while R is decreased from 50 microns (Case A) to 10 microns (Case G). One can see that the temperature differences between cases A and G are relatively small. The change in the advective flow effect is also noteworthy. For the capillary with a smaller internal radius, the axial temperature profile becomes more symmetric (see Case B in Figure 4.47), even if the applied electric field is greatly increased (see Case G in Figure 4.47) which implies a higher electroosmotic velocity. The two modes of axial heat transport are, respectively, advection and diffusion. As the capillary internal diameter is decreased, the diffusion portion, which is the symmetric end cooling, is essentially constant. The advection portion that leads to an asymmetric axial temperature profile is, therefore, decreased even at a higher liquid velocity.
Figure 4.47. Comparison of centerline temperature profile 80s after the voltage is applied when only one working parameter (with respect to the general case A) is changed for each case, which has been labeled beside the corresponding curves. For Case E with molar concentration C= lmM, £,= -60 mV is used. For case F with heat transfer coefficient h=100 Wm~2K~', the temperature profile is at the time t = 20s. All other parameters are the same as those in Figure 4.39.
200
Electrokinetics in Microfluidics
For a longer capillary, the role of conductive cooling due to the presence of reservoirs (end effects) is reduced, and the role of convective cooling around the capillary is increased, extending the time to reach the steady state and resulting in higher liquid temperatures. This can be seen from Case C compared with Case A in Figure 4.46. It is also found that the thermal end effects less alter the electroosmotic velocity profile because the induced pressure gradient is less significant in the central portion of a longer capillary. Applied electric field From the energy equation (151), we know that the Joule heat is related to the square of the electric field intensity. A reduced electric field produces less temperature rise, and thus less increase in the bulk liquid velocity and a shorter time to reach the state of thermal equilibrium (see Case D compared with Case A in Figure 4.46). The advective flow effect is reduced as well (see Case D in Figure 4.47). Concentration The molar concentration C determines the solution's electric conductivity by G(T) = X(T)C . If the solution is diluted, the electric conductivity is lowered while typically the zeta potential is slightly increased. As a result, the normalized centerline velocity starts at a higher value than Case A, and increases less (see Case E in Figure 4.46) as time goes on. Due to the reduction in Joule heat generation, the flow effect is similar to the general case (see Case E in Figure 4.47). Convective heat transfer coefficient Improving the capability of heat dissipation will definitely diminish the effect of Joule heating and hence thermal end effects. As shown in Figure 4.46, the liquid attains a low, steady state velocity very quickly when the heat transfer coefficient h is increased from 10 Wm"2K'' in the general case A to 100 Wnf 2K ' (Case F). Case A corresponds to the free air convection and Case F corresponds to the forced air convection. We also see that the temperature field is nearly symmetric (see Case F in Figure 4.47). 4-8.3 Summary A complete model and numerical analysis of electroosmotic flow in a full capillary with the Joule heating effect and the thermal end effects are presented in this section. The consideration of the temperature dependent liquid properties couples the flow field, the electric potential field and the temperature field. The axial temperature gradients result in a non-uniform electric field intensity and thus non-ideal (concave/convex/concave) electroosmotic velocity profiles.
Electroosmotic Flows in Microchannels
201
Except in the regions near the capillary ends the cross-stream velocity profile is convex. This convex velocity profile is not caused by radial temperature gradients. It is due to the negative pressure gradient induced by reservoir-based thermal end effects. The convex curvature can be diminished by either decreasing the Joule heating effect (e.g., lowering the applied electric field, reducing the capillary internal diameter or reducing the electric conductivity of the solution) or increasing the capillary length.
202
Electrokinetics in Microfluidics
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37]
F.F. Reuss, Mem. Soc. Impr. Natural, 2 (1809) 327. H.K. Tsao, J. Colloid Interface Sci., 225 (2000) 247. Y.J. Kang, C.Yang and X.Y. Huang, J. Colloid Interface Sci., 253 (2001) 285-294. W.H. Koh and J.L. Anderson, J. AICHE, 21 (1975) 1176. J.P. Hsu, C.Y Kao, S.J. Tseng and C.J. Chen, J. Colloid Interface Sci., 248 (2002) 176. S. Arulanandam and D. Li, Colloids Surfaces A, 161 (2000) 89. N.A. Patankar and H.H. Hu, Anal. Chem., 70 (1998) 1870. L. Hu, J. Harrison and J.H. Masliyah, J. Colloid Interface Sci., 215 (1999) 300. F. Bianchi, R. Ferrigno and H.H. Girault, Anal. Chem., 72 (2000) 1987. L. Ren, C. Escobedo and D. Li, J. Colloid Interface Sci., 242 (2001) 264-271. L. Ren, J. Masliyah and D. Li, J. Colloid Interface Sci., 257 (2003) 85-92. O. Soderman and B. Jonsson, J. Chem. Phys., 105 (1996) 10300. J.G. Santiago, Anal. Chem., 73 (2001) 2353. HJ. Keh and H.C. Tseng, J. Colloid Interface Sci., 242 (2001) 450. R.J. Yang, L.M. Fu and C.C. Hwang, J. Colloid Interface Sci., 244 (2001) 173. Y. Kang, C. Yang and X. Hunag, Int. J. Eng. Sci., 40 (2002) 2203-2221. W. Barrett, Advanced Engineering Mathematics, McGraw Hill, New York, 1995. R.W. Hornbeck, Numerical Methods, Prentice Hall, New Jersey, 1975. R. Weast, M.J. Astle and W.H. Beyer, CRC Handbook of Chemistry and Physics, CRC Press, Boca Raton, 1986. G.M. Mala, D. Li, C. Werner, HJ. Jacobasch and Y.B. Ning, Int. J. Heat Fluid Flow, 1997. R.J. Hunter, Zeta Potential in Colloid Science: Principles and Applications, Academic Press, New York, 1981. D.J. Shaw, Electrophoresis, Academic Press, New York, 1969. P.S. Vincett, J. Colloid Interface Sci., 69 (1979) 354. K. Seiler, J. Harrison and A. Manz, Anal. Chem., 65 (1993) 1481-1488. J.Th.G. Overbeek, Phenomenology of Lyophobic, Chapter II In: H.R. Kruyt (ed.), Colloid Science, Amsterdam, 1952. D.P. Telionis, Unsteady Viscous Flow, Springer-Verlag, New York, 1981. J.P. Hsu, Y.C. Kuo and S.J. Tseng, J. Colloid Interface Sci., 195 (1997) 388. C. Yang, C.B. Ng and V. Chan, J. Colloid Interface Sci., 248 (2002) 524. J.R. Philip and R.A. Wooding, J. Chem. Phys., 52 (1970) 953. M.H. Oddy, J.G. Santiago and J.C. Mikkelsen, Anal. Chem., 73 (2001) 5822-5832. P. Dutta and A. Beskok, Anal. Chem., 73 (2001) 5097-5102. N.G. Green, A. Ramos, A. Gonzalez, H. Morgan and A. Castellanos, Phys. Rev. E., 61 (2000)4011-4018. A. Gonzalez, A. Ramos, N.G. Green, A. Castellanos and H. Morgan, Phys. Rev. E., 61 (2000)4019-4028. A.B.D. Brown, C.G. Smith and A.R. Rennie, Phys. Rev. E., 63 (2002) 1-8. V. Studer, A. Pepin, Y. Chen and A. Ajdari, Microelectronic Eng., 61-62 (2002) 915920. P. Selvaganapathy, Y.-S. L. Ki, P. Renaud and C.H. Mastrangelo, J. Microelectromechanical Sys., 11 (2002)448-453. M. P. Hughes, Electrophoresis, 23 (2002) 2569-2582.
Electroosmotic Flows in Microchannels
203
[38] V.M. Barragan and C.R. Bauza, J. Colloid Interface Sci., 230 (2000) 359-366. [39] D. Erickson and D. Li, Langmuir, 19 (2003) 5421-5430. [40] J.V. Beck, K.D. Cole, A. Haji-Sheikh and B. Litkouhi, "Heat Conduction Using Green's Functions", Taylor & Francis, London, 1992. [41] D. Erickson and D. Li, Langmuir, 18 (2002) 1883-1892. [42] D. Erickson and D. Li, Langmuir, 18 (2002) 8949-8959. [43] D. Erickson and D. Li, Int. J. Heat Mass Transfer, 45 (2002) 3759-3770. [44] R. Brechtel, W. Hohmann, H. Rudigger and H. Watzig, J. Chromatography A, 716 (1995)97. [45] M.G. Cikalo, K.D. Bartle and P. Myers, J. Chromatography A, 836 (1999) 35. [46] E.B. Cummings, S.K. Griffiths, R.H. Nilson and P.H. Paul, Anal. Chem., 72 (2000) 2526. [47] H-K Tso, J. Colloid Interface Sci., 225 (2000) 247-250. [48] N.L. Burns, N.L et al, Biomaterials, 19 (1998) 423. [49] J.L. Anderson and W.K. Idol, Chemical Engineering Communication, 38 (1985) 93. [50] F. Kleinjung, F.F. Bier, A. Warsinke and F.W. Scheller, Anal. Chim. Acta, 350 (1997) 51. [51] S. V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere: Washington, 1980. [52] A. Anderson, J.C. Tannehill and R.H. Pletcher, Computational Fluid Mechanics and Heat Transfer; Hemisphere: Washington, 1984. [53] A.E. Herr, J.I. Molho, J.G. Santiago, M.G. Mungal and T.W. Kenny, Anal. Chem., 72 (2000) 1053-1057. [54] J.S. Newmann, "Electrochemical Systems", 2nd ed., Prentice Hall Englewood Cliffs, N.J., 1991. [55] L. Ren, C. Escobedo and D. Li, J. Colloid Interface Sci., 250 (2002) 238-242. [56] E. Grushka, R. M. McCormick and J J . Kirkland, Anal. Chem., 61 (1989) 241. [57] A.E. Jones and E. Grushka, J. Chromatogr., 466 (1989) 219. [58] J.H. Knox, Chromatographia, 26 (1988) 329. [59] J.H. Knox and K.A. McCormack, Chromatographia, 38 (1994) 207; and Chromatographia, 38 (1994) 215. [60] R.S. Rush, A.S. Cohen and B.L. Karger, Anal. Chem, 63 (1991) 1346. [61] D. Sinton and D. Li, Colloids Surfaces A, 222 (2003) 273-283. [62] W.A. Gobbie and C.F. Ivory, J. Chromatogr, 516 (1990) 191. [63] S. Bello and P.G. Righetti, J. Chromatogr, 606 (1992) 95; and J. Chromatogr, 606 (1992) 103. [64] T.S. Zhao and Q. Liao, J. Micromech. Microeng, 12 (2002) 962. [65] G.Y. Tang, C. Yang, J.C. Chai and H.Q. Gong, Int. J. Heat Mass Transfer, 47 (2004) 215-227. [66] X. Xuan, D. Sinton and D. Li, Int. J. Heat Mass Transfer, (in press). [67] S.V. Ermakov, S.C. Jacobson and J.M. Ramsey, Anal. Chem, 70 (1998) 4494. [68] E.V. Dose and G. Guiochon, J. Chromatogr. A, 652 (1993) 263. [69] L. Rice and R. Whitehead, J. Phys. Chem, 69 (1965) 4017.
204
Electrokinetics in Microfluidics
Chapter 5
Effects of surface heterogeneity on electrokinetic flow Many microchannels don't have uniform surface properties. The surface heterogeneity may be caused by the impurities of the materials, by adhesion of solutes (e.g., surfactants, proteins and cells) of the solution, and by desired chemical surface modification. Measurements of interfacial electrokinetic properties are popular means of surface characterisation for biomedical polymers [1,2]. It is well known that electrokinetic phenomena can be used as a method of measuring the degree of surface heterogeneity, for example due to the adsorption of surface active substances [3] and colloidal particles [4,5]. More recently, works by Norde [6,7] and others [8,9] have shown that the streaming potential technique can be an effective tool for the study of protein adsorption onto polymer materials that is important to the biocompatibility of various medical devices. In this technique the surface's electrokinetic properties are altered by the presence of the surface heterogeneous regions that has a different ^-potential than the original surface. The change in the surface's average ^-potential, which is reflected by the measured streaming potential, is related back to the degree of surface heterogeneity. To study the effect of surface heterogeneity on pressuredrive liquid flows, a few analytical and numerical models were proposed [4, 5, 10, 11]. These models are primarily concerned with double layer charge transport in the pressure-driven flow field. Recently a comprehensive examination of the complete three-dimensional pressure driven flow field has been performed [12], revealing a number of interesting effects such as the presence of weak circulation regions perpendicular to the main flow axis. In the majority of lab-on-a-chip systems, electrokinetic means (i.e. electroosmotic flow, electrophoresis) are the preferred method of species transport as they offer significant advantages over traditional pressure driven flow. Although the surface heterogeneity has been long recognized as a problem leading to irregular flow patterns and non-uniform species transport, only recently have researchers begun to investigate the potential benefits the presence of surface heterogeneity (non-uniform surface ^-potential or charge density) may have to offer. Early studies examining these effects were conducted by Anderson [13] Ajdari [14,15,16] and Ghosal [17]. In Ajdari's works it was predicted that the presence of surface heterogeneity could result in regions of bulk flow
Effects of Surface Heterogeneity on Electrokinetic Flow
205
circulation, referred to as "tumbling" regions. This behaviour was later observed in slit microchannels experimentally by Stroock et. al. [18], who found excellent agreement with their flow model. In another study Herr et. al. [19] used a caged dye velocimetry technique to study electroosmotic flow in a capillary in the presence of heterogeneous surface properties and observed significant deviations from the classical plug type velocity profile, an effect which was predicted by Keely, et. al. [20]. In another clever application Johnson et. al. [21] used a UV excimer laser to introduce surface heterogeneity to the side wall of a polymeric microchannel and demonstrated how this could reduce sample band broadening around turns. Recently the use of surface heterogeneity to increase the mixing efficiency of a T-shaped micromixer was proposed [22], A general theoretical analysis of 3D electroosmotic flow structure in a slit microchannel with periodic, patchwise heterogeneous surface patterns were performed [23]. This chapter will review the surface heterogeneity effects on pressuredriven flows and on electroosmotic flows in microchannels. Enhanced solution mixing by using heterogeneous surface patches will be discussed. A general model of electrokinetic flows in a microchannel network where channel branches may have different surface properties will be analysed as well.
206
5-1
Electrokinetics in Microfluidics
PRESSURE-DRIVEN FLOW IN MICROCHANNELS STREAM-WISE HETEROGENEOUS STRIPS
WITH
Consider a slit microchannel with strip-wise surface heterogeneity parallel to the flow axis [24], as illustrated in Figure 5.1. Most flow analyses of slit microchannels were reduced to the 1 -D case, since the height to width ratio was much less than one (H/W « 1). However here the existence of a changing surface potential along the x-axis prohibits this simplification and thus a 2-D model must be used. The steady state, fully-developed laminar flow of fluids of constant viscosity and density through a slit microchannel is described by a simplified Navier-Stokes equation for momentum, (1) where uz is the velocity component in the flow direction, dP/dz is the pressure gradient in the flow direction, and Fz represents any additional body forces applied to the fluid continua. If gravity is negligible the only significant body
Figure 5.1 Symmetric and non-symmetric heterogeneous distributions along channel cross section. Flow direction (z axis) points out of the page.
Effects of Surface Heterogeneity on Electrokinetic Flow
207
Figure 5.2 (a) Cross sectional view of the EDL field in a slit microchannel with 50 percent heterogeneous strip surface coverage (£„ = -0.2V and £„ = -0.08V, C = 10~6 M KCL). (b) Close up view of EDL transition zone between the solid surface and the heterogeneous region.
208
Electrokinetics in Microfluidics
force is that caused by the flow-induced electrokinetic potential, the streaming potential, which is described by
where dE/dz is the instantaneous rate of change of the streaming potential along the z axis and pe is the local net charge density per unit volume. Thus, the flow equation becomes: (2)
Clearly, one must know the local net change density and the streaming potential in order to solve for the flow field. This requires solving the EDL field. For the 2D EDL field in such a slit microchannel, the relationship between the net local charge density, per unit volume, pe, and the electrostatic potential, i//, is given by the Poisson equation,
(3)
where s dielectric constant of the fluid and s0 is the permittivity of a vacuum. Assuming the equilibrium Boltzmann distribution is applicable, pe is related to y/ by: (4)
By defining the Deybe-Huckel parameter as K?=2z2e2nJeeokbT, the hydraulic diameter as Dh = 2HW/(H+W), and introduction the following nondimensional parameters: y/* = zey//kbT, K* = KD^ and x* = x/Dh, y* = y/Dh, the non-dimensional EDL potential profile can be described by combining Eq. (3) and Eq. (4) to yield,
(5)
Effects of Surface Heterogeneity on Electrokinetic Flow
209
Figure 5.3 (a) Steady state, fully developed, velocity profile in a slit microchannel with 50 percent heterogeneous strip surface coverage (£<, = -0.2V and C^ = -0.08V, C = 10~6 M KC1). (b) Close up of velocity transition zone between the solid surface and the heterogeneous region.
210
Electrokinetics in Microfluidics
For the heterogeneous microchannel considered here, the above equation is subject to the following Dirichlet boundary conditions at the beginning of the shear plane,
(6)
where £j* £,* Q*, £^r* are the non-dimensional zeta-potentials for the bottom, top, left and right surfaces, defined by £* = zeC/k^T. For the slit geometry, where H « W, the actual value of £* on the left and right boundary is for all practical purposes irrelevant to the final solution, however they are mathematically required to ensure that the problem is well defined. The streaming current can be found by integrating the product of the net charge density and the z component of velocity over the cross section of the channel as is shown below: (7) The streaming potential Es induces a conduction current opposite the direction of the pressure-driven flow. (8) where Ac is the effective cross sectional area of the channel and Xb is the bulk conductivity of the channel. When the flow system is at a steady state, there is no net current flow through the channel, i.e., Is + Ic = 0. Using this condition we can determine the steady state streaming potential as (9)
Effects of Surface Heterogeneity on Electrokinetic Flow
211
Figure 5.4 Effect of degree of surface heterogeneity on the (a) streaming potential, and (b) volume flow rate (C,o = -0.2V and Ch = -0.08V, C = 10"6 M KCl).
212
Electrokinetics in Microfluidics
By introducing the following dimensionless groups; Es* = EsXyiJlzUeLnaPh , uz* = u/U, z* = z/L, Eq. (9) can be non-dimentionalised to the form shown below, (10)
Defining non-dimensional pressure as P*=P/P0, and recognising that dP*/dz*= -1 if a linear pressure drop is assumed, the non-dimensional version of Eq. (2) can be written as: (11) where the non-dimensional groups Gj and G2 are defined respectively as Gj = Dh2Po/fxUL, G2 = 4(zeDh2nx)2/fj.XbAc. Of note here is that a linear pressure drop implies that the velocity profile is quasi-constant at each point in the z-direction. As a result this model assumes that the heterogeneous strips have a constant width along the flow axis. Eq. (5) and Eq. (11) are non-linear partial differential equations, which must be solved numerically. While solving for the EDL profile is straightforward, the velocity profile solution is significantly more difficult since Eq. (11) must be coupled to the streaming potential relation Eq. (10). Consequently an iterative scheme was used to successively update the value of dEs*/dz* in Eq. (11) with that calculated from Eq. (10), using the previous velocity profile [24]. For the streaming current analysis this step is not required since dEs*/dz* = 0. In cases where the streaming potential significantly effects the velocity profile, a relaxation factor less that 1 was applied to the updated value dEs*/dz*. Figures 5.2a and 5.3a show the EDL and velocity profiles in a slit microchannel for a surface of C,o = -0.2V with a 50 percent symmetric heterogeneous coverage with t,h = -0.08V. In Figure 5.2a it is apparent that the heterogeneous strips have a sharp effect on the EDL profile and there is no gradual transition from one region to another as might be expected. This is confirmed in Figure 5.2b, which shows a close-up view of the transition zone between the two heterogeneous regions. Figure 5.3a shows that indeed the surface heterogeneity can have a dramatic effect on the velocity profile along the length of the channel. While
Effects of Surface Heterogeneity on Electrokinetic Flow
213
Figure 5.5 Effect of the magnitude of the heterogeneous surface C, potential on the (a) streaming potential, (b) volume flow rate (constant 50 percent strip-wise heterogeneous coverage, C,o = -0.2V, C = 10~6 M KC1).
214
Electrokinetics in Microfluidics
this figure seems to suggest that the velocity profile also experiences sharp changes along the x-axis, a close-up look in Figure 5.3b, shows the transition to be quite gradual over a large region, greater than 20 times the half height of the channel. By comparing the Figures 5.2a and 5.3a, it can be noted that in the regions of high ^-potential, the local velocity is lower than that in the low C,potential regions. Since pe(x,y) is of greater magnitude at a given distance from the wall in the regions of high ^-potential, the applied body force is greater resulting in a more significant impact on the velocity profile than in the lower C,potential regions. Figures 5.4a and 5.4b compare the changes in the streaming potential and the volume flow rate with the degree of heterogeneity for channel heights ranging from 20 um to 5 p.m. In all cases the results have been scaled by the value predicted for a homogenous channel. It was found that there is no appreciable difference between symmetric and non-symmetric heterogeneous strips, suggesting that the distribution of the surface heterogeneity has very little effect on either the streaming potential or the flow rate in cases of no double layer overlap. In the calculations, the homogenous surface of C, = -0.2 V ( Q and the heterogeneous strips of Q= -0.08 V (Q,) in a streaming solution of l x l O 6 M KCL was considered. From Figure 5.4 it is apparent that for the same degree of the surface heterogeneity, the streaming potential and the flow rate in the smaller channels are affected more by the heterogeneous strips. As the percentage of the heterogeneity with lower zeta potential increases, the average zeta potential is lower and hence the EDL effect is weaker. That is why the streaming potential decreases. Consequently, the electrokinetic body force responsible for the electro-viscous effect weakens, the flow rate therefore increases with the increase of the coverage of the low-zeta potential strips. Figures 5.5a and 5.5b consider the effects of the magnitude of the C,potential of the heterogeneity. As before a streaming solution of lxl 0~6 M KC1 and a surface of (^-potential of-0.2 V were used. The C,h was varied from -0.2 V to -0.08 V at a constant 50 percent heterogeneous coverage in channels ranging in height from 5 to 20 urn. As the zeta potential of the heterogeneous strips decreases, the average zeta potential of the channel surface is reduced, and hence the EDL effect is smaller. This may explain why the streaming potential decreases with C,h. A reduced streaming potential means a weaker electrokinetic body force or a weaker electro-viscous effect. This is why the volume flow rate increases as the Q, decreases. Similar to Figure 5.4, the small channels are more sensitive to the zeta potential change of the heterogeneous strips.
Effects of Surface Heterogeneity on Electrokinetic Flow
5-2
215
PRESSURE-DRIVEN FLOW IN MICROCHANNELS WITH HETEROGENEOUS PATCHES
Microfluidic biosensors [25~27] often involve selective capture or adhesion of a particular organism (for example E. Coli [25]) from the solution medium onto a heterogeneous probe. The presence of the organism can then be externally detected via some fluorescence or chemiluminesence method. These probes are typically arranged in a periodic fashion and generally do not necessarily have the same electrokinetic properties of the sensing platform. In a flowing medium, the capture of the target molecules or particles can be sensitive to the velocity gradients near the probing surface, with higher velocities inducing greater shear forces tending to inhibit adhesion. Other examples of microfluidics applications of surface heterogeneity include a recent study of the ability of oppositely charged patches to enhance solution mixing under electroosmotic flow [22]. Some researchers [19,20,28] have investigated the use of heterogeneous capillaries, either chemically modified or through the use of radial electric fields, in capillary electrophoresis, while still others have investigated artificial manipulation of the surface (^-potential for a number of on chip applications [29,30]. A variety of analytical and numerical models have been proposed recently to investigate electrokinetic effects over heterogeneous surfaces. Adamczyk et. al. [5] proposed an approximate analytical model for spherical particles adsorbed onto a flat surface. Changes in the measured streaming potential were attributed to damping of the local fluid velocity in the region of the particle, and additional charge transport from the particle surface. Hayes et. al. [4,11] also presented another approximate analytical model for adsorbed particles, however their model was based on an assumed shift in the shear plane due to particle adsorption. A few techniques for modeling electroosmotic flow over heterogeneous surfaces such as those by Ajdari [14,15], Strock et. al.[18,31] and Erickson and Li [22] have also been proposed. Cohen and Radke [32] presented a 2D numerical model and a perturbation technique for modeling pressure driven flow over a flat heterogeneous region perpendicular to the flow axis. With the exception of the Cohen and Radke paper [32] all of the above studies assume that the electrical double layer field is independent of the flow field and described by the Poisson-Boltzmann equation. For pressure-driven flow over a surface with heterogeneous patches, the local electrical double layer field and the ionic concentration fields are different from patch to patch. The diffusion due to the local ionic concentration gradients, the streaming convection due to the flow, the conduction convection due to the streaming potential and the local electrical double layer fields will jointly determine the ionic concentration distribution and hence the local net charge density. Thus the local EDL field and the local flow field are coupled.
216
Electrokinetics in Microfluidics
This section will discuss the 3D flow structure associated with pressure driven flow through microchannels with an arbitrary but periodic patch-wise heterogeneous surface pattern [12]. The Nernst-Planck equation, Poisson equation and Navier-Stokes equations are solved simultaneously to determine the local ionic concentration, double layer field and the flow field. The model analysis shows the influence of the heterogeneous patches on the flow field, electrical double layer distribution and measured streaming potential. Quantitative limit on when the Poisson-Boltzmann equation can be applied will also be discussed.
Figure 5.6 Simulated surface heterogeneity patterns. (A) Close packed pattern (B) loose packed pattern. Dark regions represent the heterogeneous patches on the surface. Dashed line represents the computational domain.
Effects of Surface Heterogeneity on Electrokinetic Flow
217
Figure 5.7 (A) Computational domain for 3D spatially periodic cell, showing heterogeneous surface pattern. (B) Labels for surfaces. Fi represents heterogeneous surface plane; F3, Ts, Te are symmetry surfaces; F2 and F4 are periodic surfaces.
218
Electrokinetics in Microfluidics
5-2.1 Theory and numerical method Let's consider the two general periodically repeating surface patterns, as shown in Figure 5.6, which we refer to as the close packed pattern (Figure 5.6a) and the loose packed pattern (Figure 5.6b). In each case the smallest periodic cell for the pattern is also shown. Although the close packed pattern always has a percent heterogeneous coverage of 50%, the quantity can be adjusted for the loose packed pattern by changing the size and spacing of the patches. We consider the pressure-driven flow through a parallel plate microchannel, where both the upper and lower surfaces exhibit one of the heterogeneous surface patterns discussed above. Since the pattern is repeating, the computational domain is reduced to that over a single periodic cell, similar to those shown in Figure 5.6. To further minimize the size of the solution domain we assume that the heterogeneous surface pattern is symmetric about the channel mid-plane. The result is the 3D computational domain shown in Figure 5.7a. Generally, surfaces 3, 5 and 6 (as are labeled in Figure 5.7b and as will be referred symbolically by the symbol T in the following discussion) represent symmetry boundaries while 2 and 4 are periodic boundaries of the computational domain. Further details regarding periodicity as related to this study are provided throughout this section, for a general discussion the reader is referred to a paper by Patankar et. al.[33]. In order to model the electrokinetic flow through this heterogeneous unit, it requires a description of the ionic species distribution, the double layer potential, the flow field and the induced streaming potential. The divergence of the ion species flux, often referred to as the Nernst-Planck species conservation equations [34,35], is used to describe the positive and negative ion densities (given below in non-dimensional form), (12a) (12b) where N^ and N~ are the non-dimensional positively charged and negatively charged species concentrations (JV* = n+/nm N~= ff/n^ where nx is the bulk ionic concentration), 0 is the non-dimensional electric field strength (0 = e^/k^T where e the elemental charge, kf, is the Boltzmann constant and T is the temperature in Kelvin), V is the non-dimensional velocity (V = v/vo where vo is a reference velocity based on the Reynolds number) and the ~ sign signifies that the gradient operator has been non-dimensionalized with respect to the channel half height (/_,, in Figure 5.7). For the positively charged and the negatively charged species the Peclet numbers are given by Pe+ = vol/D+ and Pe~ = voly/D~
Effects of Surface Heterogeneity on Electrokinetic Flow
219
Figure 5.8 Influence of heterogeneous patches on flow field streamlines for the close packed pattern. Homogeneous ^-potential = -60mV (white region), heterogeneous patches (dark patches) have ^-potential (A) - 40mV, (B) - 20mV, and (C) OmV. In all cases Re = 1 and nM = lxlO" 5 M KC1.
220
Electrokinetics in Microfluidics
where D+ and D~ are the diffusion coefficients for the positively charged and the negatively charged species respectively. For completeness, it is important to note that in the derivation of the above equations the ionic species are considered to be monovalent. Along the wall (surface 1 in Figure 5.7b) and symmetry boundaries of the computational domain, zero flux boundary conditions are applied. (12c)
(12d)
(12e)
(12f) where X, Y, and Z are the non-dimensional coordinates (X = x/ly, Y = y/ly and Z = z/ly). Note that in the above boundary conditions the velocity flux components need not be considered as V perpendicular to these surfaces must be zero, as will be discussed later. Periodic conditions are applied along surfaces 2 and 4 as follows: (12g) (12h) The velocity field is described by the Navier-Stokes Equations modified to account for the electrokinetic body force, and the continuity equation as shown below: (13a) (13b)
Effects of Surface Heterogeneity on Electrokinetic Flow
221
Figure 5.9 Velocity vectors and electrostatic potential contours in the double layer region at (A) x = 25 (o.m (mid-plane) and (B) x = 50 um (end-plane) for the close packed heterogeneous surface pattern. In each case Re = 1, n» = lxlO" 5 M KC1 and the C-potential of the homogeneous region and heterogeneous patch are Co = -60mV and CP = -40mV respectively.
222
Electrokinetics in Microfluidics
where F is a non-dimensional constant (F = nJykbT/jj,vo) which accounts for the electrokinetic body force responsible for electro-osmotic and electro-viscous effects, Re is the Reynolds Number (Re = pvj/^i, where p is the fluid density and n is the viscosity) and P is the non-dimensional pressure. As in the previous case periodic boundary conditions are applied along F2 and F4 .
(13c)
A no slip condition is applied along the heterogeneous surface (the channel wall) as described by Eq. (13d), (13d) Along the upper symmetry surface we enforce a zero penetration condition for the y-component of velocity and zero gradient conditions for the x and z terms respectively, (13e) Similarly a zero penetration condition for the z-component of velocity is applied along T5 and F6 while zero gradient conditions are enforced for the x and y components, (13f) The remaining quantity that must be described is the non-dimensional electric field strength. The total electric field strength CP can be divided into two components: (14) where the first component, W(X,Y), describes the electrical double layer field and the second component, E(X), describes the induced electric field resulting
Effects of Surface Heterogeneity on Electrokinetic Flow
223
Figure 5.10 Vx and Vz at the midpoint (z = 12.5(im) of the (A) mid-plane (x = 25jam) and (B) end-plane (x = 50um) of the close packed heterogeneous surface pattern. In each case Re = 1, n« = lxl0~ 5 M KCI and the ^-potential of the homogeneous region and heterogeneous patch are C,o = -60mV and t,p = -40mV respectively.
224
Electrokinetics in Microfluidics
from the streaming potential. In general the double layer component of O is described by a Poisson equation as shown below. (15a) where K is the non-dimensional double layer thickness (K=ly2ncc e2/ki,Tsw). To solve this equation, two boundary conditions commonly applied along the solid surface for the above equation are either an enforced potential gradient proportional to the surface charge density, or a fixed potential condition equivalent to the surface ^-potential. Here a surface potential function is used. (15b) where Z is the non-dimensional (^-potential (Z = eC/kbT). Along the upper and side boundaries of the computational domain symmetry conditions, Eq. (15c) and Eq. (15d), are applied, and periodic conditions, Eq. (15e), are applied at the inlet and outlet, as below: (15c)
(15d) (15e) Extending the technique used by Cohen and Radke [32] into three dimensions, the streaming potential E(X) is evaluated from the condition that the net current flux is zero for constant x, described in general by Eq. (16a), (16a) where nx in the above is a normal vector in the x direction and f and J~ are the non-dimensional positive ion current density and negative ion current density. Assuming monovalent ions for both positive and negative ionic species and using the ionic species flux terms from Eq. (12a) and Eq. (12b), condition (16a) reduces to the following:
Effects of Surface Heterogeneity on Electrokinetic Flow
225
Figure 5.11 Velocity vectors at the mid-plane of the close packed heterogeneous surface pattern at Re = 1 with bulk ionic concentrations of (A) n0 = lxl(T 5 M, and (B) nm = 1x10" M. In each case the (^-potential of the homogeneous region and heterogeneous patch are C,o = -60mV and C,p = -20mV respectively.
226
Electrokinetics in Microfluidics
(16b)
which is a balance between the conduction (term 2) and convection (term 3) currents with an additional term to account for any induced current due to concentration gradients. Solving Eq. (16b) for the streaming potential gradient yields the final form of condition (16a) given below:
(16c)
The above system of equations was solved numerically with the finite element method [36]. The computational domain was discretized using 27-noded 3D elements, which were refined within the double layer region near the surface and further refined in locations where discontinuities in the surface ^-potential were present (i.e., the boundaries between patches). Triquadratic basis functions were used for all the unknowns N^, If, Vx, Vy, Vz, and W. In general the use of these higher order elements and basis functions (as opposed to 8-noded trilinear elements) were found to be optimal as they could better approximate the sharp gradients in velocity, species concentration and electric potential within the double layer region. In all cases periodic conditions were imposed using the technique described by Saez and Carbonell [37]. After forming the computational mesh and performing the elemental matrix integration [36] an initial parabolic velocity profile was assumed and streaming potential gradient was set to zero everywhere. The solution procedure began with a semi-implicit, non-linear technique in which the ionic species
Effects of Surface Heterogeneity on Electrokinetic Flow
227
Figure 5.12 Vz at the midpoint (z = 12.5um) of the (A) mid-plane (x = 25um) and (B) endplane (x = 50nm) of the close packed heterogeneous surface pattern for various Reynolds numbers. In each case the ^-potential of the homogeneous region and heterogeneous patch are Cp = -60mV and £ p = -40mV respectively and the bulk ionic concentration is noo = lxl0~ 5 M.
228
Electrokinetics in Microfluidics
concentrations and the double layer potential equations, Eqs. (12a), (12b) and (15a) were solved for simultaneously. In general it was found that the convergence could only be achieved when the double layer potential, W, in the non-linear terms of Eq. (12a) and Eq. (12b) were fully implicit to the solution and the N^ and JV" terms explicit. An iterative procedure was then initiated, each time updating the explicit variables, until convergence on all three variables was achieved. In general it was found that a direct solver was required, as the resulting matrix was neither symmetric nor well conditioned. Once a solution for JV4, N~ and W was obtained the forcing term in the Navier-Stokes Equations could be evaluated and Eq. (13a) and Eq. (13b) were solved concurrently using a penalty method [36] which eliminated the pressure term from the formulation. While this reduced the number of equations that had to be solved, the resulting poorly conditioned matrix dictated the use of a direct solver over potentially faster iterative solver. To facilitate convergence the static pressure (evaluated when Vx, Vy, Vz = 0) induced by the Y direction electric field gradients in the double layer region was subtracted from the result. Since the inlet velocity profile was not known ahead of time, the pressure drop across the computational domain was introduced via a pressure body force as was suggested by Saes and Carbonell [37]. After the flow solution was obtained Eq. (16c) was evaluated at each cross section to yield updated values of E (streaming potential) at each point in the domain. With this result and that of the double layer potential, W, the
Effects of Surface Heterogeneity on Electrokinetic Flow
229
Figure 5.13 Electrostatic potential contours in the double layer region parallel with the flow axis on T5 (z = 0|im) for (A) Re = 0.1, (B) Re = 1.0 and (C) Re = 10.0 for the close packed heterogeneous surface pattern. In each case the ^-potential of the homogeneous region and heterogeneous patch are C$ = -60mV and C,p = -40mV respectively and the bulk ionic concentration is n^ = lxlO"5 M.
230
Electrokinetics in Microfluidics
straight, as would be expected for typical pressure driven flow, however perpendicular to the main flow axis a distinct circular flow pattern can be observed. In all cases the circulation in this plane is such that the flow near the surface is always directed from the lower ^-potential region (in these cases the heterogeneous patch) to the higher ^-potential region. As a result the direction of circulation is constantly changing from clockwise to counterclockwise along the x-axis as the heterogeneous patch location switches from the right to the left side. It can also be observed that the center of circulation region tends to shift towards the high ^-potential region as the difference in the magnitude of the homogeneous and heterogeneous (^-potential is increased. The observed flow circulation perpendicular to the pressure driven flow axis is the result of an electroosmotic body force applied to the fluid continua caused by the differences in electrostatic potential between the homogeneous surface and the heterogeneous patch. Figure 5.9 shows the velocity vectors within the double layer region, for both the mid-plane at x = 25um (A) and endplane at x = 50um (B) along the flow direction, superimposed over a contour plot of the double layer potential for the case shown in Figure 5.8a. The solid lines on the contour plots of this figure are equipotential lines and darker regions represent a more negative potential (with respect to the neutral bulk solution). As can be seen in Figure 5.9a a negative potential gradient exists near the transition zone between the two regions for the mid-plane. As there is an excess of positive ions in the double layer region, this potential gradient induces a body force in the positive z direction, as is described by Eq. (13a). The effect is similar to that observed in traditional electroosmotic flow where the excess positive charge in the double layer region is attracted to the negative electrode, inducing flow in the direction of the negative potential gradient. The similar effect can be seen in Figure 5.9b where now a positive potential gradient exits at the outlet plane inducing a flow in the negative z direction and resulting in the observed clockwise circulation. Simulations conducted for the loose packed surface pattern also exhibited a similar circulation pattern, however the magnitude tended to be significant weaker at the inlet and exit as zeta potential is homogeneous along the z-axis in these regions. In Figures 5.10a and 5.10b we compare the near surface z-directional velocity at the midpoint (z = 12.5um) of the mid-plane (A) and end-plane (B) with that of the x-direction velocity, again for the case shown in Figure 5.8a. As can be seen the z-direction velocity exhibits a very sharp increase near the wall, where both the excess positive ion concentration is the highest and the potential gradient is the strongest, reaching a maximum and then begins to taper off. Note that the discontinuity in the velocity is a numerical artifact resulting from the refinement of the mesh in the double layer region and the actual flow would exhibit a
Effects of Surface Heterogeneity on Electrokinetic Flow
231
Figure 5.14 Electrostatic potential contours in the double layer region parallel with the flow axis on T6 (z = 25um) for (A) Re = 0.1, (B) Re = 1.0 and (C) Re = 10.0 for the close packed heterogeneous surface pattern. In each case the (^-potential of the homogeneous region and heterogeneous patch are C,o ~ -60mV and Qp = -40mV respectively and the bulk ionic concentration is n«o = lxlO" 5 M.
232
Electrokinetics in Microfluidics
smoother transition. Comparing the relative magnitude of the two velocities it can be seen that the maximum z direction velocity is 4xlO~5 mm/s versus 6x10"' mm/s for the x directional velocity at a comparable distance from the surface. This indicates that the strength of the circulation regions is significantly lower than that of the main bulk flow. Though the absolute magnitude of the velocity is relatively low, especially far away from the surface, the relatively sharp gradient within the double layer region, resulting in a relatively strong shear force, may be important to adhesion phenomena. Since, as described above, the formation of these circulation regions is a electrical double layer driven effect it is of interest to investigate the influence of the double layer thickness, or alternatively the bulk ionic concentration, on the strength of the circulation. Figure 5.11 shows vector plots of the velocities in the end-plane of the system shown in Figure 5.8b for the following ionic concentrations: (A) lxlO"5 M, and (B) lxl(T 4 M, corresponding to the double layer thickness 1/K (K = K*ly from Eq. (15a)) of 97nm, and 31nm, respectively. From Figure 5.11 it is apparent that the strength of the circulation decreases significantly with the double layer thickness. In general, over the concentration range investigated here, it was found that the magnitude of the velocity in the circulation plane was approximately proportional to the double layer thickness. Another parameter of interest, which may affect on the magnitude of the circulation velocity, is the flow Reynolds number. In Figure 5.12 the zdirectional velocity at the midpoint (z = 12.5um) of the mid-plane (a) and endplane (b), similar to Figure 5.11, is presented for Reynolds number varying from Re = 0.1 to Re = 10 for the heterogeneous case shown in Figure 5.8a. This range of Reynolds number was selected, as it is typical of those encountered in microfluidics and electrokinetic studies. As can be seen from these figures, the magnitude of the z-component of velocity in the double layer increases significantly with Reynolds number. Generally it was found that the magnitude of the velocity within this double layer regions was proportional to Reynolds number, in that the magnitude increased approximately 10 fold between Re = 0.1 and Re = 1 and similarly between Re = 1 and Re = 10. 5-2.3 Influence of the induced flow field on the double layer distribution In Figure 5.13 and 5.14 the influence of the flow field on the double layer distribution for the Fs (z = Oum, Figure 5.13) and i~^ (z = 25 um, Figure 5.14) surfaces for the close packed surface pattern at Reynolds numbers of (A) Re = 0.1, (B) Re = 1 and (C) Re = 10 respectively. In all cases the heterogeneous C, potential is -40mV and the bulk ionic concentration is lxl0" 5 M. In both figures it can be seen that the double layer field becomes significantly distorted at higher Reynolds numbers, when compared to the diffusion-dominated field at
Re = 0.\
Effects of Surface Heterogeneity on Electrokinetic Flow
233
Figure 5.15 Vy in the double layer region at various Reynolds numbers for (A) the first ^-potential step change, x = 12.5 um and (B) the second ^-potential step change, x = 37.5 u.m, along Fs for the close packed heterogeneous surface pattern. In each case the C,potential of the homogeneous region and heterogeneous patch are C,o = -60mV and C,v = -40mV respectively and the bulk ionic concentration is nw = lxl0~5 M.
234
Electrokinetics in Microfluidics
and Re = 1. The diffusion-dominated field in these cases represents a classical Poisson-Boltzmann distribution. This distortion of the double layer field is the result of a convective effect observed by Cohen and Radke [32] in their 2D numerical work and comes about as a result of an induced y-direction velocity in the transition region between the two heterogeneous surfaces. This velocity perpendicular to the surface comes about as a result of the fact that the electroviscous effect is weaker over the lower C, potential heterogeneous patch and thus the velocity in the double layer region is faster in this region. To maintain continuity then a negative y-direction velocity is induced when flow is directed from a higher ^-potential region (i.e. more negative) to a lower and a positive ydirection velocity is induced for the opposite case. The magnitude of this induced velocity is plotted in Figure 5.15 at the (A) high ^-potential to low ^-potential transition point at x = 12.5 urn and (B) low C,potential to high (^-potential transition point at x = 37.5um. In both cases the results are presented for the F5plane as shown in Figure 5.13. As can be seen the magnitude of this velocity increases proportionally to the Reynolds number. However, this induced velocity is still significantly lower than the magnitude of the main flow velocity (see Figure 5.10) and thus does not significantly influence the streamlines along the flow axis shown in Figure 5.8. Relating Figure 5.15 back to the double layer distribution shown in Figure 5.13, it is apparent that the convective influence of the negative y-velocity at the first transition point led to a compression of the double layer while the positive y-velocity at the second transition point led to an expansion. As mentioned earlier, in the vast majority of studies involving electrokinetic flow Poisson-Boltzmann equation is used to describe the electrical double layer field. In other words, the equilibrium Boltzmann distribution of ions in the electrical double layer is assumed. At this point we can examine the results presented above with respect to the assumption of Boltzmann distribution used in the classical modeling treatments. As seen from Figures 5.13 and 5.14 in the absence of strong convective effects, Re = 0.1 and even Re = 1.0, the electrical double layer field is indeed identical to the description of the PoissonBoltzmann equation. It is only at Re = 10 and a low ionic concentration of lxlO"5 M that significant convective effects on the EDL field are observed. Thus we may infer that a Boltzmann distribution is a suitable approximation so long as Re < 1 and rico > 10~5 M, in the case where pressure driven flow is directed parallel with the surface plane. In more general cases we can infer from Figure 5.15 that the velocity normal to the surface in the double layer must be on the order of 10~3 mm/s before significant a non-Poisson-Boltzmann characteristics will be observed. Though in principal the mechanisms are the same, it is not clear based on the above results if these limitations are also applicable to modeling electroosmotically-driven flows over heterogeneous surfaces.
Effects of Surface Heterogeneity on Electrokinetic Flow
235
Figure 5.16 Diagram showing net charge density within the double layer region for the case of a homogeneous surface ^-potential of -60mV and an oppositely charged patch of +60mV, where the flow is directed into the paper. The detailed image shows the applied body force on the local charges due to the induced electric potential gradient between the positive and negative surfaces.
5-2.4 Oppositely charged heterogeneous patches A special case of theoretical interest occurs when the heterogeneous patches have an equal but opposite charge to that of the homogeneous surface (i.e. the surface has a ^-potential of -60mV while the heterogeneous patches have a (^-potential of +60mV). In such a case there is an excess of positive ions over the homogeneous surface, as before, but negative ions dominate over the heterogeneous surface to balance out the positive surface charge as is shown in Figure 5.16. The difference in the surface potential results in an electrostatic potential gradient within the double layer region near the transition zone between the two regions similar to that shown in Figure 5.9b. The presence of this potential gradient results in a body force on the ions in the double layer, however in this case an opposite body force is applied to the negative ions over the positively charged patch, as shown in the detailed image in Figure 5.16.
236
Electrokinetics in Microfluidics
Consequently the net body force over the region is zero and the circulation velocity perpendicular to the main flow, as shown in Figure 5.8, becomes negligible. Figure 5.17 shows a contour plot of the double layer potential field along the F5 (A) and F6 (B) planes, as defined in Figure 5.7, for the ^-potential case described above using the close packed surface pattern with a bulk ionic concentration of lxlO"5 M and a Reynolds number of 10. Note that in this figure the darkest region corresponds to the -60 mV surface, as before, while the lightest is the +60mV heterogeneous patch. As can be seen the convective influence on the double layer region is identical for both cases unlike in previous cases where an asymmetry existed depending on the direction of the direction of the streaming potential induced y-directional velocity (Figures 5.13 and 5.14). In this case however the net streaming potential is negligible (since the average C,potential is zero) and this effect is thus much smaller. The variations in the surface potential observed here are induced as a result of oppositely charged ions being transported (through both advection and diffusion) from one region to the other. In the first transition region shown in Figure 5.17a for example positive ions are being transported from the homogeneous region to the heterogeneous patch. As a result this decreases the local net negative charge density and a disturbance of the local electric potential field is observed. Similarly for the second transition region negative ions are transported from the region above the heterogeneous patch to the homogeneous region, resulting in a decrease in the positive net charge density and an opposite disturbance of the local double layer field. 5-2.5 Summary Surface electrokinetic heterogeneity occurs in a variety of electrokinetic characterization and microfluidic applications. A theoretical model for pressuredriven electrokinetic flow over a patch-wise heterogeneous surfaces is discussed in this section. This model based on a simultaneous solution to the coupled Nernst-Planck, Poisson and Navier-Stokes Equations reveals the synergetic effects of the heterogeneous EDL fields and the flow field on the resulting 3D flow structure. The presence of a heterogeneous patch was shown to induce a circulating flow pattern perpendicular to the main flow axis, which was caused by an electrostatic potential difference between the double layer fields over the homogeneous surface and heterogeneous patch. The strength of this circulation was found to be proportional to both Reynolds number and double layer thickness. At higher Reynolds number {Re =10) convective effects were found to be significant to the double layer distribution and led to a deviation from the classical Poisson-Boltzmann distribution observed at lower Reynolds numbers.
Effects of Surface Heterogeneity on Electrokinetic Flow
237
Figure 5.17. Electrostatic potential contours in the double layer region parallel with the flow axis on (A) T5 (z = 0|J.m) and (B) T6 (z = 25 um) surfaces with Re = 10.0 for the close packed heterogeneous surface pattern. In each case the ^-potential of the homogeneous region and heterogeneous patch are C,o = -60mV and t,p = +60mV respectively and the bulk ionic concentration is no, = lxl0~5 M.
238
5-3
Electrokinetics in Microfluidics
ELECTROOSMOTIC FLOW IN MICROCHANNELS WITH CONTINUOUS VARIATION OF ZETA POTENTIAL
The driving force for electroosmotic flow in microchannels depends on the local net charge density and the strength of the externally applied electrical field. Since the net charge density is dependent on the EDL field and hence on the zeta potential, C, , the electroosmotic flow behavior in turn is dependent on the zeta potential. Generally, the zeta potential is a function of the ionic valence, the ionic concentration of the electrolyte solution and the surface properties of the microchannel wall. For a system with a simple electrolyte solution and a homogeneous channel wall, the zeta potential is considered constant. However, in practice, the liquids involved in various lab-chips are solutions containing biological particles (e.g. DNA and protein) and bio-polymers. The adhesion of these particles to the channel wall will cause the non-uniform zeta potential along the channel, depending on the distribution and the extent of the adhered particles on the wall. Therefore, the understanding of the flow behavior in such a situation is important for manipulating the flow in biochip devices. Several researchers [18,19,38,39,40] have investigated the effects of variable zeta potential on electroosmotic flow. Anderson and Idol [38] developed an infinite series solution for flow in a cylindrical microchannel with a zeta potential varying as a cosine or sine function in the flow direction. Long [39] considered the similar type of zeta potential variation in planar and cylindrical capillaries, and developed approximate solutions by considering the heterogeneity as the small perturbations to the velocity. Herr [19] experimentally investigated the flow field of the electroosmotic flow and the sample dispersion rate in open capillaries with a step change in zeta potential (i.e. £j = 0 and <^2 > 0 ) and presented a simple model for the fluid velocity and the dispersion rate. Stroock, et al [18] studied two types of surface charge change in rectangular channels. One is the surface charge variation along a direction perpendicular to the applied electrical field, which generates a two-directional electroosmotic flow. Another is the surface charge variation along a direction parallel to the electrical field, which generates a re-circulating flow. Potocek, et al [40] investigated the influence of the discrete step change in zeta potential on the velocity profile of the electroosmotic flow by considering the zeta potential change between two steps. The above mentioned works all assumed thin electrical double layer and used the electroosmotic velocity as the slip boundary condition. Because the electrical body force responsible for electroosmotic flow depends on the local net charge density, therefore, a more general treatment to the effects of zeta potential (or EDL) variation on electroosmotic flow should consider the local net charge density change along the channel. This requires solving for the Poisson-Boltzmann equation which governing the EDL field.
Effects of Surface Heterogeneity on Electrokinetic Flow
239
In this section, we consider that the zeta potential changes over a range of values along a cylindrical microchannel [41]. Without assuming a thin EDL field, the non-linear Poisson-Boltzmann equation and the momentum equation were solved numerically. The influences of the heterogeneous section size and the direction of the zeta potential change on the flow rate, the velocity profile and the induced pressure distribution are discussed. In the following, two cases of electroosmotic flow in a 10-cm long heterogeneous microchannel will be analyzed. In the first case, the microchannel has 100 equal-size sections and the zeta potential linearly changes from 10 mV to 200 mV over the 100 sections. In the second case, the microchannel has 100 unequal-size sections and the zeta potential linearly changes from 10 mV to 200 mV over the 100 sections. In this case, the section size (i.e. the length of the section) is determined by the following equation: (17) where Lj is the length of the i -th section and / is a constant chosen as 1.1. Since the objective here is to investigate the qualitative effects of the heterogeneous section size on the electroosmotic flow behavior, the specific choice of the function in Eq. (17) will not affect the generality of the results. For a cylindrical microchannel, the relationship between the electrical potential, y/(r), and the net charge density per unit volume, pe, at any point in the liquid is described by the Poisson equation, ,,8) where £ is the dielectric constant of the solution and £o is the permittivity of vacuum. Assuming that the equilibrium Boltzmann distribution is applicable, the net volumetric charge density, pe, is proportional to the concentration difference between cations and anions, for symmetric electrolyte solution such as KC1 (z: z = 1:1) solution, given by: (19) where nica and z,are the bulk ionic concentration and the valence of type / ion, respectively, e is the charge of a proton, kj, is Boltzmann constant and T is the temperature.
240
Electrokinetics in Microfluidics
Substituting the net charge density in the Poission equation by Eq.(19), and introducing the dimensionless variables
where a is the radius of the microchannel, the non-dimensional PossionBoltzmann equation can be written as follows: (20)
where K: , the Debye-Huckle parameter. Because of the symmetry of the EDL field in the cylindrical microchannel, Eq. (20) is subjected to the following nondimensional boundary conditions:
(21) Eq.(20) describing the electrical double layer potential is a nonlinear partial differential equation that must be solved numerically. Taylor series expansion is used to linearize the nonlinear source term on the right-hand side of Eq. (20) as follows: (22) where y/ is the value for y/ obtained in the previous iteration [42]. A numerical finite difference scheme was used to discretize the governing differential equation and the resulting system of algebraic equations were solved using the Gauss-Seidel iterative technique, with successive over-relaxation employed to improve the convergence time [43]. Since the electrical potential field varies dramatically within a small distance of the channel walls, variable grid spacing [44] was employed to ensure that as the surface was approached, the grid spacing was refined enough to capture the sharp gradients. Once the electrical potential field i// (r ) has been found, the net charge density can be determined by Equation (19). For a steady and fully developed flow, the equation of motion is given by
Effects of Surface Heterogeneity on Electrokinetic Flow
241
(23)
where u z is the velocity component in the channel length direction, viscosity, p is the density of the fluid, VP is the induced pressure which will be discussed shortly, pe is the local net charge density and electrical field strength applied to the microchannel. Substituting the Eq. (19) for the net charge density into the Eq. introducing the following non-dimensional variables,
/u is the gradient E is the (23) and
where U is an arbitrary reference velocity, the non-dimensional equation of motion can be obtained, (24)
Eq. (24) is subjected to the no-slip and symmetric boundary conditions:
For the electroosmotic flow of a homogeneous liquid through different sections of a microchannel, certain conditions must be satisfied. Because of the uniform bulk liquid properties along the flow path, and assuming negligible effects of the surface conductance, the electrical current and the applied field strength are constant axially. The remaining key condition is the constant volume flow rate or the continuity condition. Consider a heterogeneous channel with two equal-size sections, each has a different zeta potential, and the same external pressure at the both ends of the channel. It can be understood easily that, if the zeta potential is higher, the EDL field will be stronger and hence the electroosmotic flow rate will be larger. If, for example, the high zeta potential section is in the downstream and the low zeta potential section is in the upstream, the liquid in the downstream section would have a higher electroosmotic
242
Electrokinetics in Microfluidics
velocity, and the liquid in the upstream section would have a lower velocity. This would cause different flow rate in different sections and hence violate the continuity condition. In reality, a vacuum tends to form between these two sections since the liquid in the downstream section is moving faster than that in the upstream section. Because of the same pressure (i.e. atmospheric pressure) at the both ends of the channel, a negative pressure is induced between these two sections. This induced negative pressure will generate a positive pressure gradient for the downstream section to slow the flow and a negative pressure gradient for the upstream section to increase the flow rate. In this way, the continuity condition will be satisfied. This is also why Eq.(23) or (24) has an induced pressure gradient term. The induced pressure gradient will introduce more complexity in the numerical solution. Firstly, the Possion-Boltzmann equation has to be solved to find the electrical potential distribution in each section along the channel, y/(r,z). Once the electrical potential is found, the net charge density can be determined by Eq. (19). Secondly, a guess value of the induced pressure gradient profile is chosen for the first 99 sections (from left to right). With this guess value, the equation of motion can be solved for these sections and the volumetric flow rate in these sections can be determined by the following equation: (25) where Qj is the volumetric flow rate and Uj (r) is the local velocity for the i -th section, respectively. During this process, the pressure gradient for the first section is fixed and the pressure gradients for the other sections are adjusted to satisfy the continuity condition: (26) Finally, because of the same pressure at the both ends of the channel, the pressure gradient in the last section has to be determined by the following condition: (27)
Effects of Surface Heterogeneity on Electrokinetic Flow
243
Figure 5.18 Flow rates versus the microchannel diameter for a 1x10 5 M KC1 solution under electrical field strength of 350 V/cm.
Figure 5.19 Flow rates versus the electrical field strength for a 1x10 5 M KC1 solution in a microchannel of 100 |am in diameter.
244
Electrokinetics in Microfluidics
where VP;- is the pressure gradient of the i-th section. With this pressure gradient, the equation of motion can be solved for the last section and the volumetric flow rate for this section can be determined by Eq. (25). The continuity condition, Q\ =Q\§§, has to be checked again at this moment. If the continuity condition is not satisfied, the pressure gradient for the first section has to be adjusted and repeat the above procedure until the continuity condition is satisfied. The above equations and the matching boundary conditions for the EDL field and the flow field were solved numerically. KCl electrolyte solution is used as the testing liquid and the following physical properties of KCl electrolyte solution were used: s = 80, sQ= 8.854 x 1(T12 CV 'm ', and /u = 0.90 x 10"3 kg
Figure 5.20 The velocity field (upper), the induced pressure field and the zeta potential distribution (lower) versus the distance in flow direction for a lxl(T 5 M KCl solution in a microchannel of 100 nm in diameter and under an electrical field strength of 350 V/cm. In this case, the section size is axially constant.
Effects of Surface Heterogeneity on Electrokinetic Flow
245
m 's l [45]. An arbitrary reference velocity of U= 1 mm/s was used to nondimensionalize the velocity. Figure 5.18 shows the dependence of the electroosmotic flow rate on the diameter of the microchannel. As the diameter increases, the flow rate increases. The variations of the section size and the trend of the zeta potential change in the flow direction will not affect this result. As discussed previously, the velocity of the electroosmotic flow and hence the flow rate depend on the zeta potential. In this case, the zeta potential changes from 10 mV to 200 mV along the channel. As clearly seen from Figure 5.18, for the system with equal-size sections, the flow rate generated with the average zeta potential (the average value of the maximum and the minimum zeta potentials) is the same as the flow rate for the case with a linearly increasing zeta potential distribution and the case with a
Figure 5.21 The velocity field (upper), the induced pressure field and the zeta potential distribution (lower) versus the distance in flow direction for a lxl0~ 5 M KC1 solution in a microchannel of 100 (im in diameter and under an electrical field strength of 350 V/cm. In this case, the section size is axially constant.
246
Electrokinetics in Microfluidics
linearly decreasing zeta potential distribution in the flow direction. Therefore, for the case of equal section size, the average zeta potential can be used to evaluate the electroosmotic flow rate in a microchannel with a non-uniform zeta potential distribution. For comparison, the flow rates for homogeneous microchannels with a zeta potential of 200 mV and 10 mV, respectively, are also plotted in Figure 5.18. However, for the case of unequal section size, the flow rates are different when the trend (i.e. increase or decrease) of the zeta potential change along the flow direction changes. When the zeta potential linearly decreases in the flow direction, the flow rate is smaller than that in the system with a linearly increasing zeta potential distribution in the flow direction. This is because when the zeta potential decreases in the flow direction, the lower zeta potential takes a large portion of the microchannel as the section size increases in the flow direction (i.e. I ] = 7.26x10" 4 mm, £j=200mV and L 1 0 0 =9.1mm, ClOO =10mV). This means a weaker EDL field in this portion of the channel. On the other hand, in the system with an increasing zeta potential distribution in the flow direction, the higher zeta potential sections occupy a large portion of the channel (i.e. L\ =7.26x10" mm, £]=10mV and Z]oo=9.1mm, £l00 =200mV). This implies a stronger EDL field in this portion of the channel. The flow rate depends on the EDL field and hence will be smaller in the system with a decreasing zeta potential distribution in comparison with that in the system with an increasing zeta potential distribution. The influence of the electrical field strength on the flow rate exhibits a similar behavior as shown in Figure 5.19 and can be understood similarly. Figure 5.20 shows the velocity field (upper) and the induced pressure field (lower) in a microchannel with a linearly decreasing zeta potential distribution and the equal-size sections. For a simple electroosmotic flow in a microchannel with uniform zeta potential, the plug-like velocity profile is expected. However, if the zeta potential is non-uniform such as specified in this work, the net charge density within the liquid varies axially. Meanwhile, the applied electrical field strength (V/m), which depends on the electrical current and the conductivity of the electrolyte solution, is constant axially. Consequently, the electrical body force generating the electroosmotic flow, which depends on the net charge density and the electrical field strength, is different from section to section. This would imply different volumetric flow rates for different sections. For example, if the zeta potential is lower in a section, the electrical force exerted on the fluid would be smaller and hence the generated flow rate is lower. For an incompressible liquid, the continuity condition requires the same flow rate throughout the microchannel. In order to achieve the same flow rate as that in the next section with a higher zeta potential, a negative pressure gradient (the
Effects of Surface Heterogeneity on Electrokinetic Flow
247
pressure decreases in the flow direction) is introduced to increase the flow rate. As seen in Figure 5.20, the pressure starts decreasing from a maximum pressure to the atmospheric pressure at the exit when the zeta potential is lower than 105 mV, the average value of the maximum and the minimum zeta potential of the microchannel. The nearly parabolic velocity profile was found for the downstream sections with lower zeta potentials (e.g. £=10mV), since the dominant driving force in these sections is the induced pressure gradient. For the upstream sections with higher zeta potentials, because of the presence of the negative pressure gradient in the downstream sections and the same atmospheric pressure at the both ends of the microchannel, a positive pressure gradient (the pressure increases in the flow direction) must exist to slow down the flow in these upstream sections. As shown in the figure, for the sections with higher zeta
Figure 5.22 The velocity field (upper), the induced pressure field and the zeta potential distribution (lower) versus the distance in flow direction for a 1CT5 M KC1 solution in a microchannel of 100 |^m in diameter and under an electrical field strength of 350 V/cm. In this case, the section size increases axially.
248
Electrokinetics in Microfluidics
Figure 5.23 The induced pressure distribution versus the distance in flow direction for a 1CT5 M KC1 solution in a microchannel of 50 u.m in diameter and under an electrical field strength of 350 V/cm.
potentials (e.g. C, = 200 mV), the electroosmotic velocity profile is distorted by the positive pressure gradient. If the zeta potential increases in the flow direction, the distorted electroosmotic velocity profiles in the high zeta potential sections and the nearly parabolic velocity profiles in the low zeta potential sections were also found. As shown in Figure 5.21, the pressure distribution in the flow direction is different from that shown in Figure 5.20, since the directions of the zeta potential variation in these two cases are opposite. The variation of the induced pressure in this case can be understood in a similar way. However, for a microchannel with unequal-size sections (the section size increases in the flow direction in this paper), when the zeta potential decreases from 200 mV to 10 mV in the flow direction, the electroosmotic velocity profile in the higher zeta potential sections (e.g., the C, = 200 mV section) is distorted more significantly as shown in Figure5 in comparison with that in the system with equal-size sections. This is because a major portion of the channel has lower
Effects of Surface Heterogeneity on Electrokinetic Flow
249
zeta potentials in this case and hence the flow rate is smaller than that of the equal section size system, as seen in Figure 5.18. Therefore, for the sections with the same high zeta potentials, in order to reach this small flow rate, a bigger induced pressure gradient is required resulting in the more distorted velocity profile. Figure 5.23 shows that the induced pressure field depends on the direction of the zeta potential variation and the section size distribution. As discussed earlier, for the sections with higher zeta potentials, the induced pressure gradient is positive (the pressure increases) and for the sections with lower zeta potentials, the induced pressure gradient is negative (the pressure decreases). Therefore, when the zeta potential decreases from 200 mV to 10 mV in the flow direction, the pressure increases first from the atmospheric pressure until the zeta potential decreases to a specific value, and then the pressure decreases from a maximum value to the atmospheric pressure at the exit. When the zeta potential increases from 10 mV to 200 mV, the pressure decreases first from the atmospheric pressure until the zeta potential reaches a specific value, and then the pressure increases from a minimum value to the atmospheric pressure at the exit. This specific zeta potential depends on the section size distribution. For the case of equal section size, this specific zeta potential is the average value of the maximum and the minimum zeta potentials; however, for the case of unequal section size, this value is between the maximum and the minimum zeta potential and can be determined by the numerical results. Also in Figure 5.23, we can see that the induced pressure in the case of equal section size is bigger than that in the case of unequal section size. This is because for unequal section size system, the zeta potential changes more quickly over a very small distance. For example, in the case of decreasing zeta potential, the zeta potential decreases from 200 mV to 105 mV over 0.845% of the total length of the microchannel. Therefore, the variation of the zeta potential over almost the total length of the microchannel (i.e. 99.155% of the total length) is smaller (i.e. from 105 mV to 10 mV) in comparison with that in the case of equal section size, where the variation of the zeta potential over the same distance is approximately 190 mV. Recall that the induced pressure gradient is introduced because of the non-uniform zeta potential distribution, it is easy to understand that the smaller the variation of the zeta potential, the smaller the induced pressure gradient. From the above analysis, we see that the different types of velocity profiles in the upstream and the downstream sections are caused by the nonuniform zeta potential distribution. The effects of the unequal section size and the direction of the zeta potential change in the flow direction on the flow rate are demonstrated. For the equal section size system, the trend of the zeta potential change has no effect on the flow rate and the average zeta potential can be used to evaluate the electroosmotic flow rate. However, for the unequal
250
Electrokinetics in Microfluidics
section size system, the flow rate in the case of increasing zeta potential is bigger than that in the case of decreasing zeta potential. This is because the higher zeta potential takes a larger portion in the former case and hence the EDL field is stronger in this case. The effects of the unequal section size and the direction of the zeta potential change in the flow direction on the induced pressure distribution are also demonstrated. The induced pressure distribution is different when the trend of the zeta potential variation in the flow direction changes. For the unequal section size system, the induced pressure is smaller than that in the system of equal section size, because the variation of the zeta potential over almost the total length of the channel in this case is smaller than that in the system with equalsize sections.
Effects of Surface Heterogeneity on Electrokinetic Flow
5-4
251
ELECTROOSMOTIC FLOW IN MICROCHANNELS WITH HETEROGENEOUS PATCHES
Patches-wise surface heterogeneity is the configuration most close to the reality. In order to fully understand and potentially exploit the effects of patcheswise surface heterogeneity on electroosmotic flow, this section will discuss the complex three-dimensional electroosmotic flow structure in the presence of patches-wise surface heterogeneity [23]. The model used here is based on a simultaneous solution to the Nernst-Planck, Poisson and Navier-Stokes equations, which also allows us to avoid the assumption of Boltzmann double layer distribution and to adopt a more general approach. Let's consider the electroosmotically driven flow through a slit microchannel (i.e. a channel formed between two parallel plates) exhibiting one of the three periodically repeating heterogeneous surface patterns shown in Figure 5.24. The first two of these patterns, Figures 5.24(a) and 5.24(b), are in essence one-dimensional surface patterns and are considered here in order to provide a validation of the results via a qualitative comparison with other published studies [14,18]. Since the pattern is repeating, the computational domain is reduced to that over a single periodic cell, demonstrated by the dashed lines in Figure 5.24. To further minimize the size of the solution domain it has
Figure 5.24 Periodic surface patterns (a) lengthwise strips, (b) crosswise strips, (c) patchwise pattern. Dark regions represent heterogeneous patches, light regions are the homogeneous surface. In each case the percent heterogeneous coverage is 50%.
252
Electrokinetics in Microfluidics
been assumed that the heterogeneous surface pattern is symmetric about the channel mid-plane, resulting in the computational domain shown in Figure 5.25. As a result of these two simplifications, the inflow and outflow boundaries, surfaces 2 and 4 (as labelled in Figure 5.25 and referred by the symbol F in the following discussion), represent periodic boundaries on the computational domain, while surface 3 at the channel mid-plane represents a symmetry boundary. From Figure 5.24 it can be seen that in all cases the surface pattern is symmetric about F5 and F6 and thus these surfaces also represent symmetry boundaries. Further details regarding periodicity related to this study are provided throughout this section. For a general discussion of periodic boundary conditions, however, the reader is referred to a paper by Patankar, Liu and Sparrow [33]. As mentioned above, electroosmotic flow occurs when an applied driving voltage interacts with the net charge in the electrical double layer near the surface resulting in a local net body force that causes fluid motion. In order to model the flow through this periodic unit, it requires a description of the ionic species distribution, the double layer potential, the flow field and the applied potential. To begin the divergence of the ion species flux (here we consider a monovalent, symmetric electrolyte as our model species), often referred to as the Nernst-Planck conservation equations, is used to describe the positive and negative ion densities (given below in non-dimensional form), (28a) (28b) where A^ and N~ are the non-dimensional positive and negative species concentrations (N* = n+/nm N~ = rT/n^, where n^ is the bulk ionic concentration), O is the non-dimensional electric field strength (& = e(j)/kbT where e the elemental charge, kb is the Boltzmann constant and T is the temperature in Kelvin), V is the non-dimensional velocity (V = v/v0 where v0 is a reference velocity) and the ~ sign signifies that the gradient operator has been non-dimensionalised with respect to the channel half height (ly in Figure 5.25). The Peclet number for the two species are given by Pe+ = vol/D+ and Pe~ voly/D~ where D+ and D~ are the diffusion coefficients for the positively charged and negatively charged species respectively. Along the heterogeneous surface (Fj in Figure 5.25) and symmetry boundaries of the computational domain (T 3 , F5 and F6), zero flux boundary conditions are applied to both Eq. 28(a) and Eq. 28(b), see reference [12,22] for explicit equations. As mentioned earlier, along faces F2 and F4, periodic
Effects of Surface Heterogeneity on Electrokinetic Flow
Figure 5.25 boundaries.
253
The periodical computational cell showing location of the computational
conditions are applied which take the form shown below, (28c) (28d) The velocity field is described by the Navier-Stokes equations for momentum, modified to account for the electrokinetic body force, and the continuity equation as shown below, (29a) (29b) where F is a non-dimensional constant (F = nx ly kb T / JJ. v0) which accounts for the electrokinetic body force, Re is the Reynolds Number (Re = p v0 ly/ /u, where
254
Electrokinetics in Microfluidics
p is the fluid density and n is the viscosity) and P is the non-dimensional pressure. Along the heterogeneous surface, Fj, a no slip boundary condition is applied. At the upper symmetry surface, F3, we enforce a zero penetration condition for the y-component of velocity and zero gradient conditions for the x and z terms respectively. Similarly a zero penetration condition for the z-component of velocity is applied along F5 and F6 while zero gradient conditions are enforced for the x and y components. As with the Nernst-Planck equations, periodic boundary conditions are applied along F2 and F4. For the potential field, we choose to separate, without loss of generality, the total non-dimensional electric field strength, Q, into two components (30) where the first component, *F(X,Y,Z), describes the electrical double layer field and the second component, E(X), represents the applied electric field. In pressure driven flows, E(X) is the flow induced streaming potential [12,22] and is generally quite weak. In this case however the gradient of E(X) is of similar order of magnitude to the gradient of f(^7,Z) and thus cannot be decoupled to simplify the solution to the Poisson equation. As a result for this case the Poisson equation has the form shown below, (31a) where K is the non-dimensional double layer thickness (K2 = JC2 ly2/2 where K = (2 rioo e2lswkb T)m is the Debye-Huckel parameter). Two boundary conditions commonly applied along the solid surface for the above equation are either an enforced potential gradient proportional to the surface charge density, or a fixed potential condition equivalent to the surface ^-potential. In this case we choose the former resulting in the following boundary condition applied along the heterogeneous surface (31b) where © is the non-dimensional surface charge density ( 0 = a lye / Sy, kb T, where a is the surface charge density). Along the upper and side boundaries of the computational domain, F3, F5 and F6, symmetry conditions are applied while periodic conditions are again are applied at the inlet and outlet, F2 and F4. In order to determine the applied electric field, E(X), we enforce a constant current condition at each cross section along the x-axis as shown in Eq. 32(a):
Effects of Surface Heterogeneity on Electrokinetic Flow
255
Figure 5.26 Electroosmotic flow over a surface with lengthwise heterogeneous strips with a/,omD = -4xl(T 4 C/m2 and c/,etero = +4xl(T4 C/m2 in a lxlO"5 M (1/K = 0.1 urn) monovalent electrolyte solution, (a) top-view of the velocity vectors, (b) velocity variation in the channel width direction, (c) velocity profile in channel height direction.
256
Electrokinetics in Microfluidics
(32a) where nx in the above is a normal vector in the x direction and J* and J~ are the non-dimensional positive and negative current densities. Assuming monovalent ions for both positive and negative species (as was discussed earlier) and using the ionic species flux terms from Eq. 28(a) and Eq. 28(b), condition 32(a) reduces to the following,
(32b)
which is a balance between the conduction (term 2) and convection (term 3) currents with an additional term to account for any induced current due to concentration gradients. Solving Eq. 32(b) for the applied potential gradient yields:
(32c)
The above system of equations was solved by the finite element method [36] using the BLOCS (Bio-Lab-on-a-Chip Simulation) software developed by the Laboratory of Microfluidics and Lab-on-a-Chip at the University of Toronto. In the numerical calculations, the computational domain was discretized using 27-noded three-dimensional elements, which were refined within the double layer region near the surface and further refined in locations where discontinuities in the surface charge density were present. Triquadratic basis
Effects of Surface Heterogeneity on Electrokinetic Flow
257
Figure 5.27 Electroosmotic flow over a surface with crosswise heterogeneous strips with <ShOmo = -4xl(T 4 C/m2 and ahelero = +4xl(T 4 C/m2 in a lxlO"5 M monovalent electrolyte solution (1/K = 0.1 |j.m) (a) side view of velocity vectors in x-y plane (b) Magnitude of x direction velocity at various locations along the channel height.
258
Electrokinetics in Microfluidics
functions were used for all the unknowns ht', N~, Vx, Vy, Vz, and f. In general the use of these higher order elements and basis functions (as opposed to 8-noded trilinear elements) were found to be optimal as they could better approximate the sharp gradients in velocity, species concentration and electric potential within the double layer region. In all cases periodic conditions were imposed using the technique described by Saez and Carbonell [37]. The solution algorithm begins by forming the computational mesh and performing the elemental matrix integration. All solution variables were then set to zero with the exception of N+, N~ and W which were initially approximated with a classical Poisson-Boltzmann distribution [34,35]. Using these as initial guesses a semi-implicit, Newtonian iteration technique was used to solve Eqs. 28(a), 28(b) and 31 (a) simultaneously. In general it was found that the convergence could only be achieved when the double layer potential, W, in the non-linear terms of Eqs. 28(a) and 28(b) were fully implicit to the solution and the N^ and N~ terms explicit. Once convergence of all 3 variables was obtained, the forcing term in the Navier-Stokes Equations could be evaluated and Eqs. 29(a) and 29(b) were solved concurrently using a penalty method [36]. Following convergence of the flow solution, Eq. 32(c) was evaluated at each cross section to yield updated values of E(X). The above process was repeated until convergence of E(X) was obtained. First, let's apply the above model to the cases of length-wise and crosswise strip patterns as shown in Figures 5.24(a) and 5.24(b). [14] considered a case with a sinusoidally varying surface charge density as opposed to the step changes examined here, however, as will be shown, a similar flow structure exists. To do this we consider a computational domain with dimensions lx = 50^im, ly = 4 = 25 urn (the x-y-z co-ordinate system is the same as in Figure 5.25) containing a 10~5 M KC1 solution (thus a double layer thickness, 1/K, of 0.1 urn) and an applied driving voltage of 500 V cm"1. Chemical properties of K+ and Cl" ions were taken from Vanysek [46]. We consider a homogeneous surface charge density of o/,omo = -4x10~4 C m~2 and the heterogeneous strips with Ghetero = +4x10"4 C nf2 (corresponding to ^-potentials of approximately -50 mV and +50 mV, respectively). This implies an excess of positive ions over the homogeneous region and negative ions over the heterogeneous region. The results for the length-wise strips are shown in Figure 5.26 and Figure 5.27 for the crosswise strips. Figure 5.26(a) shows a top view of the velocity vectors at the symmetry plane (F3, ly = 25 urn). As was observed by Stroock et. al. [18] the flow is directed along the negative x-axis over the heterogeneous patch and the positive x-axis over the homogeneous patch, in both cases becoming weaker and approaching zero near the transition point. Also of interest is the change in the
Effects of Surface Heterogeneity on Electrokinetic Flow
259
Figure 5.28 Electroosmotic flow streamlines over a patchwise heterogeneous surface pattern with ohomo - -4x10~ 4 C/m2 and (a) a/,etera = -2x1 (T4 C/m2 (b) ahetem = +2x1 CT4 C/m2 (c) <Shetem = +4x10"4 C/m2. The grey scale represents magnitude of velocity perpendicular to the applied potential field, scaled by the maximum velocity in a homogeneous channel. Arrow represents direction of applied electric field.
260
Electrokinetics in Microfluidics
magnitude of the induced flow velocity with increasing distance from the surface. As can be seen in Figures 5.26(b) and 5.26(c) the velocity reaches a maximum at a distance of approximately 3/K from the surface (which is approximately equivalent to the edge of the double layer where the net charge density is essentially zero) and then becomes lower and lower as the distance from the double layer increases. From Figure 5.26(b) it is also apparent that velocity profile changes from a near step change within the double layer, where the electroosmotic force is strongest, to a significantly more curved and flatter profile as the distance from the surface becomes greater and viscous forces begin to dominate. Extrapolating from these results, one would expect the flow, in a very large channel, to eventually reach a near stagnant flow at the mid-plane, which was an observation originally made by Ajdari [14] for the sinusoidal varying surface pattern. As is shown in by the velocity vectors in Figure 5.27(a), the presence of crosswise heterogeneous strips results in a series of alternating clockwise and counter-clockwise tumbling regions, or regions of flow circulation whose centre axis is perpendicular to the direction of the applied electric field. An analogous effect was predicted by Ajdari [14] for the sinusoidal surface pattern and observed experimentally by Stroock et. al. [18] for a similar strip pattern, lending further credibility to the proposed model. Figure 5.27(b) shows the magnitude of the x-direction velocity at increasing distances from the surface. Similar to the previous case a near step change in vx is observed at the discontinuity in surface charge density within the double layer and the maximum velocity is obtained at a distance of 3/K from the surface. At further distances from the surface, viscous forces begin to dominate and momentum diffusion tends to broaden this step change while reducing the magnitude of the velocity eventually resulting in the sinusoidal type pattern above the circulation axis at the symmetry plane (y = 25um). Next, let's consider the two-dimensional patchwise surface pattern shown in Figure 5.24(c) and look into the resulting three-dimensional flow structure. We consider a lx = 50um, ly = l7 = 25 urn computational domain, as shown in Figure 5.25, containing a 10~5 M KC1 solution and with an applied driving voltage of 500 V cm"1 and a ahomo = -4xlO~4 C/m2. The simulations revealed three distinct flow patterns, depending on the degree of surface heterogeneity, as shown in Figure 5.28. The grey scales in these figures represent the magnitude of the velocity perpendicular to the direction of the applied electric field (v2 = vy2 + v/) scaled by the maximum velocity in a homogeneous channel, which in this case is 1.7 mm/s. At low degrees of surface heterogeneity (o),e/ero < -2x10"4 C/m2) the streamline pattern shown in Figure 5.28(a) was obtained. As can be seen a net
Effects of Surface Heterogeneity on Electrokinetic Flow
261
Figure 5.29 Velocity vectors in the x-z plane over a patchwise heterogeneous surface with Ghomo = -4x10" 4 C/m2 and ahelero = +2x10"4 C/m2 at distances of (a) 0.05 um, (b) 1.5 um and (c) 10.5 Um above the surface.
262
Electrokinetics in Microfluidics
counter-clockwise flow perpendicular to the applied electric field is present at the first transition plane {i.e. at the initial discontinuity in the heterogeneous surface pattern) and a clockwise flow at the second transition plane. This flow circulation is a pressure-induced effect. It results from the transition from the higher local fluid velocity (particularly in the double layer) region over the homogeneous surface on the right hand side at the entrance to the left hand side after the first transition plane (and vice versa at the second transition plane). To satisfy continuity then there must be a net flow from right to left at this point, which in this case takes the form of the circulation discussed above. The relatively straight streamlines parallel with the applied electric field indicate that at this level the heterogeneity is too weak to significantly disrupt the main flow. While a similar circulation pattern at the transition planes was observed as the degree of surface heterogeneity was increased into the intermediate range (-2x1 (T4 C/m2 < Ohetero < +2x10~4 C/m2), it is apparent from the darker contours shown in Figure 5.28(b) that the strength of the flow perpendicular to the applied electric field is significantly stronger reaching nearly 50% of the velocity in the homogeneous channel. Unlike in the previous case it is now apparent that the streamlines parallel with the applied electric field are significantly distorted due to the much slower or even oppositely directed velocity over the heterogeneous patch. Figure 5.29 shows the evolution of the flow pattern for this case with increasing distance from the surface. As was noted above, with respect to the length-wise strip pattern, the flow patterns parallel with the surface tend to degrade as the distance from the double layer region increases. This effect is again visible here, as a clear circulation pattern can be observed very near the surface with the flow in the opposite direction to the applied electric field over the heterogeneous patch, Figure 5.29(a). As the distance from the surface is increased the circulation becomes stronger reaching a maximum near the edge of the double layer region, Figure 5.29(b). Farther out from the surface, Figure 5.29(c), the circulation pattern is lost leaving a more uniform flow field with only a slightly disturbed velocity over the heterogeneous regions. At even higher degrees of heterogeneity (+2x1 CT4 C/m2< uhetew< +4x10"4 2 C/m ) a third flow structure is observed in which a dominant circulatory flow pattern exists along all three co-ordinate axes, as shown in Figure 5.28(c). This results in a negligible, or even non-existent, bulk flow in the direction of the applied electric field (which is to be expected since the average surface charge density for these cases is very near zero). As indicated by the grey scales, the velocity perpendicular to the flow axis has again increased in magnitude, reaching a maximum at the edge of the double layer near the symmetry planes at the location where a step change in the surface charge density has been imposed. As electroosmotically driven flow is an inherently surface driven effect, it is of significant practical interest (in mixing applications for example) to
Effects of Surface Heterogeneity on Electrokinetic Flow
263
Figure 5.30 Influence of heterogeneous patch size on velocity vectors in the F(, plane over a patchwise heterogeneous surface with a/,omo = -4x10" 4 C/m2 and <3haero = +4x10~4 C/m2. (a) heiem = 25um (b) \heten = 15um and (c) \helem = 5um.
264
Electrokinetics in Microfluidics
examine how deeply the circulation patterns observed above will penetrate into the bulk flow. We have already shown in Figure 5.29 that the near surface flow structure tends to degrade as the distance from the double layer becomes greater and we now seek to better quantify this effect in terms of the size of the the heterogeneous patches and the channel height. In Figure 5.30 the velocity vectors along the- F6 plane are shown for the case of oppositely charged patches (as in the case shown in Figure 5.28(c)) with sizes ranging from 25um by 25um square to 5um by 5(j.m square. For all cases a 25 um channel half height (ly) is maintained, however the length (lx) and width (/z) of the computational cell were appropriately adjusted as the patch size was changed. From these figures two interesting effects can be observed. Firstly it can be seen that as the size of the heterogeneity is decreased, the flow disturbance is confined to a thinner region near the wall, leaving a near stagnant flow in the centre of the channel (as one would expect since the average surface charge density is zero). In general it was observed that for the oppositely charged case the penetration depth of the flow disturbance is approximately equivalent to the size of the heterogeneous patch. It was also observed (not shown) that as the degree of heterogeneity was decreased so was the penetration depth and thus the oppositely charged case shown in Figure 5.30 represents a maximum. Secondly it is also apparent from the size of the vectors that as the heterogeneous patch size is decreased so is the magnitude of the flow disturbance. The coupling of these two effects suggest that significant flow disturbance is likely to be observed only when the heterogeneous patch size is of the same order of magnitude as the channel height. It is well known that the velocity of an electroosmotic flow is linearly proportional to magnitude of the applied electric field (as can be deduced from Eq. 29(a)) and there is no need to expand on this point here, other than to say that the fluid velocity in the flow patterns shown above obeyed identical scaling laws. Similarly, proportionally increasing the magnitude of o>,omo and Ohetem, increased the magnitude of the flow circulation, but had no effect on the overall flow structure. As a result the flow structures shown in the preceding figures are quite general and are not significantly affected by either of these quantities. Further simulations revealed that, for the cases examined here where the thickness of the double layer is significantly less than the channel height and the heterogeneous patch size, the double layer thickness (or equivalently the bulk ionic concentration, n^) had only a minor effect on the flow field. For thicker double layers (for example n^ = 10~6 M, 1/K = 1 um) a slight reduction in the magnitude of the flow field was observed (at equivalent surface potentials), however the overall flow structure was not significantly altered. The full Nernst-Planck-Poisson formulation used here (as opposed to the traditional Poisson-Boltzmann equation) allows us to examine how the electrophoretic influence of the applied electric field and the convective effects
Effects of Surface Heterogeneity on Electrokinetic Flow
265
Figure 5.31 Influence of applied electric field on net charge density distribution in the electric double layer with c t e m o = -4xlO~4 C/m2, <jhetero = +lxlO~4 C/m2 in a l x l t r 5 M (1/K = 0.1 u.m) electrolyte solution, (a) y = 0 |im (b) y = 0.05 u.m and (c) y = 0.2 u.m.
266
Electrokinetics in Microfluidics
of the irregular flow structures will disturb the double layer structure. Figure 5.31 shows the scaled net charge density (n+ - n~)lnx distribution along the F6 surface at different y positions for the computational domain discussed above with Ghomo = -4xl(T 4 C/m2 and ohetero = +4xlO"4 C/m2. From Figure 5.3l(a) it is apparent that the near surface net charge density field becomes slightly distorted compared with the Poisson-Boltzmann distribution (which is represented by the Vapp = 0 case) as the applied electric field is increased. In an earlier work concerning pressure driven flow over heterogeneous surfaces [12,22] it was observed that at Reynolds numbers greater than 10 and a bulk ionic strength less than 10~5 M, convective effects can significantly distort the double layer field from the traditional Poisson-Boltzmann distribution. For the electroosmotic case examined here, the highest Reynolds number obtained was Re = 0.09 and thus the convective terms in Eqs. 28(a) and 28(b) could not have significantly influenced the result. The cause of the disturbance observed here is the electrophoretic influence of the applied electric field (which is orders of magnitude stronger than the induced streaming potential in the pressure driven case). In this case the applied field induces a net flux of positive ions from left to right and negative ions from right to left (see Eqs. 28(a) and 28(b)). As a result an increase in the magnitude of the net charge density is observed near the first step change in surface potential as the ion flux from the left (where positive ions are in excess) meets the ion flux from the right (where negative ions are in excess). In contrast at the second transition point a decrease in the net charge density is observed since in both cases the net ion flux is away from this point. As is apparent from Figures 5.31(b) and 5.31(c) however this effect is limited to the region very near the surface as double layer field is shown to resemble very closely a Poisson-Boltzmann distribution at distances of 0.5/K and 2/K from the surface. In general this rearrangement of the net charge density field had only a small affect on the local fluid velocity, in that the calculated velocity was slightly higher in the regions where the net charge density was increased and vice versa. As was noted above, however this did not have a significant influence on the overall flow structure. In summary, the presence of periodically heterogeneous patches is shown to induce three distinct flow structures depending on relative difference between the surface charge density of the homogeneous and the heterogeneous regions. Small differences in the charge density are shown to induce fluid motion perpendicular to the applied electric field however the bulk flow remains largely unaffected. As the degree of heterogeneity is increased the streamlines in the direction of the applied electric field become significantly distorted, however the effect is minimized as the distance from the surface is increased. When oppositely charged surfaces are encountered a strong circulatory flow regime is observed in which flow perpendicular to the applied electric field is of the same
Effects of Surface Heterogeneity on Electrokinetic Flow
267
order as that parallel with it and the net volume flow rate becomes negligible. It is also shown that the induced flow patterns are limited to a layer near the surface with a thickness equivalent to the length scale of the heterogeneous patch and that as the average size of the heterogeneous region decreases the effect becomes smaller in magnitude and more localized. In addition it was demonstrated that while convective effects are small, the electrophoretic influence of the applied electric field could distort the net charge density field near the surface, resulting in a significant deviation from the traditional Poisson-Boltzmann double layer distribution.
268
5-5
Electrokinetics in Microfluidics
SOLUTION MIXING IN T-SHAPED MICROCHANNELS WITH HETEROGENEOUS PATCHES
As demonstrated in the previous sections, the presence of heterogeneous patches on a surface can result in local circulation within the bulk flow near these heterogeneous patches. Creating these bulk flow circulation regions can be explained by considering a simple model illustrated in Figure 5.32. In this figure, electroosmotically induced flow over a homogeneous surface with a zeta potential C, = -\q o\ is compared with that for a surface with an oppositely charged heterogeneous patch characterized by C, = + \q 0\. For the homogeneous case, Figure 5.32a, the electroosmotic body force applied to the liquid continua within the double layer is equivalent at each point along the flow axis and results in a constant bulk liquid velocity at the edge of the diffuse double layer, veo, which can be described by [47]: (33) where neo = (ewg I JJ.) is the electroosmotic mobility, Sy, is the electrical permittivity of the solution, JJ. is the viscosity, q is the zeta potential of the channel wall, and <j> is the applied electric field strength. The same relation holds true for the heterogeneous surface in Figure 5.32b except the electroosmotic body force is now applied to a region with excess negative ions and thus is in the opposite direction to that in the homogeneous regions. The interaction of these local flow fields with the bulk flow results in the regional circulation zones, as is illustrated in Figure 5.32b. In this section, the effect of surface heterogeneous patches on microfluidic mixing [22] will be discussed. Let's consider a simple T-shaped microfluidic mixing system, as shown in Figure 5.33. This simple arrangement has been used for numerous applications including the dilution of a sample in a buffer [48], the development of complex species gradients [49,50], and measurement of the diffusion coefficient [51]. In addition the T-sensor has also been exploited to make diagnostic determinations of analyte concentrations [52] by introducing the analyte of interest in one stream to a receptor molecule in the other (e.g. a pH [53] or fluorescence [54] indicator) producing a measurable signal which can be related to the parameter of interest [51]. Generally most microfluidic mixing systems, whether pressure or electrokinetically driven, are limited to the low Reynolds number regime and thus species mixing is strongly diffusion dominated, as opposed to convection or turbulence dominated at higher Reynolds numbers. Consequently mixing tends to be slow and occur over
Effects of Surface Heterogeneity on Electrokinetic Flow
269
relatively long distances and time. As an example the concentration gradient generator presented by Dertinger et. al. [49] required a mixing channel length on the order of 9.25mm for a 45um x 45(im cross sectional channel or approximately 200 times the channel width to achieve nearly complete mixing. Enhanced microfluidic mixing over a short flow distance is highly desirable for lab-on-a-chip applications. One possibility of doing so is to utilize the local circulation flow caused by the surface heterogeneous patches. Without loss of the generality, we will consider that two electrolyte solutions of the same flow rate enter a T-junction separately from two horizontal microchannels, and then start mixing while flowing along the vertical microchannel, as illustrated in Figure 5.33. The flow is generated by the applied electrical field via electrodes at the upstream and the downstream positions.
Figure 5.32. Electroosmotic flow near the double layer region for (a) a homogeneous surface (C, = -\C,0\) and (b) a homogeneous surface with a heterogeneous patch (C, = +\C,0\). Over the heterogeneous patch, the excess cations are attracted to the positive electrode resulting in an electroosmotic flow in the opposite direction to that over the homogeneous regions with an excess anion concentration. Arrows represent streamlines and 1/K refers to the characteristic thickness of the electrical double layer.
270
Electrokinetics in Microfluidics
Figure 5.33. T-Shaped micromixer formed by the intersection of two microchannels, showing a schematic of the mixing or dilution process.
Electroosmotic flow results from the interaction of an applied electric field with the unbalanced ions in the electrical double layer (EDL) and is described by the Navier-Stokes Equations subject to an electroosmotic body force and the continuity Equation (given below in non-dimensional form): (34a) (34b) where Fis the non-dimensional velocity (V = v/veo, where veo is calculated using Eq. (33)), P is the non-dimensional pressure, r is the non-dimensional time and Re is the Reynolds number given by Re = pveoL/r\ where L is a length scale taken
Effects of Surface Heterogeneity on Electrokinetic Flow
271
as the channel width (w from Figure 5.33) in this case. In general the high voltage requirements limit most practical electroosmotically driven flows in microchannels to small Reynolds numbers, therefore to simplify Eq. (34a) we ignore transient and convective terms and limit ourselves to cases where Re < 0.1. Fe represents the non-dimensional electroosmotic body force give by, (34c)
where W represents the non-dimensional EDL field strength (W = zy/ e/kbT, where kb is the Boltzmann constant, T is the absolute temperature, e is the charge of an electron, and z is the ionic valence), <2> is the non-dimensional applied electric field strength (0 = §/§max, where ^>max is the maximum applied voltage) and the ~ symbol over the V operator indicates the gradient with respect to the non-dimensional coordinates (X = x/w, Y = y/w and Z = z/w). Apparently from Eq. (34c), evaluation of the electroosmotic body force requires a description of both f a n d 0, which can be given by Eq. (35) and Eq. (36) respectively, (35) (36) where K is the non-dimensional double layer thickness (K = KW, where K is the Debye-Huckel parameter, given by K: = (2naz2e2/swkbT)1/2 where n^ is the bulk ionic concentration of the aqueous medium. 1/K is usually referred to as the characteristic thickness of the double layer as shown in Figure 5.32.). Inherent in Eqs. (35) and (36) are a few assumption that are worth discussing. The description of the applied electric field by the homogeneous Poisson equation is a simplification valid only in cases where the bulk conductivity of the aqueous solution does not change significantly within the channel. In the majority of cases of fluidic transport in microchip devices the species of interest are transported in a buffer solution with an ionic concentration 100 times greater than that of the species. In these cases the buffer conductivity is assumed to dominate and thus remains constant, independent of the local species concentration. This assumption is also applicable to Eq. (35) that has been derived for a symmetric ionic species with concentration nx . n^ is referred to the buffer concentration and the contribution of the more dilute species of interest have been ignored. By decoupling the equations for the EDL and applied electric field, it has been assumed that the charge distribution near the
272
Electrokinetics in Microfluidics
wall is unaffected by the externally applied field. Along the same lines, the description of the EDL field by the non-linear Poisson-Boltzmann equation assumes that the double layer distribution is also undisturbed by external convective influences. These assumptions is generally valid so long as double layer thickness is not large, or equivalently the ionic concentration of the solution is not very low and the flow remains in the limit of low Re. Numerical simulation of electrokinetic flow and species mixing in a Tshaped mixer is complicated by the simultaneous presence of three separate length scales; the mixing channel length (mm), the channel cross sectional dimensions (um) and the double layer thickness, 1/K (nm), which we will refer to as L\, L2 and L3 respectively. In general the amount of computational time and memory required to fully capture the complete solution on all three length scales would make such a problem nearly intractable. Since the channel length and cross sectional dimensions (Lj, L2) are required to fully define the problem, most previous studies [55-57] have resolved this problem by either eliminating or increasing the length scale associated with the double layer thickness (L3). Both Bianchi et. al. [56] and Patankar et. al. [57] accomplished this by artificially inflating the double layer thickness to bring its length scale nearer that of the channel dimensions. This allowed them to fully solve for the EDL field, calculate the electroosmotic forcing term, and incorporate it into the NavierStokes Equation without any further simplification. It did however not fully eliminate the third length scale and considerable mesh refinement was still required near the channel wall. In these cases this approach was tractable largely because the geometries were selected such that the channel length (Lj) was not as large as that examined here. In a different approach Ermakov et. al. [55] set Fe = 0 in Eq. (34a) and applied a slip boundary condition at the channel, veo from Eq. (33). By this method the double layer length scale (L3) is completely eliminated from the solution domain, and a description of the double layer region is no longer required. A similar slip condition approach was used by Stroock et. al. [18,31] in their simulations of electrokinetically induced circulating flows. In both cases excellent agreement was obtained with experimental results. For the mixing problem interested here, the large mixing channel length and the extension to 3D mixing require that we follow the slip condition approach. As shown in Figure 5.33, we consider the mixing of equal portions of two buffer solutions, one of which contains a species of interest at a concentration, co. Species transport by electrokinetic means is accomplished by 3 mechanisms: convection, diffusion and electrophoresis, and can be described by Eq. (37), (37)
Effects of Surface Heterogeneity on Electrokinetic Flow
273
Figure 5.34. Electroosmotic streamlines at the midplane of a 50|im T-shaped micromixer for the a) homogeneous case with C, = -42 mV, b) heterogeneous case with six symmetrically distributed heterogeneous patches on the left and right channel walls and c) heterogeneous case with six offset patches on the left and right channel walls. All heterogeneous patches are represented by the crosshatched regions and have a C, = + 42mV. The applied voltage is <j>app = 500 V/cm.
274
Electrokinetics in Microfluidics
where C is the non-dimensional species concentration (C = c/c0, where c0 is original concentration of the interested species in the buffer solution.), Pe is the Peclet number {Pe = veow/D, where D is the diffusion coefficient), and Vep is the non-dimensional electrophoretic velocity equal to vep/veo where vep is given by: (38)
and juep ~(swgp/ri)
is the electrophoretic mobility (£w is the electrical
permittivity of the solution, \i is the viscosity, gp is the zeta potential of the tobe-mixed molecules or particles) [47]. As we are interested in the steady state solution, the transient term in Eq. (37) can be ignored. Before continuing it would be useful to clarify the relationship between the three different velocities discussed above: the electroosmotic velocity Veo (Eq. (33)), the bulk flow velocity V (Eqs. (34a) and (34b)), and the electrophoretic velocity Vep (Eq. (38)). As shown in Figure 5.32, Veo, is the induced velocity at the edge of the double layer caused by the application of the external electric field. As described above this quantity is used as a boundary condition on the Navier-Stokes equations, which are then used to determine the local velocity of the bulk flow at each point, V. While velocity at the boundary, Veo, governs the magnitude V, it is important to distinguish between these two quantities. The velocity Vep, represents the mobility of the species in the applied electric field. For a neutral species the electrophoretic velocity is zero, leaving V as the only velocity in Eq. (37). A charged species however will be attracted to either the positive or negative electrode, depending on the sign of the charge, at a speed described by Vep. In such a case the total species convection is described by the superposition of V and Vep as shown in Eq. (37). For further information on species mobility and its contribution to the convection diffusion equation (or more precisely the Nernst-Planck equation), the reader is referred to the texts by Lyklema [34,35]. In all simulations a square geometry was used with the depth equaling the width of the channel, w, and for consistency the arm length, Larm, was also assigned the value of w. The length of the mixing channel, Lmix, was dictated by that required to obtain a uniform concentration (i.e. fully mixed) at the outflow boundary. Depending on the simulation conditions this required Lmix to be on the order of 200 times the channel width. The applied electric field strength, Eq. (36) was solved subject to
Effects of Surface Heterogeneity on Electrokinetic Flow
275
Figure 5.35. 3D species concentration contours (upper image) and the midplane contours (lower image) for the 50|am T-shaped micromixer resulting from the flow fields shown in Figure 5.34. (a) homogeneous case, (b) heterogeneous case with symmetrically distributed heterogeneous patches, and (c) heterogeneous case with offset patches. Species diffusivity is 3x10~10 m2/s and zero electrophoretic mobility is assumed.
276
Electrokinetics in Microfluidics
simultaneously. Zero gradient inflow conditions were applied at the two inlet regions (dVJdX= 0, Vy = 0, Vz = 0) and similarly at the outflow boundary (Vx = 0, dVy/dZ = 0, Vz = 0). As mentioned above we consider the case of a stream with species concentration c = c0 mixing with a buffer, thus we assign C = 1 to the right inlet boundary (inlet stream 2) and C = 0 to the left inlet (inlet stream I)Eqs. (34a), (34b), (36) and (37) were all solved over the computational domain via the Finite Element method using 27-noded biquadratic brick elements for 0, V and C and 8-noded trilinear brick elements for pressure, using the BLOCS (Bio-Lab-on-a-Chip Simulation) software developed by the Laboratory of Microfluidics and Lab-on-a-Chip at the University of Toronto. In all cases the discretized systems of equations were solved using a quasi-minimal residual method solver and were preconditioned using an incomplete LU factorization. In addition to testing the program against standard test problems the electrokinetic transport simulations were validated against the numerical/experimental results published by Ermakov et. al. [55] The objectives here are to investigate the formation of electroosmotically induced circulation regions near surface heterogeneities in a T-shaped micromixer and to determine the influence of these regions on the mixing effectiveness. We consider a channel of 50um in width and 50um in depth, an applied voltage of (j)app = 500 V/cm (§app =<j>max/(Lmix + Lam)), and a mixing channel length of 15mm. These conditions are applied to all simulations discussed below unless otherwise specified. A homogeneous electroosmotic mobility of 4.0x10"8 mVVs is chosen, corresponding to a homogeneous C, potential of -42 mV. In Figure 5.34 we compare the mid-plane flow field near the T-intersection of (a) a homogeneous mixing channel, with that of a mixing channel having (b) a series of 6 symmetrically distributed heterogeneous patches on the left and right channel walls and (c) a series of offset patches also located on the left and right walls respectively. For clarity the heterogeneous regions are marked as the crosshatched regions in this and all subsequent figures. A C,potential of t, = +42mV was assumed for the heterogeneous patches. As expected both Figures 5.34b and 5.34c do exhibit regions of local flow circulation near these heterogeneous patches, however their respective effects on the overall flow fields are dramatically different. In Figure 5.34b it is apparent that the symmetric circulation regions force the bulk flow streamlines to converge into a narrow stream through the middle of the channel. The curved streamlines shown in Figure 5.34c show the more tortuous path through which the bulk flow passes as a result of the offset, non-symmetric circulation regions. Figure 5.35 compares both the 3D and channel midplane concentration profiles for the three heterogeneous arrangements shown in Figure 5.34. In all these figures a neutral mixing species (i.e. jj,ep = 0, thereby ignoring any
Effects of Surface Heterogeneity on Electrokinetic Flow
277
Figure 5.36. Species concentration contours in the cross section of the T-shaped micromixer at 200p.m downstream (upper image) and at 750u,m downstream (lower image) for the 3 cases shown in Figure 5.35; (a) homogeneous case, (b) heterogeneous case with symmetrically distributed heterogeneous patches, and (c) heterogeneous case with offset patches.
electrophoretic transport for the time being) with a diffusion coefficient D = 3x10~10 m2/s is considered. As expected both symmetric flow fields discussed above have yielded symmetric concentration profiles as shown in Figures 5.35a and 5.35b. While mixing in the homogeneous case is purely diffusive in nature, the presence of the symmetric circulation regions, Figure 5.35b, enables enhanced mixing by two mechanisms, firstly through convective means by circulating a portion of the mixed downstream fluid to the unmixed upstream region, and secondly by forcing the bulk flow through a significantly narrower region, as shown by the convergence of the streamlines in Figure 5.35b. In this second mechanism, the local concentration gradients are increased leading to a greater diffusive flux along the channel width (x-axis in Figure 5.33), N, given byEq.(39),
278
Electrokinetics in Microfluidics
(39)
where Ldi/f is the diffusion length scale. In the homogeneous case Ldiff is equivalent to the channel width, w. For the case shown in Figure 5.34b however the heterogeneous regions force the bulk flow to pass through a narrower region, resulting in L^ff < w and thereby increasing JV. This enhancement of the concentration gradients is apparent from Figures 5.36a and 5.36b, which compare the cross sectional concentration contours for these two cases at distances of 200u.ni (top plot) and 750um (bottom plot) downstream from the intersection. At 200um downstream (within the heterogeneous region) the convergence of the bulk flow into a narrow stream has led to the greater concentration gradients, as compared to Figure 5.36a, within this narrow band at the center of the channel. This results in much stronger diffusion and a more uniform species distribution further downstream, and thus better mixing, as shown in the 750um downstream contour plots. In Figure 5.35c the convective effects on the local species concentration is apparent from the concentration contours generated for the non-symmetrical, offset patch arrangement. In Figure 5.36c the influence of these offset patches on the mixing is apparent again leading to a similar enhancements in the concentration gradients at the 200um downstream point and a more uniform distribution at 750um downstream. In this case however the concentration distribution is no longer symmetric about the center axis. Influence of the heterogeneous ^-potential In the above, we have demonstrated qualitatively how the presence of heterogeneous regions can enhance the mixing process in a T-shaped micromixer. To quantify the degree of enhanced mixing, we introduce a mixing efficiency, e, given by Eq. (40),
(40)
where c is the cross sectional concentration profile at a distance y downstream and cx and c0 are the concentration profiles associated with a completely mixed and completely unmixed states respectively. A fully mixed state therefore would
Effects of Surface Heterogeneity on Electrokinetic Flow
279
have a 100% mixing efficiency while the unmixed state would have a 0% mixing efficiency. Figure 5.37 demonstrates the influence of the (^-potential of the heterogeneous region on the mixing efficiency for a 50um mixing channel with a 500um long heterogeneous region on all four channel faces (beginning at the intersection) for an electrically neutral species with D = 3x10~10 m2/s. The applied electric field strength is <j)app = 500 V/cm and the total mixing channel length is 15mm. As in the previous cases the channel has a (^-potential of 42mV in all regions outside the heterogeneous patches. From Figure 5.37 it is apparent that in all cases the introduction of heterogeneous patches does lead to an increase in the mixing efficiency over the homogeneous case. The largest improvement comes from the oppositely charged patch arrangement (C, = +42mV) where the length to achieve 50% mixing efficiency is decreased by nearly 30% (390|am vs. 550um). Similarly, the length to achieve 70% mixing is also decreased by 30% (820um vs. 1230urn).
Figure 5.37. Effect of the heterogeneous patch ^-potential on the mixing efficiency of a 50|o.m T-shaped micromixer with §ipp = 500V/cm. A single 500u,m patch on all four sides of the channel was used and the homogeneous (^-potential is -42m V. Species diffusivity is 3xlO~10 m2/s and zero electrophoretic mobility is assumed.
280
Electrokinetics in Microfluidics
In general it was observed that the circulation regions shown in Figures 5.34b and 5.34c were only present when the heterogeneous (^-potential was of opposite sign to that of the homogeneous surface. Additionally the size of the circulation regions was observed to increase as the magnitude of the heterogeneous (^-potential was increased. Referring to Figures 5.34, 5.35, and 5.36, larger circulation regions would force the bulk flow to pass through a narrower region leading to shorter local diffusion length and thus by Eq. (38) enhanced mixing. Despite the lack of circulation zones in the £ = -21mV and C, = OmV cases, a slight convergence of the flow streamlines to a narrower region at the center of the channel, caused by the localized slowing of the velocity field near the channel walls, was observed and proved to be sufficient to provide some degree of enhanced mixing. In Figure 5.37 it can also be seen that all cases exhibit an approximate step change in the mixing efficiency immediately after the heterogeneous region. This occurs as a result of the higher concentration gradients in the center of the channel making the mixing efficiency appear artificially low. Referring to Figure 5.36b it can be seen that within the heterogeneous region the gradients in the concentration profile occur over a relatively thin band in the middle of the channel cross section, leading to the improved mixing, as discussed above. Outside of this narrow band the species appear nearly unmixed resulting in an artificially low mixing efficiency. When the streamlines diverge rapidly at the termination of the heterogeneous patch, see Figure 5.34b, the concentration gradients are expanded to encompass the full channel width, as shown in the lower image in Figure 5.36b. At this point the true mixing efficiency is obtained, demonstrating the overall effect of the heterogeneous patch. This effect is also well demonstrated on the 2D image in Figure 5.35b where the expansion of the concentration gradients after the termination of a heterogeneous patch and the contraction at the beginning of the heterogeneous patch both are well shown. Also in Figure 5.37 it can be observed that the C, = +42mV case has a significantly higher mixing efficiency at the intersection (downstream distance = 0|am) than the other cases. This immediate enhancement is a result stronger circulation zone convecting a greater portion of the mixed downstream solution upstream to the entrance. Interestingly however this effect appears to be localized to the heterogeneous region, as the absolute difference between the mixing efficiencies for the C, = +42 mV and C, = +21mV cases outside the heterogeneous region is no greater than one would expect due to the enhanced concentration gradients (based on the trend from the lower C, potential simulations). Therefore, while the re-circulation mechanism does provide significant improvement of the initial mixing efficiency (at the beginning of the heterogeneous region), it appears as though the enhanced concentration gradients are the dominant mechanism for overall enhanced mixing.
Effects of Surface Heterogeneity on Electrokinetic Flow
281
Influence of heterogeneous patch size and arrangement Figures 5.38a and 5.38b show the influence of the heterogeneous patch size on the mixing efficiency for two different applied voltages, (a) <j)app = 500 V/cm and (b) tj>app = 200 V/cm. As before we consider the case of a single heterogeneous patch with £ = +42mV on all four sides of a homogeneous surface with C, = -42mV, a 50u.m by 50um cross section channel and D = 3xlO~10 m2/s. Apparent in both cases is that the larger patches, exposing the solution to the enhanced concentration gradients for a longer time, will increase the overall mixing efficiency. It is not surprising that as the app is decreased, resulting in slower flow in the lengthwise direction and allowing greater time for diffusion, the mixing efficiency is improved. Additionally the longer retention time in the heterogeneous region for the lower voltage case also serves to improve the effectiveness of the heterogeneous region, reducing the required mixing length to attain both 50% and 70% mixing efficiency by about 70%, compared with 30% for the larger applied voltage case. To investigate the effects of patch distribution we consider the two general arrangements shown in Figures 5.34b (symmetric) and 5.34c (offset) and the same transport conditions as those detailed above with app = 500V/cm. We consider both four patch and six patch arrangements (Figures 5.34b and 5.34c represent the 6 patch arrangement). The patch sizes are adjusted so that in each case a total of 250um of heterogeneous surface is spread out over the first 500um distance downstream of the intersection on both the left and right sides (e.g. in Figure 5.34c there are three 83.3 um patches on the right surface for a total of 250um of heterogeneous surface). The resulting mixing efficiencies for these cases are compared in Figure 5.39. In Figure 5.39 it is apparent that, in terms of the mixing efficiency outside the heterogeneous region, for both cases there is almost no difference between the 4 and 6 patch arrangements. In addition, very little difference between the symmetric and offset distributions is observed, suggesting that mixing efficiency is more a function of patch size, rather than the arrangement. The most striking feature of the results displayed in Figure 5.39 is the large oscillations in the mixing efficiency within the heterogeneous region, most significantly for the symmetric patch case. As discussed above, near step changes in the mixing efficiency are observed at the termination of a heterogeneous patch due to the expansion of the concentration gradients. Referring to Figure 5.35b it is apparent that as the flow passes through multiple heterogeneous regions the concentration gradients experience numerous expansion and contraction zones resulting in the observed oscillations in the mixing efficiency. Since the concentration gradients are compressed into a significantly smaller region than is observed in the homogenous case, as is
282
Electrokinetics in Microfluidics
Figure 5.38. Effect of the heterogeneous patch size on the mixing efficiency of a 50u.m T-shaped micromixer with (a) (j>app = 500V/cm and (b) §app = 200V/cm. A single patch on all four sides of the channel with C, = +42mV was used and the homogeneous (^-potential is -42mV. Species diffusivity is 3xlO~10 m2/s and zero electrophoretic mobility is assumed.
Effects of Surface Heterogeneity on Electrokinetic Flow
283
shown in the upper images of Figures 5.36a and 5.36b, it is not surprising that in some cases the mixing efficiency in the heterogeneous region drops below that of the homogenous channel. Influence of channel size Figure 5.40 compares the mixing efficiency in a 50 um x 50 urn channel with that in a 25 urn x 25 um channel for both the homogeneous and heterogeneous cases. As expected the smaller diffusion length, Ldiff in Eq. (39), in the 25 um channel leads to an increased mixing efficiency over that exhibited by the 50um homogeneous channel. Also in Figure 5.40, it is apparent that the influence of the heterogeneous patches is significantly greater for the 25 um channel, with 50%
Figure 5.39. Influence of the heterogeneous patch arrangement on the mixing efficiency of a 50um T-shaped micromixer with <|>app=500V/cm. The symmetric and offset patch arrangements are as shown in Figures 5.34b and 5.34c respectively (6 patch arrangement shown in both cases). The homogeneous ^-potential is -42mV and all heterogeneous patches have C, = +42mV. Species diffusivity is 3xlO~'° m2/s and zero electrophoretic mobility is assumed.
284
Electrokinetics in Microfluidics
Figure 5.40. Effect of channel size on the mixing efficiency for both homogeneous and heterogeneous cases with <j>apP = 500V/cm. For the heterogeneous cases a single 500um patch on all four sides of the channel was assumed with C, = +42mV and a homogeneous ^-potential of -42mV was used. Species diffusivity is 3x1 (T10 m2/s and zero electrophoretic mobility is assumed.
reductions in Lmix at both the 50% and 70% mixing efficiency points (compared to 30% for the larger channel) and resulting in nearly complete mixing at only 500um distance downstream. Influence of species diffusivity and electrophoretic mobility In the above discussions, we have confined the analysis to electrically neutral molecules, where the electrophoretic transport component in Eq. (37) has been neglected, with constant diffusion coefficient of D = 3xl0~10 m2/s. Now we seek to understand how the enhanced mixing efficiency is affected by the diffusivity and electrophoretic mobility of the mixing species. In Figures 5.41a and 5.41b we reexamine the case of a 50um x 50u.m channel and a single heterogeneous region with C, = +42mV on all four sides of the first 500u.m of the mixing channel. Figure 5.41a shows the effect of species diffusivity by
Effects of Surface Heterogeneity on Electrokinetic Flow
285
Figure 5.41. Influence of (a) species diffusivity and (b) electrophoretic mobility on the mixing efficiency in a 50nm channel for both homogeneous and heterogeneous cases with (j>app = 500V/cm. For the heterogeneous cases a single 500um patch on all four sides of the channel was assumed with t, = +42mV and a homogeneous ^-potential of -42m V was used.
286
Electrokinetics in Microfluidics
examining diffusion coefficients ranging from 1x10 10 m2/s to3xl0 10 m2/s. As in the previous cases an almost step change in the mixing efficiency is observed at the termination of the heterogeneous region (both Figures 5.41a and 5.41b) due to the expansion of the channel concentration gradients as discussed in detail previously. From Eq. (39) it is apparent that for lower D values the species flux will decrease and thus the observed lower mixing efficiency should result. This is clearly observed in Figure 10a where significantly lower mixing efficiencies are observed for the D = lxlO"10 m2/s case for both homogenous and heterogeneous channels. Also apparent in Figure 5.41a is that while in all cases the introduction of surface heterogeneity does improve the mixing efficiency, the absolute reduction in the mixing length is approximately equal for all pairs. This suggests that the effects of a higher diffusion coefficient on the mixing efficiency is neither enhanced nor degraded due to the presence surface heterogeneity. Figure 5.41b illustrates the influence of the electrophoretic mobility on the mixing in both homogeneous and heterogeneous channels, for the identical simulation conditions listed above and D = 3xl0~10 m2/s. Since a negative electrophoretic mobility increases the magnitude of the convection term in Eq. (37) (since Fand Vep will be in the same direction) whereas a positive mobility tends to reduce its magnitude, it would be expected that at a fixed downstream distance the mixing efficiency of the later of these would be considerably higher than that of the former. This is well demonstrated in Figure 5.41b where both heterogeneous and homogenous surfaces exhibit significant increases in the mixing efficiency as fiep is increased from -1.5xl0~8 m2/Vs to +1.5xl0~8 m2/Vs. Of interest in this plot is the significant increase in the initial mixing efficiency exhibited by the positive mobility case with the heterogeneous region. This significant initial enhancement is a result of the superposition of the electrophoretic and electroosmotic transport in the heterogeneous region near the wall (where the local flow direction is opposite that of the bulk flow) allowing a greater circulation of downstream ions into the unmixed initial solution. In summary, for low Reynolds number electroosmotically driven flows in microfluidic devices, species mixing is inherently diffusion dominated, resulting in poor mixing efficiency or requiring long transport distances and retention time. This section discussed the effects of surface electrokinetic heterogeneity on the electroosmotic flow mixing efficiency of a T-shaped micromixer. While all cases of surface heterogeneity were shown to enhance species mixing, the greatest improvements were found when the ^-potential of the heterogeneous surface was of opposite sign to that of the homogeneous surface, resulting in localized circulation zones within the bulk flow field. The local flow circulation enhances the mixing by firstly convection means by circulating a portion of the mixed downstream fluid to the unmixed upstream region, and secondly forcing the bulk flow through a significantly narrower region to increase the local
Effects of Surface Heterogeneity on Electrokinetic Flow
287
concentration gradients. For a fixed ^-potential, the length of the heterogeneous patch was found the most important factor while the patch distribution had only marginal effects on the mixing efficiency. In general the mixing efficiency improvement by decreasing the applied voltage and the channel size will be enhanced through the introduction of surface heterogeneity, in some cases resulting in a 70% reduction in the required mixing length. The combined effect of the heterogeneous patch and a positive electrophoretic mobility was shown to be significant in enhancing the initial mixing efficiency at the T-intersection.
288
5-6
Electrokinetics in Microfluidics
HETEROGENEOUS SURFACE CHARGE ENHANCED MICRO-MIXER
Enhancing the diffusion-dominated microscale mixing processes is key to numerous lab-on-a-chip applications. For applications where the use of external power supplies, actuators and auxiliary equipment are permitted, mixing schemes based on a variety of technologies such piezoelectrics [58,59], pneumatics [60], acoustic radiation [61], and oscillating voltage [62], have been successfully developed for microfluidic devices. In many applications, however, where ease of manufacturability, simplicity of design and the portability of the devices take precedence, passive mixing schemes, which do not rely on external stimulation or intricate components, are preferable. Two approaches to flow manipulation are prevalent in passive micromixers, the first relying on channel geometry to generate chaotic advection and increased circulation, and the second on channel surface properties. Notable initiatives, amenable to intermediate Reynolds number flows (i.e. >1) of the geometric category, have included helical or zigzag microchannels, in which mixing occurs as a result of eddies present at the channel bends [63-65]. Alternatively, oblique ridges formed on the bottom of the channel and designed to introduce a transverse component of flow have also proven effective for pressure-driven flows with Reynolds numbers between 0 and 100 and channel lengths on the order of centimetres [66] This technique has also demonstrated a negligible increase in bulk flow resistance, a constraint limiting the effectiveness of mixers based on parallel lamination for electrokinetic applications [67]. Recently a passive electrokinetic micro-mixer based on the use of surface charge heterogeneity was developed [68]. The surface charge patterns are designed to create localized flow disturbances thereby inducing an advective component in the flow to enhance the mixing. This section will discuss the experimentally observed effects of surface charge heterogeneity on microscale species transport, and the performance of the developed micro-mixer in terms of mixing efficiencies and required channel lengths. The micro-mixer is a T-shaped microchannel structure made from Polydimethylsiloxane (PDMS) and glass. Microchannels were fabricated using a rapid prototyping/soft-lithography technique [69]. Specifically, photomasks were designed in AutoCAD, exported as PDF files, and printed on a 3500 dpi image setter. Glass slides were soaked overnight in acetone, dried on a hot plate at 200°C, exposed to oxygen plasma (Harrick Plasma Cleaner model PDC-32G) for 5 minutes and again heated to 200°C subsequently, in order to prepare the surface for coating. 1.5 mL of SU-8 25 photoresist was then distributed onto each slide and degassed in a high vacuum. The photoresist was spin-coated at 1800 rpm for 10s, and at 4000 rpm for 40s with a ramping phase of 5 s between stages in order to obtain a smooth film with a thickness of 8 microns (Special
Effects of Surface Heterogeneity on Electrokinetic Flow
289
Coating System Spin Coat model G3P-8). Films were baked at 65°C for 3 minutes and at 95 °C for 7 minutes to harden. The photomasks were positioned on the photoresist films and exposed to UV light for 1 min. A two-stage postexposure bake at 65°C for 1 minute and 95°C for 2 minutes was then conducted. Masters were developed in 4-Hydroxy-4-methyl-2-pentanone for approximately 2 minutes or until the photoresist rinsed cleanly off. Subsequent to developing, masters were placed under a heat lamp for several hours. To form the microchannels, PDMS was poured over the masters and cured at 65 °C for approximately 2 hours at a pressure of-34 kPa (gauge). The rapid prototyping/soft lithography technique as described above allowed for control of and flexibility in surface charge patterning configurations, examples of which are detailed in Figure 5.42. To selectively pattern the surface charge heterogeneity, the following procedure was followed (Figure 5.43). The PDMS master, featuring a channel configuration corresponding with the pattern of heterogeneities to be examined, was reversibly sealed to a glass slide. Using suction forces with an approximate flow rate of 2.5mL/min, the PDMS master was flushed sequentially with 0.1M sodium hydroxide for 2 minutes, deionized water for 4 minutes, and 5% Polybrene solution for 2 minutes, resulting in selective regions of positive surface charge while leaving the majority of the glass slide with its native negative charge. This polyelectrolyte coating
Figure 5.42 Surface charge patterning configurations with a mixing region length L consistent for all configurations, a patch length 1, a patch width w, and patch spacing s for (A) In-line pattern (B) Staggered pattern (C) Serpentine pattern (D) Herringbone pattern and (E) Diagonal pattern.
290
Electrokinetics in Microfluidics
procedure is based on that developed by Liu et al. [70] for capillary electrophoresis microchips. All fluid was then removed from the channel and the system was left undisturbed for 40 minutes before flushing again with water for 20 minutes. To promote bonding of the polymer to the surface, the slide was aged in air for 24 hours prior to use. Before removing the PDMS master, the location of the heterogeneities was landmarked. A T-shaped microchannel, 200um in width and approximately 8 um in depth, was then permanently sealed to the glass slide such that the patterned surface heterogeneities were appropriately positioned within the mixing channel. Using the well-established current monitoring technique [71], the zeta potentials of the native-oxidized PDMS and the polybrene-coated surface were determined using a sodium carbonate/bicarbonate buffer of pH 9.0. Briefly, a dilute buffer was introduced into a uniaxial PDMS channel with the surface treatment of interest. The current was monitored and allowed to stabilize under a constant applied voltage potential. A concentrated buffer was then introduced through one reservoir and the time required for the current to reach a new plateau corresponding to the higher concentration of buffer was recorded. The electroosmotic mobility of the native-oxidized PDMS was determined to be -5.9 x 10 4 cm2/V-s compared with 2.3 x 10~4 cm2/V-s for the polybrene-coated surface, results that compared well with previous findings [70].
Figure 5.43 Schematic of selective patch patterning procedure (1) PDMS microchannel master featuring a channel configuration corresponding to a staggered pattern of heterogeneities and reversibly sealed to a glass slide. (2) PDMS master flushed with chargereversing solutions to selectively pattern charge heterogeneities. (3) Landmarking of surface heterogeneities achieved by aligning the corresponding photomask under 5x magnification with the PDMS master and firmly attaching it to the opposite side of the slide. (4) Removal of PDMS master and permanent sealing of oxidized T-channel to the glass slide such that the surface heterogeneities are appropriately positioned within the mixing channel. (5) Removal of attached photomask and completion of micro-mixer fabrication.
Effects of Surface Heterogeneity on Electrokinetic Flow
291
The effect of the surface charge patterns on mixing efficiency was then examined in T-shaped microfluidic chips. To perform the mixing experiments, 25 mM sodium carbonate/bicarbonate buffer and 100 uM fluorescein, were introduced through either inlet channel. The system was illuminated by a mercury arc lamp equipped with a fluorescein filter set. The steady state transport of the dye was observed using a Leica DM LM fluorescence microscope with a 5x objective, and captured using a Retiga 12-bit cooled CCD camera. Digital images were obtained by QCapture 1394 and OpenLab 3.1.5 imaging software at an exposure time of Is. The acquired images were of resolution 1280 x 1024 pixels with each pixel representing a 2.5 micron square in the object plane. Images were exported in TIF format to MATLAB for digital processing. Dark field subtraction and bright field normalization was performed to eliminate anomalies introduced by the image acquisition system. Following image processing, concentration profiles prior to and subsequent to the heterogeneous region were developed directly from values of pixel intensities. Profiles were smoothed using a convolution filter and linearly scaled to range between 0 and 1. Mixing efficiency, 8, was calculated and compared with homogeneous systems (e.g., a uniform and negatively charged PDMS/glass system) using the following definition:
where Cx =0.5 corresponds with perfect mixing on a normalized scale, Co is the concentration distribution over the channel width, W, at the channel inlet and C is the concentration distribution at some distance downstream. In order to facilitate optimization of surface charge patterning, the BLOCS (Bio-Lab-On-a-Chip Simulation) finite element code (see discussions in the previous section) was used to simulate the experimental conditions and surface charge heterogeneity. Specifically, channel dimensions of 200 um in width and 8 um in depth were modelled along with the physiochemical properties of fluorescein, namely a diffusion coefficient D = 4.37 x 10~10 m2/s, and a electrophoretic mobility uep = 3.3 x 10~8 m2/Vs. The numerical model described in the previous section was used to optimize the pattern of surface charge heterogeneity and to characterize the effects of channel depth, zeta potentials, and diffusivity on mixing enhancement. Simulations indicated that a nonsymmetrical configuration of oppositely charge surface heterogeneities is optimal for mixing enhancement. Figure 5.44 presents the numerical results for
292
Electrokinetics in Microfluidics
the variety of configurations as detailed in Figure 5.42. For each configuration, the patch length and spacing parameters were selected to maintain a constant ratio of heterogeneous to homogeneous surface areas over a channel length of 1.8 mm. Evidently, the non-symmetrical patterns, namely the staggered and the diagonal, generated better mixing distributions in comparison to the symmetrical herringbone and in-line arrangements. With a theoretical mixing efficiency of 96%, the staggered configuration provided the greatest degree of mixing, outperforming the diagonal, the herringbone and the serpentine configurations by 8%, 31% and 36% respectively. In comparison with the homogenous case, the staggered configuration provided a 61 % increase in mixing efficiency. The staggered configuration was thus examined in more detail to determine the optimal patch length and spacing parameters. For a 1.8 mm mixing region length, the optimal patch length was determined to be 300 urn with a two-fold increase or decrease from this optimal value resulting in a 10% reduction in mixing efficiency. In comparison, additional design factors proved insignificant, with a two-fold and five-fold increase in channel depth resulting in a 1 % and 9% decrease in mixing efficiency respectively, while a difference of
Figure 5.44 Numerically simulated concentration profiles across the channel width for the staggered pattern subsequent to the mixing region for an applied potential of 280 V/cm for varying configurations of surface charge heterogeneity as defined in Figure 5.42.
Effects of Surface Heterogeneity on Electrokinetic Flow
293
2.6% resulted from variations in spacing parameters. Values of electroosmotic motility in both the homogeneous surface and patch regions were critical for defining flow characteristics. As has been discussed in the previous section, for significant flow circulation, the patch electroosmotic mobility must be of opposite sign. Simulations revealed however that increasing the magnitude of the patch mobility above that generated by the polybrene layer does not significantly increase mixing efficiency, with a two fold increase resulting in a less than 2% change in mixing efficiency. Challenges of manufacturing may limit precise control of zeta-potentials, however as the presence of oppositely charged surface heterogeneities rather than their magnitude is the defining factor for mixing enhancement, charge altering treatments such as the polybrene coating used here should prove adequate. Based on the above analysis, the optimized micro-mixer consisting of 6 offset staggered patches (Figure 5.42B) spanning 1.8 mm downstream and offset
Figure 5.45 Images of steady state species concentration fields under an applied potential of 280 V/cm for (A) the homogeneous microchannel and (B) the heterogeneous microchannel with 6 offset staggered patches as obtained through numerical and experimental analysis. Arrows show the flow directions.
294
Electrokinetics in Microfluidics
10 um from the channel centerline with a width of 90 jam and a length of 300 um, was analyzed experimentally. Mixing experiments were conducted at applied voltage potentials ranging between 70 V/cm and 555 V/cm as corresponded to Reynolds numbers of 0.08 and 0.7 and Peclet numbers of 190 and 1500. As can be seen in Figure 5.45, experimental results compared well with numerical simulations with images of the steady state flow for the homogenous and heterogeneous cases exhibiting near identical flow characteristics and circulation at 280 V/cm. Qualitatively, experimental visualization of a staggered configuration of heterogeneities exhibited the formation of highly unsymmetrical concentration gradients indicative of flow constriction and localized circulation in the patterned region. Bulk flow was forced to follow a significantly narrower and more intricate route thereby increasing the rate of diffusion by means of local concentration gradients. Convective mechanisms were also introduced by local flow circulation that transported a portion of the mixed downstream flow upstream. Additionally, sharp, lengthwise gradients absent in the homogeneous case resulted in an additional diffusive direction and enhanced mixing as exhibited in Figure 5.46 which plots the centerline concentration against the
Figure 5.46 Experimentally measured centerline normalized concentration along the channel length with an applied potential of 280 V/cm for the optimized offset staggered micro-mixer design.
Effects of Surface Heterogeneity on Electrokinetic Flow
295
Figure 5.47 Experimental concentration profiles (homogeneous channel: dotted lines; micromixer: dashed lines) and numerical (solid lines) across the channel width for both the homogeneous case prior to the mixing region (circular markers) and subsequent to the mixing region (star markers), and for the developed micro-mixer prior to the mixing region (triangular markers) and subsequent (square markers) for (A) 70 V/cm and (B) 280 V/cm.
296
Electrokinetics in Microfluidics
downstream channel length for an applied voltage potential of 280 V/cm. Clearly evident are these oscillating local concentrations induced by the surface charge patterning. Oscillations become substantially lower in magnitude at greater downstream distances as increased species homogeneity is achieved. Concentration profiles comparing the homogeneous and heterogeneous mixing channels at 70V/cm (Figure 5.47A) and 280V/cm (Figure 5.47B) indicate a substantial increase in species mixing with consistent correspondence between experimental and numerical results. Mixing was visibly enhanced for all voltages in comparison with a homogeneous T-channel, with a more marked improvement at increasingly higher voltages. Experimental and numerical profiles were consistent with values of mixing efficiencies within 2.5% at low voltage potentials (70 V/cm) and within 5% at higher potentials (555 V/cm). Using the developed micro-mixer, mixing efficiencies were improved from 75.3% to 97.2% at 70V/cm and from 22.7% to 90.2% at 555V/cm in comparison with the homogeneous system. Figure 5.48 presents a graph of the mixing efficiencies for varying voltage potentials for the developed micro-mixer compared with the homogeneous channel. Mixing is consistently enhanced with a more substantial improvement observable at higher
Figure 5.48 Mixing efficiencies for varying applied potentials for the homogeneous (square markers) channel and for the developed micro-mixer (triangular markers) as determined by experimental (dotted lines) and numerical (solid lines) analysis.
Effects of Surface Heterogeneity on Electrokinetic Flow
297
voltage potentials where diffusive mechanisms are increasingly inefficient. As the mixing efficiency appears to become asymptotic at increasingly higher voltages in both experimental and numerical results, a maximum applied voltage potential of 555 V/cm was deemed sufficient to fully characterize micro-mixer performance. In terms of required channel lengths, numerical analysis indicated that at 2 80V/cm, a homogeneous microchannel would require a channel mixing length of 22mm for reaching a 95% mixture. By implementing the developed micromixer, an 88% reduction in required channel length to 2.6 mm was experimentally demonstrated. Practical applications of reductions in required channel lengths include improvements in portability and shorter retention times, both of which are valuable advancements applicable to many microfluidic devices. This enhanced micro-mixer technology can be applied to a wide range of lab-on-a-chip applications involving a variety of fluids ranging from fluorescent dyes with diffusion constants on the order of 10~10 m2/s to large molecules of DNA and proteins with diffusion constants on the order of 10~12 m2/s. Hence, the sensitivity to diffusion constants is of interest for the characterization of micro-mixer performance. This study was based on a fluorescein dye with diffusion constant D = 4.37 x 10~10 m2/s. With a staggered patch configuration, mixing efficiency was increased from 36 to 96% in comparison with the homogeneous case for an applied potential of 280 V/cm. Numerical simulations indicated that for a diffusion constant two orders of magnitude lower, mixing efficiencies are improved from 3.6% in the homogeneous case to 70% with staggered surface charge heterogeneities. Regardless of diffusion constant, the improvement in mixing efficiency over the purely diffusive case is substantial. As results are highly sensitive to diffusion coefficients however, required channels lengths for complete mixing will be largely application dependent. In summary, the passive electrokinetic micro-mixer with an optimized arrangement of surface charge heterogeneities can increase flow narrowing and circulation, thereby increasing the diffusive flux and introducing an advective component of mixing. Mixing efficiencies were improved by 22-68% for voltages ranging from 70 to 555 V/cm. For producing a 95% mixture, this technology can reduce the required mixing channel length by 88% for flows with Pec let numbers between 190 and 1500 and Reynolds numbers between 0.08 and 0.7.
298
5-7
Electrokinetics in Microfluidics
ANALYSIS OF ELECTROKINETIC FLOW IN MICROCHANNEL NETWORKS
Many microfluidic devices have a complex network of microchannels [72,73]. Different channel branches may have different size and surface properties. Controlling liquid flow in such microfluidic networks is critical to the functionality and performance of these microfluidic devices. Therefore, understanding of the electrokinetic flow characteristics in complicated microchannel networks is highly desirable. As a first step, a general analytical model [74] for one-to-multi-branch microchannels (a fundamental element in microfluidic networks) will be reviewed in this section. Two important parameters, the hydrodynamic conductance and the electrokinetic power, are defined to simplify the analysis and facilitate the understanding. The applications of this analytical model to a two-section heterogeneous microchannel and a oneto-two-branch microchannel system will be discussed. 5-7.1 General equations of electrokinetic flow in a single microchannel First, let's consider electrokinetic flow in a homogeneous microchannel of circular cross section of radius a (see Figure 5.49). For simplicity, we consider a microchannel filled with a symmetric electrolyte solution. The valence and the bulk ionic concentration are thus identical for the cations and anions. The ionic mobility is also assumed to be the same for the two types of ions. At steady state, the electrokinetic flow is described by the axial Navier-Stokes equation with an electrical body force
Figure 5.49. Electrokinetic flow in a homogeneous cylindrical microchannel of radius a and length /. The channel wall has a zeta potential C,. The volume flow rate is Q. The electrical current is /.
Effects of Surface Heterogeneity on Electrokinetic Flow
299
where fi is the fluid viscosity, r the radial coordinate, u the axial fluid velocity, Vp the axial pressure gradient, and V^ the electric potential gradient parallel to the axis. The net charge density pe is given by the Poisson equation (42) where s is the dielectric constant of the fluid, s0 the permittivity of vacuum (8.854xlO~12 CV"1), and i// the EDL potential induced by the surface charge or the zeta potential C, at the channel wall. By substituting Eq. (42) into Eq. (41) and then integrating twice, the axial velocity u can be derived as Eq. (43) with the following boundary conditions: u(a) = 0, i//(a) = £ and «'(o) = y/'(o)= 0, where the prime indicates the first derivative with respect to r, (43) The current density j in an electrokinetic flow consists of two components: one is the convection current density due to the transport of ions with the bulk fluid flow, and the other is the conduction current density due to the motion of ions relative to the bulk fluid, (44) where m is the ionic mobility, z the valence of ions, e the electron charge (1.602xl0~19 C), n+ and n_ the volume densities of the cations and anions, respectively. The local ionic densities can be expressed in the form of the Boltzmann distribution, (45)
where n^ is the bulk ionic density, k^ the Boltzmann's constant (1.381x10 JK"1) and T the absolute temperature. Thus, Eq. (44) can be rewritten as
23
(46)
300
Electrokinetics in Microfluidics
where a = Imzen^ is the bulk conductivity of the liquid, and pe = -2ze« 00 sinh(zey// kfrT) is the very definition of the net charge density. Integrating Eqs. (43) and (46) over the channel cross-section gives the volume flow rate Q and the total current / (47) (48) where Ap = Vpl and A0 = V0/ are, respectively, the pressure difference and electric potential difference between the two ends of a microchannel with / being the channel length. L^ are the phenomenological coefficients from nonequilibrium thermodynamics of electrokinetic phenomena, and are expressed in terms of the properties of the microchannel and the liquid, (49a) (49b) (49c) where Gj (i= 1, 2, 3) are dependent only on the liquid properties and the channel size, the definitions of G, are given by [3]:
Effects of Surface Heterogeneity on Electrokinetic Flow
301
where IQ and /j denote the zero order and the first order modified Bessel ( 1 1 /
V2
function of the first kind; K-\2Z e n^jes^k^T] is the Debye-Huckel parameter whose reciprocal {Ilk) indicates the characteristic thickness of EDL. As seen clearly from the above equations, Gy and G2 depends only on the nondimensional electrokinetic radius, K a. However, G3 is associated with the zeta potential of the channel wall, and is definitely no less than 1. As a result, the apparent conductivity for electrokinetic flow in a microchannel is higher than the bulk conductivity of the electrolyte (see the RHS terms of Eq. (49c)). This difference is attributed to the surface conductance due to the EDL effect. The equality of cross-coupling coefficients Z,12 a n d -^21 *s m accordance with Onsager's theorem. It should be noted that Eqs. (47) and (48) are applicable to electrokinetic flow in microchannels of arbitrary cross-section geometry. The cross-sectional shape of microchannels affects only the definitions of Li;. The model presented later in this paper, therefore, can be generalized to a variety of microchannels.
Figure 5.50. Electrokinetic flow in a one (superscript index 0) to n-branch microchannel (superscript indices from 1 to n) system. The subscript t represents the interface region from the main-channel to branch-channels. The hollow arrows show the moving directions (assumed positive) of the liauid in each branch of the channel.
302
Electrokinetics in Microfluidics
5-7.2 Electrokinetic flow in one-to-multi-branch microchannels Generally, a microfluidic network is composed of several one-to-multibranch channels and/or the opposite. Therefore, as a first step, we investigate only the electrokinetic flow in a one-to-multi-branch microchannel system. The same approach can be applied to more complicated microchannel networks. Figure 5.50 shows schematically such a one-to-n-branch microchannel. All branches (superscript indices from 0 to n) in this channel are assumed to be homogeneous and of circular cross-sections. In each branch, the phenomenological equations of the flow rate and the current, i.e., Eqs. (47) and (48), are transformed to (50a) (50b) (51a) (51b) where i varies from 1 to n, and all the symbols are labeled in Figure 5.50. According to Kirchoff s principle, the mass and electric current should be conserved when the liquid flows from the main-channel (superscript index 0) into n branches, i.e., (52)
(53) Pressure-driven flow For steady-state pressure-driven electrokinetic flow in a one-to-n-branch microchannel system, zero total electric current in each branch should be obeyed. This requires n constraints (i.e., one branch-channel gives one constraint) on the flow system, (54)
Effects of Surface Heterogeneity on Electrokinetic Flow
303
Substituting Eqs. (50) and (51) into Eqs. (52)-(54), we obtain the following equations in the matrix form
(55)
where /^? and t are, respectively, the hydrodynamic pressure and electric potential at the junction from the main-channel to branch-channels (see Figure 5.50). <j>V' (i=l...n) denote the flow-induced electric potential at the end of each branch-channels. For this case, pt, (j)t and $S1' are the (n+2) independent variables to be determined. Using Cramer's rule, solutions of these quantities are derived as
(56)
(57)
(58)
The physical meanings of these parameters will be discussed later. It should be noted that <j>V' - ){ > gives the streaming potential generated in the passage from the main-
304
Electrokinetics in Microfluidics
channel to the fh branch-channel. The volume flow rates through the mainchannel and each branch-channel of this microchannel system are given by
(59a)
(59b)
From the above equations, we see that the flow rate is completely determined by the parameter gV' (z=O,l,...,w). If a conventional Poiseuille flow takes place in the same one-to-n-branch microchannel system, Eqs. (56)-(59) still hold except that g^1' is replaced by -L\\
in Eq. (49a) because the cross-coupling coefficients Z ^ vanishes.
Electroosmotic flow In electroosmotic flow, an externally applied electric field is used to transport the liquid while all ends of the microchannel system are usually exposed to atmosphere (i.e., p^' = 0 for i=0,\,..n). Consequently, Eqs. (52) and (53) are sufficient to determine the two independent variables pt and <j)t. As a result, the volume flow rate and the electric current through the /* branch of this microchannel system can be obtained as:
(60)
Effects of Surface Heterogeneity on Electrokinetic Flow
305
5-7.3 Model analysis In the above derivations, we see two key parameters f*' and g^' involved in the general model. In order to understand the physical meanings of
and the Ohm's law (i.e., / = -^22^0 --na crA<j)/l) are readily recovered from Eqs. (47) and (48), respectively. It should be noted that, for a pressure-driven flow, A0 in Eqs. (47) and (48) is the flow-induced electrokinetic potential, i.e., the streaming potential. In this case, on the right hand side of Eq. (48), the first term represents the streaming current, and the second term represents the conduction current. At steady state, the net current along the microchannel should be zero, i.e., / = 0 in Eq. (48). This condition allows finding an expression for the streaming potential. That is,
channel system. For a pure electroosmotic flow with Ap = 0, combining Eqs. (47) and (48) gives the flow rate (63) As seen from the above two equations, / is an important parameter in electrokinetic flow. By analogy to the thermoelectric power (i.e., the so-called
306
Electrokinetics in Microfluidics
Seebeck coefficient) defined in thermoelectrics, we may name this parameter as the electrokinetic power because electrokinetic effects can also be viewed as energy conversion (i.e., from pressure difference to electrical potential in Eq. (62)). From those definitions of phenomenological coefficients L;j in Eq. (49), we see that this electrokinetic power / can be negative or positive depending on the sign of zeta potential (negative C, corresponds to positive / ) , similar to that the sign of thermoelectric power is related to the polarity of ions (negative electrons correspond to negative thermoelectric power). Furthermore, Eq. (62) clearly indicates that the streaming potential is a linear function of the pressure drop. After substitution of Eq. (62) into Eq.(47), the flow rate in a pressure-driven flow can be expressed by (64) where g = (ij2^21 ~ A\^2l)l^22 > identical to the definition of g^> for the multi-branch channel system, can therefore be viewed as the hydrodynamic conductance at zero electric current in electrokinetic flow. The inequality LjjZ22 > ^12^21 must hold because of the positive entropy production requirement according to Prigogine's theorem. Therefore, g is a positive quantity. Also, g is independent of the sign of zeta potential. Combining Eqs. (62) and (64) yields a relationship between the streaming potential and the volume flow rate, (65) Under Debye-Huckel linear approximation and neglecting the variation of electric conductivity across the channel cross-section, the analytical forms of the electrokinetic power / and the hydrodynamic conductance g are given by [74]:
where
Effects of Surface Heterogeneity on Electrokinetic Flow
307
The first term on the RHS of the above denotes the hydrodynamic conductance of Poiseuille flow, and the second term reveals the negative effect due to electroviscosity resulting from EDL. The function F{KQ,C,) is dependent on both the electrokinetic radius and the zeta potential. Figure 5.51 shows the curves of / against the electrokinetic radius at different zeta potentials. The non-dimensional electrokinetic radius,
Figure 5.51. Electrokinetic power/with respect to the electrokinetic radius Ka at different zeta potentials. Pure water (lxl0~ 5 M concentration) is selected as the electrolyte solution. The working parameters include 7=298 K, Woo=6.022xl021 rrf3, £=80, ^=0.9xl0' 3 Kgm'V 1 ando=10~ 4 Sm~'.
308
Electrokinetics in Microfluidics
Ka = a /(I/ k), is a measure of the channel size relative to the EDL thickness 1/K where K is the Debye-Huckel parameter. At C, = 0, / vanishes, equivalent to zero streaming potential in pressure-driven flow (Eq. (62)) or zero flow rate in electroosmotic flow (Eq.(63)). When C, ^ 0 , the magnitude of / increases with that of the zeta potential. Moreover, / becomes higher at a larger electrokinetic radius (e.g., a larger channel size or a higher ionic density or a thinner EDL), and approaches to constant for very large electrokinetic radius. In addition, / is independent of the channel length. Figure 5.52 displays the curves of hydrodynamic conductance g scaled by the one at £ = 0 (see the inset of Figure 5.52) against the electrokinetic radius. We see that g increases with the electrokinetic radius, but becomes lower in the presence of EDL (£ ^ 0 ) . The reason that g decreases with the zeta potential is due to the electro-viscous effect as mentioned above. The g curve attains a minimum when the electrokinetic radius is approximately 2. The g curves approach to one when the electrokinetic radius is larger than 100, indicating the electro-viscous effect becomes
Figure 5.52. Scaled hydrodynamic conductance with respect to the electrokinetic radius Ka at different zeta potentials. The microchannel is 1 cm long. The hydrodynamic conductance g is scaled by the one in the absence of EDL effect (C=0) which is plotted in the inset.
Effects of Surface Heterogeneity on Electrokinetic Flow
309
negligible at high electrokinetic radii. Next, we will apply the general analytical model to two special channels. Pure water (lxl(T 5 M concentration) is selected as the electrolyte solution. 5-7.4 Two-section heterogeneous microchannel The two sections of a microchannel under investigation are assumed to have different zeta potentials. The microchannel of circular cross section has the same radius a in both sections, but the length of each section is not necessarily the same. Pressure-driven flow Setting n =1 in Eqs. (59) and (58), we can obtain the measurable volume flow rate and the streaming potential for pressure-driven flow in such a heterogeneous microchannel,
The denominator of the RHS term in Eq. (66) represents the total hydrodynamic resistance of the two-section microchannel, while the numerator is the total pressure drop applied onto the two ends of the channel. Moreover, the streaming potential in Eq. (67) is similar to that of a homogeneous channel (see Eq. (65)), and is the sum of the streaming potentials in the two sections. If we put Eq. (67) back into the form of Eq.(62), we find that the equivalent electrokinetic power /two-section of the two-section microchannel is given by
(68)
Figure 5.53 shows the distributions of flow-induced electric potential. Both the sign and the magnitude of zeta potential strongly influence this electric potential. This influence is due to the dependence of electrokinetic power / on the zeta potential. The lengths of homogeneous sections also affect the electric potential profile because they are involved in function of the hydrodynamic conductance g(l> (see Eq. (67)). The effect of zeta potential on the flow rate and
310
Electrokinetics in Microfluidics
Figure 5.53. Distributions of flow-induced electric potential for pressure-driven flow in two-section heterogeneous microchannels: (a) /°WM).5 cm; (b) ^0)=0.25 cm and /(I>=0.75 cm. £0)= -50 mV, while: Case A, gl)= -50 mV; Case B, ^°= -100 mV; Case C, ^'WSO mV; Case D, i^'WlOO mV. a ( °W 1 ) = 5 urn. The applied pressure drop between the two ends of the channels is 1 atm.
Effects of Surface Heterogeneity on Electrokinetic Flow
311
the streaming potential are illustrated in Figure 5.54. The flow rate scaled by the flow rate without EDL effect is independent of the sign of zeta potential (symmetric about £X)= 0), but becomes lower at a higher zeta potential. For the cases shown here, the flow rate is only slightly affected by the magnitude of zeta potential (less than 2% in Figure 5.54. smaller channels should amplify this effect). As the zeta potential varies, the streaming potential changes significantly and even switches its polarity. It is not surprising that the streaming potential is almost a linear function of the zeta potential in one section of the heterogeneous channel. At a relatively large electrokinetic radius (e.g. 51.5 in Figure 5.54), the zeta potential has a minor effect on the hydrodynamic conductance g (see Figure 5.52). Hence, the streaming potential in Eq.(67) is mainly determined by the electrokinetic power/ which is approximately proportional to the zeta potential (see Figure 5.51). We also find that the pressure gradients in the two sections of the heterogeneous microchannel are not identical although very close to each other.
Figure 5.54. Scaled pressure-driven flow rate (thick lines) and the streaming potential (thin lines) against ^ of a two-section heterogeneous microchannel with 4<0>=-50 mV. Case A, ^0)=0.5 cm and ^ 0 =0.5 cm; Case B, /<0)=0.25 cm and P=0.15 cm; Case C, / 0) =0.75 cm and P=0.25 cm. a —a '=5 urn. The applied pressure drop between the two ends of the channels is 1 atm. The flow rate is scaled by the one in the absence of EDL effect (£=0) which is 0.166 jxl-mirf'.
312
Electrokinetics in Microfluidics
Electroosmotic flow For the case of electroosmotic flow in a two-section microchannel, the volume flow rate and the electric current can be derived from Eqs. (60) and (61), (69)
(70)
Figure 5.55. Velocity profiles for electroosmotic flow in two-section heterogeneous microchannels. Thin lines stand for the case with f°^=f^-0.5 cm and thick lines for f°^=0.25 cm and f1}=0J5 cm. <^0)= -50 mV, while: Case A, £"= -50 mV; Case B, 4°>= -100 mV; Case C, 4<1)=-t-5O mV; Case D, ^^=+100 mV. a(0)=o(1)=5 nm. The applied potential drop between the two ends of the channels is 100 V.
Effects of Surface Heterogeneity on Electrokinetic Flow
313
(71)
Comparing Eqs.(68) and (71) gives (72)
Figure 5.56. Electroosmotic flow rate (thick lines) and scaled electric current (thin lines) against ^ of a two-section heterogeneous microchannel with <^°)= -50 mV. Case A, r ) = 0 . 5 cm and /°=0.5 cm; Case B, r ) = 0 . 2 5 cm and / 1} =0.75 cm; Case C, / 0) =0.75 cm and / 1 ) = 0.25 cm. a(0)=a(1)=5 \xm. The applied potential drop between the two ends of the channels is 100 V. The electric current is scaled by the one calculated under the Ohm's law (C=0), which is 0.0785 nA.
314
Electrokinetics in Microfluidics
In other words, the electrokinetic flow in the two-section heterogeneous microchannel still satisfies Onsager's reciprocal relations. Figure 5.55 presents the velocity profiles of electroosmotic flow in twosection microchannels. When zeta potentials are identical in the two sections (Case A), i.e., a homogeneous channel, the electroosmotic velocity profile is plug-like, as expected. If zeta potentials in the two sections are different (Cases B, C and D), the electroosmotic velocities in the two sections will be different. Thus, pressure gradients are induced along the heterogeneous channel in order to satisfy the constant flow rate throughout the heterogeneous channel. Consequently, the velocity profiles in the bulk liquid in both sections become parabolic, although the velocities in EDL regions (or close to the channel wall) remain consistent with the local zeta potentials. That is, a higher zeta potential gives rise to a higher electroosmotic velocity, and a positive zeta potential will generate an electroosmotic flow in a direction opposite to the electroosmotic flow generated by a negative zeta potential. Therefore, these bulk
Figure 5.57. Effect of the main channel length f0) on the volume flow rate for pressure-driven flow in one-to-two-branch microchannel systems. It is assumed that / ' W 2 ' while i(0)+/(1)=l cm. a(0)=5 urn. Case A, a(1)=a(2)=5 nm; Case B, a(1)=a(2)=2.5 urn; Case C, «(1)=5 urn and a(2)=2.5 urn. The applied pressures are // 0 ) =l atm and // 1)= p (2) =0. The zeta potential is -50 mV in all branches.
Effects of Surface Heterogeneity on Electrokinetic Flow
315
liquid velocity profiles are strongly dependent on the zeta potentials in both downstream and upstream sections of the channel. Figure 5.56 illustrates the effect of zeta potential on the flow rate and the electric current. The flow rate varies significantly and even the flow direction changes with the zeta potential (see also Figure 5.55). The electric current scaled by the one calculated from the Ohm's law only has a slight change with zeta potential. This is attributed to the reciprocal relation in Eq.(72) (see also Figure 5.54). Another point worthy of note is that the scaled electric current larger than unity results from the contribution of convection current. This contribution (can be as much as 8% in Figure 5.56) is the very reason that we cannot assume a uniform electric field in a two-section heterogeneous microchannel. 5-7.5 One-to-two-branch microchannel All branches (superscript indices from 0 to 2, as shown in Figure 5.50) of the one-to-two-branch microchannel are of circular shapes. Their radii, lengths and zeta potentials are not necessarily the same. The cases of dispensing (i.e., flowing from the main-channel into branch-channels and the flow rate is considered to be positive) and mixing (i.e., flowing from branch-channels into the main-channel and the flow rate is considered to be negative) can be realized by adjusting the relative magnitude of p^' or $\1' (z—0,1,2). Pressure-driven flow Despite that Eqs.(59) and (58) at n = 2 give the general formulae of volume flow rate and streaming potential in a one-to-two-branch microchannel, we restrict ourselves to only the special pumping or mixing cases with Then, we have pv) =pv).
(73a)
(73b)
(74)
where / may be 1 or 2. One can see that Eqs.(73) and (74) are analogous to those for the two-section channel just discussed. We only need to replace the
316
Electrokinetics in Microfluidics
hydrodynamic resistance 1/ g^'
in Eqs.(66) and (67) with the total resistance
rme
V\S + S } f° current two branch-channels. Figure 5.57 shows the effect of the main-channel length (superscript index 0) on the total flow rate. Three different channel structures are calculated. One can see that the channel size (both the radius and the length) strongly affects the flow rate, which provides another method to control the flow behavior other than using the applied pressures or electric potentials. Electroosmotic flow The formulae for the electroosmotic flow rate and the electric current through the main-channel of a one-to-two-branch microchannel are given by
(75)
(76)
assumed. Those formulae for the two branch-channels are omitted here for compactness. Figure 5.58 shows the curves of flow rate in branch-channel 2, Q\ ', and the ratio of flow rate g
/Q
with respect to the radius of the
branch-channel 2, a^ >. Similar to Figure 5.57 for the pressure-driven flow, the channel size also strongly affects the electroosmotic flow rate. However, this flow rate is approximately proportional to the square of channel radius, while the pressure-driven flow rate is proportional to the fourth power of channel radius. This difference can be identified from the definitions of those phenomenological coefficients L\\ and L\i appearing in Eq.(47) (see also Eqs.(49a) and (49b)).
Effects of Surface Heterogeneity on Electrokinetic Flow
317
Figure 5.58. Effect of a{2) on the electroosmotic flow rate Q{2) (thin lines) and the ratio Q( )/g(0) ^ h i ^ ] j n e S j for t n e faree c a s e s ^ B and C) in one-to-two-branch microchannel systems. a (0) =a (1) =5 urn. Case A, / 0) =0.25 cm and fl)=f2)=0.75 cm; Case B, fO)=fl)=f2)=O.5 cm; Case C, Z^O.75 cm and / ' W ^ O ^ cm. The applied potentials are 0(O)=1OO V and ^ (l) =^ (2) =0. The zeta potential is -50 mV in all branches.
318
Electrokinetics in Microfluidics
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37]
C. Werner, U. Konig, A. Augsburg, C. Arnhold, H. Korber, R. Zimmermann and H-J Jacobash, Colloids Surfaces A, 159 (1999) 519. C. Werner and H-J Jacobash, Int. J. Artificial Organs., 22 (3) (1999) 160. D.W. Fuerstenau, J. Phys. Chem., 60 (1956) 981. R. Hayes, M. Bohmer and L. Fokkink, Langmuir, 15 (1999) 2865-2870. M. Zembala and Z. Adamczyk, Langmuir, 16 (2000) 1593-1601. W. Norde and E. Rouwendal, J. Colloid Interface Sci., 139 (1990) 169. A.V. Elgersma, R.L.J. Zsom, J. Lyklema and W. Norde, Colloids Surfaces, 65 (1992) 17. M. Zembala and P. Dejardin, Colloids Surfaces B, 3 (1994) 119. C. Werner and H.J. Jacobasch, Macromol. Symp., 103 (1996) 43. Z. Adamczyk, P. Warszynzki and M. Zembala, Bull. Polish Acad. Sci. Chem., 47 (1999) 239-258. R. Hayes, Colloids Surfaces A, 146 (1999) 89-94. D. Erickson and D. Li, Langmuir, 18 (2002) 8949-8959. J. L. Anderson, J. Colloid Interface Sci., 105 (1985) 45-54. A. Ajdari, Phys. Rev. Lett., 75 (1995) 755-758. A. Ajdari, Phys. Rev. E., 53 (1996) 4996-5005. A. Ajdari, Phys. Rev. E , 65 (2001) 016301. S. Ghosal, J. Fluid Mech., 459 (2002)103-128. A.D. Stroock, M. Week, D.T. Chiu, W.T.S. Huck, P.J.A. Kenis, R.F. Ismagilov and G.M. Whitesides, Phys. Rev. Lett., 84 (2000) 3314-3317. A. Herr, J. Molho, J. Santiago, M. Mungal and T. Kenny, Anal. Chem., 72 (2000) 10531057. C. Keely, T. van de Goor and D. McManigill, Anal. Chem., 66 (1994) 4236-4242. T.J. Johnson, D. Ross, M. Gaitan and L.E. Locascio, Anal. Chem., 73 (2001) 3656-3661. D. Erickson and D. Li, Langmuir, 18 (2002) 1883-1892. D. Erickson and D. Li, J. Physical Chemistry, (in press). D. Erickson and D. Li, J. Colloid Interface Sci., 237 (2001) 283-289. D. Stokes, G. Griffin and T. Vo-Dinh, Fresenius J. Anal. Chem., 369 (2001) 295-301. B. Cheek, A. Steel, M. Torres, Y-Y Yu and H.Yang, Anal. Chem., 73 (2001) 5777-5783. A. Dodge, K. Fluri, E. Verpoorte, N. de Roojj, Anal. Chem., 73 (2001) 3400-3409. C-T Wu, T-L Huang, C. Lee and C. Miller, Anal. Chem., 65 (1993) 568-571. J.K. Towns and F.E. Regnier, Anal. Chem., 64 (1992) 2473. W. Nashbesh and Z. El Rassi, J. High Resolut. Chromatogr., 15 (1992) 289-292. A.D. Stroock, M. Week, D.T. Chiu, W.T.S. Huck, P.J.A. Kenis, R.F. Ismagilov and G.M. Whitesides, Phys. Rev. Lett., 86(26) (2001) 6050. R. Cohen and C. Radke, J. Colloid Interface Sci, 141 (1991) 338-347. S. Patankar, C. Liu and E. Sparrow, J. Heat Trans, 99(2) (1977) 180-186. J. Lyklema, Fundamentals of Interface and Colloid Science, Volume 1: Fundamentals, Academic Press: San Diego, 1991. J. Lyklema, Fundamentals of Interface and Colloid Science, Volume 2: Solid-Liquid Interfaces, Academic Press: San Diego, 1995. J.C. Heinrich and D.W. Pepper, Intermediate Finite Element Method, Taylor & Francis: Philadelphia, 1999. A. Saez and R. Carbonell, Int. J. Num. Meth. Fluids, 5 (1985) 601-614.
Effects of Surface Heterogeneity on Electrokinetic Flow
319
[38] J.L. Anderson and W.K. Idol, Chemical Engineering Communications, 38 (1985) 93106. [39] D. Long, H.A. Stone and A. Ajdari, J. Colloid Interface Sci, 212 (1999) 338-349. [40] B. Potocek, B. Gas, E. Kenndler and M. Stedry, J. Chromatography A, 709 (1995) 5162. [41] L. Ren and D. Li, J. Colloid Interface Sci., 243 (2001) 255-261. [42] C.R. Wylie and L.C. Barrett, Advanced Engineering Mathematics, McGraw Hill, New York, 1995. [43] R.W. Hornbeck, Numerical Methods, Prentice Hall, New Jersey, 1975. [44] D.A. Anderson, J.C. Tannehill and R.H. Pletcher, Computational Fluid Mechanics and Heat Transfer, Hemisphere, Washington, 1984. [45] R. Weast, M.J. Astle and W.H. Beyer, CRC Handbook of Chemistry and Physics, CRC Press, Boca Raton, 1986. [46] P. Vanysek, "Ionic Conductivity and Diffusion at Infinite Dilution", in CRC Handbook of Chemistry and Physics. [47] R.J. Hunter, Zeta Potential in Colloid Science, Academic Press: London, 1981. [48] J.D. Harrison, K. Fluri, K. Seiler, Z. Fan, C. Effenhauser and A. Manz, Science, 261 (1993)895-897. [49] S.K.W. Dertinger, D.T. Chiu, N.L. Jeon and G.M. Whitesides, Anal. Chem., 73 (2001) 1240-1246. [50] N.L. Jeon, S.K.W. Dertinger, D.T. Chiu, I.S. Choi, A.D. Stroock and G.M. Whitesides, Langmuir, 16 (2000) 8311-8316. [51] A.E. Kamholz, E.A. Schiling and P. Yager, J. Biophys., 80 (2001) 1967-1972. [52] B. Weigl and P. Yager, Science, 283 (1999) 346-347. [53] P. Galambos, F.K. Forster and B. Weigl, Proceedings of the international conference on solid state sensors and actuators: IEEE, Piscataway, NJ, 1997. [54] A.E. Kamholz, G.H. Weigl, B.A. Finlayson and P. Yager, Anal. Chem., 71 (1999) 53405347. [55] S.V. Ermakov, S.C. Jacobson and J.M. Ramsey, Anal. Chem., 70 (1998) 4494-4504. [56] F. Bianchi, R. Ferrigno and H.H. Girault, Anal. Chem., 72 (2000) 1987-1993. [57] N.A. Patankar and H.H. Hu, Anal. Chem., 70 (1998) 1870-1881. [58] Z. Yang, H. Goto, M. Matsumoto and R. Maeda, Electrophoresis, 21 (2000) 116-119. [59] P. Woias, K. Hauser and E. Yacoub-George, Micro Total Analysis Systems 2000, Kluwer Academic Publishers, Dordrecht, (2000) 277. [60] K. Hosokawa, T. Fujii and I. Endo, Micro Total Analysis Systems 2000, Kluwer Academic Publishers: Dordrecht, 481. [61] K. Yasuda, Sens. Actuators B, 64 (2000) 128-135. [62] M.H. Oddy, J.G. Santiago and J.C. Mikkelsen, Anal. Chem., 73 (2001) 5822-5832. [63] R.H. Liu, K.V. Sharp, M.G. Olsen, M. Stremler, J.G. Santiago, R.J. Adrian, H. Aref and D.J. Beebe, J. Microelectromech. Syst, 9 (2000) 190-198. [64] V. Mengeaud, J. Josserand and H.H. Girault, Anal. Chem., 74 (2002) 4279-4286. [65] A.D. Stoock, S.K.W. Dertinger, A. Ajdari, I. Mezi, H.A. Stone and G.M. Whitesides, Science, 295 (2002) 647-651. [66] V. Haverkamp, W. Ehrfeld, K. Gebauer, V. Hessel, H. Lowe, T. Richter and C. Wille, Fresenius J. Anal. Chem., 364 (1999) 617-624. [67] D.J. Beebe, R.J. Adrian, M.G. Olsen, M.A. Stremler, H. Aref and B. Jo, Mec. Ind., 2 (2001)343-348. [68] E. Biddiss, D. Erickson and D. Li, Anal. Chem., (in press).
320
Electrokinetics in Microfluidics
[69] J. C. McDonald, D.C. Duffy, J.R. Anderson, D.T. Chiu, H. Wu, O.J.A. Schueller and G.M. Whitesides, Electrophoresis, 21 (2000) 27-40. [70] Y. Liu, J.C. Fanguy, J.M. Bledsoe and C.S. Henry, Anal. Chem., 72 (2000) 5939-5944. [71] A. Sze, D. Erickson, L. Ren and D. Li, J. Colloid Interface Sci., 261 (2003) 402-410. [72] S.C. Jacobson, T.E. McKnight and J.M. Ramsey, Anal. Chem., 71 (1999) 4455-9. [73] T. Thorsen, S.J. Maerkl and S.R. Quake, Science, 298 (2002) 580-4. [74] X. Xuan and D. Li, J. Micromechanics Microeng., 14 (2004) 290-298.
Effects of Surface Roughness on Electrokinetic Flow
321
Chapter 6
Effects of surface roughness on electrokinetic flow In addition to electrokinetic properties related to the intrinsic surface chemistry of the channel walls, surface roughness plays an important role in various microfluidic processes. Generally, microchannel surfaces may exhibit certain degrees of roughness generated by the manufacturing techniques or by adhesion of biological particles from the liquids. The height of the surface roughness usually ranges from a few hundreds nanometers to several microns. There are some reported works dealing with more realistic channel surfaces, such as the studies of electroosmotic flows in capillaries with surface defects [1,2] and in microchannels with surface heterogeneity [3,4]. Ajdari's works [5-7] predicted that the presence of surface heterogeneity could result in regions of bulk flow circulation. This behavior was later observed experimentally in slit microchannels by Stroock et. al. [8]. Long et. al. [1] also developed an analytical model for an isolated heterogeneous spot in a flat plate. However, so far there has been little published work of electrokinetic flows in rough microchannels.
Figure 6.1. An example of a silicon surface with microfabricated, symmetrically arranged prism elements.
322
Electrokinetics in Microfluidics
The sensitivity and the efficiency of a wide range of chemical or biochemical reactions in lab-on-a-chip applications depend on the area of reacting surfaces. One way to enhance the bio-chemical reactions is to increase the reaction surface area. In a microchannel, creating many three-dimensional roughness elements on the microchannel walls can significantly increase the surface area. These 3D roughness elements can be produced by photolithography based microfabrication techniques, and an example is shown in Figure 6.1. However, these 3D roughness elements will inevitably influence liquid flow and sample transfer in the microchannel. Therefore it is necessary to examine the influence of micron-sized rough elements on electroosmotic flow and the associated molecule transport in microchannels.
Effects of Surface Roughness on Electrokinetic Flow
6-1
323
ELECTROOSMOTIC TRANSPORT IN A SLIT MICROCHANNEL WITH 3D ROUGH ELEMENTS
In this section, we consider electroosmotic flow through slit microchannels where the microchannel wall is a homogeneous surface with uniformly distributed 3D rough elements [9]. To simplify the modeling and the numerical simulation, we assume the rough elements to be cylinders with a square cross section (i.e., rectangular prisms), and consider two types of distributions: symmetrical and asymmetrical arrangements, as illustrated in Figure 6.2a and 6.2b. The electroosmotic flow field and the sample concentration field are investigated in terms of the influence of the rough element size, height, density, the element arrangement, the channel size and the electrokinetic mobility. 6-1.1 Mathematical model The rough microchannel studied here is formed by two parallel surfaces with rectangular prismatic rough elements. The parameters describing the rough microchannel include: (i) the roughness element's size, a; (ii) the separation distance, b, between the roughness elements in both x and z directions; (iii) the roughness element's height, h; (iv) the microchannel height, H; and (v) the rough
Figure 6.2. The symmetrical (a) and asymmetrical (b) roughness arrangements on the homogeneous microchannel wall and the corresponding computational domains (enclosed by the dashed lines) from the top view.
324
Electrokinetics in Microfluidics
elements' arrangement: symmetrical and asymmetrical arrangements. As our goal is to examine the roughness effects on the electrokinetic transport processes through microchannels, the flow through a virtual smooth microchannel will be used as a comparison base. This virtual smooth channel is defined as the smooth channel having the same channel volume, the same channel width and the same channel length as the rough channel. Under these conditions, the virtual smooth channel height, Hs, and the virtual smooth channel cross-section area, Ac<s can be obtained. In this way, the comparison can avoid the effect of the reduction in the flow passage due to the presence of the roughness elements. The same electric potential difference and zero pressure difference are applied to the inlets and the outlets of both the virtual smooth channel and of the rough microchannel. Further simplifications are required for developing a proper model. Firstly, in the study, the flow is limited to a low-Reynolds regime (0.001 < Re < 10), turbulence and wake region at the rough elements' tail region can be neglected. Secondly, we assume that the influences of the applied electric field and the flow field on the EDL field can be neglected. This means that the ion distributions in the EDL region are determined by the EDL field only. This is a necessary condition to derive the Smoluchowski equation: Veof = jj,eo E = (sC, /p.) E for electroosmotic flow. The validity of this assumption can be understood as follows. EDL is a thin layer at the solid-liquid interface and its characteristic thickness, for example, is approximately 9.6 nm for a 10~3 M KC1 solution. Within this thin layer, the electrical potential drops from the zeta potential (on the wall) to zero (at the edge of the EDL). Generally the zeta potential is of the
Figure 6.3. Illustration of 3D coordinates, computational grid systems for the case of symmetrical arrangement with a/H= 0.6, b/H= 1 and h/H= 0.2.
Effects of Surface Roughness on Electrokinetic Flow
325
order of 100 mV. The field strength of EDL therefore is approximately of the order of 107 V/m (i.e., lOOmV/lOnm). On the other hand, the applied electrical field strength to generate electroosmotic flow in microchannels ranges approximately from 103 to 105 V/m. Therefore, the EDL field is dominant. Generally the electroosmotic flows in microchannels are low Reynolds number laminar flow (0.001 < Re < 10). A detailed study [10] demonstrated that a weak flow field (Re < 10) will not influence appreciably the ion distributions in the EDL. In addition, we assume that the entrance effect and the edge effect are neglected, and hence the flows through rough microchannels are considered to be fully-developed periodical flows, the numerical methods for simulating this kind of flow can be found elsewhere [11]. The top view of the computational domains is shown in Figure 6.2. As seen from this figure, we take the region between two symmetrical planes with one roughness separation distance as the computational domain width. The height of the computational domain is chosen as half of the microchannel height because of the symmetry. For simplicity, we consider the separation distances between roughness elements in both the x-direction and the z-direction are the same, denoted by b. It should be noted that two periods are chosen as the computational domain along the main flow direction. This treatment has particular benefit to simulation convergence [12]. Electrical field We consider a thin double layer and no net charge density in the bulk liquid. According to the theory of electrostatics, the applied electrical potential established in the rough microchannel, <j>, can be described by the Poisson equation,
(1) Since the inlet and the outlet planes of the computational domain are geometrically symmetrical planes, the electrical potential is uniformly distributed on these planes. We introduce the following dimensionless parameters: (2)
where (/),„ and <j)otl, are the electrical potential values on the inlet and outlet of one period computational domain, respectively; H is the height of the microchannel
326
Electrokinetics in Microfluidics
and chosen as the characteristic length of the system studied here. The dimensionless Poisson equation and its boundary conditions are given below:
As the dimensionless electrical potential field is identical in each period, which is half of the computational domain of length 2b, it is only necessary to solve the electrical potential field in one period. The corresponding boundary conditions *
(see Figure 6.3). Once the electrical field in the rough microchannel is known, the local electric field strength can be calculated by (4)
The dimensionless form of electrical field strength in rough channels is defined as the ratio of the local electrical field strength in the rough microchannel to the overall electrical field strength over one period:
(5)
where Ex s is the modulus of the tensor Ex s, and i , j , k are the unit tensors in x, y, z directions, respectively. Flow field The basic equations describing the electroosmotic flow through microchannels are the continuity equation and the Navier-Stokes equation with the electric body force in the right-hand side, (6)
"ill
Effects of Surface Roughness on Electrokinetic Flow
(7)
where Veo = ueo i + veo j + weo k is the electroosmotic velocity in rough microchannel, p is the pressure, pe is the net charge density, and p, \x are the density and the viscosity of the bulk liquid, respectively. Generally, numerical solution of electroosmotic flow is complicated by the simultaneous presence of three separate length scales: the channel length (mm), the channel depth or width (um) and the double layer thickness (nm). The simplest way to avoid this multiple-scale problem is to apply a slip boundary condition at the channel wall iYwaii = -He<>V where \xe0 = sQ/n is the electroosmotic mobility and C, is the zeta potential of the channel wall) to solve for the bulk liquid motion. Since the net charge density is zero within the bulk liquid, this treatment eliminates the electrical body force term in Eq. (7) and thus the need to solve the equation for the double layer field. The slip boundary velocity on the roughened microchannel wall, Veo snp, can be written as the following: (8)
Introducing p - pl{fi-Veos where
Veo s
the
IH),
following
dimensionless
group:
Veo - Veo I Veo s,
and define the Reynolds number Re = pVeosH
is the characteristic velocity determined by
E x s,
I JJL, i.e.,
me
governing equations of electroosmotic flow Ko,s ~ Veo,s ' - Veo ^x,s ' > through rough microchannels can be changed from Eqs. (6) and (7) into: (9)
OO)
328
Electrokinetics in Microfluidics
For electroosmotic flow through rough microchannels, three types of boundary conditions should be applied to Eqs. (9) and (10). They are called the fully-developed periodic conditions [12] applied at the inlet and the outlet of the computational domain, the slip boundary condition applied at the rough wall, and the symmetrical conditions applied at the two sides and the top plane of the computational domain (see Figure 6.3), respectively. It is worthwhile to note that the periodic conditions here are different from the conventional fully-developed periodic conditions. As no pressure difference is imposed on the inlet and the outlet of the rough microchannels, we have the following period conditions:
(11)
In the meantime, Veo and p at each point should satisfy the governing equations, Eqs. (9) and (10), especially at the inlet and the outlet planes of the computational domain. The slip boundary condition is applied at the channel wall (including the surface of roughness elements). According to the definition of the dimensionless governing equations, the dimensionless electroosmotic slip boundary velocity can be derived from Eq. (8): (12) At the rest computational domain boundaries, symmetrical conditions are applied. That is, for the velocity component parallel to the symmetric plane, the change of the velocity component normal to the plane is zero; for the velocity component normal to the plane, the velocity component itself is zero. Concentration field Study of the concentration field of sample species is of great importance in various microfluidics applications, for example, dye-based microflow velocity profile measurements, microscale species mixing, microfluidic dispensing, and surface reaction kinetic studies. Consider that a sample with concentration Co is released into the rough microchannel at time t0, the sample transport in time and space can be described by the law of mass conservation, which takes the form: (13)
Effects of Surface Roughness on Electrokinetic Flow
329
where Q is the concentration of /th species, Vep r- is the electrophoretic velocity of the /th species, and is given by Vepi = Hepj E, Dj and fiepj are the diffusion coefficient and the electrophoretic mobility of the z'th species, respectively. In Eq. (13), the first term on the left hand side is the transient term; the second term include two parts: the first part represents the convection effect resulting from the electroosmotic flow, the second part stands for the electrophoretic effect; the right hand side term is the diffusion term. By introducing the dimensionless group, (14) and define Peclet Number as Pe = VeosH I Dj, we can non-dimensionalize Eq. (13) as, (15)
Because releasing sample species into a microchannel is a transient process, initial condition and boundary conditions are required. The initial condition is: When t = 0, C,- (x ,y ,z ,0) = 0. The boundary conditions are: At the inlet plane, Cj (0, y , z , / ) = 1; at the outlet plane, we assume the concentration at each point is influenced only by its upwind control volume, and hence have VC7- = 0; at the symmetrical boundaries and the walls, no flux condition VC,- = 0 is applied. As seen from the above defined dimensionless parameters, the electrokinetic property of the microchannel surface (zeta potential C, ) and the properties of the solution (density and viscosity p and /S) are considered in the Reynolds number. For most electroosmotic flows in microfluidics, Re is in the range of 0.001-10. The property of the sample, the diffusion coefficient Dh is accounted in the Peclet number. Because the species electrophoresis and the liquid electroosmosis occur simultaneously under the same electrical field, the ratio nep,- / fieo instead of the individual values determines the relative electrophoretic velocity as Eq. (14) shows. Therefore, in addition to the parameters describing the surface roughness, Reynolds number is the basic
330
Electrokinetics in Microfluidics
dimensionless parameter determining the characteristic of the electroosmotic flow; Peclet number and the ratio }xepi I neo are the basic dimensionless parameters determining the characteristics of the sample transport in the microchannels. The mobility ratio fiepj /fieo represents the ratio of the electrophoretic force to the electroosmotic force. If the two mobilities have a different sign (i.e., the electrical charge of the sample and the charge of the microchannel wall have different signs, one positive and the other negative), that is, juepj /jueo< 0, the sample molecules will move in the same direction as the electroosmotic flow and in a speed faster than the liquid. If juepi /jueo> 0 (i.e., the electrical charge of the sample and the charge of the microchannel wall have the same sign), the sample molecules will move in the direction opposite to the electroosmotic flow and in a speed slower than the liquid when ignoring the movement caused by diffusion. The absolute value of this ratio, | fiePii /jueo \ > 1 or | juepJ /jueo | < 1, signifies the speed of the sample's electrophoretic motion relative to that of the bulk liquid electroosmotic flow. Here we use the properties [13] of Rhodamine 6G as a representative sample to calculate the Peclet number and the ratio liepj Ineo, where Hepj= 1.4xlO~4 cm2/(Vs), and Dt = 3.0x1 (T6
ctn/(s). The above described model for the electrical potential field, the electroosmotic flow field and the concentration field during the species transport process was solved numerically by using a 3D-computation code based on the finite volume approach [14] and SIMPLEC algorithm [14]. As a part of the computational domain, the virtual space in the roughness is also solved using the extension of computational region approach [14]. That is, when solving the electrical potential field, a conductivity function is set up so that it will has a value of zero on the control volumes inside the non-conducting roughness prisms, and a value of one on the control volumes in the bulk liquid phase. When solving the electroosmotic flow field, we consider the control volumes in the roughness elements as a kind of fluid with extremely high viscosity. When solving the species transport process, we consider the control volumes in the solid roughness elements as species with a zero diffusion coefficient. 6-1.2 Electroosmotic flow and the associated sample transport in smooth channels As the same electrical potential difference is applied to both the smooth and the rough microchannels with the same channel length, the electrical potential drop over the length of each period, A<j> = (j)out - <j)jn, is the same for the rough and the smooth microchannels. It is known that in a smooth channel, the applied electrical potential varies linearly along the length direction, and hence the electrical field strength and electroosmotic flow velocity are uniform in the whole channel, as given below:
Effects of Surface Roughness on Electrokinetic Flow
331
(16) (17) Referring to the non-dimensional groups used for the rough channels, the values Exs and Veo?s are the characteristic applied electrical field strength and the characteristic electroosmotic velocity. Therefore, the dimensionless values 0*, E and Ve0 in rough channels represent the relative values of electrical potential, the applied electrical field strength and the electroosmotic velocity at a certain point in the rough channel to the values of the smooth channel, respectively. Substituting the above Eqs. (16) and (17) into Eq. (14), the dimensionless governing equation for the associated sample transport through smooth channels can be derived from Eq. (15):
(18)
6-1.3 The applied electrical potential field and the electroosmotic flow field Due to the existence of the rough elements, the surfaces of the constant electrical potential are not parallel to each other. For example, the electrical potential field for the case of symmetrically distributed roughness elements is shown in Figure 6.4, where the colors represent the values of the dimensionless potential <j>. These isoelectric surfaces are distorted by the rough elements and therefore the electrical strength is not uniform in the microchannels. Thus the calculated local slip boundary velocities are different in the values and in the directions. As shown in Figures 6.5a and 6.5b, the top view of the slip boundary velocity on the bottom channel surface, the color scale shows the value of the dimensionless velocity component in x (the main flow) direction. The values bigger than one at certain regions indicate the slip velocity component in the main flow direction is larger than the electroosmotic velocity in the smooth channels. Figures 6.5a and 6.5b clearly shows that the slip velocity larger than one is mainly distributed in the lengthwise pathway formed between two neighboring rough elements (which will be called the "pathway" in the following), and the maximum value of ue0 is present around the corner in the pathway; while the slip velocity much less than one is mainly located at the gap between the front and the back surfaces of two neighboring rough elements (which will be called the "gap" in the following). It should also be noted that at
332
Electrokinetics in Microfluidics
the rough element's surface, especially at the front and the backside of the prism's surface, see Figure 6.5c, the distorted electrical potential builds up a varying slip boundary velocity field. On the backside surface of the roughness elements, the velocity vectors are directed towards the center of the intersection line of the roughness element and the substrate surface. The magnitude of the velocity decreases gradually as it approaches this point. On the front side, the slip boundary velocities have the same distribution except with the opposite directions. This indicates the driven force distribution around the rough element's surface. The velocity field of the electroosmotic flow and the induced pressure field in a rough microchannel with symmetrically arranged roughness elements are shown in Figure 6.6. Even though no pressure differences are applied to the inlet and outlet of the whole channels, local pressure gradients are induced by the existence of the roughness elements in order to satisfy the flow continuity. Figure 6.6 shows that there are high-pressure zones at the upwind regions and the low-pressure zones at the tail regions of the rough elements, the pressure in the central flow above the roughness has little changes. The side view of the flow
Figure 6.4. The electrical potential field in the symmetrically arranged rough microchannels with a/H- 0.6, b/H= 1 and h/H= 0.2. The surfaces of the constant electrical potential are distorted due to the presence of the rough elements. The color represents the value of the dimensionless potential <j>.
Effects of Surface Roughness on Electrokinetic Flow
333
Figure 6.5. The slip boundary velocity distribution on (a) the bottom plane in symmetrically arranged rough microchannel, (b) the bottom plane in asymmetrically arranged rough microchannel, and (c) the surface of one quarter of the rough element, with rough channel parameters: a/H= 0.6, b/H= 1 and h/H= 0.2. The arrows are local velocity vectors. The color represents the value of the dimensionless velocity.
334
Electrokinetics in Microfluidics
and the pressure fields also shows that the maximum values of the velocity component ueo at each cross section all occur on the roughened wall. The similar phenomena are found for the electroosmotic flow in microchannels with asymmetrically arranged rough elements. Figure 6.7 shows the 3D streamlines of the electroosmotic flow through the symmetrically arranged rough microchannels. One can clearly see that the streamlines are curved with the roughness shapes. This is true for all the simulations conducted for cases of symmetrically and asymmetrically arranged roughness under various conditions. There are no flow recirculation zones in the gap, no matter what roughness element's sizes, separation distances, heights, the channel heights and the Reynolds numbers are. This implies that there is no sample molecule trap or accumulation of sample molecules in the gap regions between rough elements when sample transporting with the electroosmotic flows.
Figure 6.6. The top view (upper) and the side view (lower) of the electroosmotic velocity and the induced pressure distribution in symmetrically arranged rough microchannels with parameters: a/H = 0.6, b/H= 1 and h/H = 0.2. The arrows are local velocity vectors. The color represents the value of the local dimensionless pressure.
335
Effects of Surface Roughness on Electrokinetic Flow
6-1.4 The influence on the electroosmotic flow rate and the induced pressure The roughness-induced pressure has a direct effect on the electroosmotic flow rate and the ability of removing the liquid from the gap to the flow pathway. In order to demonstrate the roughness effects on the flow rate, we define the corresponding electroosmotic flow rate in the virtual smooth channel as Qs = Veo,s^c,s • Since the pressure variations in x directions is key to the flow, we define the average dimensionless pressure in the cross section area as (19) where Ac is the cross section area in the rough microchannels. The effect of rough element's size on the induced pressure field Figure 6.8a shows the roughness size effect on the cross-section averaged dimensionless pressure (pave) in two-period-length symmetrically arranged rough channels. From Figure 6.8a, one can see that the induced pressure is periodically undulating along the rough channel, and the local maximum locates at the upwind region of the rough element, while the local minimum locates at the element's tail region. A higher maximum value of pave
indicates a larger
Figure 6.7. The 3D streamlines of the electroosmotic through symmetrically arranged rough microchannels with parameters: a/H= 0.6, b/H= 1 and h/H= 0.2.
336
Electrokinetics in Microfluidics
pushing force acting on the fluid element near the upwind region to push the fluid from the gap to the central flow. Similarly, at the tail region of the roughness element, the low-pressure attempts to suck the fluid from central flow into the gap. Such exchange of fluid between the gap and the central flows is referred as the even-out effect. From Figure 6.8a, one can clearly see that this even-out effect exhibits a maximum value at a/H = 0.4 when the roughness element's size changes. When a/H < 0.4, the electrical potential field distortion due to the presence of the roughness elements is rather small, and hence difference in the slip velocities and the induced pressure gradients at the front and the back faces of roughness elements is small, resulting in a relatively small even-out effect. When a/H > 0.4, the high-pressure zone and the low-pressure zone are very close to each other, the overlapping effect reduce the absolute value of the induced pressure in the gap, reducing the fluid exchange between the gap and the central flow. for the Figure 6.8b shows the roughness size effect on pave asymmetrically arranged rough microchannels. To show the pressure fluctuation clearly, only half flow period is shown in Figure 6.8b. Similar to the symmetrically arranged rough channels, the maximum of pave
occurs at the
upwind region and the minimum of pave occurs at the tail region of the roughness, as shown by the curve a/H = 0.2 in Figure 6.8b. It is obvious that there is a large difference between Figures 6.8a and 6.8b. This difference is originated from the roughness element's arrangement. For the asymmetrically arranged roughness as illustrated in Figure 6.5b, the locations of the high/low pressure zones on the opposite sides of the flow pathway are asymmetrical (i.e., at different x positions) along the flow direction. For the symmetrical arrangement, as seen in Figure 6.6, the high/low pressure zones on the opposite sides of the flow pathway are at the same x position. Thus the curve Pave ~ x for the asymmetrical arrangement is more complicated and the distance between the high-pressure zone and the next low-pressure zone is much smaller than that of the symmetrically arranged rough channels. The effects of the rough element's separation distance, height, and the channel height When varying each parameter of b, h and H in turn, the curves of pave ~ x have similar undulating characteristics, and thus we do not show the figures. However, it is worthwhile to note the following effects. When the separation distance between roughness elements, b, changes from lOum to 20um while keeping the other parameters fixed (a = 6um, h = 2um, H = 10am, symmetrical arrangement), the simulations show an increasing even-out effect (a larger
Effects of Surface Roughness on Electrokinetic Flow
337
Figure 6.8. The roughness size effect on the induced cross-section averaged dimensionless pressure p*ave in (a) two periods of the symmetrically arranged rough microchannel, and in (b) half period of the asymmetrically arranged rough microchannel with parameters: b/H= 1 and h/H= 0.2.
338
Electrokinetics in Microfluidics
maximum value of pave) caused by the pressure-pushing force in the pathway. When the roughness element's height, h, increases, for example from 2um to 5|^m, while the other parameters are fixed ( # = 1 0 um, b = 10 um, a = 6um), the even-out effect is smaller because the velocity component on the backside of the rough element's is smaller. When the rough channel's height, H, increases, for instance from 10um to 30u.m, the curves of pave~ x overlap with each other, this indicates the induced cross-section averaged pressure resistance, dpaveldx, is proportional to the \IH under the same Reynolds number. As the central flow above the roughness contributes little to the pressure undulation, the induced high (or low) pressure at the front (or back) side of the rough elements does not vary so much when changing only the rough channel's height. The roughness effects on the electroosmotic flow rate The above-mentioned roughness effects on the electrical field and on the induced pressure field ultimately influence the electroosmotic flow rate. Figure 8 shows the influences of the rough microchannel geometry parameters on the electroosmotic flow rate, where Q/Qs is the ratio of the flow rate through the rough channels to that of the virtual smooth channels. Figure 6.9 shows that the
Figure 6.9. The effect of roughness parameters a, b, h, H on the electroosmotic flow rate with default parameters: a = 6(^m, b = lO^m, H= lO^im, and h = 2um.
Effects of Surface Roughness on Electrokinetic Flow
339
induced pressures dramatically reduce the electroosmotic flow rates through rough microchannels, and the effect increases as the roughness size or the roughness height increases, and decreases as the roughness separation distance increases. As the rough channel's height increases, the ratio of the flow rate to that of the smooth channel decreases at first and then increases. When H decreases from 20um to lOum, relatively more electroosmotic driven force is produced due to the relatively larger rough element's surface and the larger electrical field distortion. These two effects contribute to the slight increase of the flow rate compared to the electroosmotic flow rate in smooth channel. For all the asymmetrically arranged roughness, the flow rates are all larger than those of the symmetrically arranged roughness. This is because the asymmetrical arrangement produces more streamlined flows, the induced pressure resistance is divided into more steps while the fluctuation amplitude is smaller and the pressure resisting flow mostly happens in each step. 6-1.5 Sample mass transport in the rough microchannels The characteristic of sample mass transport in the rough microchannels Generally, sample mass transport is the result of three major mass transport mechanisms: convection, electrophoretic transport and diffusion. In view of the characteristics of the electroosmotic flow field and the electrical potential field in rough microchannels, sample mass transport in rough microchannels is very different from most reported studies where smooth channel walls are considered. Figure 6.10a shows the sample concentration field in the computational domain when a certain amount of sample is released into the computational domain at the inlet plane. In the simulations, the electrokinetic properties are chosen as /ieo = 4x1 CT8 m2/Vs and fj.ep = 1.4xl(T8 rn/Vs. The color scale from dark to light green represents the value of dimensionless concentration from zero to one. For a given quantity of sample released into the computational domain, Figure 6.10a shows the concentration field in the computational domain during a period of/* = 9.15. At the end of this period, 71.5% of the sample remains in the computational domain and the rest has flown out. As seen from Figure 6.10a, there's no 3D enclosed iso-concentration surface in the gap region between the roughness elements. This implies that there will be no micro sample molecules trapped or accumulated in the gap regions. Finite sized particle may be trapped in the gaps. However, such particle-liquid systems are well beyond the scope of this paper. The model presented here generally is not applicable to the (micron sized) particle-liquid systems. The sample-liquid systems in this study are limited to small molecule-solution systems. If there is a 3D enclosed recirculation flow zone in the gaps between roughness elements, such a recirculation flow zone is referred to as the liquid molecule trap. The even-out effect under the electroosmotic driven flow as discussed in this work avoids such
340
Electrokinetics in Microfluidics
a recirculation flow zone in the gaps. It is also noticed from the outlet plane of the computational domain that the sample concentration at the near wall region is higher than the region above. This is due to the larger values of the electroosmotic velocity near the channel wall region, while the central velocity field above the roughness region was impaired by the induced pressure gradient. Figure 6.10b shows the sample concentration field on the rough microchannel surface under the same conditions as in Figure 6.10a. From Figure 6.10b one can see that the concentration in the pathway surface is the highest. The effect of electrokinetic mobility on sample transport To examine the effects of electrokinetic mobility, we will compare the
Figure 6.10. The sample concentration field in and h/H = 0.2 at the time when a given microchannels. The concentration field (a) in rough surface with fxeo = 4x10~8 nf/Vs and \iep =
a rough microchannel: a/H = 0.4, b/H = 1 amount of sample is released into the the computational domain and (b) on the 1.4xlO"8 m2/Vs.
Effects of Surface Roughness on Electrokinetic Flow
341
sample mass transport through the microchannel for a given quantity of sample flowing into the computation domain. The quantity of sample flowing into the computation domain is determined by the original sample concentration Co, and the time of releasing. The amount of sample flowing into the computation domain is defined by: (20)
where Nin is the quantity of the sample molecule flowing into the computation domain during a period At; Nin is its dimensionless form, Nin = Njn / CQH ; Ain
Figure 6.10. The sample concentration field in a rough microchannel: a/H= 0.4, b/H= 1 and h/H = 0.2 at the time when a given amount of sample is released into the microchannels. The concentration field (c) in the computational domain and (d) on the rough surface with neo =
6xlQ-8 rn/Vs and nep = lAxl0's tn/Vs.
342
Electrokinetics in Microfluidics
is the cross section area at the inlet of computational domain, At is the duration that the released sample Nin reaches a given value. It should be noted here that the amount of sample transferred into the inlet of the computational domain by diffusion is much smaller than that by electroosmotic convection and electrophoresis, and therefore is neglected. We focus on the effect of electroosmotic mobility neo and the electrophoretic mobility jj.ep on the sample transport through rough microchannels. Figure 6.10c shows the sample concentration field in the whole computational domain with an increased electroosmotic mobility value, jieo = 6x10~8 m/Vs, while all other parameters are the same as in Figures 6.10a and 6.10b. As the electroosmotic mobility increases the electroosmotic velocity increases proportionally, therefore the duration of releasing the same quantity of sample into the computational domain is shorter, t = 6.13 in Figure 6.10c (note
Figure 6.10. The sample concentration field in a rough microchannel: a/H = 0.4, b/H= 1 and h/H = 0.2 at the time when a given amount of sample is released into the microchannels. The concentration field (e) in the computational domain and (f) on the rough surface with jieo = 4xlO"8 rn/Vs and /iep = -1.4xl0" 8 m2/Vs.
Effects of Surface Roughness on Electrokinetic Flow
343
t* = 9.15 in Figure 6.10a). From Figure 6.10c one can see that the concentration in the gap region is higher than that in Figure 6.10a, and there are overall more sample molecules (76% comparing with 71.5% in Figure 6.10a) in the whole computational domain. Figure 6.10d shows the corresponding concentration field on the roughened microchannel surface. Clearly the surface concentration is more uniform. When \xep = -1.4xlO~8 rn/Vs, i.e., the same absolute value of the electrophoretic mobility as in Figures 6.10a and 6.10c but with a different sign, while keeping the rest parameters the same, one sees a further improved concentration filed in the whole computational domain as shown in Figure 6.10e, and on the channel surface as shown in Figure 6.1 Of. When the electric charge of sample molecules has a different sign from that of the channel wall, the electrophoretic motion is in the same direction as that of the electroosmotic flow, and the sample molecule velocity is larger than the electroosmotic velocity of the bulk liquid. This mechanism contributes significantly to enhance the concentration uniformity in the channel's cross section area, and to increase the amount of the sample molecules in the gap regions, and the surface concentration. The duration of releasing the same quantity of sample into the computational domain is reduced further to / = 2.36, and more sample molecules, 81.7%, are in the computational domain. 6-1.6 Summary The electroosmotic flow through microchannels is greatly influenced by the presence and characteristics of roughness on the channel walls. The electrical field is distorted by surface roughness elements, which makes the electroosmotic slip velocity non-uniform. Due to the presence of roughness elements in the flow passage, the electroosmotic flow in the rough microchannels induces a periodic pressure field that makes the central flow velocity smaller than that in the near wall region and hence the flow rate through the rough microchannel is significantly reduced. The induced high-pressure zone occurs at the upwind region of the rough elements and the low-pressure zone is at the tail region. The induced pressure field causes the exchange of liquid between the gaps and the central flow, which is referred to as the even-out effect. The roughness element's size, height and the separation distance have large influences on the even-out effect, the microchannel height has a small influence on the even-out effect. There is no sample concentration trap in the gap regions between roughness elements during the electroosmotic transport. The increase of the electroosmotic mobility or the decrease of the electrophoretic mobility can significantly improve the uniformity of the sample concentration field in a rough microchannel.
344
6-2
Electrokinetics in Microfluidics
EFFECTS OF 3D HETEROGENEOUS ROUGH ELEMENTS
As we have seen in the previous section, the 3D roughness elements will influence liquid flow and sample transfer in microchannels. If the surface of these 3D roughness elements is chemically modified, or the rough elements are made of different materials, the electrokinetic properties of these rough elements will be different from that of the substrate surface. This will bring more complication to the electrokinetic transport processes. This section will examine the influence of micron-sized heterogeneous rough elements on electroosmotic flow and the associated mass transfer in microchannels [15]. We consider electroosmotic transport in a slit microchannel (formed between two parallel plates) with numerous heterogeneous prismatic roughness elements on the microchannel walls. All 3D prismatic rough elements have the same surface charge or zeta potential, the substrate (the microchannel wall) surface has a different zeta potential. We consider two types of distributions: symmetrical and asymmetrical arrangements, as illustrated in Figure 6.2 in the last section. The electroosmotic flow field and the sample concentration field are investigated in terms of the influence of the rough element size, height, density, the element arrangement, the channel size and the ratio of the electroosmotic mobility of the rough element's surface to that of the substrate. 6-2.1 Mathematical model The model assumptions and the computational domain (see Figures 6.2 and 6.3) are the same as in the previous section. The mathematical models of the electrical field, flow field and the concentration field are also the same as in the case of homogeneous roughness channels in the last section, except the slip flow boundary velocities. Because the roughness elements in this case have a different zeta potential from the substrate surface, the slip flow boundary conditions are different between the substrate (channel wall) surface and the roughness elements' surface. This will make the characteristics of the electroosmotic flow in this case very different from that shown in the previous section. The slip boundary velocity on the substrate surface is given by:
The slip boundary velocity on the roughness elements' surface is given by:
Effects of Surface Roughness on Electrokinetic Flow
345
where s^ = jj,eOiR /fxeo is the ratio of the electroosmotic mobility of the rough elements' surface to that of the substrate surface. The influence of electrophoretic mobility on the electrokinetic flow through microchannels was studied in the previous section. Therefore, in this section, we focus on an approximately electric neutral species [16], Rhodamine B, as a representative sample in the electrokinetic species transport in heterogeneous rough microchannels. The electrophoretic mobility of Rhodamine B is negligible under certain pH values, therefore we simply take juepj/jueo = 0. The diffusion coefficient of Rhodamine B is taken as [16] £),• = 4.37x10" cm /(s).
Figure 6.11 The slip boundary velocity distribution on (top) the bottom plane in a symmetrically arranged heterogeneous rough microchannel, and (bottom) the surface of one quarter of the rough element, with roughness parameters: a/H= 0.6, b/H= 1, h/H= 0.2 and e^ = -0.5. The arrows are local velocity vectors. The color represents the value of the dimensionless velocity.
346
Electrokinetics in Microfluidics
6-2.2 The applied electrical potential field and the electroosmotic flow field Due to the existence of the rough elements, the surfaces of the constant electrical potential are not parallel to each other. The isoelectric surfaces are distorted by the rough elements and therefore the electrical strength is not uniform in the microchannels. Thus the local slip boundary velocities are different in values and in directions. As shown in Figure 6.11, the top view of the slip boundary velocity on the bottom channel surface and 3D view of the slip boundary velocity on the rough element surface, the color scale shows the value of the dimensionless velocity component in x (the main flow) direction. The values bigger than one at certain regions indicate the slip velocity component in the main flow direction is larger than the electroosmotic velocity in the smooth channels. Figure 6.1 l(top) clearly shows that the slip velocity larger than one is mainly distributed in the lengthwise pathway formed between two neighboring rough elements (the "pathway"), and the maximum value of ueo is present around the corner in the pathway; while the slip velocity much less than one is mainly located at the gap between the front and the back surfaces of two neighbor rough elements (the "gap"). It should also be noted that at the rough element's surface, especially at the front and the backside of the prism's surface, refer to Figure 6.11 (bottom), the distorted electrical potential builds up a varying slip boundary velocity field. On the front and the backside surface of the roughness elements, the velocity vectors have a centrifugal or centripetal distribution with the center at the central point of the intersection line of the roughness element and the substrate. The value of the local slip boundary velocity on the roughness element's surface is proportional to the electroosmotic mobility. The velocity field of the electroosmotic flow and the induced pressure field in a symmetrically arranged rough microchannel with ^ = -0.5 and s^= 1.5 are shown in Figure 6.12 and Figure 6.13, respectively. The electroosmotic mobility ratio sM = -0.5 implies that the surface charge of the rough elements has an opposite sign to that of the substrate surface, and the value of the zeta potential of the rough elements is 50% smaller than that of the substrate surface, which forms the negative (opposite to the x-axis or main flow direction) slip boundary velocities on the roughness surface (e^ = -0.5) and the positive (in the x-axis or main flow direction) slip boundary velocities on the substrate. Figure 6.12(a) shows that when e^ = -0.5 there are high pressure zones located near the front of the roughness and the entrance of the pathway, and low pressure zones located near the backside of the roughness and the exit of the pathway. As seen in this figure, the first high pressure zones are near the rough element's front surface. This is because the flow from upwind driven by the positive substrate electroosmotic force and the counter-flow caused by the negative electroosmotic driving force on the rough elements' surface congest in front of the roughness
Effects of Surface Roughness on Electrokinetic Flow
347
Figure 6.12 The electroosmotic velocity and the induced pressure distributions in symmetrically arranged heterogeneous rough microchannels with parameters: a/H= 0.6, b/H = 1, h/H= 0.2 and e^ = -0.5 on (a) the plane y = 0.007 (the bottom substrate surface); (b) the plane y* = 0.035 (a horizontal plane above the substrate surface); (c) the plane z = 0.18 (a cross-section cutting through two rough elements and the gap in between); and (d) the middle vertical plane z* = 0.5. The arrows are local velocity vectors. The color represents the value of the local dimensionless pressure.
348
Electrokinetics in Microfluidics
elements and form the high pressure zones. The next high pressure zones are located at the entrance corners of the pathway near the roughness element, which is caused by the countercurrents on the roughness side surfaces forming the pathway. In the gap between two rough elements (Figure 6.12(c), the vertical surfaces of the rough elements are perpendicular to the horizontally applied electrical field, and hence there is no active electroosmotic flow on these surfaces. However, the positive (moving in right or x-axis direction) flow on the substrate surface generates a high pressure zone in front of the second rough element and a negative pressure zone near the back surface of the first rough element. These high and low pressure zones produce a counter-clock-wise circulation flow in the gap, as shows in the side view of the flow field, Figure 6.12(c). Further investigation shows that the circulation not only locates in the gap, it also propagates into the intersection. Figure 6.12(d) shows the velocity and pressure distributions in the plane z = 0.5, i.e., the middle vertical plane of the computational domain. One can see a larger circulation with the center a little higher than that of the circulation in the gap. Therefore, there is a circulation flow crossing over the width of the heterogeneous rough microchannels. As the negative s^ value increases, lower circulation center and higher circulation strength were found. The electroosmotic mobility ratio s^ - 1.5, implies that, although the rough elements' surface has the same type of surface charge, the zeta potential of the rough element surface is 50% higher than that of the substrate surface. As seen in Figure 6.13, there are high pressure zones located near the front of the roughness and the exit of the pathway, low pressure zones located near the backside of the roughness and the entrance of the pathway at the plane close to the substrate surface. In comparison with Figure 6.12, the different
Figure 6.13 The electroosmotic velocity and the induced pressure distribution in symmetrically arranged heterogeneous rough microchannels with parameters: a/H = 0.6, b/H = 1, h/H= 0.2 and e^ = 1.5 on the plane y = 0.007 (the bottom substrate surface).
Effects of Surface Roughness on Electrokinetic Flow
349
position of the high pressure zone in this case is caused by the larger slip boundary velocity (e^ = 1.5) on the roughness element's surface, which forms the flow congestion in the pathway and induces the pressure resistance in the pathway. In the electroosmotic flow through heterogeneous rough channel with a positive roughness heterogeneity s^ >1, no flow circulation were found. 6-2.3 The influence of the heterogeneous roughness on the electroosmotic flow rate The synergic roughness and heterogeneity effect on the slip boundary velocity and on the induced pressure field ultimately influence the electroosmotic flow rate. The value Q/Qs is the ratio of the flow rate through the heterogeneous rough channels to that of the virtual smooth and homogeneous channels. Similar to the homogeneous rough microchannels, the heterogeneous rough elements' size, height, separation distance, arrangement and the channel height determine the excess driven force at the solid-liquid interface relative to that of the smooth channel, the position and the frequency of the local pressure zones and the total resistance to the flow. Additionally, the electroosmotic mobility ratio sM determines the relative value and the direction of the electroosmotic driven force on surfaces, which also influence the flow patterns, local pressure distribution and the total resistance to the electroosmotic flow and hence the flow rate in the heterogeneous rough channels. The coupled effect of the heterogeneity and the roughness on the flow rate is shown in Figure 6.14, with s^ = -0.5. From Figure
Figure 6.14 The effect of the heterogeneous roughness parameters a, b, h, H on the electroosmotic flow rate with parameters as: a = 6(im, b = lOum, H = lOum, h = 2|im and £„ = -0.5.
350
Electrokinetics in Microfluidics
6.14, one can see that the two roughness arrangements have almost the same flow rate for the given set of rough microchannel geometry parameters. The relative flow rate to that of the smooth channel increases linearly when the roughness separation distance or the channel height increases. This is because when the roughness separation distance or the channel height increases, the roughness resistance effect decreases and hence the flow and the flow rate becomes closer to that in the smooth microchannels. The relative flow rate decreases when the roughness size increases. These behaviors are similar to the flow in the homogeneous rough microchannels. It is worthy to note that when the roughness height increases, the value Q/Qs increases from negative to a value close to zero. This is because when increases the roughness height, the increase of the induced pressure resistance on the opposite flow exceeds the effects of the counter flow driven force on the roughness surface. The studies of the effects of the degree of the roughness heterogeneity or the electroosmotic mobility ratio e^ with identical rough microchannel geometry parameters show that the relative flow rate increases linearly when increases the electroosmotic mobility ratio eM in the range of (-1 ~ 1.5). For example, with the rough microchannel geometry parameters set as a/H'= 0.6, b/H = 1.0, h/H = 0.2 and H = lOum for symmetrical and asymmetrical arrangements, Q/Q value ranges from -0.2 to 0.5 when s^ increases in the range of (-1 ~ 1.5). This indicates the roughness always acts as a hindrance of the flow, no matter the flow is in its normal direction with positive s^, or the flow change its direction due to large negative driven force on the rough elements when the roughness heterogeneity £M is less than a certain negative value. 6-2.4 Sample mass transport in the heterogeneous rough microchannels Generally, sample molecule transport is the result of three major mass transport mechanisms: convection, electrophoretic transport and diffusion. When electrically neutral species are released into the heterogeneous rough microchannels, sample transport is controlled by two mechanisms: convection and diffusion. Figures 6.15(a), (b) and (c) show the concentration fields of an electrically neutral sample (Rhodamine B in this case) in the microchannels at a moment when a given (the same) amount of sample molecules enters the inlet surface of the computation domain. The non-dimensional time /* is defined in Eq. 14. It should be noted that sM = -0.5 in Figures 6.15(a) and (b), and £^ = 1.5 in Figure 6.15(c). The electroosmotic flow is much faster in the case of Figure 6.15(c), consequently the time required to transport the same amount of sample molecules is shorter in Figure 6.15(c). Figure 6.15 shows the concentration fields of the vertical middle plane z = 0.5 and the bottom plane y = 0 to present the sample transport in the pathway and in the intersection region. From Figures
Effects of Surface Roughness on Electrokinetic Flow
351
Figure 6.15 The sample concentration field in a heterogeneous rough microchannel on the bottom plane y = 0 and the middle vertical plane z = 0.5 with roughness parameters as a/H = 0.6, b/H = 1 and h/H = 0.2 for (a) symmetrical arrangement and e^ = -0.5 at time t = 39.9; (b) asymmetrical arrangement and en = -0.5 at time t = 39.9; (c) symmetrical arrangement and £„ = 1.5 at time t = 1.03.
352
Electrokinetics in Microfluidics
6.15(a) and 6.15(b), one can clearly see the flow circulation, as indicated by the concentration field contours. The sample transport in the heterogeneous rough microchannel forms a tidal wave like concentration distribution in the plane z = 0.5. For asymmetrical arrangement, it forms four circulation zones with less circulation strength in the computation domain. This is caused by the staggerly distributed roughness elements that provide one more T-shape intersection region in each flow period. For symmetrical arrangement, it forms two circulation zones with higher circulation strength. This is caused by the large cross-shape intersection and stronger counter flow driven force due to the larger roughness surface area than that of the asymmetrical arrangement. When the electroosmotic mobility ratio e^ is increased to 1.5, there is no circulation exist in the flow and the sample concentration profiles in the vertical plane is similar to that of the smooth channel. The concentration field in the gap region between two rough elements is shown in Figure 6.16 for the case of Figure 6.15(a), i.e., a/H'= 0.6, b/H'= 1 and h/H = 0.2 and e^ = -0.5 in symmetrical arrangement at t = 39.9. Figure 6.16 shows that the sample circulation in the gap region occurs close to the substrate. It also indicates that the sample molecules moving into the gap is mostly from the main flow pathway, few sample molecules move over the roughness elements and enter the gap. This behavior is very different from the sample transport in homogeneous rough channels, where sample transports from both top and side of the rough elements and fills in the gap region.
Figure 6.16 The sample concentration field in a heterogeneous rough microchannel on the vertical plane z* = 0.28 with roughness parameters as a/H = 0.6, b/H = 1 and h/H = 0.2 and e^ = -0.5 for symmetrical arrangement at time t = 39.9.
Effects of Surface Roughness on Electrokinetic Flow
353
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16]
D. Long, H.A. Stone and A. Ajdari, J. Colloid Interface Sci., 212 (1999) 338. A.D. Stroock, S.K. Dertinger, G.M. Whitesides and A. Ajdari, Anal. Chem., 74 (2002) 5306. D. Erickson and D. Li, Langmuir, 18 (2002) 1883. B. Potocek, B. Gas, E. Kenndler and M. Stedry, J. Chromatogr. A, 709 (1995) 51. A. Ajdari, Phys. Rev. Lett., 75 (1995) 755. A. Ajdari, Phys. Rev. E, 53 (1996) 4996. A. Ajdari, Phys. Rev. E, 65 (2002) 016301. A.D. Stroock, M. Week, D.T. Chiu, W.T.S. Huck, P.J.A. Kenis, R.F. Ismagilov and G.M. Whitesides, Phys. Rev. Lett., 84 (2000) 3314. Y. Hu, C. Werner and D. Li, Anal. Chem, 75 (2003) 5747-5758. D. Erickson and D. Li, Langmuir, 18 (2002) 8949. S.V. Patankar, C.H. Liu and E.M. Sparrow, ASME J. Heat Transfer, 99 (1977) 180. R.S. Amano, ASME J. Heat Transfer, 107 (1985) 564. S.V. Ermakov, S.C. Jacobson and J.M. Ramsey, Anal. Chem., 70 (1998) 4494. S.V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere, New York, 1980. Y. Hu, C. Werner and D. Li, Proceedings of 2nd International Conference on Microchannels and Minichannels, Rochester, New York, June 17-19, 2004. D. Sinton, D. Erickson and D. Li, J. Micromech. Microeng., 12 (2002) 898.
354
Electrokinetics in Microfluidics
Chapter 7
Experimental studies of electroosmotic flow Electro-osmotic pumping (i.e., transporting liquids in microchannels by electroosmosis) is a preferred method in a variety of Lab-on-a-Chip devices, because the electroosmotic pump has no moving mechanical parts, zero-volume (i.e., it requires only electrodes that can be embedded in the chip) and provides uniform cross-section flow (i.e., plug-type velocity profile). To design and to operate an electroosmotic pump, it is critical to know the volumetric flow rate or the velocity of the electroosmotically pumped flow. As the flow rates involved in the lab-chip devices usually are very small (e.g., 0.1 uL/min.), and the size of the microchannels is very small (e.g., 10-100 urn), it is extremely difficult to measure directly the flow rate or velocity of the electroosmotic flow in these microchannels. To study liquid flow in microchannels, various microflow visualization methods have evolved. Micro particle image velocimetry (microPIV) is a method that was adapted from well-developed PIV techniques for flows in macro-sized systems [1-5]. In the microPIV technique, the fluid motion is inferred from the motion of sub-micron tracer particles. To eliminate the effect of Brownian motion, temporal or spatial averaging must be employed. Particle affinities for other particles, channel walls, and free surfaces must also be considered. In electrokinetic flows, the electrophoretic motion of the tracer particles (relative to the bulk flow) is an additional consideration that must be taken. These are the disadvantages of the microPIV technique. Dye-based microflow visualization methods have also evolved from their macro-sized counterparts. However, traditional mechanical dye injection techniques are difficult to apply to the microchannel flow systems. Specialized caged fluorescent dyes have been employed to facilitate the dye injection by using selective light exposure (i.e., the photo-injection of the dye). The photo-injection is accomplished by exposing an initially non-fluorescent solution seeded with caged fluorescent dye to a beam or a sheet of ultraviolet light. As a result of the ultraviolet exposure, caging groups are broken and fluorescent dye is released. Since the caged fluorescent dye method was first employed for flow tagging velocimetry in macro-sized flows in 1995 [6], this technique has since been used to study a variety of liquid flow phenomena in microstructures [7-15]. The
Experimental Studies of Electroosmotic Flow
355
disadvantages of this technique are that it requires expensive specialized caged dye and extensive infrastructure to facilitate the photo injection. As an alternative to photo-injection of fluorescent dye, a flow marker can be created in a uniform solution of fluorescent dye by local photobleaching [11,15,16]. In additional to these PIV and dye-based techniques, the electroosmotic flow velocity can also be estimated indirectly by monitoring the electrical current change while one solution is replaced by another similar solution during electroosmotic flow [12,17,18]. In this chapter, we will discuss the indirect techniques and direct techniques of measuring the electroosmotic flow velocity.
356
7-1
Electrokinetics in Microfluidics
MEASUREMENT OF AVERAGE ELECTROOSMOTIC VELOCITY BY A CURRENT METHOD
An experimental approach was developed [17] to evaluate the electroosmotic flow velocity by monitoring the current change in a process where one solution replaces another similar solution in a microchannel. In this method, a capillary tube is filled with an electrolyte solution, then brought into contact with another solution of the same electrolyte but with a slightly different ionic concentration. Once the two solutions are in contact, an electrical field is applied along the capillary in such a way that the second solution is pumped into the capillary and the 1st solution flows out of the capillary from the other end. As more and more of the second solution is pumped into the capillary and the first solution flows out of the capillary, the overall liquid conductivity in the capillary is changed, and hence the electrical current through the capillary is changed. When the second solution completely replaces the first solution, the current will reach a constant value. Knowing the time required for this current change and the length of the capillary tube, the average electroosmotic flow velocity can be calculated by (1) where L is the length of the capillary and At is the time required for the higher (or lower)-concentration electrolyte solution to completely displace the lower (or higher)-concentration electrolyte solution in the capillary tube.
Figure 7.1. Illustration of an experimental setup for measuring the current change during an electroosmotic displacing process.
Experimental Studies of Electroosmotic Flow
357
Figure 7.1 shows an experimental set-up used to measure the average velocity by monitoring the current change in the electroosmotic displacing process. In the experiments, polyamide coated silica capillary tubes of various diameters (Polymicro Technologies Inc., Phoenix, AZ) were cut to 10cm in lengths and used to connect the electrolyte solution in reservoir 1 to the electrolyte solution in reservoir 2. The reservoir 2 was filled with an electrolyte solution at a desired concentration. The capillary tube and reservoir 1 were filled with an electrolyte solution at a concentration lower than the concentration in reservoir 2. For example, the capillary tube and reservoir 1 were filled with an electrolyte solution at a concentration that is 95% of the concentration in reservoir 2. Immediately after connecting the two reservoirs by the capillary tube, a voltage difference between the two reservoirs was applied by setting reservoir 1 at ground potential and reservoir 2 to a high voltage power supply unit (CZE1000R, Spellman, NY) via Platinum electrodes. The applied electrical field results in an electroosmotic flow in the capillary tube. During electroosmosis, the higher-concentration electrolyte solution from reservoir 2 gradually displaces the lower-concentration electrolyte solution in the capillary tube. As a result, the overall electrical resistance of the liquid in the capillary tube changes. By monitoring the change in the electrical current under a constant applied voltage difference during electroosmosis, the change in electrical resistance can be monitored. An L-DAS8 data acquisition chip (Kieffley) was used to record the voltage (kV) and current (uA) as a function of time (seconds). Once the lower-concentration solution in the tube is completely replaced by the higher-concentration solution from reservoir 2, the
Figure 7.2. A typical result of current versus time for the specific case of capillary diameter D=100 urn and Ex=3500V/10 cm. KCL concentration in Reservoir 1 is C95%=0.95xl0"4 M. KCL concentration in Reservoir 2 is Cioo%= lxl(T 4 M.
358
Electrokinetics in Microfluidics
Figure 7.3 The measured average velocity versus the applied voltage and the capillary diameter, for a 10 cm long polyamide coated silica capillary tube.
Experimental Studies of Electroosmotic Flow
359
current will reach a maximum and constant value. The measured time for the current to reach such a plateau value is the time required for the solution from reservoir to travel through the entire capillary tube. The average velocity of the liquid flow can then be calculated by using Eq.(l). The average velocity of the liquid during electroosmosis is determined by Eq. (1). Figures 7.3a, 3b and 3c are plots of the average velocity versus the applied voltage and the capillary diameter for a polyamide coated silica glass capillary of 10 cm in length, for 3 different ionic concentrations. As seen in the figures, an almost linear relationship between the applied voltage and the average velocity is evident. In addition, for each plot it is clear that the average velocity is independent of the diameter of the capillary tubes, a relationship that has also been shown for rectangular microchannels in a numerical study [19]. In the experimental studies, each measurement was repeated at least three times for a given set of conditions. All experiments were conducted at room temperature (22°C). In Chapter 4, we have discussed the general electroosmotic solution displacing processes between two solutions of different electrolytes and very different concentrations. We may apply the model developed in Chapter 4 to examine the electroosmotic flow of one solution replacing another similar solution. The current-time relationship, the time evolution of diffusing zone between the two solutions, and the average velocity of the electroosmotic flow can be predicted in this way. The model predictions will be compared with the experimental data. When the two solutions are brought into contact in a microchannel under an applied electrical field, the mixing (involving convection and diffusion) occurs near the interface of two solutions. As discussed in Chapter 4, during the solution replacing process, the capillary can be divided into three sections according to the concentration distribution. The first section is filled with one solution, the second section is the diffusing zone and the third section is filled with another solution. The total electrical resistance of the capillary and the
Figure 7.4. Coordinator system for a cylindrical capillary tube. Z is the flow direction.
360
Electrokinetics in Microfluidics
electrical resistance of each section depend on the type of ions and the ionic concentration distribution in the capillary and vary with time. The electrical current, which is generated by the ions' motion under the applied electrical field, is axially uniform and varies with time. Consequently, the electrical field strength along the capillary (volts per meter) is different from location to location. If the two solutions are very different in terms of the ion type, the ionic valence, and the ionic concentration, the zeta potential, £ , and the ion distributions in the EDL field will be different in different sections. Therefore, the conditions of the electroosmotic flow are different in different sections. However, here we consider the electroosmotic replacing flow of two solutions that contain the same kind electrolytes but have a small difference in ionic concentration. For such a system, the zeta potential and the EDL field can be approximated as the same along the microchannel. As discussed before, if we consider a symmetrical electrolyte solution ( z : z = 1:1), the net charge density pe and the electrical potential \j/{r) in the electrical double layer field in a cylindrical capillary tube are described by the Boltzmann equation and the Poisson-Boltzmann equation, respectively. (2)
(3) where s is the dielectric constant of the solution and SQ is the permittivity of vacuum; ni<x> and zt are the bulk ionic concentration and the valence of type i ion, respectively; e is the charge of a proton, k^ is Boltzmann constant and T is the temperature. Introducing the dimensionless variables, (4)
where d is the diameter of the capillary tube and r is the radius of the capillary tube, the non-dimensional Poisson-Boltzmann equation can be written as
Experimental Studies of Electroosmotic Flow
361
(5)
Because of the symmetry of the EDL field in the cylindrical capillary, Eq. (5) is subjected to the following boundary conditions: (6)
where C, is the zeta potential. During the displacing flow, there are three sections of different ionic concentrations in this capillary. As shown in Figure 7.1, the solution on the left side (entering the capillary) has a relatively higher concentration, C2, the solution on the right side (leaving the capillary) has a relatively lower concentration, Cj, and between these two section, there is a diffusion zone due to the concentration difference. For the case interested here, we consider (C2 - Ci)IC2 < 10%. The liquid is considered as an incompressible fluid and there are no velocity components in r and 9 directions. That is,
Substituting the above conditions into the continuity equation for incompressible fluid: (7) yields, (8) It is obvious that the velocity component in z-direction uz is constant axially and can be calculated based on any section of the capillary. For simplicity, we calculate the velocity for a pure solution section. At a given time, the flow of a pure solution can be considered as a fully developed, onedimensional flow. The momentum equation for such a case is given by:
362
Electrokinetics in Microfluidics
Figure 7.5. Comparison of the experimentally determined current - time relationship with the model prediction under 35 KV/m.
Experimental Studies of Electroosmotic Flow
363
(9) where ^i is the viscosity, pei is the local net charge density and Ej is the electrical field strength applied to the z'-th section. Substituting the Eq. (2) for the net charge density into the Eq. (9) and introducing the following non-dimensional variables,
the non-dimensional equation of motion can be obtained as: (10)
Figure 7.6. Comparison of the experimentally determined average velocity - concentration relationship with the model prediction for KCl solution under an electrical field of 35 KV/m.
364
Electrokinetics in Microfluidics
which is subjected to the no-slip boundary conditions: (lla) (lib) Consider the electroosmotic flow for the system shown in Figure 7.1. During the electroosmosis, the high concentration solution gradually displaces the lower concentration solution in the capillary tube. Because there is difference in chemical potential between the high concentration solution and the low concentration solution, the diffusion appears across the interface, meanwhile, the convection also occurs due to the movement of the liquid. Consequently, the concentration in diffusion zone changes with time, and the diffusing zone length changes. The concentration distribution is governed by the conservation law, which for this case takes the form: (12) where C is the bulk ionic concentration in capillary tube, uz is the electroosmosis velocity of the liquid which can be determined by the equation of motion and D is the diffusion coefficient of the electrolyte in the solution. Introducing the following non-dimensional parameters:
where z represents the flow direction as shown in Figure 7.4. The equation of concentration can be non-dimensionalized as follows: (13)
The above equation is subjected to the following initial and boundary conditions: (14a)
Experimental Studies of Electroosmotic Flow
365
(14b)
(14c)
where Ltoto/ and rf are the total length and the internal diameter of the capillary tube, respectively. In this case, Ltoto/ is 10 cm and d is 0.01 cm. L\ is the length of the section filled with the low concentration (C/) electrolyte solution. Now we have the complete set of equations, Poisson-Boltzmann equation (5), motion equation (10) and concentration equation (13) and the matching initial and boundary conditions. In this model, the diffusing zone length is defined as follows: The position where C = 99.99%C2 will be the left boundary of the diffusing zone. The position where C - 99.99%Cj will be the right boundary of the diffusing zone. It is obvious that the length of the diffusing zone and the lengths of the other two sections change with time. Consequently, according to the equation of electrical resistance, (15) where Lj is the length of the z'-th section, C,- is the concentration of the i-th section, Xj is the bulk conductivity of the i-th section and A is the cross section area of the capillary, the overall electrical resistance of the liquid in the capillary tube will change with time during the electro-osmosis. At a given time, the total resistance is the sum of the resistance of the three sections, given by: (16) where R\ is the electrical resistance of the section with the low concentration (C/) electrolyte solution; R2 is the electrical resistance of the section with the higher concentration solution (C2), and Rmix is the electrical resistance of the diffusing zone where the ionic concentration is between C] and C2. When the total resistance and electrical voltage applied to the capillary are known, the electrical current through the capillary can be determined by:
366
Electrokinetics in Microfluidics
(17) where Vtotai is the total electrical voltage applied to the capillary. As the current keeps constant along the capillary and the electrical resistance varies from section to section (see Eq. (15)), the electrical field strength for each section will be different and can be calculated by (18) Using the above-defined equations, one can determine the concentration distribution, the electrical field strength for each section and the diffusing zone length. First, at a given time, a guessing value for the concentration is chosen, the equation of electrical potential Eq. (5) can be numerically solved to find the EDL potential y/(r). Once the EDL potential is known, the local net charge density pe(r) can be determined according to Eq. (3). With the guessing value for the concentration field again, the electrical resistance, the electrical current and the electrical field strength for each section can be determined by Eqs. (15) (18). The equation of motion, Eq. (10), can be numerically solved to find the velocity, uz(r). Using this velocity profile, the equation of concentration, Eq. (13), can be solved to obtain the concentration distribution in capillary, and hence the length of the diffusing zone can be determined based on the obtained concentration profile. Repeat this iteration procedure until the convergence of concentration is reached. Finally, with the results obtained in this time step, the above procedure is repeated for the next time step until the high concentration solution completely replaces low concentration solution in the capillary tube {i.e., the left boundary of the diffusing zone reaches the exit of the capillary tube). Figure 7.5 shows the experimentally measured current-time relationships. Because the electrical current depends on the applied electrical field and the electrical resistance of the solution, which in turn depends on the concentration and the conductivity, when the concentration in the capillary change, the current will change. Therefore, the electrical current in the capillary increases with time as the high concentration solution gradually replaces the low concentration solution. The increase in current continues until the capillary is completely filled with the high concentration solution. Thereafter the current reaches a steady and maximum value. Using the aforementioned numerical model, the complete set of non-linear differential equations, governing the diffusion and convection process of the two solutions in capillary, were solved to predict the diffusing process. As
Experimental Studies of Electroosmotic Flow
367
Figure 7.7. Concentration distribution in capillary tube at time t = 12 s, t = 24 s and t = 36s for 1.1 x 10"4 M KCl solution under an electrical field of 35 KV/m.
368
Electrokinetics in Microfluidics
seen from Figure 7.5, there is a good agreement between the experimental data and the model prediction. As explained before, the time for the current reaches a plateau value is the time required for the high concentration electrolyte solution to travel through the entire capillary tube. The average velocity of the liquid flow can then be calculated by using Eq. (1). From the data shown in Figure 7.5, the required time can be determined, therefore, the average velocity of the electro-osmotic flow can be obtained with the known capillary length. Figure 7.6 is the plot of the average velocity versus the ionic concentrations. As seen in Figure 7.6, an essentially linear relationship between the average velocity and the concentration is evident and a good agreement between the model predictions and experimental data was found. The model presented in this section also reveals the diffusing process. The concentration difference across the interface generates the diffusion. Initially, the length of the diffusing zone increases with time after the two solutions are brought into contact. However, the length of the diffusing zone will decrease when the low concentration solution completely leaves the capillary tube and the high concentration solution continues to move towards the end of capillary. This can be seen from Figure 7.7. The length of the diffusing zone at 24 seconds is bigger than that at 12 seconds, which shows the expansion of the diffusing zone. The length of the diffusing zone at 36 seconds is smaller than that at 24 seconds, showing the reduction of the diffusing zone at the end of the displacing process.
Experimental Studies of Electroosmotic Flow
7-2
369
MEASUREMENT OF AVERAGE ELECTROOSMOTIC VELOCITY BY A SLOPE METHOD
In the last section, we discussed an experimental approach to evaluate the electroosmotic flow velocity by monitoring the current change in a process where one solution replaces another similar solution in a microchannel. However, practically it is difficult to determine the exact time required for one solution completely replacing another solution. This is because the gradual changing of the current with time and the small current fluctuations exist at both the beginning and the end of the replacing process. Figure 7.8 shows one of the typical results of the measured current-time relationship. These make it very difficult to determine the exact beginning and ending time of the current change from the experimental results. Consequently, significant errors could be introduced in the average velocity determined in this way. It has been observed that, despite the curved beginning and ending sections, the measured current-time relationship is essentially linear in most part of the process, as long as the concentration difference is small. Therefore, a new method was developed to determine the average electroosmotic velocity by using the slope of the current-time relationship [18].
Figure 7.8. Typical results of the measured current-time relationship for lxlO" 4 M KC1 solution in a microchannel of 10cm in length and 200 u.m in diameter, under an applied electrical field of 30 kV/m.
370
Electrokinetics in Microfluidics
When a high concentration electrolyte solution gradually replaces another solution of the same electrolyte with a slightly lower concentration in a microchannel under an applied electric field, the current increases until the high concentration solution completely replaces the low concentration solution, at which time the current reaches a constant value. If the concentration difference is small, a linear relationship between the current and the time is observed. This may be understood in the following way. In such a process, because the concentration difference between these two solutions is small, e.g., 5 %, the zeta potential, which determines the net charge density and hence the liquid flow, can be considered as constant along the capillary. Consequently, the average velocity can be considered as a constant during the replacing process. As one solution flows into the capillary and the other flows out of the capillary at essentially the same speed, and the two solutions have different electrical conductivity, the overall electrical resistance of the liquid in the capillary will change linearly, and hence the slope of current-time relationship is constant during this process. The slope of the linear current-time relationship can be described as: (19) where Ltotal is the total length of the capillary and uave is the average electroosmotic velocity, E is the applied electrical field strength, A is the cross section area of the capillary, ( A ^ - ^ J I ) *s t n e bulk conductivity difference between the high concentration solution and the low concentration solution. In Eq. (19), all the parameters are known and constant during the replacing process. If we group all the known parameters and denote 77 =EA(Xj,2 - H\)l^totah Eq(19) can be rewritten as: (20)
Eq.(20) indicates that the average electro-osmotic flow velocity can be determined by measuring the slope of the current-time relationship. To measure the slope of the current-time relationship, experiments similar to what described in the last section were performed. In the experiment, Polyamide-coated silica capillary tubes of different internal diameters (Polyamide Technologies Inc., Phoenix, AZ) were cut to 10 cm in length and used to connect the electrolyte solutions in reservoir 1 and reservoir 2. In the experiment, reservoir 2 was filled with a higher concentration solution 2. The capillary tube and reservoir 1 were filled with a lower concentration solution 1. All the solutions in our experiments were prepared by using Deionized
Experimental Studies of Electroosmotic Flow
371
ultrafiltered water (Fisher Scientific, Ontario), KC1 (Anachemia Science, Quebec), and LaCl3 (Fisher Scientific Company, New Jersey). Immediately after connecting the two reservoirs by the capillary tube, a voltage difference between the two reservoirs was applied by setting reservoir 1 at a ground potential and reservoir 2 to a higher voltage (HV power source: CZE 1000 R, Spellman, NY) via platinum electrodes.
Figure 7.9. Examples of the measured current—time relationship, for 1x10 3 M KC1 solution (top) and lxlO" 3 M LaCl3 solution (bottom) in a 10cm long capillary of 100 urn in diameter under an electrical field of 10 kV/m.
372
Electrokinetics in Microfluidics
Figure 7.10. The average velocity versus the applied voltage and the capillary diameter, for 1x10 3 M KC1 solution (top) and lxlO"4 M KC1 solution (bottom) in a 10cm long capillary tube.
Experimental Studies of Electroosmotic Flow
373
The applied electrical field results in an electroosmotic flow in the capillary tube. The electrolyte solution from reservoir 2 gradually displaces the electrolyte solution in the capillary tube. As a result, the overall electrical resistance of the liquid in the capillary tube changes, A PGA-DAS 08 data acquisition chip (OMEGA Engineering, Quebec) was used to record the voltage (KV) and current (uA) as a function of time (second). Once the solution in the capillary tube is completely replaced by the solution from the reservoir 2, the current reaches a constant value. In the experimental studies, each measurement was repeated at least five times for a given set of conditions. All experiments were conducted at a room temperature (25°C). Because the electrical current depends on the applied electrical field strength and the conductivity of the solution, when the concentration in the capillary changes, the conductivity and hence the current will change. Therefore, the electrical current through the capillary increases with time as the high concentration solution gradually replaces the low concentration solution in the capillary. The increase in current continues until the capillary is completely filled with the high concentration solution. Thereafter the current reaches a constant and maximum value. Some typical results of the current change with time are shown in Figure 7.9. All the experimental data shown in the figure is an average value of 5 independent measurements with an error approximately 5%. The average electroosmotic velocity can be determined by using the Eq. (2) and the measured slope. Figure 7.10 shows that the average velocity versus the applied voltage and the capillary diameter for lxlO"3 M and lxlO"4 M KC1 solutions in a glass capillary of 10 cm in length. As seen from the figure, a linear relationship between the applied voltage and the average velocity is evident. In addition, it is clear that the average velocity is essentially independent of the size of the capillary. Figure 7.11 shows that the comparison of the average velocity for lxlO"3 M KC1 solution and for lxl0" 3 M LaCl3 solution between the theoretical model prediction and the experimental results determined by the current—time method and by the slope method. Here the theoretical model has been described in the previous section. It can be seen from this figure that there is a better agreement between the model predictions and the experimental results determined by the slope method. The experimental results determined by using the estimated time for the complete displacing process deviate from the above two results. This is because of the inaccuracy in determining the required time (particularly the ending time) for one solution replacing another solution. In the current—time method presented in the last section, the total time required for one solution completely replacing another solution has to be found, which often involves some inaccuracy of choosing the beginning and the ending points of the replacing process. In this slope method, instead of identifying the
374
Electrokinetics in Microfluidics
Figure 7.11. Comparison of the model predicted average velocity with the experimentally determined average velocity by the current—time method and by the slope method for (top) lxlO"3 M KC1 solution, and (bottom) lxlO' 3 M LaCl3 solution.
Experimental Studies of Electroosmotic Flow
375
total period of time, only a middle section of the current-time data is required to determine the slope of the current-time relationship. This method is easier and more accurate than finding the exact beginning and the ending points of the replacing process. In addition, for some cases of high concentration solutions under low electrical field strength, the time required for the high concentration solution completely replacing low concentration solution is very long (e.g. 700 seconds). Therefore, the joule heating effects become significant in these cases. However, Using this new slope method, the joule-heating problem can be avoided since it is not necessary to complete the displacing process. Usually, a sufficient amount of data to determine the linear relationship can be obtained in less than 100 seconds, the average velocity can thus be determined and the experiments can be stopped before appreciable heating effect takes place.
376
7-3
Electrokinetics in Microfluidics
MICROFLUIDIC VISUALIZATION BY A LASER-INDUCED DYE METHOD
Although the current-monitoring methods discussed in the previous sections provide simple means to evaluate the average electroosmotic flow velocity in microchannels, they are indirect methods and cannot provide direct insight of microflows. This section will introduce a direct microfluidic visualization technique to measure the microflow velocity profiles. A major macroflow visualization technique is the particle image velocimetry (PIV) [1-5]. Application of PIV to microflows typically involves use of fluorescent microspheres. Taylor and Yeung [1] calculated fluid velocity from the streaks produced from fluorescent particles in a liquid flow in a capillary. PIV techniques involving double-frame cross-correlation algorithms have been successfully applied to microflows [2,3]. The hydrophobicity of latex particles, a common choice of marker in PIV, can cause particle aggregation and adhesion to the channel surface. The use of hydrophilic markers, such as the unilamellar liposomes employed by Singh et al. [4], can mitigate these effects. Extracting an unaltered local fluid velocity from an observed marker velocity, however, remains a fundamental challenge in PIV and flow visualization techniques in general. This is further complicated on the microscale by the addition of electrokinetic effects including electrophoresis, electrophoretic relaxation, and particle-wall interaction [20]. Dye-based microflow visualization techniques have also evolved from their macro-sized counterparts. A major difficulty with the dye techniques is how to directly inject the dye into the microscale flows without significantly altering the flow pattern. The conventional mechanical injection techniques are difficult to apply on the microscale without interfering with the flow field. One option is to replace a solution with a dye solution and track the concentration front [1]. Such a method is complicated by the relatively fast diffusion at the dye front, and the changes in the EDL due to differences between the dye solution and the original solution. In addition, accurate tracking usually requires high concentration gradients, which in turn generate high rates of diffusion relative to the mean flow velocity. Tracking points of zero concentration gradient (maxima) would minimize diffusion effects, however, these points are not clearly definable when one solution is gradually replaced with another. Specialized caged fluorescent dyes can serve to eliminate the problems associated with the mechanical dye injection. A photo-injection is accomplished by exposing an initially non-fluorescent solution seeded with caged fluorescent dye to a beam or a sheet of ultraviolet light. As a result of ultraviolet exposure, caging groups are broken and fluorescent dye molecules are released [21]. The free fluorescent dye may then be imaged using classical fluorescence techniques. Lempert et al. [6] applied caged dyes to flow visualization in a lcm internal
Experimental Studies of Electroosmotic Flow
377
diameter pipe as well as a large water channel facility. Paul et al. [7] applied the method to microflow visualization, they presented images of dye transport in pressure-driven flow and electroosmotic flow. Velocity profiles were calculated for the pressure-driven case using image pairs and a scalar image velocimetry (SIV) technique. That study outlined several key issues associated with cageddye techniques on the microscale, namely: decreasing signal to noise ratio near the wall; and predicting electrophoretic motion of the charged dye molecules (especially problematic in large molecular weight, multi-component dyes). Dyebased velocimetry requires knowledge of the electrophoretic mobility of the dye molecules when the method is applied to electrokinetic flows (analogous to PIV). Herr et al. [8] applied the method of Paul et al. [7] to study capillary flow with nonuniform zeta potential. Herr et al.[8] tracked the motion of the dye by fitting Gaussian curves to the axial concentration profiles. Dahm et al. [9] presented a SIV technique for four-dimensional, turbulent velocity field measurements. Recently, Sinton and Li [13] developed a microchannel flow visualization system and complimentary analysis technique by using caged fluorescent dyes. Both pressure-driven and electrokinetically driven velocity profiles determined by this technique compare well with analytical results and those of previous experimental studies. Particularly, this method achieved a high degree of near-wall resolution. This technique will be discussed below.
Figure 7.12. An imaging apparatus of the laser-caged dye microflow visualization method.
378
Electrokinetics in Microfluidics
Generally, in the experiment, a caged fluorescent dye is dissolved in an aqueous solution in a capillary or microchannel. The caged dye molecule cannot emit fluorescent light because it is encapsulated by a polymer group. Ultraviolet laser light is focused into a sheet crossing the capillary (perpendicular to the flow direction). The caged fluorescent dye molecules exposed to the UV light are uncaged and thus are able to shine. The resulting fluorescent dye is continuously excited by an argon laser and the emission is transmitted through a laser-powered epi-illumination microscope. Full frame images of the dye transport are recorded by a progressive scan CCD camera and saved automatically on the computer. In the numerical analysis, the images are processed and cross-stream velocity profiles are calculated based on tracking the dye concentration maxima through a sequence of several consecutive images. Several sequential images are used to improve the signal to noise ratio. Points of concentration maxima make convenient velocimetry markers as they are resistant to diffusion. In many ways, the presence of clearly definable, zeroconcentration-gradient markers is a luxury afforded by the photo-injection process. Imaging apparatus A labeled photograph of the system is provided in Figure 7.12. The two required laser beams were fixed and ran horizontal and parallel: the blue, excitation beam above; and the ultraviolet beam below. The ultraviolet light was provided by a 300uJ, A = 337nm, pulsed nitrogen laser (LSI, VSL -337NDS). This beam was reflected upwards by a stationary mirror and through a vertical optical rail assembly containing two consecutive, counter-oriented, cylindrical optics followed by a blank optic on which the capillary rested (all of which were fused silica). The first cylindrical optic (f= 100mm piano convex) focused the beam into a sheet. The second cylindrical optic (f= 50mm piano convex) condensed this sheet into the capillary in the cross-stream direction. All three optics were adjusted relative to each other on the vertical optical rail. The rail itself, including the flow module was mounted on a precision three-axis stage. A single-line, 200mW, A = 488nm argon laser (American Laser Corp., LS300B) provided a continuous flood of excitation light through the objective. The laser beam passed through a lOx beam expander before entering the epifluorescent microscope head (Leica DMLM). The original, lamp-based optics in the microscope was removed to facilitate collimated light, and no excitation filter was required as the light was monochromatic. The beam is reflected off a dichroic mirror (A = 510nm long pass) directly into the oil immersion, 25x objective with numerical aperture, NA = 0.75 (Leica PL Fluotar). High numerical aperture objectives are especially desirable in weak signal, epi-
Experimental Studies of Electroosmotic Flow
379
illumination applications. The received signal intensities vary with numerical aperture (A^) and magnification (mag) as follows:
(2D Other advantages of high NA include reduced depth of field and increased resolution, often at the expense of working distance [22]. The Abbe spatial resolution of this apparatus was approximately 1.2|am in the object plane, and the depth of field was calculated as 2.1 urn. A single drop of optical oil surrounded the windowed portion of the capillary and connected directly to both the objective and the blank base optic. This full immersion design provided index matching for imaging and both the excitation and ultraviolet light. Cargille (FF) optical oil was used due to its ultra-low fluorescence and an index of refraction, « re /= 1.485, similar to that of the fused silica, nre/ = 1.46. The received fluorescent signal passed through the dichroic mirror and through an additional suppression filter (A = 515nm long pass) to the 2/3" progressive scanning interline transfer CCD camera (Pulnix, TM-9701). Progressive scan architecture enabled all pixels in the CCD array to be exposed concurrently with all data being subsequently transferred (in contrast to even and odd field interlacing). This is desirable from an image quality perspective, however, when running in this mode the signal could only be viewed through the computer. The video monitor required signals in interlaced mode and could only be used for focusing and alignment. Images received by the camera were digitized and saved automatically through a PC-based image acquisition system. Full frame images could be collected at rates up to 30Hz. Image acquisition A four-channel delay generator (SRS, DG535) controlled the firing of the nitrogen laser, run frequency, and synchronization with the camera when necessary. The system could be run in two modes: 1.) The camera was asynchronously reset by the delay generator a set time after the laser had been fired; and 2.) The camera was run in video mode and the delay generator fired the laser at convenient intervals. To increase the amount of uncaged dye, the nitrogen laser could be pulsed multiple times (to a maximum rate of 60Hz) at each uncaging event. The number of pulses per uncaging event is restricted only by fluid velocity and the desired sheet thickness. The excitation laser ran unshuttered, continuously flood-illuminating the capillary. Image exposure durations were controlled using electronic shuttering modes in the camera. The acquired images had a resolution of 640 x 484 pixels. This corresponded to a viewing area of 346 x 261 um with each pixel representing a 0.54jj.m square. A
380
Electrokinetics in Microfluidics
lx c-mount was employed such that this magnification is only a consequence of the objective magnification and the size of the CCD array. Image Processing To remove any non-uniformities present in the imaging system, dark field image subtraction and bright field image normalization were performed [22]. A dark field image (or background image) for each run was taken immediately before the laser was fired. This dark field image contained a faint signal from stray background light as well as fluorescent emission from trace amounts of uncaged dye in the bulk fluid. Before a bright field image was taken, the flow was stopped and the nitrogen laser was fired repeatedly until uncaged dye had diffused well beyond the camera's field of view. Using a Matlab program, the dark field image was subtracted from all images including the bright field image. These images were then normalized with respect to this bright field image and smoothed using a distance-based 7x7 pixel kernel. This kernel size corresponds to a 3.2um square in the object plane, such that pixel intensity values could influence neighbors up to 1.62um away in each radial and axial direction. Finally, the resulting image series was linearly scaled by a single factor so as to fill the grayscale range. Flow systems A custom designed, ultra-low flow rate syringe pump provided pressuredriven flow. A DC micromotor was required to ensure a continuous flow rate, not achievable with commercial designs based on stepper motors. For electroosmotic flow, a separate, two-reservoir flow module was used. A 14cm long capillary liquid junction joined the small reservoirs with embedded platinum electrodes. The capillary and both reservoirs were filled with the same caged dye solution before any voltage was applied. All capillaries were fused silica with exterior polymide coating as supplied by Polymicro Tech. The nominal inner diameters of the capillaries were lOOum and 205um. However, using an electron microscope the inner diameters were found to be 102 urn and 205 um. Each capillary was prepared by flushing with pure water, followed by filtered buffer. The exterior polyimide coating was oxidized and removed to create a viewing window. Since only minute amounts were involved in uncaging events, most dye was used in filling the flow system. The cost of specialized caged dyes was such that careful design of these systems can reduce the operating costs significantly. Chemicals Two caged fluorescent dyes, supplied by Molecular Probes, were employed here: 5-carboxymethoxy-2-nitrobenzyl (CMNB)-caged fluorescein (826.81 MW); and 4,5-dimethoxy-2-nitrobenzyl (DMNB)-caged fluorescein
Experimental Studies of Electroosmotic Flow
381
dextran (10000 MW). Both dyes were dissolved in sodium carbonate buffer of pH = 9.0. The buffer was prepared by dissolving 39.8xlO~3 mol of NaHCC>3 and 3.41xlO~3 mol of Na2CO3 in 1 litre of pure water resulting in a solution of ionic strength, / = 0.05. Stock solutions of DMNB-caged fluorescein dextran and CMNB-caged fluorescein were prepared in 0.2 mM and lmM concentrations respectively. The solutions were aliquoted and stored in darkness at -20°C. Immediately before use, all solutions were filtered using 0.2um pore size syringe filters. Both caged dyes release fluorescein upon photolysis which has an absorption maximum at approximately X = 490nm and is well suited to excitation with the X = 488nm output line of the argon laser. DMNB caging groups absorb light most efficiently from X = 340nm to X = 360nm. CMNB caging groups, however, exhibit an absorption maximum at X = 334nm which is very well suited to the X = 337nm output of a nitrogen laser. To facilitate the multiple-image analysis technique, all results presented here were acquired using mode 2 in which the camera ran in progressive scan video mode and the delay generator controlled the ultraviolet laser. The nitrogen laser was triple-pulsed at 30Hz at each uncaging event. The overall run frequency was set to 0.15 Hz, providing more than sufficient time for uncaged dye to exit the field of view. The camera was run at 15 Hz resulting in 100 stored images per uncaging event. Individual exposure times were 1/125 sec (corresponding to shutter mode 8 on the TM-9701). The timescale over which the uncaging process occurs is also an important consideration. In general, this delay can vary from microseconds to seconds [21]. Here, dye release was effectively complete 50msec after the uncaging event, and only images taken after this period were used in velocimetry calculations. This delay corresponds to the first major timescale of dye release found by Lempert et al. [6]. Numerical Analysis Technique Although the mass-weighted average velocity, u, is of interest, only the transport of the uncaged dye ions is observed. The numerical analysis technique must infer the mass-weighted average velocity, u, from images of ion transport which can differ from the advective transport due to forced and ordinary —* diffusion. The molar flux of uncaged dye ions, _/,• , is governed by the NernstPlanck equation in the form (22)
382
Electrokinetics in Microfluidics
where ¥ is the electrostatic potential. If the flow is fully developed, and the axial gradient of the electrostatic potential is a known constant, Ex, the axial electrophoretic velocity, uPh, becomes a known constant and the ionic flux takes the form (23)
(with respect to the special case of pressure-driven flow in the presence of a thin EDL, the streaming potential becomes insignificant and the electrophoretic velocity, uPh, may be neglected.) In locations where the concentration gradient is zero, the observed ionic transport is only the addition of a known electrophoretic velocity and the desired, mass-weighted averaged velocity. In this method, each of these locations of zero axial dye concentration gradient were chosen as marker points, such that the cross-sectional velocity profile could be calculated directly from the observed velocity of these points as follows: (24)
Before each run, the camera was rotated such that the pixel grid was aligned with the radial and axial directions. After each run, the received images were processed and read into the analysis program in matrix format. Along each axial line of pixels, a zero gradient marker point was located using a weighted average of the highest intensity values on that line. The set of markers from each image formed a cross-stream, maximum concentration profile. Any pair of these profiles could provide a velocity profile by dividing the distance between them, Ax, by the corresponding time step, At. With 8 images, 28 velocity profiles could be constructed. In general, for a set of N images a set of n velocity profiles could be constructed where (25)
Here, all n displacements were calculated and an error-weighted average technique was applied to calculate the velocity profile. The basis of the weighting was that displacement measurements, Axh would all have similar error due to finite, random, signal noise. The size of the corresponding timestep, however, would scale the effect of that error on the velocity value. Weighting each measurement such that each contributed equally to the overall error resulted in the straightforward calculation below
Experimental Studies of Electroosmotic Flow
383
(26)
Pressure-Driven Flow In the absence of an applied electric field and any electro-viscous effects, pressure-driven flow in a circular cross-section capillary results in a parabolic profile of the form
(27)
where uP is the velocity at the centerline. Images of a pressure-driven flow in a 205 um i.d. capillary are shown in Figure 7.13. The DMNB-caged fluorescein dextran in the sodium carbonate buffer was used. Although it is difficult to calculate the exact concentration of dissociated sodium ions, the ionic strength of / = 0.05 would indicate a bulk counter-ion concentration on the order of 10~2 M at minimum. This is a relatively high ion concentration with respect to electrokinetic effects. An estimate of the Debye length is about 3nm. A conservative estimate of EDL thickness is four Debye lengths. This corresponds to a thickness on the order of lOnm, which is significantly smaller than the maximum achievable resolution of a light microscope. Such an EDL thickness is sufficiently thin to ensure negligible electro-viscous effects, so that the velocity profile can be expected to remain parabolic right up to the wall. Figure 7.14(a) shows plots of the relative signal intensity (along the centerline of the capillary) vs. axial distance for 8 consecutive images. The points of zero axial concentration gradient correspond to peaks in signal intensity. Profiles of
Figure 7.13. Images of the uncaged dye in a pressure-driven flow through 205u,m i.d. capillary at 133msec intervals.
384
Electrokinetics in Microfluidics
Figure 7.14. Data from 8 consecutive images (taken at 15Hz.) of uncaged dye in a pressuredriven flow through a 205um i.d. capillary: (a) Plots of relative signal intensity (along the centerline) vs. axial distance for all images; (b) Plots of concentration maxima for all images; and (c) Calculated velocity (points) plotted with the parabolic analytical solution.
Experimental Studies of Electroosmotic Flow
385
concentration maxima calculated along each row of pixels in each image are shown in Figure 7.14(b). Applying the numerical analysis technique to these profiles resulted in the velocity data (points) plotted with the parabolic analytical solution (line) in Figure 7.14(c). The parabola was calculated by using conventional laminar flow equation assuming a zero-slip velocity at the wall and a centerline velocity equal to that determined experimentally. The velocity profile is in good agreement with the parabolic analytical solution up to 8 microns from each wall (r/R = 0.92). Paul et al. [7] also showed a strong parabolic trend but only data for -0.5 < r/R < 1 is displayed due to an insufficient signal-to-noise ratio near the one wall. Near the other wall the reported profiles show a large scatter beyond r/R = 0.80. The symmetry of the full cross-section velocity profile given here in Figure 7.14(c) is also noteworthy and demonstrates the effectiveness of this method. Electro-osmotic Flow The two reservoirs in the flow module were carefully balanced in order to avoid pressure-driven flow. In addition, both reservoirs contained the dye solution so that the system could be left for several minutes and allowed to come into equilibrium. In early tests, it was found that the fluorescein dextran dye performed poorly in electroosmotic flows. Paul et al. [7] suggested that the multi-component nature of this dye make it unsuitable for these applications. Here, the CMNB-caged fluorescein with the sodium carbonate buffer and 102(im i.d. capillaries were used in all electroosmotic flows. Images of the uncaged dye transport in four different electroosmotic flows are displayed in vertical sequence in Figure 7.15. The dye diffused symmetrically as shown in Figure 7.15(a). Image sequences given in Figures. 7.15(b), (c), and (d) were taken with voltages of 1000V, 1500V, and 2000V respectively (over the 14cm
Figure 7.15. Images of the uncaged dye in electroosmotic flows through a 102u,m i.d. capillary at 133 msec intervals with applied electric field strength: (a) 0V/14cm; (b) 1000V/14cm; (c) 1500/14cm; and (d) 2000V/14cm.
386
Electrokinetics in Microfluidics
length of capillary). The field was applied with positive at left and negative at right. The resulting plug-like motion of the dye is characteristic of electroosmotic flow in the presence of a negatively charged surface at high ionic concentration. The cup-shape of the dye profile was observed in cases 7.15(b), (c), and (d) within the first 50msec following the ultraviolet light exposure. This period corresponded to the uncaging time scale in which the most significant rise in uncaged dye concentration occurs. Although the exact reason for the formation of this shape is unknown, it is likely that it was an artifact of the uncaging process in the presence of the electric field. Fortunately, however, the method is relatively insensitive to the shape of the dye concentration profile. Once formed, it is the transport of the maximum concentration profile that provides the velocity data. This also makes the method relatively insensitive to beam geometry and power intensity distribution. As in the pressure-driven flow case, the major counter-ions to the negatively charged wall were those offered by the buffer solution. The resulting EDL thickness on the order lOnm ensured that a flat electroosmotic velocity profile could be expected to extend right to the wall. Admittedly, there would be a steep velocity gradient in the EDL region, however, the thickness of the EDL prohibits resolution of this gradient with this or any other known velocimetry technique. This realization provided an opportunity to test the nearwall resolution of the present technique as follows: The point near the wall where the velocity profile falls from the electroosmotic velocity could be considered the closest significant datum. A diffusion coefficient of D = 4.37xlO~6 cm2/s was assigned, giving an electrophoretic mobility of vPf, = -3.3x10 4 c m V ' s " ' as suggested by Harrison et al. [23] and Paul et al. [7]. Due to the negative charge of the uncaged dye ions, their net velocity was necessarily less than the underlying electroosmotic velocity. Figure 7.16 shows velocity data for the four flows corresponding to the image sequences in Figure 7.15. Each velocity profile was calculated using an 8-image sequence and the numerical analysis technique. The velocity profile resulting from no applied field, appearing on the far left in Figure 7.15, corresponds closely to stagnation as expected. This run also serves to illustrate that, despite significant transport of dye due to diffusion, the analysis method is able to recover the underlying stagnant flow velocity. Although the other velocity profiles resemble that of classical electroosmotic flow [5], a slight parabolic velocity deficit of approximately 4% was detected in all three flows. This was caused by a small back-pressure induced by the electroosmotic fluid motion. In general, several factors can cause concave velocity profiles including: a narrowing or partial blockage anywhere in the capillary; the presence of charged particles in the fluid; capillary forces resulting from different curvatures of the
Experimental Studies of Electroosmotic Flow
387
free surfaces in the reservoirs; contact angle hysteresis of a bubble entrapped in the channel; a non-uniform wall surface charge distribution; and slight height differences between the free surfaces in the reservoirs. Sensitivity to these factors is highly correlated with the capillary radius. In this case, it was evaluated that an adverse pressure gradient of only 50 N/m3 would have been sufficient to generate the observed back-flow. Such a pressure gradient could be easily generated if the height of one reservoir exceeded the other by less than 1 mm. Velocity profiles plotted in Figure 7.16 indicate electroosmotic wall velocities of 3.8xlO"4 m/s, 6.0xl(T4 m/s, and 8.2xlO~4 m/s (from left to right) giving electroosmotic mobility values of vE = 5.32xlO~4 c m V s " 1 , vE = 5.60x10 4 cm2V"1s"',vE= 5.74xlO"4 c m ^ v V 1 respectively. These values are in good agreement with vE = 5.07xl0"4 cmV^s" 1 in pH 8.0 reported by Paul et al. [7], and vE = 5.87xlO"4 c m V s l in pH 8.5 reported by Harrison et al. [23] (both in fused silica). From the average mobility calculated here, an effective zeta potential for this system is calculated to be C= -80 mV. This is in keeping with silica oxide surfaces at KNO3 concentrations near 10"2 M at pH 9.0 as reported by Hunter [20]. It should be noted that the resolution of the determined velocity profiles in the electroosmotic flows is better than that in the pressuredriven flows. This is because the two flows differ with respect to cross-stream velocity gradients, which induce radial concentration gradients and radial diffusion. In pressure-driven flow, the dye sample is transported with the parabolic velocity field, which leads to an increased radial diffusion (in addition
Figure 7.16. Plots of velocity data from four electroosmotic flow experiments through a 102u.ni i.d. capillary with applied electric field strengths of: 0V/14cm; 1000V/14cm; 1500/14cm; and 2000V/14cm (from left to right).
388
Electrokinetics in Microfluidics
Figure 7.17. Velocity data of three electroosmotic flow experiments through a 102^m i.d. capillary with an applied electric field strength of 1000V/14cm: (a) Plotted over capillary cross-section; (b) Plotted in lO^m leading up to the wall (-0.8 > r/R > -1).
Experimental Studies of Electroosmotic Flow
389
to axial diffusion) and a loss of resolution. In electroosmotic flow however, the sample is mostly translated with the plug-like flow filed and only axial diffusion acts to disperse the sample. As shown in these figures, all electroosmotic flow profiles exhibit a strong degree of symmetry. The velocity profile remains relatively flat for 0.95< r/R <0.95 in all cases. Due to the known, sub-micron thickness of the EDL, this range can be interpreted as confident velocimetry measurements up to r/R = 0.95, or 2.5 um from the capillary wall. These results also provide experimental verification of the electroosmotic flow profile predicted in classical theory [20]. To illustrate repeatability, velocity data from three electroosmotic flows at an applied electrical field strength of 1000V/0.14m are superimposed in Figure 7.17(a). Data near r/R = - 1 is plotted in an expanded view in Figure 7.17(b). The grouping of the velocity data from the three separate runs as well as the 2.5 urn near-wall resolution is readily apparent. In future, retrieving fluorescent signal closer to the wall may be possible using channels of non-circular cross-section. Due to the angle between the inner sidewall and the optical axis, a closer index match between the solution and channel may also improve data acquisition near the wall (currently the capillary and the optical oil are well matched at « re /= 1.46 and wre/ = 1.485 respectively, but the solution remains approximately that of water at nref ~ 1.33). As mentioned previously, a caged dye releasing a neutral marker would also be beneficial to the method. In lieu of that, however, caged dyes releasing single component fluorescent molecules with predictable electrophoretic mobility may be used. In summary, the laser-caged-dye based microflow visualization technique with near-wall resolution described here can be employed to directly measure the local velocity in microchannels. An improved near-wall resolution is the result of the optical system design and the multi-image analysis technique. The described optical system was designed to maximize signal reception and minimize distortion using high numerical aperture optics and complete oil immersion. The described analysis technique utilizes a multi-image sequence and calculates the cross-stream velocity profile based on an error-weighted average of the displacements of the concentration maxima profiles. When applied to pressure-driven flow through a 205 um i.d. capillary, this method provided the measured velocity profiles in good agreement with the parabolic analytical solution up to 8um from each wall. For electroosmotic flow though a 102um i.d. capillary, the ability of the technique to determine fluid velocities up to 2.5 urn from the wall was demonstrated. This degree of accuracy was made possible by the combination of the optical system design and the image analysis technique described in this section.
390
7-4
Electrokinetics in Microfluidics
VELOCITY PROFILES OF ELECTROOSMOTIC FLOW IN MICROCHANNELS
In a microchannel, electroosmotic flow results from the motion of mobile counter-ions in the diffuse portion of the electrical double layer in response to an axially applied electric field. When the channel is much wider than the electrical double layer, this results in a plug-like velocity profile in the bulk liquid region. However, this is the ideal case. In practice, several factors can induce velocity gradients in the bulk liquid and hence distort the plug-like velocity profile. These factors include: a narrowing or partial blockage anywhere in the microchannel; the presence of charged particles in the fluid; capillary forces resulting from different curvatures of the free surfaces in the reservoirs connecting to the microchannel; and slight height differences between the free surfaces in the reservoirs; a non-uniform wall surface charge distribution; and Joule heating effects. Sensitivity to these factors is highly correlated with capillary radius or the microchannel size. This section will presents some recent experimental work of electroosmotic velocity profiles in both square and circular cross-section channels (in widths ranging from 20 to 200um) measured by the laser-caged dye method described in the previous section [14]. The disturbance to the measured velocity profiles by the pressure effect and by the Joule heating will be discussed. As discussed in the last section, determining the local flow velocity by analyzing a sequence of images of dye transport falls under the broad category of scalar image velocimetry [9]. Analysis is greatly simplified by the presence of a maximum concentration line, when a discrete dye sample is used as a flow marker. The motion of this locus of concentration maxima is, for the most part, not affected by diffusion, and reflects the bulk motion of the fluid. If the dye chosen is non-neutral, however, the observed velocity, VQ^, of the concentration maxima will be the summation of the desired bulk fluid velocity, V^Q , and the electrophoretic velocity vpjj of the dye as follows: (28)
In the case of the electroosmotic flow of a relatively high ionic strength solution in a straight channel, the electrophoretic velocity, vPH, may be calculated directly from the electrophoretic mobility and the axial applied field strength. The bulk electroosmotic velocity profile can then be calculated from Eq. (28). In many cases, a constant mobility value determined experimentally or specified through the diffusion coefficient has proven sufficient. However, in cases where Joule heating leads to significant fluid temperature increases, mobility values can change dramatically from their reference room-temperature
Experimental Studies of Electroosmotic Flow
391
values. To account for this, Ross et al. [24] estimated the temperature in the channel by comparing the measured conductivity to a reference value. In this section, a different method will be used to consider the compensation for the dye mobility temperature dependence without explicit knowledge of the fluid temperature. A schematic diagram showing the flow module and the three lightinteractions (the ultraviolet uncaging light pulse, the continuous blue excitation light, and the resulting green fluorescent emission) is given in Figure 7.18. The ultraviolet light was provided by a 300uJ, 337nm, pulsed nitrogen laser (Laser Science Inc.). This beam was reflected upwards through a vertical optical rail assembly containing two, counter-oriented cylindrical optics. The continuous flood of excitation light was provided by a single-line, 200mW, 488nm argon laser (American Laser Corp.), through the 25x, NA = 0.75, oil immersion microscope objective. The index matching illustrated in Figure 7.19 (bottom right) is a key feature of this setup. This design facilitates a degree of near-wall resolution not achieved in other dye-based microflow visualization studies.
Figure 7.18.
Schematic illustration of experimental setup.
392
Electrokinetics in Microfluidics
A four-channel delay generator controlled the firing of the nitrogen laser and run frequency. The camera was run in video mode and the delay generator fired the laser at convenient intervals. To increase the amount of uncaged dye, the nitrogen laser was pulsed multiple times (at 30Hz) at each uncaging event. The excitation laser ran unshuttered, continuously flood-illuminating the capillary. The camera was run at 15 Hz with individual exposure times of 1/125 s. The acquired images had a resolution of 640 x 484 pixels. This corresponded to a 543|am visible length of capillary, with each pixel representing a 0.85um square in the object plane. The camera orientation was carefully adjusted such that the pixel grid was aligned with the radial and axial directions. The image processing was discussed in the previous section. To remove any non-uniformity present in the imaging system, dark field image subtraction and bright field image normalization were performed with each image. These images were then smoothed using a distance-based 7x7 pixel kernel. From these processed images, dye concentrations were calculated directly from pixel intensity values based on a previously determined camera response characteristic. To determine the observed dye velocity profile, a zero gradient marker point was located along each axial line of pixels, using a weighted average of the highest concentration values on that line. The set of markers from
Figure 7.19. Schematic diagram of the microchannel geometry and corresponding images of the channels with the photo-injected dye plugs used to visualize the flow.
Experimental Studies of Electroosmotic Flow
393
Figure 7.20. Cross-steam velocity profiles of electroosmotic flows with applied field strengths of 5, 10, 15, and 20 kV/m in: a.) lOOum diameter circular cross-section channel; and b.) 100|j.m width square cross-section channel.
394
Electrokinetics in Microfluidics
each image formed a cross-stream, concentration maxima profile. In a given sequence, any pair of the concentration maximum profiles could provide a velocity profile by dividing the distance between the profiles by the corresponding time-step. In cases where an axially applied electrical field is present, the electrophoretic motion of the dye must also be considered. This may be done by applying Eq. (28), or, in cases where Joule heating leads to significant temperature increases, a method presented below has to be used. To facilitate electroosmotic flow, a 15cm length of capillary joined two small reservoirs with embedded platinum electrodes. The capillary and the upstream reservoir were filled with the caged dye solution, and the downstream reservoir was filled with buffer. All capillaries were fused silica. Schematic diagrams of each capillary and corresponding images of photo-injected fluorescent dye plugs are given in Figure 7.19. New capillaries (as supplied by Polymicro Tech.) were prepared by flushing with pure water followed by flushing with buffer. At the capillary midpoint, the exterior polyimide coating was oxidized and removed to create a viewing window. The caged fluorescent dye employed in this study was 5-carboxymethoxy2-nitrobenzyl (CMNB)-caged fluorescein (826.81 MW) as supplied by Molecular Probes. The dye was dissolved to a concentration of 1.5 mM, in TAE buffer at pH = 9.0. A buffer concentration of 0.5x (~50mM) was used in all cases except in the experiments used to verify the temperature compensation method. In those cases, lx concentration was used to increase the conductivity of the solution. Immediately before use, all solutions were filtered using 0.2um pore size syringe filters. Cross-stream velocity profiles of electroosmotic flows in both a 100 um i.d. circular cross-section channel, and a lOO^im side length square cross-section channel are given in Figures 7.20 (a) and (b) respectively. EDL theory predicts that the diffuse double layer in this case extends less than 0.01 um from the wall. The remaining bulk fluid experiences no applied forces and exhibits no velocity gradients (i.e. a plug-like velocity profile). Also, since the channel width is much larger that the electrical double layer, the velocity reached is independent of channel cross-sectional shape. The velocity profiles in Figure 7.20, obtained at 5 kV/m, 10 kV/m, 15 kV/m and 20 kV/m, are in good agreement with these theoretical predictions. Although individual profiles have some variability, the profiles, in general, remain flat across the bulk fluid region. Also, the electroosmotic velocities correspond closely in both the square and circular cross-section channels. In the worst case, the square-channel profile exhibited a velocity 5% lower than the corresponding circular-channel profile. This level of error may be attributed to differences in surface charge due to manufacturing, or slight differences in surface history.
Experimental Studies of Electroosmotic Flow
395
The measured electroosmotic velocities in all channels are plotted versus applied electrical field strength in Figure 7.21. The data show a linear trend extending through the origin. This indicates constant electroosmotic and electrophoretic mobility and negligible temperature increases due to Joule heating. The average electroosmotic mobility is V^Q = 5.7X1 0~8 n ^ V ^ V , which is in close agreement with VEQ= 5.87X10~ 8 m V ' V 1 reported by Harrison et al. [23] under similar circumstances. Velocity profiles of electroosmotic flow in 20^m and 200um i.d. circular cross-section capillaries are given in Figure 7.22 (a) and (b) respectively. Although the average electroosmotic velocities in the 20 um channel are consistent to those of Figure 7.20, the profiles show more variability. This is a result of the small channel width and the high degree of curvature at the solutioncapillary interface. In contrast, the resolution in the 200um channel is quite high as evidenced by the large number of data points spanning the channel. In both the 5 kV/m and 10 kV/m cases, however, a parabolic backward flow component is readily apparent.
Figure 7.21. Electroosmotic velocities measured in 100|im width square channel, 50(j.m width square channel, 20u.m diameter circular channel, 100u.m diameter circular channel, and 200u.m diameter circular cross-section channels versus applied field strength.
396
Electrokinetics in Microfluidics
The velocity deficit at the centerline is just under 10% of the electroosmotic velocity in both profiles. As expected, the electroosmotic velocities (measured near the wall) are in good agreement with other channels as shown in Figure 7.21. However, the bulk velocity profile is not flat as expected for ideal electroosmotic flow. Two potential causes for the curvature, the pressure difference and the Joule heating, which occurs in larger channels and correlates with applied field strength, are discussed below. The sensitivity of microfluidic systems to pressure effects can be determined from classical theory of hydrodynamics [25], and capillarity [26]. Hydrostatic pressure resulting from height differences between the reservoirs is one potential source of pressure-driven flow. It is informative to calculate the
Figure 7.22. Cross-steam velocity profiles of electroosmotic flows in: (a) 20nm diameter circular cross-section channel with applied field strengths of 10, 15 and 20 kV/m; and (b) 200(xm diameter circular cross-section channel with applied field strengths of 5, and 10 kV/m.
Experimental Studies of Electroosmotic Flow
397
effective slope of the capillary system required to generate a significant pressuredriven flow component as follows, (29)
where, Ah is the height difference between reservoirs, L is the length of the channel, VPD is the velocity developed at the centerline, and g is the acceleration due to gravity. The slope required to generate a velocity disturbance of VPD = 100 um/s is plotted versus channel diameter in Figure 7.23. As illustrated in the figure, a 20um i.d. channel requires a tremendous height difference, where as a 200(j.m channel requires a height difference of only 700um over a 15cm channel length to generate this level of disturbance. This result was verified experimentally and the results are presented in Figure 7.24. With the reservoirs balanced to the best of current abilities, the electroosmotic velocity profile in Figure 7.24(a) was obtained. Elevating the downstream reservoir by less than lmm, a pressure-driven back-flow was generated (with a velocity of-lOOum/s at the centerline). The electric field was turned off, and this back-flow was imaged,
Figure 7.23. The effective slope (h/L) required to induce a pressure driven flow disturbance of strength 100 u.m/s at the centerline of the channel, versus channel diameter. Channel and reservoir schematics illustrate the slopes required for the 20(im, 100(xm, and 200um channel diameters.
398
Electrokinetics in Microfluidics
resulting in the velocity profile in Figure 7.24(b). Reapplying the electric field lead to the flow profile given in Figure 7.24(c), which is a linear combination of the profiles given in Figures 7.24(a) and (b). By integrating this velocity profile, it was found that these disturbances lead to a 17% reduction in flow rate from the ideal electroosmotic case. A second source of pressure disturbances is Laplace pressure acting on differentially curved free surfaces in the reservoirs. This can be especially critical in microfluidic chips with short channels (
Figure 7.24. Velocity profiles in a 200um circular cross-section channel of: (a) an electroosmotic flow with a slightly concave profile; (b) a pressure driven back-flow induced by raising one reservoir less than 1 mm; (c) a combination of the electroosmotic and pressure driven back-flow resulting in a 17% reduction in flow rate from the ideal electroosmotic case.
Experimental Studies of Electroosmotic Flow
399
Like pressure disturbances, Joule heating is strongly correlated with channel size as well as applied field strength. Joule heating (or resistive heating) results in a uniform generation of heat throughout the liquid in the channel. The rate of heat generation is equivalent to the total power dissipated in the channel, which can be expressed (for a capillary of circular cross-section) as follows [29], (30)
where, q is the heat release per unit length of channel, r is the channel radius, E is the applied electrical field strength, X is the molar conductivity of the solution, c its concentration, and s the permitivity of the medium. Noteworthy is the strong dependence on electric field and radius in Eq. (30). The temperature increase resulting from this heat generation is a function of the heat transfer environment. The heat dissipation characteristics of planar glass chips are much improved over capillaries in air or forced air configurations. This permits the use of much higher electrical fields, resulting in faster separation times. This increased capability is a major factor driving the replacement of traditional capillary electrophoresis with on-chip electrophoresis techniques. Through application of a one-dimensional heat transfer model [29-33], it was determined that the centerline temperature of the 200um channel at 10 kV/m in free air convection can rise 8.6°C over the ambient temperature. Under the same conditions, fluid in a 100am channel and a 20um channel would rise over the ambient temperature by only 2.6°C and 0.2°C respectively. If the electric field were increased from 10 kV/m to 20 kV/m, the 200um channel would rise over the ambient by more than 25°C. Although the presence of the parabolic velocity component correlates with increased effects of Joule heating, the exact mechanism is not known. One possible cause is a non-uniform cross-stream temperature distribution. Since heat is generated uniformly throughout the liquid, the temperature profile is parabolic, with the maximum temperature occurring at the centerline. This would cause the magnitude of the dye mobility to increase toward the center of the channel. Since the dye used here is negatively charged, this would cause the observed velocity to be reduced in the center of the channel, as was observed experimentally. This, however, is only a partial explanation as it was found that if the entire velocity deficit was attributed to mobility, the required temperature difference between the center of the channel and the wall was much higher than could be attributed to Joule heating. Another possible contributing factor is the enhanced heat transfer in the oilimmersed portion of the capillary (see Figure 7.18). This can be attributed to the increased conductivity of the optical oil (k = 0.07 Wm"'K"') over air (k = 0.026 Wm"'K"'). In order to fully investigate this phenomenon, particular to large microchannels at moderate to high applied field strengths, a flow module with a
400
Electrokinetics in Microfluidics
temperature controlled microchannel is required. However, the interpretation of images of dye transport at elevated temperatures requires special treatment to include the effect of temperature on the dye mobility. An image velocimetry method, which includes temperature compensation, is presented next. In cases where Joule heating leads to significant fluid temperature increases, the interpretation of dye transport images is complicated by the temperature dependence of dye mobility. However, Eq. (28) may be modified to compensate for this by first expressing the electroosmotic and electrophoretic velocities as follows: (31)
A*
u
Ph= y
ze 1 Ex z -
(32)
L67U2J H
where e is the permitivity of the medium, (the zeta potential, Ex the axially applied electric field strength, fx the dynamic viscosity, z and a the valence and hydrated radius of the dye respectively, and e the charge of a proton. The viscosity is the most temperature dependent parameter in Eqs. (31) and (32) with a relative sensitivity of 2.2 (% / °C) at 25°C. Key to this method is the common inverse proportionality with viscosity in both equations. Combining Eqs. (31) and (32), the electrophoretic and electroosmotic velocity can be related as follows,
U UEO
^ [^)
(33)
Although both the electrophoretic and electroosmotic velocities depend on temperature, the bracketed term in Eq. (33) is relatively constant with temperature as the viscosity dependence of both terms has canceled. Substituting Eq. (33) into Eq. (28) gives the relation,
U
EO = 7
UOB ~r
7T =
n C
• UOB
(34)
where C can, in most cases, be considered a constant for a given liquid-surfacedye combination, and calculated from property values and a velocity
Experimental Studies of Electroosmotic Flow
401
Figure 7.25. Electro-osmotic velocity versus applied field strength with velocity measurements by the independent current-time method, by the flow visualization assuming constant mobility (Eq.(28))., and by the flow visualization method using the temperature compensating method (Eq. (34)).
measurement using a low-strength applied field at a known reference temperature (ambient temperature). If a neutral dye is used, C collapses to unity as expected. However, if the dye carries a charge, the bulk electroosmotic velocity may be determined by scaling the observed velocity by C as in Eq. (34). In the case of negligible Joule heating, Eq. (34) is equivalent to Eq. (28). If Joule heating gives rise to significant fluid temperature increases, however, Eq. (34) includes compensation for electrophoretic mobility increases where as the assumption of constant mobility in Eq. (28) can cause significant errors. This is evidenced in Figure 7.25 where the electroosmotic velocities in a 100|am circular cross-section channel are determined by applying Eq. (28) (no temperature compensation), by using Eq. (34) (temperature compensation method), and by the current-time method (as discussed in the previous sections). Considering the results of the current-time method, a change in mobilities (due to a change in temperature) is evident in the non-linear increase in velocity with increasing applied field. At 10 kV/m the three methods agree, indicating that the channel temperature in that case was close to that assumed by the mobility value used in
402
Electrokinetics in Microfluidics
Eq. (28). At higher applied fields, the velocities calculated with the temperature compensating method are in close agreement with the current-based measurements. The use of Eq. (28), however, is shown to lead to a significant underestimation of the bulk flow velocity. As a final note, if the temperature compensating method is to be applied to flows exhibiting significant crossstream velocity gradients, Eq. (34) should first be applied near the wall to determine the bulk electroosmotic velocity. Then, additional velocities in the bulk region relative to the near-wall velocity (such as a parabolic pressure-driven component) can be added directly to the electroosmotic velocity profile. In summary, electroosmotic velocities in a range of channel geometry were measured by the laser-caged dye microfluidic visualization technique. Plug-like velocity profiles, in agreement with electrokinetic theory, were presented for circular and square cross-section channels of lOOum width. A parabolic velocity deficit was found in electroosmotic flows in a 200um i.d. circular cross-section channel at moderate applied field strengths. Causes for this including pressure effects and Joule heating were discussed. A method to obtain accurate dye-based velocimetry data in channels with elevated temperatures was proposed and independently verified. This temperature compensation method was shown to perform significantly better at high electric fields than a traditional method that assumes a constant dye mobility.
Experimental Studies of Electroosmotic Flow
7-5
403
COMPARISON OF THE CURRENT METHOD AND THE VISUALIZATION TECHNIQUE
The foregoing sections introduced two types of experimental techniques for measuring the electroosmotic velocity in microchannels: the indirect current based methods (either using the full range of the current—time data or using the slope of the measured current—time relationship), and the direct laser-caged dye visualization method. The main advantage of the current based techniques is its simplicity. The current-based experiments described in this chapter can run quickly and with relatively little infrastructure. The main disadvantages of the current-based technique are indirect and lack of spatial and temporal resolution. The microfluidic visualization technique based on "photo-injection" of the fluorescent dye can directly measure the local velocity. The main disadvantages of this direct flow visualization-based method are the required use of specialized, expansive caged dyes and the specialized hardware. The amount of hardware and the expenditure required are essentially the cost for the spatial and temporal resolution this method affords. Employing microfluidics in lab-on-a-chip applications requires use of solutions other than simple and dilute electrolyte solutions. During electrophoresis in various lab-on-a-chip processes, such as DNA sequencing, gene expression, and genotyping, water can be electrolyzed, generating protons at the anode, and hydroxyl ions at the cathode. The cathode end of the electrophoresis chamber then becomes basic and the anodal end acidic. Buffer solutions are used to overcome this problem. Two commonly used buffers are Tris-acetate with EDTA (TAE) and Tris-borate with EDTA (TBE). For controlling various on-chip microfluidic processes, it is desirable to know the electroosmotic mobilities and general electrokinetic properties of TAE and TBE buffers. In this section, both the direct and indirect velocity measurement techniques were applied (concurrently and independently), and electroosmotic velocities of both TAE and TBE at lx concentrations in fused silica capillaries, at a range of electrical field strengths are presented here [12]. A schematic diagram showing the flow module and the supporting hardware for both the indirect and direct measurements is given in Figure 7.18 of the last section. During the experiments, a 10cm length of lOOum i.d. polyitnide-coated fused silica capillary joined the two small reservoirs. Platinum electrodes were embedded into each reservoir. Current data for the indirect method was recorded from the power source by a personal computer-based data acquisition system. Direct, flow visualization measurements were taken through a viewing window at the midpoint between the two reservoirs.
404
Electrokinetics in Microfluidics
Indirect, current-based method Initially, the upstream reservoir (reservoir 1) was filled with buffer. The capillary and the downstream reservoir (reservoir 2) were connected and filled with the same buffer as reservoir 1, but at a 5% lower concentration. Reservoir 2 was connected to ground, and reservoir 1 was connected to a high voltage power supply via platinum electrodes. Then, the capillary (already connected to reservoir 2) was connected to reservoir 1, the current jumped to a non-zero value and electroosmotic flow was initiated. The higher concentration buffer from reservoir 1 gradually filled the capillary, replacing the lower concentration solution. The data acquisition computer recorded the voltage (kV) and current (uA) as a function of time (s). When the replacement was complete, the current reached a plateau value. As an example, Figure 7.26 shows the current-time record obtained during a TAE solution replacement process. In the figure, the sloped, replacement region and the plateau are well defined. There is a small curvature at the start of the process corresponding to the initiation of flow and settling of the electrical instruments. After the replacement, the average velocity, uave, was calculated directly by dividing the length of the capillary by the measured displacement time. Direct, flow visualization-based method The microfluidic flow visualization system with the flow module is shown in Figure 7.18 of the last section. The focused pulse of ultraviolet light enters the capillary from below, and both the excitation light and resulting fluorescent emission are transmitted through the objective. To facilitate direct flow
Figure 7.26. A typical current-time relationship measured for lx TAE buffer solution in a 10cm length of lOOum i.d. capillary, under an applied electrical field of 7.5 kV/m.
Experimental Studies of Electroosmotic Flow
405
visualization, all solutions involved were seeded with caged fluorescent dye. During the experiments, ultraviolet light was focused into a sheet crossing the capillary (perpendicular to the flow direction). The resulting uncaged fluorescent dye was continuously excited and it's fluorescent emission was collected with a laser-powered epi-illumination microscope. Full frame images of the dye transport were recorded by a progressive scan CCD camera and saved automatically on the image acquisition computer. Figure 7.27 shows a sequence of images (at 267 msec intervals) of dye uncaged in the electroosmotic flow of TAE with 7.5 kV/m. The camera was run in open video mode and a delay generator fired the laser at convenient intervals (the laser was pulsed 3 times at 30Hz at each uncaging event). The camera was run at 15Hz with individual exposure times of l/125sec. The acquired images had a resolution of 640x484 pixels. This corresponded to a 346u.ni visible length of capillary, with each pixel representing a 0.54um square in the object plane. To remove any non-uniformity present in the imaging system, dark field subtraction and brightfield normalization were performed. Although the mass-weighted average velocity of the fluid, u, is of interest, only the transport of the uncaged dye ions is observed. The numerical analysis technique must infer the mass-weighted average velocity, u, from images of ion transport which can differ from the advective transport due to forced and ordinary diffusion. In locations where the dye concentration gradient is zero, the observed ionic transport is only the addition of a known electrophoretic velocity and the desired, mass-weighted average velocity. In this method, each of these locations of zero axial dye concentration gradient were chosen as marker points, such that the cross-sectional velocity profile could be calculated directly from the observed velocity of these points as follows:
where u is the mass-averaged velocity of the fluid, wo£ is the observed velocity of the dye, and u „/,, is the electrophoretic velocity of the dye.
Figure 7.27. Images of uncaged dye in an electroosmotic flow of lx TAE buffer, though a lOOum capillary at 267 ms intervals with applied electric field strength of 7.5 kV/m.
406
Electrokinetics in Microfluidics
In the numerical image analysis, a zero gradient marker point was located along each axial line of pixels, using a weighted average of the highest intensity values on that line. The set of markers from each image formed a cross-stream, concentration maxima profile. Figure 7.28(a), (b), and (c) each show concentration maxima profiles calculated from each of eight images, in sequence, obtained with applied electric field strengths of 5kV/m, 7.5kV/m, and lOkV/m respectively. In a given sequence, any pair of profiles could provide a velocity profile by dividing the distance between them, Ax, by the corresponding time step, At. With 8 images, 28 velocity profiles could be constructed. Here, all 28 were calculated and an error-weighted averaging technique was applied. Velocity profiles of the fluid (calculated from the maximum dye concentration profiles in Figure 7.28) are shown in Figure 7.29.
Figure 7.28. Plots of concentration maxima for sequences of images obtained from electroosmotic flow of TAE lx in a lOOum i.d. capillary at: a.) 5kV/m; b.) 7.5kV/m; and c.) lOkV/m.
Experimental Studies of Electroosmotic Flow
407
In this experiment, TAE buffer lx contained 40mM Tris base, 40mM glacial acetic acid, and lmM EDTA in demineralized and distilled water, resulting in a pH of 8. The TBE buffer lx contained 89mM Tris base, 89 mM boric acid, and 2 mM EDTA in demineralized and distilled water, resulting in a pH of 8.3. Although termed 'lx', these buffers differ substantially in ionic concentrations. Although it is difficult to calculate the exact concentration of dissociated ions, the initial concentrations would indicate a bulk counter-ion concentration on the order of 10~2 M for TAE and 10 ! M for TBE at minimum. When the direct, flow visualization method was applied, all solutions included a 1.5xlO~3 M concentration of caged dye. The caged fluorescent dye employed here was 5-carboxymethoxy-2-nitrobenzyl (CMNB)-caged fluorescein (826.81 MW). Using the Nernst-Einstein relation [20], a measured value of the diffusion coefficient of the uncaged dye was used to determine the electrophoretic mobility to be -3.3xl0~8 r r ^ v V . Immediately before use, all solutions were filtered using 0.2um pore size syringe filters, and capillaries were rinsed with pure water and then with the appropriate buffer. The concurrent (direct and indirect) experiments were performed for each buffer. During the experiments, direct velocimetry measurements were taken by firing the laser at ten second intervals with the delay generator. Figure 7.27 shows a sequence of images of the resulting uncaged dye at 267 msec intervals. The plug-like motion, characteristic of electroosmotic flow, is apparent from the one-dimensional transport of the dye. Other image sequences taken during the same replacement process displayed similar behavior and similar electroosmotic velocity. Maximum concentration profiles and calculated velocity profile corresponding to these images are given in Figure 7.28(b) and Figure 7.29(b) respectively. The velocity profiles remain fairly flat right to the wall. This is to be expected as the relatively high ionic concentration results in a compact electrical double layer (EDL). A conservative estimate of the Debye length gives 1/K = 3 nm, which corresponds to an EDL thickness on the order of 10 nm. This EDL thickness is well below the spatial resolution of light microscopy. Thus, the small degree of curvature at the edges of the velocity profiles is an artifact of the vanishing signal-to-noise ratio near the wall. For TAE, the indirect method gave a velocity of 0.374 mm/s, as compared to 0.375 mm/s given by the direct method. For TBE, the indirect method gave a velocity of 0.204 mm/s as compared to 0.213 given by the direct method. The results of these and multiple independent measurements (from each of the direct and indirect methods) for TAE and TBE are presented in Figure 7.30 and Figure 7.31 respectively. Several aspects of the results presented in Figures 7.30 and 7.31 are noteworthy. In general, the agreement between the two methods is strong, and the methods themselves are mutually validated in this sense. For TBE, the direct and indirect measurements differ by less that 5% in most cases (5
408
Electrokinetics in Microfluidics
out of 7), and by 10% or less in all cases. For TAE, values obtained using the two techniques differed by less than 9% in all cases with the exception of the lowest velocity value at 5 kV/m. The difference between methods in this one case (over 50%) is believed to be due to errors encountered using the indirect, current-based method. This measurement involved the lowest electric field strength, corresponding to the lowest current values and the lowest velocities measured. As the velocity is decreased, the replacement process takes longer to complete and the once sharp concentration front between the two solutions becomes progressively blurred. This mixing causes the sloped and the plateau region of the current-time relationship to be smoothed together, which can lead to erroneous measurements, either by estimation of the end time, or through use of the slope method. Although many factors affect the resolution of the currentbased measurements, 5 kV/m was found here to be a lower limit of applicability for this particular solution-capillary system. One possible way to measure velocities at lower electric field strengths would be to decrease the capillary length, and hence decrease the replacement time.
Figure 7.29. Velocity profiles from electroosmotic flow of TAE lx in a 100(j.m i.d. capillary at: a.) 5kV/m; b.) 7.5kV/m; and c.) lOkV/m.
Experimental Studies of Electroosmotic Flow
409
Figure 7.30. Average electroosmotic velocities versus applied electric field strength for lx TAE buffer. Results from both the indirect, current-based method and the direct, flow visualization based method are shown.
Figure 7.31. Average electroosmotic velocities versus applied electric field strength for lx TBE buffer. Results from both the indirect, current-based method and the direct, flow visualization based method are shown.
410
Electrokinetics in Microfluidics
The electroosmotic velocity of both buffers was shown to be directly proportional to the applied electrical field strength. Thus, the electroosmotic behavior of these buffers, under these circumstances, may be characterized by single electroosmotic mobility values. Mobilities were calculated in both cases from the slope of a least squares fit to the data obtained with the direct method. Electroosmotic mobilities of 4.90xl0~8 n^V's" 1 and 3.10xl0~8 m W 1 were determined for TAE and TBE respectively. Applying classical theory, these mobilities indicate zeta potentials of CTAE = -63 mV, and CTBE = -40 mV. The difference in zeta potential, corresponds to the difference in ionic concentrations between lx TAE and lx TBE. This is consistent with data reported by Hunter [20], for vitreous silica surfaces in contact with solutions at high pH. The trend of the data presented in Figures 7.30 and 7.31 is linear and extends through the origin. Although this is to be expected from the classical theory, it is often not found to be the case due to Joule heating effects. If the fluid temperature were raised above the ambient temperature, the electroosmotic mobility would increase. An overall increase in the mobility would be the combined effect of viscosity decreasing (increasing mobility) and decreasing relative permitivity (slightly decreasing mobility). Unfortunately, specific Joule heating effects cannot be studied in isolation from the heat transfer characteristics of the channel and the surroundings. This will be left for another study. Here, it is important to note that using a fused silica capillary in open air (which is a relatively well insulated channel in the context of most microfluidic applications), Joule heating effects were shown to be negligible for both buffers with applied electrical fields up to 20kV/m in the short period of time required (typically, a measurement by the indirect method is complete in less than 200 seconds). Concerning applicability to microfluidic applications, the TAE buffer was found to behave very well. The significantly higher ionic concentration in the TBE buffer, however, was found to present some practical problems. Firstly, the electroosmotic flow was quite slow. Secondly, and more importantly, some bubbles were detected at the electrodes when even moderate electric fields were applied. Although the pH in the reservoir would be maintained by the action of the buffer, bubbles themselves are highly undesirable in most microfluidic applications. One way to avoid this may be to use 0.5x or 0.25x TBE buffer strength as opposed to the more common lx concentration. The strong agreement between data obtained by these two methods provides mutual validation of both methods, and establishes confidence in the electroosmotic mobility values obtained. A linear flow rate increase with applied field strength indicates constant mobility and negligible Joule heating effects.
Experimental Studies of Electroosmotic Flow
7-6
411
FLOW VISUALIZATION BY A MICRO-BUBBLE LENSING INDUCED PHOTOBLEACHING METHOD
As introduced in the previous sections, to visualize the flow field in a microchannel, specialized caged fluorescent dyes have been employed. The photo-injection of dye is accomplished by exposing an initially non-fluorescent solution seeded with caged fluorescent dye to ultraviolet light. As a result of the ultraviolet exposure, caging groups are broken and fluorescent dye is released. However, this technique requires specialized caged dye (which is expensive), and extensive infrastructure to facilitate the photo injection. In this section, we will introduce a micro-bubble lensing induced photobleaching ((J.-BLIP) method for microfluidic visualization [15]. In essence, the micro bubble's surface is used as a lens to intensify the light exposure of the dye in the liquid near the bubble to cause a photobleaching effect. The photobleached dye is then used as a dark marker in the liquid for flow visualization. In the experiment, a capillary was filled with an aqueous fluoresceinbuffer solution such that it contained a single bubble at the midpoint of the capillary. Each end of the capillary was connected to a reservoir and the bubble region was viewed under an oil-immersion, epi-illumination fluorescent microscope. The fluorescent emission of the liquid was imaged with a progressive scan CCD camera, digitized, and stored on the computer. Following the experiment, image processing was performed, and the images were analyzed. In cases where fluid flow was involved, scalar image velocimetry was performed to determine liquid velocities. Imaging was performed with a fluorescent microscope system as described in the previous sections. A continuous, uniform flood of excitation light was provided by a single-line, 200mW, 488nm argon laser (American Laser Corp.), through the 25x, NA = 0.75, oil immersion microscope objective (Leica). The index of refraction of the oil {now = 1.48) was matched to that of the capillary ("fused silica= 1 -46), to avoid lensing caused by the curvature of the capillary. The camera was run in video mode at 15 Hz with individual exposure times of 1/60 s. Due to the geometry of the CCD chip, the camera captured only a square central portion of the field of view illuminated by the microscope optics. The acquired images had a resolution of 640 x 484 pixels. This corresponded to a 543 um visible length of capillary, with each pixel representing a 0.85um square in the object plane. The camera orientation was carefully adjusted such that the pixel grid was aligned with the radial and axial directions. To remove any non-uniformity present in the imaging system, dark-field image subtraction and bright-field image normalization were performed with each image. The bright-field image was obtained by imaging the channel filled with a uniform concentration of fluorescent dye without a bubble present. The
412
Electrokinetics in Microfluidics
pixel intensity values were then scaled linearly by a single factor such that the degree of photobleaching that occurred spanned the grayscale range. Finally the images were smoothed with a distance based 10x10 pixel kernel. Pixel intensities of post-processed images were directly interpreted as dye concentration based on a previously determined, linear, camera response characteristic. The microchannels were 100 urn i.d., 15cm long, circular cross-section, fused silica capillaries (Polymicro Tech.). At the capillary midpoint, the exterior polyimide coating was oxidized and removed to create a viewing window. New capillaries were prepared by flushing with pure water followed by buffer, and then buffer with fluorescent dye. To inject the bubble, air was drawn into the capillary by suction provided by a low-volume, high-pressure syringe (10 p.L, Hamilton Gastight). The air was then slowly pumped out until only the bubble length desired remained in the capillary. At that point, the capillary end was reinserted into the solution and the bubble was drawn to the capillary midpoint. Each end of the capillary was then connected to a small reservoir with embedded platinum electrodes. The fluorescent dye employed here was fluorescein, 332 MW (Molecular Probes). The dye was dissolved to a concentration of c0 = 50 uM in sodium (bi) carbonate buffer at pH = 9.0, / = 0.50. Rhodamine B was also used in preliminary tests. Although u-BLIP was clearly observed with rhodamine B, its
Figure 7.32. Images of a bubble: (a) with backlighting, and (b) with fluorescent emission from the liquid phase.
Experimental Studies of Electroosmotic Flow
413
fluorescence intensity is strongly temperature dependant. To isolate photobleaching from thermal effects, fluorescein (which is relatively insensitive to temperature) was used. Immediately before use, all solutions were filtered using 0.2um pore size syringe filters. When light is incident on an interface, it will refract according to SnelPs Law [34]:
where n and 0 are the index of refraction and transmission angle (relative to the local interface normal) respectively, and subscripts 1 and 2 denote the medium. Thus, when the light is traveling from a higher index medium (ni) to a lower index medium (n2), the light bends away from the normal. Light incident at the critical angle, 0i = 0C, will be refracted along the interface (02 = 90°) where, (37) Any light incident at an angle greater than the critical angle will be totally reflected back into the higher index medium. The effect of this phenomenon, termed total internal reflection (TIR), is illustrated in the back-lit image shown in Figure 7.32(a). The edges of the air bubble are visible in the transparent aqueous solution because the index of refraction of air (nair = 1.0) is less than that of water (iiwater = 1-3)- The edges of the bubble appear dark because the light (from below) is reflected in this region, where the angle of incidence relative to the local interface normal is greater than the critical angle of 0C = 50°. This darkening does not occur in the center of the bubble where the angle between the local interfacial normal and the incident light is less than the critical angle (i.e. the interface is more perpendicular to the incident light). The same phenomena, though to a lesser extent, darkens the channel-liquid interface where (richannei = 1.46 < nwater)- This darkening makes the channel walls visible in Figure 7.32(a), and is one of the key factors that limit near-wall resolution of microflow visualization methods. When the liquid is seeded with fluorescent dye and excitation light is applied, the bubble interface will also reflect the green fluorescent emission from the liquid. The 'bright-spot' optical artifacts shown in Figure 7.32(b) evidence the reflection of fluorescent light generated in the liquid off the bubble caps. Note that this is not the blue (k = 488 nm) excitation light because, through filtering, only green fluorescent emission (A, > 510 nm) is passed to the camera. In a similar manner, the blue excitation light from the microscope will be reflected at the bubble interface as illustrated in Figure 7.33. Figure 7.33(a) illustrates that uniform excitation light exposure in a liquid-only
414
Electrokinetics in Microfluidics
capillary would cause a spatially uniform rate of fluorescence, and in turn, a spatially uniform rate of photobleaching. The addition of an air bubble will alter the path of the excitation light depending on its angle of incidence with respect to the local interface normal. Thus, excitation light may be refracted or reflected at the interface as illustrated in Figure 7.33(b). Light incident at angles greater than the critical angle is reflected back into the liquid as shown on the right in Figure 7.33(b). Light incident at angles less than the critical angle is refracted at both interfaces and transmitted back into the liquid as shown on the left in Figure 7.33(b). In both ways the intensity of the excitation light is increased in the liquid near the bubble. This results in higher total excitation light exposure, initially higher fluorescence intensity, and a higher rate of photobleaching, which in time, results in lower fluorescence intensity. This local photo-destruction is the result of micro-bubble lensing induced photobleaching ((x-BLIP).
Figure 7.33. Illustrative schematics of optical phenomena: (a) uniform photobleaching in a liquid-filled channel, and (b) increased photobleaching in the near-bubble liquid due to the presence of a bubble.
Experimental Studies of Electroosmotic Flow
415
It is important to note that Figure 7.33 and the above explanation are simplifications of the real case. The two-dimensional model of the bubble cap is reasonable considering the hemi-spherical geometry of the interface. However, the excitation light beam will be, in reality, conical, due to the high numerical aperture objective used. This would add a layer of complexity to the optical analysis but the physics remains the same as that illustrated in Figure 7.33. In addition, lensing and increased photobleaching in the liquid film between the bubble and the wall (similar to wave-guide or fiber optic light transmission) is also expected. Due to the small thickness of the film however, the volume of liquid photobleached in this region does not contribute significantly compared to that photobleached at the bubble caps. For the same reason, the extent of photobleaching is essentially independent of bubble length. The image sequence in Figure 7.34 demonstrates the u-BLIP process. Figure 7.34(a) is an image of the uniform fluorescent emission of the dye-filled channel without a bubble. The image in Figure 7.34(b) was taken shortly after the bubble was moved into position. Although both caps contribute equally to the photobleaching, the camera's field of view is focused on the right cap, the point of which has a 'brightspot' reflection as discussed previously. In Figure 7.34(b) the fluorescent emission appears fairly uniform throughout the liquid at the level of the original dye concentration without a bubble (Figure 7.34(a)). The images in Figure 7.34(b-f) were taken in sequence at 20/15 s (1.33 s) intervals, and processed identically. Significant darkening in the near-bubble liquid is apparent. The radial orientation of the dark/bright ray pattern (believed to be caused by interference) is further evidence of the bubble lensing phenomena. After 5.3 s (Figure 7.34(f)), a significant photobleached region is formed around the bubble, where as the dye at the right-hand side of the camera view has retained the original intensity level. This is more apparent in the axial concentration profiles shown in Figure 7.35 where the five profiles correspond to the image sequence in Figure 7.34(b-f). As shown in Figure 7.35, the bulk of the photobleaching occurs during the first 2 seconds. The broadening of the dark region beyond that is mostly due to diffusion of the photobleached dye. Increasing the excitation light energy and correspondingly decreasing the exposure time could reduce this axial broadening. Allowing time for radial diffusion, however, creates a more uniform, axially symmetric photobleached region. For the most part, experimental parameters such as these (solutions, light intensities, exposure duration) were chosen out of convenience and it is likely that performance could be improved with optimization. One option would be to use high molecular weight dyes such as fluorescein-dextran conjugates (Molecular Probes) which are less susceptible to diffusion than standard dyes. The image in Figure 7.34(g) was taken after an axial electric field was applied to the channel. The resulting electroosmotic liquid flow has transformed the dye photobleached at both bubble caps into a dark, cross-stream, liquid flow
416
Electrokinetics in Microfluidics
Figure 7.34. (a) Image of fluorescent emission from a liquid-filled capillary, (b)-(f) Image sequence demonstrating the |a-BLIP process, (g) Image of photobleached dye advected with electroosmotic flow.
Experimental Studies of Electroosmotic Flow
All
marker. The electrokinetic transport of the bulk liquid in this system is calculated numerically. Scalar image velocimetry is then applied to a (i-BLIP generated flow marker for experimental determination of the cross-stream liquid velocity profile in the microchannel. In order to understand the underlying physics of the u-BLIP process, a series of numerical simulations were conducted using the BLOCS (Bio-Lab-Ona-Chip Simulation) finite element code. In these simulations we have considered a long, axially symmetric capillary, having a diameter equivalent to those used in the aforementioned experiments (lOOum), with a 200|0.m long stationary bubble located at the midpoint. The liquid domain geometry was discretized using 9-noded bi-quadratic elements, which were significantly refined in the region near the bubble and coarsened near the capillary entrance and exit. The first stage of the numerical analysis is the determination of the applied potential field in the liquid system which, for the case of a constant conductivity electrolyte, can be determined from the Laplace equation, (38) where (j> is the applied potential field. Requiring that the solution remain finite at the capillary axis and applying insulation conditions, d(j)/dn where n is the normal
Figure 7.35. Axial fluorescence intensity profiles corresponding to the image sequence in Figure 7.34(b)-(f). The bulk of the photobleaching occurs close to the interface and in the first few seconds of exposure.
418
Electrokinetics in Microfluidics
to the surface, were used along the bubble and capillary walls. The electric field lines generated for the case equivalent to 2600 V applied over the 15cm capillary (consistent with the experiments discussed here) are shown in Figure 7.36(a). Since the liquid cross sectional area is significantly reduced, and thus the local channel resistance is greatly increased, the electric field lines are concentrated in the thin film that surrounds the bubble. This has the effect of increasing the gradient of within this region and reducing it far away from the bubble. This is demonstrated in Figure 7.36(b) which shows the change in <j> along the length of the capillary for the no bubble case, a bubble with a .5jxm film thickness and a bubble with a . 1 ^im film thickness. The presence of the bubble reduces dtydx. to 86% and 96% of that for the no bubble case for the 0.1 u.m film thickness and 0.5um film thickness respectively. Also of interest in Figure 7.36(a) is the shape of the iso-potential lines that assume a slightly curved shape beyond the thin film on either side of the bubble. At a distance less that one capillary diameter from the edge of the bubble however, they are nearly perpendicular to the channel wall. This limited influence of the bubble on the shape of the iso-potential field lines will facilitate the use of u-BLIP as a micro-flow visualisation technique. With the potential field solution developed the flow field can be determined using the low Reynolds number Stokes equations, Eq. (39a) and the compressibility condition Eq. (39b), (39a) (39b) where u is the fluid velocity, P is the pressure and n is the viscosity. In principal, Eq. (39a) should contain an electrical body force term resulting from the application of the electric field to the net charge density in the double layer. In this analysis however, we assume an electroosmotic slip velocity, ueo=veoE where veo is the electroosmotic mobility (taken as 6.2 xlO~8 m2/Vs [35]) and E is the gradient of 0 tangential to the surface, applied along the capillary wall. This simplification eliminates the double layer formulation and has proven accurate in transport cases, such as that examined here, where the double layer thickness is very thin compared with the capillary diameter. In this case the condition is slightly more stringent as the double layer thickness must be thinner than the film surrounding the bubble. In the case examined the ionic strength of the buffers is on the order of 10~2 M which gives us a double layer thickness on the order of 3 ran. This is several orders of magnitude smaller than the film thickness encounter here. Applying this slip boundary condition along the capillary wall, a boundedness condition along the capillary axis and a zero tangential shear
Experimental Studies of Electroosmotic Flow
419
Figure 7.36. (a) Numerically determined iso-potential lines in the presence of an insulating bubble, (b) Influence of bubble film thickness on the global applied electric field.
condition along the bubble [36] yields the near-bubble flow field shown in Figure 7.37(a). The velocity profiles at 0.1, 0.4 and 1.0 capillary diameters away from the edge of the bubble are shown in more detail in Figure 7.37(b). As can be seen, very near the bubble the velocity close to the capillary wall is slightly higher than that at the center, however it very rapidly evolves to the traditional plug type velocity profile expected for electroosmotic flow. This result will be used as a verification of the microflow visualization technique developed in the following section.
420
Electrokinetics in Microfluidics
Methods for inferring a bulk fluid velocity by analyzing a sequence of images of dye transport fall under the broad category of scalar image velocimetry. In general, the goal of these methods is to extract the bulk fluid velocity vector, u, from the advective terms of the mass-conservation equation satisfied by the imaged dye species: (40) where c is the concentration of dye species and D is diffusion coefficient. The analysis may be greatly simplified when applied to internal, fully-developed flows by using a discrete dye sample. This is because the motion of the locus of dye-concentration maxima is, for the most part, not affected by diffusion, and reflects the bulk motion of the fluid. This maximum concentration tracking method has been applied in caged-dye based microfluidic flow visualization studies as discussed in the previous sections. In those cases, a photo-injection of fluorescent dye provided the only non-zero dye concentration, c, in an otherwise non-fluorescent solution. In the case of u-BLIP, however, the solution contains a uniform dye concentration, c = c0, and a portion of the dye is photo-chemically destroyed. Since the fluorescent dye concentration, c, is a conserved quantity, the concentration of photo-destroyed dye, c , is also conserved. Thus Eq. (40) is satisfied for c = c', where, (41) and F is any constant. Thus 'maximum' concentration tracking methods can be directly applied to inverted and linearly scaled intensity images of a photobleached sample. Equivalently one may think of the conserved quantity as the photodestroyed fluorophores whose concentration is quantified by a lack of fluorescence. Once the near-bubble liquid was photobleached (as shown in Figure 7.34), an axial electric field was applied by setting the upstream (left) reservoir potential to 2600V, and connecting the downstream (right) reservoir to ground. The camera recorded the resulting transport of the photobleached sample. A five-image sequence taken at 1/15s intervals is given in Figure 7.38(a). The electroosmotic liquid flow has combined the dye photobleached at both bubble caps into a dark, cross-stream, band, which is advected downstream with the electroosmotic flow. The bubble itself is shown to have a finite velocity in the direction opposite to the electroosmotic liquid flow.
Experimental Studies of Electroosmotic Flow
All
Figure 7.37. (a) Numerically determined electroosmotic flow field in the near bubble region, (b) X-direction velocity profiles at various distances from the bubble edge. L denotes distance from the edge of the bubble.
422
Electrokinetics in Microfluidics
Figure 7.38. (a) An image sequence of the p.-BLIP generated flow marker in the electroosmotic flow, (b) Corresponding axial concentration profiles of the flow marker, (c) Corresponding cross-stream profiles of marker concentration maxima.
Experimental Studies of Electroosmotic Flow
423
The concentration field of photobleached dye (c ) was calculated for each image using in-house developed image processing software. Axial marker concentration profiles of the images shown in Figure 7.3 8(a) are given in Figure 7.38(b). The vertical lines on the left in Figure 7.38(b) indicate the motion of the interface, and the profiles on the right show the marker transport. A relatively clear concentration maximum is apparent in each profile. To determine the bubble velocity the distance between the vertical lines may be divided by the corresponding time step. Here a bubble velocity of iibubbie = —190 um/s was determined. To determine the velocity profile in the liquid region, the point of maximum concentration was located along each axial line of pixels, using a weighted average of the highest concentration values in the liquid phase. The set of maximum concentration points from each image formed the cross-stream, concentration maxima profiles shown in Figure 7.38(c). In a given sequence, any pair of concentration profiles could provide a velocity distribution by dividing the distance between them by the corresponding time-step. Here, all five profiles were used in an error-weighted average to determine the velocity data. Once calculated, this velocity data represents the 'observed' velocity of the marker, uob. Since the dye molecules are charged, the observed velocity of the concentration maxima will be the summation of the bulk fluid velocity, ueo, and the electrophoretic velocity of the dye, uph, as follows, (42) The electrophoretic velocity, uph, must be calculated directly from the electrophoretic mobility, vph, and the applied electrical field strength, E, as: (43) The electrophoretic mobility of fluorescein may be taken as vph = -3.3x10 8 (m2/Vs). The local axial applied electric field, however, requires special consideration here. Through the numerical simulations it was shown that the isopotential lines became almost totally radial less than one diameter away from each bubble cap (Figure 7.36(a)). Thus it is reasonable to assume that the electrophoretic velocity of the marker is purely axial (aligned with the potential gradient vector, E). In the case of a channel containing no bubble and filled with a uniform electrolyte solution, the magnitude of the axial electric field may be calculated simple by dividing the applied voltage differential, AV, by the length of the channel, L. As shown in the numerically determined axial potential profiles, Figure 7.36(b), the bubble adds resistance to the channel and hence causes a field reduction in the liquid far away from the bubble. To experimentally determine this reduced field strength in the liquid the electrical
424
Electrokinetics in Microfluidics
current was measured both with and without the bubble present and the following calculation performed: (44)
where / is the electrical current draw, subscript B indicates the presence of a bubble, and subscript NB indicates that no bubble is present. The bracketed current ratio in Eq. (44) is the factor by which the electric field is reduced from the overall value (AV/L) due to the presence of the bubble. A plot of the currents measured with and without the bubble present vs. overall applied electrical field strength is given in Figure 7.39. Both curves trend slightly upwards due to Joule heating induced increases in fluid temperature [14,32]. The current ratio (in Eq. (44)), however, was found to be relatively constant at 0.85 ±0.025 over this range of applied field. Thus according to these current measurements and Eq. (44), the applied voltage drop of AV = 2600 V over the L = 0.15 m length of capillary generated an electrical field strength of E = 14.7 kV/m in the liquid region. This is consistent with the with the decrease in E observed for the 0.1 um film thickness case shown in Figure 7.36(b), suggesting the film thickness is of
Figure 7.39. A plot of the current values measured vs. the overall electrical fields applied to the channel with and without a bubble present.
Experimental Studies of Electroosmotic Flow
425
this order. Using this E value to calculate the electrophoretic marker velocity in Eq. (43), and substituting the result into Eq. (42), gave the bulk liquid electroosmotic velocity profile shown in Figure 7.40. The flat plug-like profile observed is characteristic of electroosmotic flows and is in keeping with the corresponding numerically predicted velocity profile in Figure 7.37. From the wall velocity, the electroosmotic mobility was calculated to be veo = 6.8 xl(T 8 m2/Vs, in reasonable agreement with the value of veo = 6.2 xlO~8 m2/Vs reported by Duffy et al. [35]. The plug-like velocity profile achieved with the (a-BLIP generated flow marker extends to within 4 urn from each wall. This degree of near-wall resolution is comparable to, and in many cases improved over, that achieved with caged-dye based micro-flow visualization techniques that involve increased infrastructure and specialized chemicals. Another advantage of this technique is the ability to concurrently image the bubble geometry, bubble velocity and the local fluid velocity in multiphase systems. In this case, the effective film thickness was measured (through current measurements) and calculated (through numerical simulations shown in Figure 7.36) to be less than 1 um. Near the bubble caps, the interface is optically indistinguishable from the wall, indicating that the film in that area it is less than 1 um. However, the images in Figure 7.38(a) also show significant film thickening in the middle
Figure 7.40. Cross-stream electroosmotic liquid velocity profile obtained by applying scalar image velocimetry to the transport of the n-BLIP generated flow marker.
426
Electrokinetics in Microfluidics
region of the bubble. This thickening will result in lower liquid velocities in that portion of the film. By integrating the cross-stream velocity profile (Figure 7.40), and determining the bubble velocity from the interface movement in Figure 7.38(b), the liquid film velocity was determined to be 3.3 mm/s at the bubble midpoint and an estimated 30 mm/s near the bubble caps. This demonstrates the applicability of this technique to the study of coupled dynamic transport phenomena, characteristic of microscale multiphase systems. In summary, the micro-bubble lensing induced photobleaching ((i-BLIP) method takes advantage of the curvature and the step change in properties across a gas-liquid interface to create a lens/mirror optical arrangement in which light incident on the bubble is focused into the surrounding liquid resulting in a locally increased total light exposure. The effect is demonstrated experimentally by imaging the increased photobleaching rate of fluorophores in the near-bubble region. Based on these phenomena, a micro-bubble lensing induced photobleaching (u-BLIP) technique can then be applied as a method to inject a marker for flow visualization. A series of numerical simulations on the multiphase system demonstrated that both the iso-potential lines and the flow field are disturbed by the presence of the bubble however the effect is limited to the near bubble region (typically less than 1 capillary diameter from the edge of the bubble). Using the u-BLIP technique, the electrokinetic transport of the photobleached marker is analyzed to determine the cross-channel velocity profile of the liquid phase, and the liquid velocity in the film. The results are in good agreement with the numerical predictions and are consistent with velocity measurements from previous studies. This is a new technique for microfluidic flow visualization, particularly applicable to the study of multiphase microchannel flows.
Experimental Studies of Electroosmotic Flow
1-1
427
JOULE HEATING AND HEAT TRANSFER IN CHIPS WITH T-SHAPED MICROCHANNELS
A side effect of microchannel transport by electrokinetic means is the internal heat generation, i.e., Joule heating, caused by electrical current flow through the liquid (e.g., a buffer solution). Maintaining uniform and controlled solution temperatures is important for minimizing dispersion in electrokinetic separations [30,31] and for controlling temperature sensitive chemical reactions such as DNA hybridization [37]. Therefore, it is essential to understand Joule heating and the related heat transfer in microfluidic or lab-on-a-chip systems. Microfluidic or lab-chip systems must have the ability to rapidly reject this heat to the surroundings. Generally it is the ability to dissipate this heat that limits the strength of the applied electric field and thus the maximum flow rate, etc. Recently, more and more microfluidic systems and lab-chips are made from low-cost polymeric materials such as poly(dimethsiloxane) (PDMS) [3841], poly(methylmethacrylate) (PMMA) [42-44], and others (see Becker and Locascio [45] for a comprehensive review) as opposed to traditional materials such as glass or silicon. The primary attractiveness of using these materials is that they tend to involve simpler and significantly less expensive manufacturing techniques, they are also amenable to surface modification [46,47] and the wide variety of physiochemical properties allows the matching of specific polymers to particular applications. While the development of these systems has reduced the time from idea to chip from weeks to days, and the per unit cost by a similar ratio (particularly with the advent of rapid prototyping techniques [35]), the low thermal conductivity inherent in these materials (0.18 W/mK for PDMS which is an order of magnitude lower than that of glass) retards the rejection of internally generated heat during electroosmosis [48]. Studying the on-chip Joule heating and heat transfer requires techniques of temperature measurements in microchannels. A few techniques have been recently developed for making direct "in-channel" measurements of buffer temperatures in microscale systems, the advantages and disadvantages of many of which are discussed in detail by Ross et. al. [48]. While NMR [33], Raman spectroscopy [49], and the recently developed on-chip interferometric backscatter detector technique [32] have been used, the most popular techniques involve the addition of temperature sensitive probes (for example: thermochromic liquid crystals [50], nanocrystals [51], or special fluorescent dyes [52]) to the buffer solution and an observation of the spatial and temporal changes in the thermal field via some type of microscopy technique. Rhodamine B is a fluorescent dye whose quantum yield is strongly dependent on temperature in the range of 0°C to 100°C, making it ideal for liquid based systems. Recently Ross et. al. [48,53] developed a Rhodamine B based thermometry technique for monitoring temperature profiles in microfluidic systems, based on that developed
428
Electrokinetics in Microfluidics
by Sakakibara et. al. [54] for macroscale systems. A similar technique was used by Guijt et. al. [55] to experimentally examine chemical and physical processes for temperature control in microfiuidic systems. In general very little work has been done concerning microscale thermal analysis of microfiuidic based lab-chips on a "whole-chip" level (as opposed to examining just the fluidic domain for example), particularly with respect to the recently developed polymeric systems. This section will present a detailed experimental and numerical analysis of the dynamic changes in the in-channel temperature and flow profiles during electrokinetic pumping at a T-shaped microchannel intersection, using pure PDMS/PDMS and PDMS/Glass hybrid microfiuidic systems [56]. A T-intersection was selected as it represents a general structure of a microfluidic system and it provides an interesting theoretical case due to the inherent spatial gradients in current density and volumetric flow rate. Using a fluorescence based thermometry technique, direct measurements of the in-channel temperature profile are taken and the results are compared with a detailed "whole-chip" finite element model, which accounts for the effects of the temperature field on the local solution conductivity and fluid viscosity as well as thermal conduction through the substrate.
Figure 7.41. Computational geometry for PDMS/PDMS or PDMS/Glass hybrid microfiuidic system.
Experimental Studies of Electroosmotic Flow
429
Thermal model and simulation domain Unlike momentum and species transport analysis, which is confined to the fluidic domain, thermal modeling presents some unique challenges as the heat transfer necessarily requires the simulation domain to include not only the liquid domain but also a significant portion of the chip, if not the entire chip. Different from a macroscale system, where the fluidic domain is most often of comparable size to the solid regions, a microchannel system typically encompasses only a very small fraction of the substrate and thus heat transfer is typically governed by a large time scale thermal diffusion process through the solid region. As shown in Figure 7.41, the system of interest here comprises three coupled domains, the lower substrate (made from either glass or PDMS), fluidic domain (buffer solution) and the upper substrate that contains the channel (made from PDMS). Electroosmotic flow occurs when an applied driving voltage interacts with the net charge in the electrical double layer near the liquid/solid interface resulting in a local net body force that induces the bulk liquid motion. When this voltage is applied to a buffer solution with a finite conductivity, the resulting current induces an internal heat generation effect often referred to as Joule heating. Within the fluidic domain the non-dimensional energy equation takes the form, (45)
where PeF is the Peclet number for the fluidic domain (Pe = pCpLvJk where p is the density, Cp is the specific heat, L is a length scale taken to be the channel height in this case, v0 is the reference velocity and k is the thermal conductivity), 6 is the non-dimensional temperature (6 - (T - To)kL/X00max where Xo is the electrical conductivity at the reference temperature, To, <j)max is the maximum applied voltage and the subscript L signifies properties of the liquid domain), 8 is the non-dimensional time, V is the non-dimensional velocity (V = v/vo) which will be obtained via the fluidic analysis discussed below, A is the nondimensional buffer electrical conductivity (A = A/Ao) and is in general a function of temperature, (P is the non-dimensional applied electric field strength (CP = (t>/<j>max), and the ~ symbol over the V operator indicates the gradient with respect to the non-dimensional coordinates (X = x/L, Y = y/L and Z = z/L). While most thermal properties in the above formulation (i.e. Cp, k, etc.) remain relatively constant over the temperature range of interest, and thus have been assumed so for the purposes of this study, the buffer electrical conductivity has a strong temperature dependence which cannot be ignored. Here we assume that the
430
Electrokinetics in Microfluidics
buffer conductivity is linearly proportional to temperature (having slope a) as shown below, (46) Within the upper substrate (subscript US) and lower substrate (subscript LS) the energy equation takes on a simplified form in that convective effects and the internal heat generation term is absent leaving only the transient and diffusion terms as shown below, (47)
The dominant mechanism of heat rejection from the fluidic domain is diffusion into the solid substrate. For very short times the temperature field is confined to a small region around the channel. However, at longer times (approaching those which are required for the system to reach equilibrium) the temperature field can span an area several orders of magnitude larger than the
Figure 7.42. Measured temperature as a function of scaled fluorescence intensity of Rhodamine B dye. Solid line represents a second order polynomial fit to the data.
Experimental Studies of Electroosmotic Flow
431
channel size, due to lateral thermal diffusion. This poses significant computational problems as it introduces another length scale into the problem. While the height of the computational domain is fixed by the system geometry, the required width of the domain was found to be approximately 50 times the channel width through numerical experimentation. Choosing a smaller domain necessarily led to a significant underestimation of the lateral heat transfer. Boundary conditions on the thermal domain were selected to conform to how most chip-based microfluidic experiments are conducted. The lower surface of the substrate was assumed to have a fixed room temperature (as would be the case for a chip sitting on a relatively large flat surface) while the upper surface was assigned a free convection boundary condition, (48) where Bi is the Biot number (Bi = hL/kus where h is the heat transfer coefficient computed from the "heated upper plate" relation from Incropera and DeWitt [57] to be h = 10 W/m2K). Zero flux conditions were used along the side surfaces. Flow model Generally, the high voltage requirements limit most practical electroosmotically driven flows in microchannels to small Reynolds numbers therefore transient and momentum convection terms in the Navier-Stokes equations can be ignored and thus the fluid motion is governed by the Stokes and continuity equations as shown below, (49)
(50) where a is the non-dimensional shear stress (
432
Electrokinetics in Microfluidics
Figure 7.43. Comparison between numerically (a) and experimentally (b) obtained temperature profiles in a T intersection for the PDMS/PDMS chip. A 2.05 kV/m electric field strength was applied.
Experimental Studies of Electroosmotic Flow
433
Figure 7.44. Comparison between numerically (a) and experimentally (b) obtained temperature profiles in a T intersection for a PDMS/Glass hybrid chip. A 2.05 kV/m electric field strength was applied.
434
Electrokinetics in Microfluidics
simplest way to alleviate this multi-scale problem is to apply a slip boundary condition at the channel wall (Vwau = /^eoV<j) /vo where ^eo = s^C, /r\ is the electroosmotic mobility and C, is the zeta potential) to solve for the bulk liquid motion. Since W = 0 within the bulk liquid, this treatment eliminates the electrical body force term in Eq. (49) and thus the equation for the double layer field from the formulation. As is well known, and was directly observed recently by Ross et. al. [24], neo tends to increase with temperature mostly due to its inverse relationship with viscosity. As with electrical conductivity, the temperature dependence of viscosity is too significant to be ignored and thus we model the changes in viscosity using the following relationship [58],
(51) where ry0 is the viscosity at 0°C (1.788x10 3 kg/ms). The potential field, involved in Eq. (45) and Eq. (49), within the fluidic domain, is described by (52)
which differs from the traditional Laplace equation in that the temperature dependence of the conductivity has been accounted for. Insulation boundary conditions are assigned along the edges of the domain while fixed conditions are applied at the upstream inlets (0=1) and downstream outlets (0=0). Numerical method The above sets of equations have been solved using an in-house written finite element code. The code discretizes the computational domain using 27noded 3D brick elements and makes use of tri-quadratic basis functions for integration of the unknowns. The coupled Navier-Stokes and continuity equations were solved using a penalty method approach, which eliminates the pressure variable from the formulation. The transient convection diffusion problems associated with the thermal analysis were discretized using an implicit first order Euler scheme and solved using an iterative bi-conditioned stabilized conjugate gradient method. As discussed above both the velocity and potential fields are strongly dependent on the changes in the temperature field and thus need to be updated periodically. In the solution scheme used here both these fields were updated every Is (or approximately every 5 time-steps) which was found to be a good compromise between solution accuracy and computational time.
Experimental Studies of Electroosmotic Flow
435
Figure 7.45. Simulated temperature contours for the (a) PDMS/PDMS and (b) PDMS/Glass composite systems 30s after a 2050 kV/m voltage was applied (identical to cases shown in Figures 7.43 and 7.44). Upper image shows the 3D temperature contours in the substrates in the region very near the fluidic region while lower figure details the 2D temperature profile 2.5 mm downstream of the intersection.
436
Electrokinetics in Microfluidics
Experimental technique and image analysis In this study, the chips with a T-shaped microchannel were made by using the soft lithography technique developed by Duffy et. al. [35]. Some of these chips were made of PDMS (both upper and lower substrates), others were hybrid chips, i.e., PDMS upper substrate and Glass lower substrate. Rhodamine B dye is one of a class of fluorescent dyes whose quantum yield is strongly dependent on temperature [52]. As such the in-channel temperature profile can be obtained by observing the relative spatial and temporal changes in the local intensity of the dye using a fluorescence imaging technique. The fluorescent imaging system include a Leica DM-LB fluorescence microscope (Leica Microsystems (Canada), Richmond Hill, Ontario) with a 1 Ox, 0.3 N.A. long working distance objective, a rhodamine B filter set (excitation: band pass 546nm/12nm, emission: band pass 600nm/40nm) and a broadband mercury illumination source was used. 12-bit, 1024 x 1280 pixel grey scale intensity images were captured every 0.25 s for approximately 40s using a Retiga-1300 cooled digital CCD camera (Qimaging, Burnaby, British Columbia) at a typical exposure time of 10 ms. Image acquisition and storage was controlled by Openlab software (Improvision, Guelph, Ontario). Laser grade rhodamine B dye (Acros Organics, Pittsburgh, Pennsylvania) was initially dissolved in pure water at a concentration of lmM and stored at -30°C. Prior to the experiments the dye was further diluted to a concentration of 50 uM in 25 mM carbonate buffer solution at a pH of 9.4. All solutions were filtered before use with a 0.2 urn syringe filters (Whatman, Ann Arbor, Michigan). Prior to each experiment the microchannel system was allowed to cool to the room temperature and an isothermal "cold field" intensity image of the system was taken. Following the acquisition of the cold field image, the electric field was switched on (inducing electroosmotic flow and joule heating in the microchannel) and full speed image acquisition was initiated. In all cases a uniform potential was applied at all upstream inlet reservoirs while the downstream waste reservoir was grounded. Current/Voltage monitoring was done through duel 0--10V signals output from the high voltage source (Spellman, Hauppauge, New York), and captured using a data acquisition card and a custom designed software interface. Following the capture of 150 high resolution images the excitation light was blocked and the electric field was turned off, allowing the system to cool back down to room temperature. After cooling was complete a second cold field image was acquired. The second cold field image was then compared to initial image and in general it was found that the intensity values of the two images were identical, suggesting that any photobleaching of the dye during the experiment was not significant. To extract the in-channel temperature profiles from the captured intensity images each was first normalized by the cold field image as described above. A
Experimental Studies of Electroosmotic Flow
437
Wiener type adaptive filter was then applied in order to smooth the images and reduce the effect of any background noise. The intensity values of treated images were then converted to temperature using the intensity vs. temperature calibration discussed in the proceeding section. To calibrate the intensity vs. temperature behavior of the dye, a PDMS vessel, containing approximately lmL rhodamine B dye solution, with two embedded type-J thermocouples was constructed. Initially the vessel was loaded with dye solution at room temperature and the entire system was heated to approximately 80°C. The system was allowed to cool in air while intensity images were taken at specified intervals and the data acquisition system recorded the instantaneous thermocouple readings. The low thermal conductivity of the PDMS vessel ensured a uniform temperature profile within the higher thermal conductivity liquid region. Random locations in the calibration images were then selected and the intensities at each point were scaled by the cold field image at 30°C (a 30°C cold field image was used here since below that the error in the
Figure 7.46. Effect of applied potential field strength on the temperature profile in the mixing channel of the PDMS/PDMS system 30 seconds after the indicated potential has been applied. Hollow markers represent experimental results and solid lines show numerical predictions.
438
Electrokinetics in Microfluidics
thermocouple measurement became significant compared with the temperature difference). The data was then normalized such that an intensity of 1 corresponded to room temperature. The final results are shown in Figure 7.42. When compared, these results were found to compare very well (within ± 1°C) with those presented by Ross et. al. [48] for similar buffer concentration and pH. The T-shaped microchannel intersection used in this study is illustrated in Figure 7.41. The two inlet channels have a width of lOOum and a length of 6mm, the single outlet channel has a 120um width and is 24mm long, and all channels are 20um deep. In Figures 7.43 and 7.44 we consider two construction configurations, a PDMS/PDMS system (where both substrates are composed of PDMS, Figure 7.43) and a PDMS/Glass hybrid system (where the bottom substrate is glass and the upper substrate is PDMS, Figure 7.44). For both cases the thickness of the upper PDMS piece (containing the channel design) is
Figure 7.47. Comparison between experimentally measured and numerically predicted current load for the experiments shown in Figures 7.43 and 7.44. In all cases a 2.05 kV/m electric field strength was applied. For clarity not all experimental data points are shown.
Experimental Studies of Electroosmotic Flow
439
1.0mm, while the lower substrate for the PDMS/PDMS chip is 1.75mm and 1.0mm for the PDMS/Glass chip. In each case a potential of 2.05 kV was applied to the inlet reservoirs and the downstream outlet was grounded. In all simulations the thermophysical properties of PDMS are: k = 0.18 W/mk, Cp = 1100 kJ/kgK, p = 1030 kg/m3 [59], the thermophysical properties of glass are: k = 1.4 W/mk, Cp = 835 kJ/kgK, p = 2225 kg/m3 [57], and the thermophysical properties of buffer solution are: k = 0.61 W/mk, Cp = 4179 kJ/kgK, p = 1030 kg/m3 [57]. The buffer conductivity at 25°C was measured as 0.22 1/Qm and was found to increase approximately 3%/°C (this relationship was used in determining a from Eq. (46)). neo for the PDMS/PDMS and PDMS/Glass systems were taken as 6.2 xlO~4 cm2/Vs and 6.0xl0~4 cm2/Vs respectively [24]. Thermal analysis Prior to examining the results it is useful to compare the numerical and experimental predictions. As can be seen in Figures 7.43 and 7.44, the experimental and numerical in general agree quite well, typically within ±3°C, with the numerical predictions tending to slightly overestimate the temperature profile. The likely causes of the overestimation are the unknowns in the thermal
Figure 7.48. Computed non-dimensional conductivity profiles in the (a) PDMS/PDMS system and (b) PDMS/Glass system after a 2.05 kV/m voltage has been applied for 30s. Conditions are identical to those discussed for the experiments shown in Figures 7.43 and 7.44.
440
Electrokinetics in Microfluidics
properties of PDMS for which only approximate values could be obtained [59]. The average noise level in the experimental results was approximately ±1°C, which was reduced through the smoothing procedure discussed above, and the observed repeatability was approximately ±2°C. The experimental results shown in Figures 7.43 and 7.44 represent the median case of all experiments preformed. Immediately apparent by comparing either the numerical or experimental results is the dramatic difference between the temperature rise observed in the PDMS/PDMS system compared with the PDMS/Glass system. As can be seen, after 30s the PDMS/PDMS system obtained a maximum temperature in the mixing channel of 58°C while at the same point the hybrid system reached only 32°C. This suggests that indeed the much higher thermal conductivity of the glass substrate (1.4 W/mK as opposed to 0.18 W/mK for PDMS) does significantly improve the heat transfer qualities of the microfluidic system. To analyze the primary cooling mechanisms involved in this effect, consider the dominant heat transfer mechanism on two different timescales. In the initial heating stages, the primary fluid cooling mechanism is the transient but relatively slow heating of the surrounding substrate. During these times an exponential
Figure 7.49. Computed volume flow rate in mixing channel for PDMS/PDMS and PDMS/Glass system 30s after the voltage is applied. Conditions are the same as those for results shown in Figures 7.43 and 7.44.
Experimental Studies of Electroosmotic Flow
441
rise in the temperature profile can be expected (i.e. behavior consistent with a semi-infinite solid [57]), as was observed by Ross et. al. [48]. Beyond this initial substrate heating stage (i.e. at times approaching those required to reach a steady state) the system temperature is governed by the rate of heat rejection from the substrate to the surroundings. To examine this consider Figure 7.45 which shows the computed temperature contours within the solid substrates and the fluidic region, after the 2.05kV potential has been applied for 30s, for the identical (a) PDMS/PDMS and (b) PDMS/Glass systems discussed above. As can be seen for the PDMS/PDMS system, the temperature profile in the substrate is centered around the fluid region and then spreads radially outwards. The obvious temperature gradients seen here demonstrate the inapplicability of previous chip heat rejection models, which assume isothermal conditions [60] for polymeric substrates. Farther away from the fluidic region, boundary effects begin to influence the temperature profile and it can be noted that the temperature gradients become sharper in the lower substrate than in the upper. The sharper gradients suggest that at these longer time periods the majority of the heat is rejected through the lower substrate to the room temperature reservoir on the underside of the chip (despite the larger thickness of the lower substrate) as opposed to free convection from the upper surface. We can examine this effect further by considering a simple steady state thermal resistance model of heat transfer through the upper and lower substrates as shown in the equations below, (53a)
(53b) where R is the resistance to heat transfer and / is the thickness of the substrate. It is important to note here that these equations ignore lateral heat transfer and transient effects and thus cannot be used to make direct predictions of the channel temperature. As is discussed below, they do however allow for order of magnitude estimates of how chip design changes will affect the heat transfer properties of a microfluidic system. For the PDMS/PDMS system, the RLS/Rus ratio is approximately 0.09 suggesting that over 10 times as much heat is transferred through the lower substrate to the room temperature reservoir than through the upper substrate. This has significant implications for the design of PDMS microsystems in that the effects of Joule heating can be dramatically reduced by (see Eq. (53 a)) either reducing the thickness of the lower substrate or
442
Electrokinetics in Microfluidics
increasing its thermal conductivity. The coupling of these effects are shown in Figure 7.45b (PDMS/Glass system) where the smaller lower substrate thickness and higher thermal conductivity has greatly improved the chips ability to regulate the buffer temperature. The upper substrate thickness and material properties are considerably less important since the thermal resistance, from Eq. (53b), is dominated by the convective heat transfer term (i.e. 1/h). Thus significant enhanced heat transfer through the upper substrate can only be obtained by incorporating forced convection or enhanced heat transfer surface (e.g. fins). In both cases shown in Figure 7.43 and Figure 7.44 a significant change in the buffer temperature is observed as the two inlet channels converge into the single outlet channel. This is because of the differences in the rate of internal heat generation in the inlet channels where both the current and potential field gradient are approximately half that of the mixing (outlet) channel. As can be
Figure 7.50. Computed velocity vectors at the channel midplane in (a) PDMS/PDMS system and (b) PDMS/Glass systems 30s after 2.05kV/m electrical field has been applied. Conditions are identical to those discussed for the experiments shown in Figures 7.43 and 7.44.
Experimental Studies of Electroosmotic Flow
443
seen in Figure 7.44, these gradients tend to be smoothed out by the presence of the glass lower substrate, due to the lower lateral thermal resistance. As such, while a hybrid system is much more effective at maintaining a constant buffer temperature, the pure PDMS system is more effective at maintaining on-chip temperature differences (such as those that would be required for the aforementioned thermal cycling technique). The influence of the electric field strength on the temperature rise is demonstrated in Figure 7.46 which shows the temperature profile along the centerline of the mixing channel for the PDMS/PDMS system, 30s after the potential field was applied. As expected the decrease in the potential field leads to a significantly lower buffer temperature, however we also note a slightly more uniform temperature field near the intersection. This is a result of the reduction in the magnitude of the convective term in Eq. (45), which tends to push the colder liquid in the inlet arms into the mixing (outlet) channel, since the lower applied electric field also reduces the magnitude of the electroosmotic velocity. It is also informative to examine the current load on the system for the two cases. Figure 7.47 compares both the experimentally measured and numerically calculated total current draw through the two systems discussed above. As before we see good agreement between the experimental and numerical predictions, with the numerical results tending to slightly underestimate the current, but exhibiting a similar trend. Of interest here is the dramatic difference between the current draw between the PDMS/PDMS and PDMS/Glass system, which can be can be attributed to the increase in solution conductivity with temperature. Figure 7.48 compares the non-dimensional conductivity, A, at the midplane of the channel 30s after the 2.05kV potential has been applied for the (a) PDMS/PDMS and (b) PDMS/Glass systems discussed above. As expected a dramatic change in the conductivity profile is observed near the intersection (corresponding to the sharp temperature gradients also observed there) nearly doubling the buffer conductivity in the mixing channel after 30s while the PDMS/Glass systems maintains a much more uniform conductivity. Such rapid changes in the solution conductivity can have dramatic effects on the local potential field, via Eq. (52), and thus are very important in, for example, sample transport. These local Joule heating induced conductivity differences present an interesting alternative to concentration induced conductivity differences employed previously to mitigate or induce separation effects [61,62]. This increasing of the electrical conductivity necessarily results in an increase in the rate of internal heat generation which, as can be seen in Figure 7.47, results in the continual increase in the current load for the PDMS/PDMS system whereas the hybrid system tends to level off much sooner. This continual increase is reflective of the fact that the large substrate to fluid ratio coupled with this temperature dependent conductivity significantly increases the time required to reach a steady state.
444
Electrokinetics in Microfluidics
Fluidic analysis As mentioned above the electroosmotic mobility tends to increase with temperature primarily due to its inverse relationship with viscosity. As such it is of interest here to examine how significantly the system flow rate will be affected by the joule heating discussed earlier. Figure 7.49 compares the computed volume flow rate (i.e. in the mixing channel) 30s after the voltage has been applied for the two cases shown in Figures 7.43 and 7.44. As can be seen, despite having nearly identical electroosmotic motilities at room temperature (6.2 xlO'4 cm2/Vs for the PDMS/PDMS and 6.0x10"4 cm2/Vs PDMS/Glass), the higher temperature rise in the PDMS/PDMS system significantly increases the volume flow rate over that in the PDMS/Glass system. This could be of substantial importance when attempting to deliver precise quantities of reactants for downstream separation or chemical reactions. Figure 7.50 examines in more detail the flow field at the channel mid-plane 30s after the potential is applied for the 2.05kV case. As discussed, the velocity vectors are significantly larger in the PDMS/PDMS system, however despite the relatively large changes in the local
Figure 7.51. Influence of channel aspect ratio on temperature profile in the mixing channel of a PDMS/PDMS system 30s after a 2.05 kV/m voltage is applied. All channels have an equivalent cross sectional area of 3600 (j.m . W/H is the ratio of the mixing channel width to the mixing channel height.
Experimental Studies of Electroosmotic Flow
445
temperature (and thus viscosity) the flow structure and velocity profile do not change greatly. In the mixing channel, where the temperature is the highest, the viscosity is the lowest (tending to induce a higher velocity), however the higher conductivity tends to reduce the potential field gradient (tending to induce a lower velocity). As a result the viscosity and potential fields tend to have a somewhat counterbalancing effect which tends to maintain the overall flow structure. Influence of channel aspect ratio It is of interest to examine how the channel aspect ratio can influence the temperature profile in the channel. In general it is relatively well understood that microchannels with a larger surface area to volume ratio are better at rejecting internally generated heat to the substrate; but it is of interest here to quantify how significant this effect will be in low thermal conductivity polymeric substrates. Figure 7.51 shows the mixing channel temperature profile for the PDMS/PDMS system, as shown in Figure 7.43, 30s after the 2.05kV potential is applied. In all cases the channels have a cross sectional area of 3600 (am2 such that the rate of internal heat generation and current load in the initially isothermal system were equivalent. As expected the buffer temperatures remained lowest for the very large aspect ratios (25% lower for H/W = 36 vs. H/W =1). Such large aspect ratios however pose significant operational and construction difficulties (particularly for soft elastomers like PDMS which tend to sag in the middle of such wide, thin channels). For more practical aspect ratios however (H/W = 2 for example) it can be seen that the computed temperature difference is less than a degree Celsius and thus is not particularly significant.
446
7-8
Electrokinetics in Microfluidics
JOULE HEATING EFFECTS ON ELECTROOSMOTIC FLOW
As discussed previously, while electrokinetically driven transport processes have great advantages in the flow control and species transport in microfluidic devices, there exists inevitable Joule heating in the liquid when an axial electrical field is applied to generate the electrokinetic flow. This internal heat source can lead to significant increase and non-uniformity in the liquid temperature [29-31,63-66]. Consequently, the electrical field and the flow field are strongly affected via the temperature dependent electrical conductivity and viscosity of the liquid [56,67]. In the absence of Joule heating effects, the electroosmotic velocity profile is plug-like only in a homogeneous microchannel. In heterogeneous microchannel systems [10,68,69], however, the non-plug-like electroosmotic flow [14,70-72] or multidirectional flow or recirculating flow appears [73-77]. In addition, the applied electrical fields in the intersections of many microchannel networks are not uniform and can result in complicated local electroosmotic flow fields [11,78-83]. In the presence of appreciable Joule heating effects, the electroosmotic flow field may be influenced significantly. Generally, the Joule heating effects on the electric field, the temperature filed and the flow field in the microchannel should be considered simultaneously via temperature dependent electrical conductivity, thermal conductivity and viscosity of the liquid [56,67]. Erickson, Sinton and Li [56] examined Joule heating and heat transfer in a T-shaped microchannel intersection in poly(dimethylsiloxane) (PDMS) microfluidic systems, and experimentally observed dramatic temperature gradients as predicted by their "whole-chip" simulation. Xuan, Sinton and Li [67] conducted a comprehensive numerical simulation of the electroosmotic flow in a complete capillary with the consideration of the Joule heating and thermal end effects. A non-uniform electric field along the capillary was revealed first time. A slightly convex cross-stream velocity profile was predicted except in the regions near the capillary ends where concave profiles were found. These velocity gradients were induced by the axial temperature variations via temperature dependent liquid viscosity. A recent experimental study [84] shows that Joule heating effects can also induce perturbations to the electroosmotic velocity profiles in a homogeneous capillary. In this work, a fluorescence-based thermometry technique and lasercaged dye velocimetry technique, which have been described in the previous sections, were applied to measure the local liquid temperature and the local electroosmotic velocity, respectively, at several different points along a homogeneous capillary. A theoretical model was also developed to simulate the temperature field, electric field and flow field in the whole capillary system. The comparison between experimental results and numerical predictions are made.
Experimental Studies of Electroosmotic Flow
447
This section will review the findings of this work on the effects of Joule heating on the electroosmotic flow in a whole capillary. Theory Consider the electroosmotic flow in a glass capillary that is suspended in the air and supported at the two ends by two liquid reservoirs. The set of equations governing the temperature, electric potential and flow fields with the consideration of the Joule heating effects are summarized below. As an axial electrical field is applied to induce the electroosmotic flow in a capillary, the electric current passing through the buffer solution results in Joule heating. This Joule heat is then convectively dissipated to the surroundings after conducting through the capillary wall. The general energy equation in the whole capillary takes the form, (54) where p is the density of either the liquid or the solid (constant liquid density implies an incompressible flow), C „ the specific heat, T the absolute temperature, t the time, u the velocity vector, k(T) the temperature dependent thermal conductivity, <J(T) the temperature dependent electrical conductivity of the liquid, and E the externally applied electric field. Note that the terms containing u and <J(T) vanish in the solid domain because there is neither liquid flow nor electric current in the capillary wall. The whole capillary system is initially at equilibrium with the ambient temperature TQ. Here we imposed isothermal condition at both ends of the capillary, a symmetric condition with respect to the axis, and a convective boundary condition surrounding the capillary with a heat transfer coefficient h. Although both reservoirs, particularly the downstream reservoir, receive heat from the liquid flowing in the capillary, this heat contribution is negligible as the reservoir volumes are of the order of micro-litres. Due to the temperature dependence of liquid conductivity CT(T), the electric fiend E = -V0 becomes non-uniform along the capillary, where the externally applied electric potential <j> is determined by (55) In solving for (/>, insulation conditions were imposed along the edges of the liquid domain. The potentials at the inlet and the outlet of the capillary are EQL and
448
Electrokinetics in Microfluidics
zero, respectively, where EQ represents the externally applied electric field and L is the length of the capillary. Since electroosmotic flows are generally restricted to the range of small Reynolds numbers, the inertia terms in the Navier-Stokes equations can be ignored and thus the liquid motion is governed by the Stokes and the continuity equations, (56) (57) where p is the hydrodynamic pressure, /I(T) the temperature dependent liquid viscosity. The net charge density pe formed by the electrical double layer is zero except in the thin electrical double layer region adjacent to the capillary wall (the characteristic thickness is on the order of nanometers). Therefore, we applied a slip boundary condition Uwau =-S(T)SQ
Experimental Studies of Electroosmotic Flow
449
technique can be found the previous section. As for the measurement of electroosmotic velocity in the capillaries, the laser-caged fluorescein dye microfluidic visualization technique was used in this study. This technique was described in the previous sections of this chapter. In the experiments, laser grade rhodamine B dye (Acros Organics, Pittsburgh, PA) was initially dissolved in pure water (Fisher Scientific Canada, Ottawa, ON) at a concentration of 1 mM and stored at -30 °C. Prior to the temperature measurements, the dye was further diluted to 50 (iM in 25 mM sodium carbonate buffer solution at pH 8.5. For the velocity measurements, fluorescein bis-(5-carboxymethoxy2-nitrobenzyl) ether dipotassium salt (CMNB-caged fluorescein) was used as supplied by Molecular Probes (Eugene, OR). Stock solutions of caged dye (3 mM in pure water stored at -30 °C) were diluted to 1.5 mM in 25 mM carbonate buffer before the experiment. All solutions were filtered before use with 0.2 urn syringe filters (Whatman, Fisher Scientific Canada, Ottawa, ON).
Figure 7.52. Comparison between numerically (solid lines) and experimentally (markers) obtained temperature distributions along the capillary axis 15s after the indicated electric fields were applied. Hollow and filled markers represent measurements in different days.
450
Electrokinetics in Microfluidics
The exterior polyimide coating of a 10cm long fused-silica capillary (Polymicro Technologies Inc., Phoenix, AZ) was oxidized and totally removed to make it transparent to both ultraviolet and visible lights. The inner and outer diameters of the bare capillary are 200 um and 320 um, respectively. In order to make the heat transfer conditions uniform around the whole capillary, an assembly of poly(dimethylsiloxane) (PDMS)—glass reservoirs was custom designed, in which the capillary was suspended only at the two ends by the reservoirs such that the condition of free air convection was realized. Prior to experiments, the capillary was rinsed with pure water, base (1 M NaOH) and pure water in order, followed by filtered buffer. For the velocity measurements, the capillary and the upstream reservoir were filled with the solution of caged fluorescein dye, and the downstream reservoir contained only pure buffer. For the temperature measurements, however, both the capillary and the two reservoirs were filled with the solution of rhodamine B dye. Two platinum electrodes were placed in the solution in each of the reservoirs, and connected to a high-voltage DC power source (CZE1000R, Spellman, Hauppauge, NY). At each end there was around 5 mm long capillary in contact with the solution. Hence, the effective length of the capillary surrounded by air is 9 cm. The whole capillary flow system was mounted on a three-axis translation stage, so that the thermal and fluidic behaviors at several different points along the capillary could be investigated through adjusting the stage.
Figure 7.53. Computed temperature contour (not to scale) in the whole capillary 15s after an electric field of 15 KV/m has been applied. The grey scale represent temperature changes. The dashed line indicates the interface between the liquid and capillary wall. All numbers labelled in the contour refer to the absolute temperatures in K units. The temperature differences between adjacent contour levels are all 2 K except those specially labelled.
Experimental Studies of Electroosmotic Flow
451
Uncaging of the caged fluorescein dye was performed with a pulse of ultraviolet light (provided by a pulsed nitrogen laser, 337 nm, pulse energy 300 uJ, pulse width <4 ns, Laser Science Inc., Franklin, MA) focused into a sheet perpendicular to the flow direction. A four-channel delay generator (Stanford Research Systems, Sunnyvale, CA) controlled the firing of the nitrogen laser. A single-line argon laser (488 nm, 200 mW, American Laser Corp., Salt Lake City, UT) continuously excited the uncaged fluorescein dye or the rhodamine B dye. The resulting fluorescence emission was transmitted first through an epifluorescent microscope head (Leica Microsystems (Canada), Richmond Hill, ON) equipped with a 32x, NA=0.6 (Numerical Aperture) air objective, and then through a 0.63x C mount into a progressive CCD camera (Pulnix America Inc., Sunnyvale, CA). The camera was run at 15 Hz with individual exposure times of 1/60 s. The acquired images had a resolution of 640 x 484 pixels corresponding to 550 um visible length of capillary. While the mismatch of refractive indices of the capillary wall (fused silica, n=1.46) and the air (n=l) causes image distortions in the radial direction, the extraction of axial velocity data remains intact due to the zero curvature along the flow direction. Since the internal
Figure 7.54. Comparison between numerically (solid lines) and experimentally (markers) obtained temperature transients at two indicated points along the capillary axis. An electric field of 15 KV/m was applied in both cases.
452
Electrokinetics in Microfluidics
diameter of the performed capillary is known, we can determine the radial position of dyes and thus the radial velocity profile. In both temperature and flow field measurements, several points close to the two ends and in the middle region of the capillary were investigated. At each point, three different electric fields (5, 10 and 15 KV/m) were applied, between which approximately two minutes were allowed to let the whole capillary system cool to room temperature. For the temperature measurements, the direction of electric field was switched once at every point. Hence, the same point in focus could be close to the inlet or the outlet of the capillary depending on the direction of the applied electric field. The electric field and the CCD camera were turned on simultaneously. When the liquid flow was visualized, however, the electric field was always run in one direction to ensure that the uncaged dye went into the waste reservoir (initially filled with pure buffer). Furthermore, before and after the velocity measurement at each applied electric field, the flow measurements under zero applied voltage were conducted to verify that the magnitude of the mean velocity resulting from a hydrostatic pressure gradient (due to the levelling of the capillary) along the length direction was minimum (i.e., less than 10 um/s). In the flow measurements, the uncaging pulse of ultraviolet light was triggered by a TTL signal every 5 s to ensure that the uncaged dyes from last uncaging event could go sufficiently far away from the field of view. The electric field was applied immediately after the dye was uncaged. For both temperature and velocity measurements, the electric field was left on for about 20 s. These two experiments were conducted twice at different days to verify their repeatability. To extract the temperature data from the captured fluorescent images of rhodamine B dyes, the average intensity of a square of 20x20 pixels in the middle of each image was calculated (the background intensity was subtracted) and then normalized by the one read from the first image under the same electric field (i.e., the very first image immediately after the electric field and the camera were simultaneously turned on, whose intensity corresponds to the room temperature and may be called the reference intensity). This treatment is necessary for different points under observation. The calibration curve (a thirdorder polynomial fit) provided in Ross et aPs work [48] was employed to convert the above obtained intensity values to temperature. Erickson et al. [56] have verified the applicability of this treatment to the buffer and dye concentrations used in this work. To determine the electroosmotic velocity profile from the images of uncaged fluorescein dyes, a zero gradient marker point was located along each axial line of pixels, using a weighted average of the highest concentration (reflected by the pixel intensity) values on that line. The resulting profiles of the highest concentration markers in a series of images could then provide the
Experimental Studies of Electroosmotic Flow
453
velocity profile by dividing the distance between the profiles by the corresponding time-step. This velocity, however, is not the true electroosmotic velocity but the vector addition of liquid electroosmotic velocity and the dye electrophoretic velocity. In order to correctly extract the electroosmotic velocity profile in the presence of Joule heating effects, Sinton and Li [14] proposed a scalar image velocimetry to compensate for the temperature dependence of the electrophoretic mobility of dyes without explicit knowledge of the liquid temperature. This velocimetry method was applied here, where the dye mobility was taken as -3.3xlO^8m2V~Is~1 at room temperature [3,86]. Joule heating effects Joule heating effects cause temperature rises and temperature gradients inside the whole capillary. The axial temperature gradients will make the electric field and the flow field non-uniform along the length direction via temperature
Figure 7.55. Numerically predicted distributions of the axial electric field strength (dashed lines, scaled by the indicated electric fields,) and induced pressure (solid lines, normalized by pUlall where p is the liquid density and £/°o;/ is the slip wall velocity at room temperature To) along the capillary 15s after the indicated electric fields were applied. The arrow shows the increase of electric field from 5 to 15 KV/m.
454
Electrokinetics in Microfluidics
dependent electric conductivity and viscosity, and hence induce axial pressure gradients in the liquid in order for the continuity equation to be fulfilled. Consequently, the electroosmotic velocity profile is curved differently in different regions of the capillary. The temperature and flow fields were examined by using the above-described experimental and numerical approaches. Table 1 gives the physico-chemical properties of the buffer solution (assumed to be identical with pure water, except for the electrical conductivity at room temperature which was measured at a low electric field before experiments) and the capillary wall (fused silica). The heat transfer coefficient for the free air convection around the capillary is estimated [87] as h =55 Wm"2K"!. Temperature field As can be seen in Figure 7.52, the numerically and experimentally predicted temperature distributions (15s after electric fields were applied) agree well along the whole capillary. Each experimental temperature data point is the average value of 15 images. As numerically predicted, sharp temperature drops close to the two ends and a high-temperature plateau in the main body of the capillary were observed. However, the numerical simulation tends to slightly overestimate the flow effect in the downstream region. In the presence of electroosmotic flow, the axial temperature profile of the liquid is inclined to the downstream. The faster the liquid flow is, the more significant the inclination becomes. The experimental results indicated very small radial temperature difference in the capillary. This can also be seen from the model predicted temperature contour (see Figure 7.53).
Table 1 Physico-chemical properties of materials used in the numerical simulation. 7b=298 K denotes room temperature. Liquid Capillary Wall Density p 103 kg/m3
1.00
2.15
Heat capacity Cp 103 J /(kgsxK)
4.18
1.00
Thermal conductivity k W/(msxK)
0.61+0.0012(7- To)
1.38+0.0013(7- To)
Electric conductivity a S/m
0.21 [1+0.02(7- 7b)]
Dynamic viscosity /i kg/(msxs)
2.761exp(1713/7)xl0" 6
Dielectric constant e
305.7exp(-7/219)
Experimental Studies of Electroosmotic Flow
455
The comparison between numerical and experimental predictions of temperature transients at two different points of the capillary is displayed in Figure 7.54. Each experimental temperature data point is an average value obtained from 6 images. This treatment could smooth the inevitable temperature fluctuations during the measurements. Overall, the numerical simulation predicted the trend of temperature development. However, it seems that the total heat capacity of the capillary system was overestimated in the simulation. As a result, the experimentally obtained liquid temperatures rose more quickly in the first several seconds (about 5 seconds in Figure 7.54) and then slowly approached to their final steady values. Flow field The electrical conductivity of the liquid is linearly increased with the rise of temperature (see Table 1), resulting in locally higher electric fields in the inlet and outlet regions of the capillary. Figure 7.55 illustrates the profiles of axial electric field strength scaled by their nominal values. As expected, the scaled electric field strength in the main body of the channel is lower than unity. While both the dielectric constant E and viscosity /j are inversely proportional to the temperature, the electroosmotic slip velocity at the wall
Figure 7.56. Images of electroosmotic flow in different regions along the capillary 15s after an electric field of 15 KV/m was applied, (a) Close to the inlet, (b) In the middle, (c) Close to the outlet. The time interval between images was 1/15 s for all cases.
456
Electrokinetics in Microfluidics
is increased at lower temperatures near the two ends of the capillary. As a result, positive pressure gradients, as shown in Figure 7.55, have to be induced near the capillary ends to reduce the liquid velocity in the center of channel cross-section. Correspondingly, a negative pressure gradient is generated in the middle part of the channel to balance the positive pressure gradients at the ends. The resulting electroosmotic velocity profile is thus convex in the main body of the capillary while becomes concave in its inlet and outlet regions. A series of images of electroosmotic flow (revealed by the motion of uncaged fluorescein dyes) at a point close to the inlet, a point in the middle section and a point close to the exit of the capillary is shown in Figure 7.56 (15s after applying an electric field of 15 kV/m). The moving, curving and dispersing of the dye bands are clearly seen despite that an air objective was used in the flow visualization without any refractive indices matching. As theoretically anticipated, the dye bands (the net migration velocity of dyes is the vector addition of the liquid electroosmotic velocity and the dye electrophoretic velocity) became concave near the capillary ends while convex in the middle region. As seen in Figure 7.56, the uncaging pulse of ultraviolet laser (see the first image at each column) was not perfectly adjusted to be perpendicular to the flow direction. However, the slightly inclined dye bands do not affect the
Figure 7.57. Comparison of numerically (thin solid lines) and experimentally (thick, rough solid lines) obtained profiles of electroosmotic flow in different regions along the capillary 15s after electric fields were applied. The applied electric fields were 5, 10 and 15 KV/m from left to right in all sub-figures (top ones). The bottom figure (not in scale) shows the computed velocity vector at 15 KV/m.
Experimental Studies of Electroosmotic Flow
457
velocity profile to be extracted because it is the relative distances between profiles of the highest concentration markers that determine the final velocity profile. Figure 7.57 compares the profiles of electroosmotic velocity obtained from the numerical simulations and measurements, respectively. As mentioned above, the numerical and experimental predictions of concave velocity profiles are in good agreement in both the inlet and the outlet regions of the capillary (see the top left and top right plots in Figure 7.57). Only at the highest electric field (15 kV/m) is the concave curvature a little underestimated in the simulation. This underestimation might be due to the slight increase in the electrical conductivity of buffer solution mixed with 1.5 mM caged fiuorescein dyes. However, this increase was not considered in the numerical simulation, resulting in less temperature rises in the liquid. The concave curvature, however, is underestimated by the simulations in Figure 7.57, particularly at high electric fields (15 KV/m). One cause for this discrepancy is an increase in conductivity of the solution (found to be approximately 10%) over the duration of the velocity experiments. The conductivity increased during the measurements for two reasons. Firstly, evaporation of water at the free surfaces in the fluid reservoirs
Figure 7.58. Comparison of numerically (solid line) and experimentally (markers) obtained average electroosmotic velocities at different electric fields. All data were extracted 15 s after the electric fields have been applied. Dashed line gives the electroosmotic velocity in the absence of Joule heating effects.
458
Electrokinetics in Microfluidics
concentrated the solution over time. Secondly, the progressive uncaging release of fluorescent dye increased the ionic concentration of the solution. Although experiments were conducted in a timely manner, these two effects are believed to have contributed to the discrepancy between experimental results and numerical results (which did not include these effects). Another potential source of error in the numerical simulations lies in the assumption that the physical buffer properties are identical to those of pure water. Since these properties are temperature dependent (see Table 1), they all affect the curvature of velocity profiles. A considerably curved profile of electroosmotic velocity was observed in the middle region of the capillary (see the top middle plot in Figure 7.57). Even at a low electric field (e.g., 5 kV/m) the convex profile is noticeable. According to the numerical simulation, however, the velocity profile is only slightly convex in this region (see the plot of velocity vector at the bottom of Figure 7.57, not in scale). Johnson et al. [10] saw a slightly convex velocity profile in a straight microchannel. They argued that this slight curvature was due to a non-uniform density of surface charges and a non-uniform radial temperature profile. Indeed, the latter could make the electrophoretic velocity parabolie-like. However, the electroosmotic flow profile remains plug-like if we
Figure 7.59. Comparison between numerically (solid lines) and experimentally (markers) obtained transients of average electroosmotic velocities at different electric fields.
Experimental Studies of Electroosmotic Flow
459
account for only the radial temperature gradient. Hence, the resulting net migration velocity should be concave because the uncaged dyes carry negative charges opposite to those of the channel wall [7,13,14,23,24]. Figure 7.58 compares numerically calculated average velocities with those obtained from different measuring points. The electroosmotic velocities in the absence of Joule heating effects are also shown in Figure 7.58. Note that the average velocity equals the volume flow rate divided by the cross-sectional area of the capillary. One can see that Joule heating effects significantly increase the average velocity (i.e., the flow rate) at high electric fields (for example, more than 50% increase at 15 kV/m as demonstrated in Figure 7.58). The close agreement of average velocities measured from three different points (i.e., inlet, outlet and middle points) verifies that the velocimetry method with temperature compensation works well. It has been shown in Figure 7.54 that the liquid temperature quickly rises in the first several seconds immediately after the electric field is applied. Similarly, we can predict a quick rise in the flow rate. The comparison of the computed and the measured average electroosmotic velocities is presented in Figure 7.59. In accordance with the temperature transients, the numerical simulation overestimates the transient time required for the flow rate to be stable. One can also see that the measured average velocities at time zero are slightly higher than those computed values. This difference is attributed to the finite time delay on measuring the velocity at time zero, during which the Joule heating has taken effects. It should be noted that the electroosmotic flow velocity with no Joule heating effect should be a horizontal line, i.e., independent of time. For graphic clarity, these horizontal lines were not plotted in Figure 7.59. In summary, Joule heating effects can induce perturbations to the electroosmotic velocity profiles in a homogeneous capillary. The liquid temperature and volume flow rate can be significantly increased due to Joule heating effects. Sharp temperature drops were observed close to the two ends of the capillary, and the concave-convex-concave velocity profiles from the inlet to the outlet of the capillary were also observed. These phenomena are due to the Joule heating effects via the temperature dependence of electrical conductivity and viscosity of the liquid.
460
Electrokinetics in Microfluidics
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21]
[22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32]
J.A. Taylor and E.S. Yeung, Anal. Chem., 65 (1993) 2928-2932. J.G. Santiago, S.T. Wereley, C D . Meinhart, D.J. Beebe and R.J. Adrian, Experiments in Fluids, 25(1998)316-319. CD. Meinhart, S.T. Wereley and J.G. Santiago, Experiments in Fluids, 27 (1999) 414419. A.K. Singh, E.B. Cummings and D.J. Throckmorton, Anal. Chem., 73 (2001) 10571061. S.T. Wereley and CD. Meinhart, Experiments in Fluids, 31 (2001) 258-268. W.R. Lempert, K. Magee, P. Ronney, K.R. Gee and R.P. Haugland, Experiments in Fluids, 18(1995)249-257. P.H. Paul, M.G. Garguilo and D.J. Rakestraw, Anal. Chem., 70 (1998) 2459-2467. A.E. Herr, J.I. Molho, J.G. Santiago, M.G. Mungal, T.W. Kenny and M.G. Garguilo, Anal. Chem, 72 (2000) 1053-1057. W. J. A Dahm, L.K. Su and K.B. Southerland, Physics of Fluids A, 4( 10) (1992) 2191 2206. T.J. Johnson, D. Ross, M. Gaitan and L.E. Locascio, Anal. Chem., 73 (2001) 3656. J.I. Molho, A.E. Herr, B.P. Mosier, J.G. Santiago, T.W. Kenny, R.A. Breenen, G.B. Gordon and B. Mohammadi, Anal. Chem., 73 (2001) 1350-1360. D. Sinton, C. Escobedo, L. Ren and D. Li, J. Colloid Interface Sci, 254 (2002) 184-189. D. Sinton and D. Li, Int. J. Thermal Sciences, 42 (2003) 847-855. D. Sinton and D. Li, Colloids Surfaces A, 222 (2003) 273-283. D. Sinton, D. Erickson and D. Li, Experiments in Fluids, 35 (2003) 178-187. B.P. Mosier, J.I. Molho, J.G. Santiago, Experiments in Fluids, 33 (2002) 545-554. S. Arulanandam and D. Li, J. Colloid Interface Sci., 225 (2000) 421-428. L. Ren, C. Escobedo and D. Li, J. Colloid Interface Sci, 250 (2002) 238-242. S. Arulanandam and D. Li, Colloids Surfaces A, 161 (2000) 89. R.J. Hunter, Zeta Potential in Colloid Science: Principles and Applications, Academic Press; New York, 1981. T.J. Mitchison, K.E. Sawin, J.A. Theriot, K. Gee and A. Mallavarapu, Caged fluorescent probes, in Caged Compounds, Vol 291 G. Marriot (Ed.), in: Methods in Enzymology, Academic Press, New York, 1998. S. Inoue and K.R. Spring, Video Microscopy, the Fundamentals, second ed. Plenum, New York, 1997. D.J. Harrison, A. Manz, Z. Fan, H. Ludi and M. Widmer, Anal. Chem, 64 (1992) 19261932. D. Ross, T. J. Johnson and L.E. Locascio, Anal. Chem, 73 (2001) 2509. R.B. Bird, W.E. Stewart and E.N. Lightfoot, Transport Phenomena, John Wiley & Sons, New York, 1960. A.W. Adamson, Physical Chemistry of Surfaces, John Wiley & Sons, New York, 1990. J.H. Crabtree, E.C.S. Cheong, D.A. Tilroe and CJ. Backhouse, Anal. Chem, 73 (2001) 4079. D. Sinton, L. Ren and D. Li, J. Colloid Interface Sci, 260 (2003) 431-439. J.H. Knox, Chromatographia, 26 (1988) 329. J.H. Knox and K.A. McCormack, Chromatographia, 38 (1994) 215. J.H. Knox and K.A. McCormack, Chromatographia, 38 (1994) 207. K. Swinney and D.J. Bornhop, Electrophoresis, 23 (2002) 613-620.
Experimental Studies of Electroosmotic Flow
461
[33] M.E. Lacey, A.G. Webb and J.V. Sweedler, Anal. Chem., 72 (2000) 4991. [34] E. Hecht, Optics, 4 th ed., New York: Addison-Wesley, 2001. [35] D.C. Duffy, J.C. McDonald, O.J.A. Schueller and G.M. Whitesides, Anal. Chem., 70 (1998)4974-4984. [36] D. Erickson, D. Li and C. Park, J. Colloid Interface Sci., 250 (2202) 422-430. [37] D. Erickson, D. Li and U.J. Krull, Anal. Biochem., 317 (2003) 186-200. [38] J.M.K. Ng, I. Gitlin, A.D. Stroock and G.M. Whitesides, Electrophoresis 23 (2002) 3461-3473. [39] J.C. McDonald and G.M. Whitesides, Ace. Chem. Res., 35(7) (2002) 491-499. [40] M.L. Chabinyc, D.T. Chiu, J.C. McDonald, A.D. Stroock, J.F. Christian, A.M. Karger and G.M. Whitesides, Anal. Chem., 73 (2001) 4491-4498. [41] E. Delamarche, A. Bernard, H. Schmid, B. Michel and H. Biebuyck, Science, 276 (1997) 779-781. [42] R.D. Johnson, I.H.A. Badr, G. Barrett, S. Lai, Y. Lu, M.J. Madou and L.G. Bachas, Anal. Chem., 73 (2001) 3940-3946. [43] G.B. Lee, S.H. Chen, G.R. Huang, W.C. Sung and Y.H. Lin, Sensors Actuators B, (2001) 142-148. [44] D.L. Pugmire, E.A. Waddell, R. Haasch, M.J. Tarlov and E. Locascio, Anal. Chem., 74 (2002)871-878. [45] H. Becker and L.E. Locascio, Talanta, 56 (2002) 267-287. [46] Y. Liu, J.C. Fanguy, J.M. Bledsoe and C.S. Henry, Anal. Chem, 72 (2000) 5939-5944. [47] S.L.R. Barker, D. Ross, M.J. Tarlov, M. Gaitan and L.E. Locascio, Anal. Chem., 72 (2000) 5925-5929. [48] D. Ross, M. Gaitan and L.E. Locascio, Anal. Chem, 73 (2001) 4119-4123. [49] K.L. Davis, K.L.K. Liu, M. Lanan and M.D. Morris, Anal. Chem, 66 (1994) 3744-3750. [50] A.M. Chaudhari, T.M. Woudenberg, M. Albin and K.E. Goodson, J. MEMS, 7(4) (1998) 345-355. [51] H. Mao, T. Yang and P.S. Cremer, J. Am. Chem. Soc, 124 (2002) 4432-4435. [52] J. Lou, T.M. Finegan, P. Mohsen, T.A. Hatton and P.E. Laibinis, Rev. Anal. Chem, 18 (1999) 235-284. [53] D. Ross and L.E. Locascio, Anal. Chem, 74 (2002) 2556-2564. [54] J. Sakakibara and R.J. Adrian, Exp. Fluids, 26 (1999) 7-15. [55] R.M. Guijt, A. Dodge, G.W.K. van Dedem, N.F. de Rooij and E. Verpoorte, Lab on Chip, 3 (2003) 1-4. [56] D. Erickson, D. Sinton and D. Li, Lab on Chip, 3 (2003) 141-149. [57] F.P. Incropera and D.P. DeWitt, Introduction to Heat Transfer; John Wiley & Sons: New York, 1996. [58] F.M. White, Fluid Mechanics; McGraw-Hill: New York, 1994. [59] Thermophysical Properties of Sylgard 184 Poly(dimethylsiloxane) btained from manufacturers "Product Information" and through personal communications. [60] M. Jansson, A. Emmer and J. Roeraade, J. High Resolut. Chromatogr, 12 (1989) 797801. [61] R.L. Chien, Electrophoresis, 24(3) (2003) 486-497. [62] M. Urbanek, L. Krivankova and P. Bocek, Electrophoresis, 24(3) (2003) 466-485. [63] S. Bello and P.G. Righetti, J. Chromatogr, 606 (1992) 95-102, 103-111. [64] A. E. Jones and E. Grushka, J. Chromatogr, 466 (1989) 219-225. [65] W.A. Gobie and C.F. Ivory, J. Chromatogr, 516 (1990) 191-210. [66] E. Grushka, R.M. McCormick and J.J. Kirkland, Anal. Chem, 61 (1989) 241-246. [67] X. Xuan, D. Sinton and D. Li, Int. J. Heat Mass Transfer, (2004) (in press).
462
[68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83] [84] [85] [86] [87]
Electrokinetics in Microfluidics
W. Norde and E. Rouwendal, J. Colloid Interface Sci., 139 (1990) 169-176. A. D. Stroock and G. M. Whitesides, Account. Chem. Res, 36 (2003) 597-604. D. Erickson and D. Li, J. Colloid Interface Sci., 237 (2001) 283-289. L. Ren and D. Li, J. Colloid Interface Sci., 243 (2001) 255-261. L.M. Fu, J.Y. Lin and R.J. Yang, J. Colloid Interface Sci, 258 (2003) 266-275. R.R. Cohen and C.J. Radke, J. Colloid Interface Sci, 141 (1991) 338-347. A. Ajdari, Phys. Rev. Lett, 75 (1995) 755-758. A. Ajdari, Phys. Rev. E, 53 (1996) 4996-5005. D. Erickson and D. Li, Langmuir, 18 (2002) 1883-1892. A.D. Stroock, M. Week, D.T. Chiu, W. T. S Huck, P. J. A Kenis, R.F. Ismagilov and G.M. Whitesides, Phys. Rev. Lett, 84 (2000) 3314-3317. S.C. Jacobson, T.E. McKnight and J.M. Ramsey, Anal. Chem, 71 (1999) 4455-4459. T. Thorsen, S.J. Maerkl and S.R. Quake, Science, 298 (2002) 580-584. X. Xuan and D. Li, J. Micromech. Microeng, 14 (2004) 290-298. C.T. Culbertson, S.C. Jacobson and J.M. Ramsey, Anal. Chem, 70 (1998) 3781-3789. N.A. Patankar and"H.H. Hu, Anal. Chem, 70 (1998) 1870-1881. R. Bianchi, R. Ferrigno and H.H. Girault, Anal. Chem, 72 (2000) 1987-1993. X. Xuan, B. Xu, D. Sinton and D. Li, Lab on Chip, (in press). S.V. Ermakov, S.C. Jacobson and J.M. Ramsey, Anal. Chem, 70 (1998) 4494-4504. S. Devasenathipathy, J.G. Santiago and K. Takenhara, Anal. Chem, 74 (2002) 37043713. F.P. Incropera and D.P. DeWitt, Fundamentals of Heat and Mass Transfer, New York, Wiley, 1990.
Electrokinetic Sample Dispensing in Crossing Microchannels
463
Chapter 8
Electrokinetic sample dispensing in crossing microchannels An important component of many bio- or chemical lab-chips is the microfluidic dispenser, which employs electroosmotic flow to dispense minute quantities of samples for chemical and biomedical analysis. The precise control of the dispensed sample in microfluidic dispensers is key to the performance of these lab-on-a-chip devices. In this chapter, we will focus on the electrokinetic sample dispensing or injection processes in a crossing microchannel structure. The perpendicular intersection of the two microchannels is of particular importance for two reasons: Firstly, it represents the most fundamental unit of more complicated, grid-like, microchannel networks; secondly, it can be used, in conjunction with electroosmotic flow, to realize discrete sample injections. The studies of the on-chip sample injection started with capillary electrophoresis separation. Harrison et. al. [1] integrated capillary electrophoresis and sample handling (sample injection and separation) system on a planar glass chip. In this study, they examined and demonstrated the feasibility of using electroosmotic pumping to transport liquids in a manifold of channels and the possibility of conducting the electrophoretic separation on a planar substrate. Seiler et. al. [2] presented improvements in the instrumentation and experimental method for the device described by Harrison et. al. [1]. Effenhauser et. al. [3] performed a high-speed separation of antisense oligonucleotides on a micromachined capillary electrophoresis device on a glass plate, where electroosmotic flow was employed to inject the sample. All these experimental studies focused on the design and testing of microfluidic devices, and did not investigate how to control the microfluidic processes in such devices. The length of the channel required to separate two analytes in capillary electrophoresis is directly proportional to the initial length of the sample [4]. Reducing the separation length reduces both the length of channel and time required to complete a separation. Towards this end, most on-chip dispensing efforts to date have been concerned with tightly focused samples, and high electric field strengths. Jacobson et al. [5] demonstrated the focusing of a sample stream to 3.3 urn in an intersection of 18um channels. Sub-millisecond electrophoresis separation has been achieved in a 200um length of channel using a custom designed chip to support very high field strengths (>1000kV/m) [6]. Through
464
Electrokinetics in Microfluidics
numerical modelling, however, Ermakov et al. [7] reported that the tightest sample focusing may not be suitable for detection purposes due to dilution (dispersion) effects. This is because intensely focused samples contain less mass and exhibit high concentration gradients, and hence are dispersed quickly by diffusion. It was also found that electrokinetic focusing confines the stream spatially but does not increase the peak sample concentration. They simulated different degrees of focusing by altering the field strengths in reasonable agreement with a previous experimental study [5]. Ermakov et. al. [7] presented a 2-D mathematical model to investigate the electrokinetic focusing in a cross intersection of microchannels and the sample mixing in a T-shape microchannel. Based on this model, they [8] studied electrokinetic injection techniques in microfluidic devices. In this paper, the effect of the electric field distribution in the channels on the injected sample concentration was studied, however, they didn't investigate how to control the volume and the uniform distribution of the dispensed sample. Patankar et. al. [9] simulated a 3-D electroosmotic flow in two crossing microchannels. The focus of this model study is the electroosmotic flow of buffer solutions, and the sample transport was not considered. The loading and dispensing of samples that can be transported along the dispensing channel (rather than be directly separated electrophoretically) is very important. These processes are required to deliver discrete samples to points where they may be chemically modified, diluted or mixed, or coupled to other analytical techniques such as chromatography or mass spectroscopy. Recently, how to control the size, shape and concentration of the dispensed samples in straight-cross microchannel dispensers has been investigated numerically and experimentally [10-14]. In general, using a on-chip crossing microchannel dispenser, the sample injecting or dispensing may be realized as follows: Electrical fields are applied to the two microchannels perpendicular to each other as illustrated in Figure 8.1 via
Figure 8.1. Illustration of a sample dispensing process in a crossing microchannel.
Electrokinetic Sample Dispensing in Crossing Microchannels
465
four electrodes. Initially the dominant electrical field along channel 1 is used to load sample solution 1 (white in Figure 8.1) into the intersection. The perpendicular channel 2 contains a solution 2 (black in Figure 8.1). Then the dominant electrical field is switched to be applied along channel 2, the solution 2 starts flowing. Because of the viscous shear force, the flowing solution 2 will "cut" a pocket of solution 1 from the intersection and carry it to the downstream in channel 2. A sample is dispensed in this way. The key in the dispensing process is how to control the size and the concentration of the dispensed samples. This chapter will outline the effects of buffer electroosmotic mobility, sample diffusion coefficient, sample electrophoretic mobility, the conductivity difference of the solutions, the applied field strength and the channel size on the loading and dispensing processes. By understanding the complicated electrokinetic processes involved in the sample dispensing processes, we can find optimal applied voltages and the improved methods to control the size, shape and concentration of the dispensed samples.
466
8-1
Electrokinetics in Microfluidics
ANALYSIS OF ELECTROKINETIC SAMPLE DISPENSING IN CROSSING MICROCHANNELS
To understand the physics of the electrokinetic sample dispensing process, this section will show how to model the loading and the dispensing processes in a crossing microchannel dispenser. The numerical simulation results of this model [10] will be discussed. The microfluidic dispenser examined here is formed by two crossing microchannels as shown in Figure 8.2. The depth and the width of all the channels are chosen to be 20 urn and 50 |uim, respectively, except indicated otherwise. There are four reservoirs connected to the four ends of the microchannels. Electrodes are inserted into these reservoirs to set up the electrical field across the channels. Initially, a sample solution (a buffer solution with sample species) is filled in Reservoir 1, the other reservoirs and the microchannels are filled with the pure buffer solution. When the chosen electrical potentials are applied to the four reservoirs, the sample solution in Reservoir 1 will be driven to flow toward Reservoir 3 passing through the intersection of the cross channels. This is the so-called loading process. After the loading process reaches a steady state, different electrical potentials are applied to the four reservoirs. The sample solution loaded in the intersection will be "cut" or dispensed into the dispensing channel by the dispensing solution flowing from Reservoir 2 to Reservoir 4. This is the so called the dispensing process. The volume and the concentration of the dispensed sample are the key parameters of this dispensing process, and they depend on the applied electrical
Figure 8.2. The schematic diagram of a crossing microchannel dispenser. Wx and Wy indicate the width of the microchannels.
Electrokinetic Sample Dispensing in Crossing Microchannels
467
field, the flow field and the concentration field during the loading and the dispensing processes. Electrical field In most on-chip applications, buffer solutions are used. Buffer solutions usually have high ionic concentration, e.g., 10~2M, correspondingly, the electrical doubly layer is very thin, on the order of several nanometers. Recall that only in the electrical double layer, there is a non-zero net charge. To simplify the analysis, we consider the electrical double layer filed with negligible thickness. In this way, the liquid in the microchannel will be considered electrically neutral, i.e., the number of positive ions is equal to the number of negative ions, or the net charge density is zero. According to the theory of electrostatics, the applied electrical potential, (f), in the liquid can be described by Poisson equation, (1) Because no electrical potential is applied in z-direction (the direction of the channel depth), the electrical potential profile in this direction is constant. Therefore, Eq. (1) can be rewritten as (2, Introducing nondimensional parameters
where O is a reference electrical potential and h is the channel width chosen as 50 (a.m, Eq. (2) can be non-dimensionalized as:
Boundary conditions are required to solve this equation. We impose the insulation condition to all the walls of microchannels, and the specific non-
468
Electrokinetics in Microfluidics
dimensional potential values to all the reservoirs. Once the electrical field in the dispenser is known, the local electric field strength can be calculated by (4)
Flow field The basic equations describing the flow field are the continuity equation, (5) and the momentum equation, (6)
where Veo is the bulk electroosmotic velocity vector, p is the density of the liquid, P is the pressure in microchannels, n is the viscosity of the liquid and pe is the net charge density in the solution. As mentioned earlier, the driving force in the liquid flow is the electrical force, peE, which appears as the third term of the right hand side of Eq. (6). The electroosmotic velocity is constant everywhere in the channel except in the region very near the wall (i.e. within the double layer). However, when the concentration is higher (i.e. C > 10" mol/l), the electrical double layer is very thin (i.e., less than 10 run) as compared to the dimensions of the microchannel's cross-section (i.e., 50 urn). Therefore, we will neglect the electrical driving force term in Eq. (6) and consider the effect of electroosmotic flow as the slip wall boundary conditions to the equation of motion. The electroosmotic mobility is assumed to be fj.eo =5.5xl0~ 8 m2 l(V-s) in this section, except indicated Vh otherwise. In the electroosmotic flow process, Reynolds number, Re = — (v is v the kinematic viscosity, V is the electroosmotic velocity and h is the channel width), the ratio between the inertial forces and the viscous forces, is very small (e.g. Re < 0.1). Hence, the viscous forces prevail and define the characteristic time scale at which the flow field reaches a steady state. Therefore, the time scale for electroosmotic flow to reach a steady state can be evaluated by
Electrokinetic Sample Dispensing in Crossing Microchannels
469
(7) This time scale is very small as compared to the characteristic time scales of the sample loading and sample dispensing. Hence, the electroosmotic flow here is approximated as steady state. In addition, the velocity component in z-direction, w, is very small as compared to the velocity component in x-direction and y -direction, wand v. Therefore, the dispensing process in the dispenser can be considered as a twodimensional problem, which has been verified by Patankar [9]. Taking into account of the above considerations, Eq. (5) and Eq. (6) can be rewritten as: (8)
(9a)
(9b)
where ueo, veo are the electroosmotic velocity component in x and y direction, respectively. Introducing non-dimensional parameters:
where Pa is the atmospheric pressure, Eqs. (8), (9a) and (9b) can be nondimensionalized as: (10)
470
Electrokinetics in Microfluidics
(lla)
(lib)
Boundary conditions are required in order to solve this set of equations numerically. The slip velocity conditions are applied to the walls of the microchannels, the fully developed velocity profile is applied to all the interfaces between the microchannels and the reservoirs, and the pressures in the four reservoirs are considered as the atmospheric pressure. Concentration field In order to obtain the information about the volume of the dispensed sample, the sample's concentration distribution has to be determined. The distribution of the sample concentration can be described by the conservation law of mass, which takes the form of
(12)
where C; is the concentration of the i-th species, ueo and veo are the components of the electroosmotic velocity of the i- th species, Dj is the diffusion coefficient of the i-th species chosen as Dj =1.0x10" 11 m /(V • s) in all the calculations unless specified otherwise, and uepi and vepi are the components of the electrophoretic velocity of the / - th species given by uepi - E[xepj, where /j.epi is the electrophoretic mobility. As will be shown in the next section, rhodamine dye will be used as a sample in the visualization experiments to verify the model predictions. Therefore, rhodamine dye is chosen as a sample in all the simulations reported here. The electrophoretic mobility for the rhodamine dye was
experimentally
determined
as
-2.46xlO~ 8 m /(V-s)[l5]
and
Vepi = -2-0 x 10~8 m /(V • s) was used in all computations in this section unless specified otherwise. We are interested in how to control the volume of the dispensed sample. This volume depends on the electroosmotic mobility of the solution, the
Electrokinetic Sample Dispensing in Crossing Microchannels
471
diffusion coefficient and the electrophoretic mobility of the sample, the applied electrical potential and the dimensions of the dispenser. Introducing nondimensional parameters:
where C is a reference concentration, we can non-dimensionalize Eq.(12) as:
(13)
Boundary conditions are required to solve the above equation. In the calculations, the concentration of sample in reservoir 1 is the applied sample concentration; the sample concentrations in reservoir 2 and 4 are zero, and the flux of the sample to reservoir 3 is zero. Since both loading and dispensing process are unsteady process, the initial conditions are required to solve the above equation. For the loading process, the initial concentration in reservoir 1 is the applied sample concentration, in the other reservoirs and the channels are zero. After the loading process reaches the steady state, the loaded sample in the cross intersection can be dispensed into the dispensing channel by adjusting the electrical potentials applied to all the reservoirs. Therefore, the concentration distribution for loading process at the steady state was used as the initial conditions for dispensing process. The complete set of non-dimensional equations, Eq. (3), Eq. (10), Eq. (lla), Eq. (lib) and Eq.(13), were solved using the semi-implicit method for pressure-linked equation (SIMPLE) algorithm developed by Patankar [16]. The algorithm is based on a finite control volume discretization of the governing equations on a staggered grid. In order to capture all the features near the four corners of the dispensers shown in Figure 8.2, the non-uniform grid is employed. The control volume size next to the wall is minimum. The size of successive control volumes away from the walls is increased by a factor of 1.2. In this implementation, the solution to this set of equations is obtained by an iterative procedure. During each iteration procedure, the discretized equations are solved by a line-by-line iteration method. Electrical field and flow field Once the electrical potentials are applied to the four reservoirs, the electrical fields are set up across the channels. These electrical fields exert the electrical force on the liquid to pump the solution containing the sample to flow
472
Electrokinetics in Microfluidics
through the cross intersection of the channels (loading process) and then dispense the sample into the dispensing channel. We assume the chip is made of electrically insulation material, and applied the insulation conditions to all the channel walls. Therefore, when the applied electrical potentials change or the channel sizes change, the electrical field will change and hence the velocity field will in turn change. Figure 8.3 shows the typical electrical field and flow field for loading and dispensing process, respectively. In this figure, the non-dimensional applied electrical potentials are:
Figure 8.3. Examples of the applied electrical field (left) and the flow field (right) at the intersection of the microchannels in a loading process (top) and in a dispensing process (bottom).
Electrokinetic Sample Dispensing in Crossing Microchannels
473
Figure 8.4. Effect of the diffusion coefficient on the concentration field of the sample in the dispensing process. The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The diffusion coefficient: a) 1.0 x 10" 11 m2/s, b)1.0 x 10" 10 m2/s .
474
Electrokinetics in Microfluidics
Figure 8.5. Effect of the electroosmotic mobility on the concentration field of the sample in the dispensing process. The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The electroosmotic mobility: a) 5.0xl0~ 8 /n 2 /(F-.s), b) 6.0xl0~ 8 m 2 /(V-s).
Electrokinetic Sample Dispensing in Crossing Microchannels
475
where (i) represents the non-dimensional applied electrical potential to / - th reservoir. For this specific case, the electrical field and the flow field for loading process are symmetric to the middle line of the horizontal channel, and the electrical field and the flow field for the dispensing process are symmetric to the middle line of the vertical channel. Diffusion coefficient effect Under the same applied electrical field as specified in the above, two cases with
two
different
diffusion
coefficients,
L\ =1.0x10~ n m Is
and
Z)2 = 1.0x10~10/w Is, were tested. Figure 8.4 shows the concentration distribution during the dispensing process. In this figure, the black regions are the buffer solution; the white regions represent the region where the sample concentration is higher than 80% of the original sample concentration (the same for all other figures in this section). Generally, the dispensed sample size decreases with the increase of diffusion coefficient, because more samples will diffuse into the buffer solution when the diffusion coefficient is larger, consequently the high concentration region of the sample is smaller. However, the effect of the diffusion coefficient is not significant for this particular case. As seen from Figure 8.4, when the diffusion coefficient increases from 1.0xl0" n w 2 Is (Figure 8.4a) to 1.0xl0" 10 w 2 Is (Figure 8.4b), the volume of the dispensed sample decreases only by 2 % (i.e., from 390 pico-liters to 383 pico-liters in this particular case). Electroosmotic mobility effect Under the same applied electrical field as specified above, two cases with the
different
electroosmotic
mobility,
^ieoj =5.0x10
m l(V-s) and
V-eoi ~ 6-0 x 10 m I(V • s), were examined. The simulation results show that when the electroosmotic mobility increases, the dispensed sample moves further downstream in the dispensing channel, as can be seen in Figure 8.5a and Figure 8.5b. This is because when the electroosmotic mobility increases, the average velocity of the bulk flow increases. Thus, during the same time period, the solution with high electroosmotic mobility is moved further than that with low electroosmotic mobility. The numerical results also reveal that when the electroosmotic mobility increases, the size of the dispensed sample is smaller. For example, at time / = 1.25s, the sample volume for the case of low electroosmotic mobility is 395 pico-liters and for the case of high electroosmotic mobility is 372 pico-liters. This is because the velocity in the upstream of the dispensing channel is higher than that in the downstream of the dispensing channel, as shown in Figure 8.3. When the sample is dispensed downstream in
476
Electrokinetics in Microfluidics
Figure 8.6. Effect of the electrophoretic mobility on the concentration field of the sample in the dispensing process. The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The electrophoretic mobility: a) -2.0x10" 8 m 2 l(V-s), b) -3.5xlO" 8 /n 2 l(V-s).
Electrokinetic Sample Dispensing in Crossing Microchannels
477
the dispensing channel, the front of the sample (in the downstream of the dispensing channel) moves relatively slower than the back of the sample (in upstream of the dispensing channel), consequently, the sample is compressed. When the electroosmotic mobility increases, the difference between the velocity in the upstream and the velocity in the downstream of the dispensing channel is larger. Indeed, the numerical results show that the difference between the velocities in the downstream and the upstream increases when the electroosmotic mobility increases. Consequently, the dispensed sample is more compressed and the volume of the dispensed sample is smaller for the case of high electroosmotic mobility. Electrophoretic mobility effect Under the same applied electrical field as described previously, two cases with different electrophoretic mobility, /uep =-2.0x10 Hep =-3.5x10
m /(V-s),
m /(V-s)
and
were examined. Figure 8.6 shows that the
concentration distribution during the dispensing process. From this figure, we can see that when the absolute value of the electrophoretic mobility increases from 2.0xl0~ 8 m2 /(V-s) to 3.5xl0" 8 m2 I(V -s), the volume of dispensed sample at the same time (i.e. t = 2.0s) increases from 332 pico-liters to 406 picoliters. This is because, with the negative electrophoretic mobility, the electrophoretic motion of the charged sample species is against the electroosmotic flow. The net effect is a reduced bulk flow velocity. When the value of the electrophoretic mobility increases (from 2.0x10 8
m /(V-s)
in
2
Figure 8.6a to 3.5xlO~ w /(V-s) in Figure 8.6b), the net flow velocity of the sample decreases, and the velocity difference between the front of the sample (in the downstream of the dispensing channel) and the back of the sample (in upstream of the dispensing channel) is smaller. Consequently, the compressing effect (as discussed in the previous section) against the diffusion is reduced, and the volume of the dispensed sample increases. Channel size effect In the simulations of the channel size effect, the non-dimensional applied electrical potentials are:
478
Electrokinetics in Microfluidics
Figure 8.7. Effect of the microchannel size on the concentration field of the sample in the dispensing process. The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The channel width ratio: a) Wy/Wx=1:1; b) Wy/Wx=2:1; c) Wy/Wx=l:2.
Electrokinetic Sample Dispensing in Crossing Microchannels
479
Three pairs of channel sizes were used to test the channel size effect on the dispensing process. The results are shown in Figure 8.7. In the first case (Figure 8.7a) the size of the horizontal channel is equal to that of the vertical channel. We describe this case by the ratio of these two channel sizes, Wy/Wx = 1:1, where Wy is the width of the horizontal channel and Wx is the width of the vertical channel. In the second case (Figure 8.7b), the size of the horizontal channel is twice that of the vertical channel, i.e., Wy/Wx = 2:1. In the third case (Figure 8.7c), the size of the horizontal channel is half that of the vertical channel, i.e., Wy/Wx = 1:2. All other parameters are same for these three cases. As seen from Figure 8.7, the channel size has significant effects on the dispensing process and on the dispensed sample volume. Under the same applied electrical field, the shape and the volume of the loaded sample at the intersection are very different for these three cases. The loaded sample volume in the case of Wy/Wx = 1:1 is the largest. Consequently, the shape and the volume of the dispensed sample are very different for these three cases. The dispensed sample volume in the case of Wy/Wx = 1:1 is the largest (the sample volume is approximately 113 pico-liters). This is because under the same applied electrical potentials and the same channel surface properties, if the channel size changes, the electrical field will change since we applied the insulation boundary conditions to all the walls. Since the flow field depends on the electrical field as shown in Figure 8.3, the concentration field, in turn, is strongly dependent on the electrical field. When the electrical field changes due to the change of channel size, the concentration distribution will change. Electrical field strength effect For a given set of properties of the sample and the buffer solutions and the specified microchannels' surface properties and dimensions, the applied electrical potentials are the key controlling parameters for the loading and dispensing processes. Two different combinations of non-dimensional applied electrical potentials were tested as shown in Figure 8.8 and Figure 8.9. The nondimensional applied electrical potentials for the two cases in Figure 8.8 are: In Figure 8.8a:
480
Electrokinetics in Microfluidics
Figure 8.8. Effect of the applied electrical potentials in the loading process on the concentration field of the sample in the dispensing process. The electrical potentials for the dispensing process are the same for both case a) and b). The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The non-dimensional applied electrical potentials are shown at the bottom of this figure.
Electrokinetic Sample Dispensing in Crossing Microchannels
481
The non-dimensional applied electrical potentials for the two cases in Figure 8.9 are: In Figure 8.9a:
The simulation results show that the loading electrical field strength has very important effects on the volume of the dispensed sample, as shown in Figures 8.8a and 8.8b. Under the same dispensing conditions, if the applied electrical potential for the loading process is changed, the size of the loaded sample in the intersection of the microchannels changes and eventually the volume of the dispensed sample changes. Figures 8.8a and 8.8b clearly demonstrate the difference. The volume of the dispensed sample in the downstream of the dispensing channel changes significantly from 390 pico-liters in Figure 8.8a to 204 pico-liters in Figure 8.8b. Comparing Figures 8.9a and 8.9b, we see that under the same loading conditions, if the applied electrical potential for the dispensing process changes, the volume of the dispensed sample changes too. For example, during the dispensing process, while (j) (2) and (j) (4) keep constant, (l) and (j> (3) are increased from 0.2 in Figure 8.9a to 1.0 in Figure 8.9b, the volume of the dispensed sample increases (i.e. from 390 pico-liters in Figure 8.9a increases to 527 pico-liters in Figure 8.9b). This is because when <j) (l) and <j) (3) are bigger, the electrical field strength between the intersection and reservoir 1 and between the intersection and reservoir 3 is weaker. Consequently, less amount of the sample is drawn back toward these two reservoirs and a bigger portion of the sample is dispensed downstream in the dispensing channel (Figure 8.9b). Practical applications such as on-chip separation and detection require that the dispensed sample must distributed uniformly cross the channel cross-section, or have an evenly cut plug shape. However, the shape of the dispensed sample strongly depends on the dimensions of the cross microchannels and the combination of the applied electrical potentials. The even plug-shape of the dispensed samples presented in most figures in this section is the results of carefully choosing the optimal controlling electrical potentials. Figure 8.7b shows an example of non-uniformly dispensed sample.
482
Electrokinetics in Microfluidics
Figure 8.9. Effect of the applied electrical potentials in the dispensing process on the concentration field of the sample in the dispensing process. The electrical potentials for the loading process are the same for both case a) and b). The black regions are the buffer solution; the white regions are the sample solution with a sample concentration higher than 80% of the original sample concentration. The non-dimensional applied electrical potentials are shown at the bottom of the figure.
Electrokinetic Sample Dispensing in Crossing Microchannels
483
In summary, the diffusion coefficient, the electroosmotic mobility, the sample's electrophoretic mobility, the applied potentials and the channel sizes all have different effects on the dispensing process. When the diffusion coefficient increases 10 times, the volume of the dispensed sample decreases by only about 2%. When the electroosmotic mobility increases, the average velocity of the bulk flow increases, and hence the sample is transported more quickly, and has less chance to diffuse into the buffer solution. When the absolute value of the sample's electrophoretic mobility increases, the region of high sample concentration increases and hence the dispensed sample volume increases. Because both the loading and the dispensing processes are electrokinetically driven processes, changes in the applied potentials and in the channel sizes will directly change the applied electrical field in the dispenser, and hence dramatically affect the effectiveness of the loading process and the size of the dispensed sample. Overall, the results discussed in this section show that the size and the shape of the dispensed sample can be controlled by the channels' dimensions and the applied electrical potentials. The model presented here can be used to find the optimal controlling parameter values for the loading and the dispensing processes.
484
8-2
Electrokinetics in Microfluidics
EXPERIMENTAL STUDIES OF ON-CHIP MICROFLUIDIC DISPENSING
The theoretical model and the numerical simulation presented in the last section have revealed the characteristics of the electrokinetic dispensing processes associated with the crossing-microchannel dispenser. These theoretical predictions were verified by an experimental study of the dispensing processes by using a fluorescent visualization technique [11]. The results demonstrated the ability to load and dispense moderate and large sizes of samples on a crossing microchannel chip. During dispensing, smaller samples were found to disperse at a significantly higher rate than larger samples in similar conditions. In addition, the sensitivity of the loading step sample geometry to pressure disturbances was shown to increase dramatically with the extent of the sample. This section will review the experimental method and results, and compare the experimental results with the theoretical model predictions.
Figure 8.10. Schematic diagram of the imaging system and the crossing microchannel chip.
Electrokinetic Sample Dispensing in Crossing Microchannels
485
In the experiments, the sample employed was a fluorescent dye, and a fluorescent epi-illumination video microscope was used to visualize the process. A schematic of the imaging apparatus and the microfluidic chip is given in Figure 8.10. The dye was excited by a continuous flood of blue light provided by a single-line, 200mW, 488nm argon laser, through the microscope objective. A 32x, NA = 0.6 objective was used for magnified views of the intersection and a 16x, NA = 0.3 objective was employed for larger fields of view. The received signal was split by the dichroic mirror (51 Onm LP/ short-reflecting) and passed through an additional filter (515nm LP) and a 0.63x c-mount before reaching the camera. The chip was clamped to a precision 3-axis stage. The two horizontal positioning axes were used to position the field of view, and the vertical axis to focus the microscope. Such arrangements are flexible, but care must be taken to ensure that the objective remains sufficiently far/insulated from the electrodes. Images were captured and saved on the computer at a rate of 15Hz with individual exposure times of 1/125s. A progressive scan CCD camera was used to avoid image defects due to field-field interlacing. The acquired images had a resolution of 640x484 pixels and an 8-bit dynamic range. This corresponds to a viewed region of 1065x805 urn, and 550x416 um using the 16x and the 32x objectives respectively. The camera orientation was carefully adjusted before each run such that the pixel grid was aligned with the coordinate directions of the intersection. Digital image processing was. performed to remove any nonuniformity present in the imaging system and to relate the pixel intensity values to dye concentration. A background noise signal was subtracted, and brightfield image normalization was performed for each image. These images were then smoothed with a distance-based kernel, and scaled by a single factor such that the image series filled the grey scale range. The glass chips with crossing microchannels used in this study were manufactured by Micralyne Inc (Edmonton, Canada). A schematic diagram of the chip from above, and the shape of the channel in cross-section are given in Figure 8.10. The channels are 20u.m deep and 50u.m across, and the reservoirs are 2 mm in diameter. The D-shaped cross-section is a product of the manufacturing technique. The sample was loaded into reservoir Rl, the sample waste reservoir was R3, and the buffer focusing reservoirs were R2 and R4. The lengths of the channels from intersection to reservoir are 5 mm, 4 mm, 80 mm, and 4 mm for channels 1,2,3, and 4 respectively. Two fluorescent dyes, supplied by Molecular Probes, were employed here: fluorescein C2oH1205 (332.31MW); and rhodamine 110 (366.8MW). Both dyes have an absorption maximum near X = 490nm which is well suited to excitation with the X = 488nm output line of the argon laser. The dyes were dissolved in sodium carbonate buffer of pH = 9.0. The buffer was prepared by dissolving 39.8xlO"3 mol of NaHCO3 and 3.41xlO"3 mol of Na2CO3 in 1 litre of pure water
486
Electrokinetics in Microfluidics
resulting in a solution of ionic strength, / = 0.05 M. Stock solutions of fluorescein and rhodamine were prepared in 0.1 mM and 0.2 mM concentrations respectively. The solutions were aliquoted and stored in darkness at -20°C. Immediately before use, all solutions were filtered using 0.2 urn pore size syringe filters. For fluorescein, a diffusion coefficient of D = 4.37x10~10 m2/s was assigned, giving an electrophoretic mobility of vphj = -3.3xl0"8 m2/(V-s). The diffusion coefficient for rhodamine was estimated to be the same as that of fluorescein, however, the electrophoretic mobility was found to be significantly less (less negative) than that of fluorescein. Both dyes were loaded and an onchip separation was performed to determine that rhodamine carried a charge of z R = - 1 at pH = 9.0 (as opposed to fluorescein, zF = -2) giving it an electrophoretic mobility of vpf, R = -1.65xlO"8 m2/(V-s).
Figure 8.11. The loading and dispensing of a focused rhodamine sample: a.) Processed images (experimental results); (b) Iso-concentration profiles at 0.1 Co> 0.3Co, 0.5Co, 0.7Co, and 0.9Co, calculated from the images; and (c) Corresponding Iso-concentration profiles calculated through numerical simulation.
Electrokinetic Sample Dispensing in Crossing Microchannels
487
Chips were prepared by rinsing with each of the following solutions for 20 minutes (in sequence): 1M filtered Nitric acid; 1M filtered Sodium hydroxide (NaOH); and filtered buffer. To pull these solutions through each channel, a vacuum was applied to R3 using a 30mL syringe. A tapered lOOOul pipette tip was cut to the appropriate diameter and friction fitted to join the syringe to the chip. Experimental and numerical results of the loading and dispensing of a focused rhodamine sample are presented together in Figure 8.11. The numerical simulation results were obtained by using the theoretical model presented in the last section under the specified experimental conditions. Processed images are displayed in Figure 8.11 (a) (the top row). The first image shows the steady state loading process. The following sequence images show the dispensing process at 2/15s intervals. Iso-concentration contours of the dye at 0.1Co, 0.3Co, 0.5Co, 0.7Co and 0.9Co (where Co is the concentration of the original sample solution) are given in Figure 8.1 l(b). Corresponding iso-concentration contours calculated using the theoretical model are given in Figure 8.1 l(c). With respect to all plots, the chip was oriented with channels from Rl, R2, R3, and R4, counter-clockwise from top. Orienting the loading process in this way took advantage of the inherent left/right symmetry such that the same voltage could be applied at R2 as R4. This simplified both the experiments and the numerical analyses. The voltage applied to Rl in the loading step in Figure 8.11 was V1L = 1295V. Applied voltages normalized with this value were 1.0, 0.91, 0, and 0.91 for s\i - s^i respectively, (where e indicates normalized applied voltage and in the subscripts the numbers correspond to reservoir number and L denotes the loading step). The triangular shape of the sample was a result of the intersection of the buffer flow from R2 and R4 and the sample solution flow from Rl. Considering the applied voltages and the length of channels 1,2, and 4, the electroosmotic flow rates from each channel were very similar (in fact, the flowrates in channels 2 and 4 were only slightly higher than that in the sample channel 1). The result is a narrowing of the sample stream as it joins the two buffer streams that feed into the sample waste channel 3. The specific geometry of the sample is a combination of advective, diffusive, and electrophoretic transport. To dispense the sample into channel 4, potentials of strength 0.022, 0.063, 0, and 0 were applied for e\j) - s^p (in the subscripts the numbers correspond to reservoir number and the D means the dispensing step) respectively. Under this field, most of the fluid pumped from channel 2 enters the dispensing channel 4, however, some fluid is also pumped into the sample and sample waste channels (1 and 3). It is preferred to drive the sample streams back in this way to avoid leakage and 'cut' the sample. At t = t3, the experimental results in Figure 8.1 l(b) indicate that significant sample diffusion has occurred and the peak
488
Electrokinetics in Microfluidics
concentration value is just over 0.7Co. The numerical results in Figure 8.1 l(c) show a similarly sized sample with a more triangular shape. At t3 there is a small amount of sample at 0.7Co, although most the peak region is between 0.5Co and 0.7Co. The discrepancies between experimental results and numerical results may be due to the assumption of two-dimensionality and/or errors in measuring the applied voltages, diffusion coefficients and mobility values (required as input to the model). Nonetheless, both numerical and experimental results indicate significant loss of sample at the peak concentration only a few channel diameters downstream of the intersection. The results in Figure 8.12 show the effect of decreasing the applied field strength in channels 2 and 4 during the loading stage. These voltages were
Figure 8.12. The loading and dispensing of a less-focused rhodamine sample: a.) Processed images (experimental results); (b) Iso-concentration profiles at 0.1 Co, 0.3Co, 0.5Co, 0.7Co, and 0.9Co, calculated from the images; and (c) Corresponding Iso-concentration profiles calculated through numerical simulation.
Electrokinetic Sample Dispensing in Crossing Microchannels
489
reduced to £21 =£\i= 0.88 (a 3% decrease) and all other run parameters were identical to those pertaining to Figure 8.11. The orientation, formatting and time-steps used in Figure 8.12 are the same as Figure 8.11. As a result, the loading sample geometry became significantly broader. When dispensed, the experimental sample profile is considerably more uniform in the cross-stream direction, and at t3 it contains over 20um of sample at the peak concentration. The numerical solution in Figure 8.12(c) also predicts a significant increase in dispensed sample, although again, the sample geometry is not as uniform in the cross-stream direction and thus not as concentration-dense as the corresponding experimental result. Both experimental and numerical results here demonstrate the ability to alter the sample size and concentration distribution through only minor adjustments to the applied potentials. Figure 8.13 shows the loading and dispensing of a sample of fluorescein dye, in contrast to the rhodamine dye (corresponding to the results in Figures 8.11 and 8.12). Although fluorescein and rhodamine are similar molecules, at pH = 9.0 fluorescein carries a higher charge and hence has a more negative mobility. Since the sum of the electroosmotic mobility and the electrophoretic mobility is greater than zero, the sample moves in the direction of the bulk flow (like rhodamine) but at a reduced rate. The voltage applied to Rl in the loading step was V1L = 1300V, and the reservoir potentials normalized with this value are 1.0, 0.88, 0, and 0.88 for s\i -£41 respectively. This field was effectively the same as that applied in the rhodamine-loading step in Figure 8.12. The fluorescein sample, however, is not as wide as the corresponding rhodamine sample during the loading step. Although it is likely that the total mass flowrates of solution are similar to that of the rhodamine case, the mass flux of the fluorescein species entering the intersection is significantly lower than that of rhodamine due to the difference in their electrophoretic mobilities. With less mass flux, diffusion into the buffer channels is reduced and the sample stream is narrowed as it enters the sample waste channel. This narrowing results in the dispensed sample geometry shown in Figure 8.13, which is smaller than that in Figure 8.12 and less uniform in the cross-stream direction. The dispensing potentials applied in this case were 0.055, 0.127, 0, and 0 for S\D -£40 respectively. The observed velocity of the fluorescein sample in the dispensing stage is similar to that of the rhodamine sample because the electric potential, e 2Z>> w a s doubled. Both numerical and experimental results presented in Figures 8.11, 8.12 and 8.13 demonstrate that: sample shape and concentration profile can be controlled through minor adjustments to the applied field; and similar sample geometry for samples with dissimilar mobility may be achieved through appropriate adjustments to the applied field. However, in all the three cases the numerically determined sample shapes were more focused than those determined
490
Electrokinetics in Microfluidics
experimentally. Thus, the numerical samples were less uniform in the crossstream direction and contained less sample. These differences were amplified in the dispensing process due to diffusion. It is expected that these systematic differences were due to an underestimation of the diffusion coefficients and the corresponding electrophoretic mobility values (related through the NernstEinstein relationship). Methods for more accurately determining these coefficients are needed. Processed images of the dispensing of a focused, large, and very large axial extent rhodamine sample are shown in Figure 8.14 (a), (b) and (c) respectively. To demonstrate the transport process, a wider field of view was used as well as higher dispensing potentials of 0.106, 0.193, 0.01, and 0 for
Figure 8.13. The loading and dispensing of a focused fluorescein sample: a.) Processed images (experimental results); (b) Iso-concentration profiles at 0.1Co, 0.3Co, 0.5Co, 0.7Co, and 0.9Co, calculated from the images; and (c) Corresponding Iso-concentration profiles calculated through numerical simulation.
Electrokinetic Sample Dispensing in Crossing Microchannels
491
6
ID ~ SAD respectively. In each case, the dispensed sample is shown (at right) at approximately t = 4/15s after the dispensing field was applied. As expected, the peak concentration of the smaller sample was reduced during transport while the large samples maintained a significant portion of the original sample concentration. The largest sample, shown in the loading stage in Figure 8.14(c), extended well beyond the intersection. The width of the sample, measured as the full-width-at-half-maximum concentration, was over two channel widths (> 100(am). It was found that this was the largest axial extent sample that could be produced repeatably under these conditions. To further increase sample size, the electroosmotic flow from the buffer reservoirs (R2 and R4) would have to be
Figure 8.14. Processed images (experimental results) of the loading step (left) and dispensing step (at t £ 4/15s, right) of: (a) A focused rhodamine sample; (b) A large axial extent rhodamine sample; and (c) A very large axial extent rhodamine sample.
492
Electrokinetics in Microfluidics
further reduced, reducing the mass-based Peclet number (the ratio of advection to diffusion). Eventually diffusive forces dominate and the sample fills both focusing channels (2 and 4). It is important to note that a modified Peclet number is require here, because it is the sample's velocity which is pertinent. This sample velocity, USAMPLE, is a summation of bulk electroosmotic velocity, uEo, and electrophoretic velocity of the sample, uPH, as follows: (14) and the modified Peclet number, PeM, may be defined as: (15) where L is the characteristic length of the sample and D is the diffusion coefficient. Experimentally it was found that the sample would first leak into one sample stream or the other rather than both concurrently. This was due to a slight velocity bias caused by pressure forces originating in the reservoirs. Like diffusive effects, these pressure disturbances take on a greater role as the electroosmotic flow rate is decreased. In Figure 8.14(c), the center of the loaded sample is shifted slightly to the right of the intersection. This breaking of symmetry is a subtle indicator of the presence of a small pressure driven flow component. These pressure effects are investigated next. To investigate the sensitivity of large axial extent samples to pressure disturbances, a pressure difference, AP, was applied between R2 and R4. Reducing the fluid contained in R2 and maintaining R4, as shown in Figure 8.15, accomplished this. With this pressure gradient in place, the potential of R2 and R4 were varied to achieve the three steady state loading sample geometries in Figure 8.16. A top voltage of V iL = 1320V was applied, with &2L =£4L = 0.924, 0.923, and 0.904 corresponding to Figure 8.16 (a), (b), and (c) respectively. The focused sample shown in Figure 8.16 (a), is not significantly effected by the pressure gradient, although the narrowed sample stream in the sample waste channel 3, is visibly shifted to the left, indicating a larger flow rate from R2 than from R4. By reducing the electroosmotic buffer flow by 1%, the sample becomes larger and significantly asymmetrical, as shown in Figure 8.16 (b). Finally, reducing the electroosmotic flow in channels 2 and 4 a further 1%, the sample is washed into the focusing channel as shown in Figure 8.16 (c). Two conclusions may be drawn here: the axial extent of the sample is very sensitive to
Electrokinetic Sample Dispensing in Crossing Microchannels
493
Figure 8.15. Schematic diagram illustrating reservoir liquid levels as a source of pressure driven flow.
changes in the focusing potentials; and larger axial extent samples are significantly more sensitive to pressure disturbances that focused samples. The source and magnitude of such pressure disturbances in these microfluidic chip applications may be analyzed as follows. Consider the pressure-driven flow between two parallel plates separated by 2h = 20iam (which is the depth of the microchannels employed here). The average velocity generated by a pressure gradient may be expressed as, (16) where ji is the dynamic viscosity, and AP is the pressure differential applied over length L, in the downstream direction. In order for such a pressure driven flow component to significantly influence the sample geometry, the velocity generated must be comparable to the combined electroosmotic and electrophoretic velocity of the sample.
494
Electrokinetics in Microfluidics
(17) Substituting Eq. (17) into Eq. (16), and substituting representative values for the velocities, viscosity, and channel dimensions gives AP = \00Pa. The hydrodynamic head, H, required to generate this pressure may be calculated as (18) where p is the mass density and g is the acceleration due to gravity. The hydrodynamic head generated by depleting R2 is shown schematically in Figure 8.15. Eq. (18) gives a required head of H = 10 mm, which is not reasonable, especially considering the reservoirs are only 1.1 mm in depth. Considering Laplace pressure effects, the pressure across a curved liquidgas interface may be calculated by:
Figure 8.16. Iso-concentration lines at 0.1Co, 0.3Co, 0.5Co, 0.7Co, and 0.9Co, calculated from images of loading sample geometries under the influence of an applied pressure gradient: (a) A focused sample; (b) A less-focused sample; and (c) A large axial extent sample causing washing of sample into the buffer stream.
Electrokinetic Sample Dispensing in Crossing Microchannels
495
(19) where yLV is the surface tension, taken here to be that of water. If R4 is maintained with negligible curvature, and the meniscus of R2 is on the order of the reservoir radius, o = 1 mm (as shown in Figure 8.15), Eq. (18) gives a pressure differential AP = 140Pa. This is more than sufficient (>100Pa) to overcome the combined electroosmotic and electrophoretic velocity in the focusing channels. Thus it is found here that differential Laplace pressures generated in the reservoirs were the most significant cause of pressure-based disturbances in these flows. One solution may be to increase reservoir size (decreasing curvature) and/or increase channel lengths (increase L). Unfortunately these size increases may mitigate some of the key benefits of onchip processing, derived from small sample volumes, and short processing times. In summary, the loading and dispensing of sub-nanolitre samples is achievable by using a microfluidic chip with a crossing microchannel. The ability to inject and transport large axial extent, high-concentration samples was demonstrated experimentally. Both experimental and numerical results indicate the shape, cross-stream uniformity, and axial extent of the samples were very sensitive to changes in the electric fields applied in the focusing channels. In the dispensing process, larger samples were shown to disperse less than focused samples, maintaining more solution with the original sample concentration. In addition, a cross-chip pressure gradient may influence the sample dispensing. Larger samples were found to be more sensitive to pressure disturbances than the more focused samples.
496
8-3
Electrokinetics in Microfluidics
DISPENSING USING DYNAMIC LOADING
In many on-chip applications, longer dispensed samples (samples of greater axial extent) are required. This section will discuss a dispensing method to produce longer dispensed samples. Most on-chip sample dispensing/injection techniques have involved a two-step process. In a cross microchannel dispenser, as described in the previous section, the sample is focused through the intersection of the crossing microchannels by steady flows of buffer from the adjacent dispensing and buffer supply channels. To dispense the sample, a field is applied along the dispensing channel. Common on-chip dispenser/injector configurations include the cross, the tee, and the offset twin-T (or equivalently the 'double-T'). In the offset twin-T configuration, the sample and dispensing channel run together for the short distance between T-intersections resulting in an effectively larger injector length.
Figure 8.17 The three-step injection process illustrated with an image sequence at left and a schematic of the flow directions at right.
Electrokinetic Sample Dispensing in Crossing Microchannels
497
This is used to obtain dispensed samples of greater length than previously achieved in crossing microchannel chips. However, the flowing of buffer and sample adjacently in the injector channel causes increased diffusion and results in a diluted dispensed sample. Furthermore the sample size is greatly determined by the specific injector channel length of the chip and cannot be varied during a given process. Liu et al. [19] conducted sequencing separations of single stranded DNA in a cross and offset twin-T chip injectors with injector lengths of 500 urn, 250 yun, and 100 yun. The best one-color and four-color separations were observed with the 250 pm, and 100 um injectors respectively. Using larger injector lengths was found to increase signal at the expense of resolution. Wallenborg [17] separated and detected a mixture of explosives by micellar electrokinetic chromatography using a straight-cross injector and offset twin-T
Figure 8.18. Image sequences of sample injections with dynamic loading periods of: (a) t D L = 0s (no dynamic loading); (b) t DL = 0.02s; and (c) t D L = 0.05s.
498
Electrokinetics in Microfluidics
injectors with injector lengths of 250 \im and 100 ^im. It was found that the 250 l^m injector gave a 35% increase in peak signal over the straight cross. Although this was an improvement, it was less than the expected based on loaded sample size. Further investigation showed a major loss of sample during the onset of the separation step due to pullback flow induced in the offset channels. Other advances in injector geometry have been reported, such as a significant improvement in resolution, column efficiency and sensitivity using narrow sample channels [20]. In the previous sections, the formation of longer samples in crossing microchannel dispenser by reducing focusing buffer flow was discussed. In that technique, reduced focusing buffer velocities permit broadening of the sample during the steady-state loading step. This was found to produce significantly longer samples than more focused sample injections. However, since the broadening was effectively diffusive, the sample concentration was non-uniform in the axial direction and the sample contained relatively little solution at the original concentration. Experimentally it was found that the reduced focusing buffer velocities also increased the sensitivity of the injection to reservoir-based pressure disturbances. Slight pressure gradients were found to cause sample distortions, and in some cases excessive sample leakage. To generate longer dispensed samples with good control over the sample's size and concentration, a three-step injection procedure for use in crossing microchannel chip was developed [12]. The technique is a variation of the injection process described in the previous section with the addition of an intermediate dynamic loading step. In Figure 8.17, the three steps are shown in an image sequence at left with the flow directions illustrated schematically at right (a relatively long sample was chosen to illustrate the process). Results for both short and long dynamic loading periods will be presented, and compared with the focused and the less-focused crossing microchannel injections. In the experiments, the experimental setup, imaging processing and analysis, the buffer solutions, the dyes, and the crossing-microchannel chips are the same as described in the previous sections. In general, a three-step process requires independent control of three potential levels for each of the four electrodes (12 settings). For this purpose, a Glassman high voltage power supply was used in conjunction with a custom-made voltage controller. The timing of each switch was controlled via a Stanford Research Systems digital delay generator. The voltage controller was programmed for default outputs corresponding to the steady-state loading step. Rising edge triggers from the delay generator commenced dynamic loading and dispensing steps. The period over which the dynamic loading voltages are applied is termed the dynamic loading period, t^i. Sample injections with short dynamic loading periods were conducted in an effort to increase peak signal strength with minimal
Electrokinetic Sample Dispensing in Crossing Microchannels
499
Figure 8.19. The measured Centerline concentration profiles corresponding to injections imaged in Figure 8.18. The profile of the steady-state loading step (common starting point of each injection) is delineated with "o" symbols. The following two profiles were obtained during the dispensing step (at 3/15 s intervals).
500
Electrokinetics in Microfluidics
sample length increase. All electrical potentials were normalized with respect the largest potential, which was 1355V. Normalized potentials applied during the steady-state loading step, sL, the dynamic loading step, e^i, and the dispensing step, S£>, are given below:
(20)
Figure 8.20. The measured iso-concentration contours of the dispensed samples corresponding to injections imaged in Figure 8.18 (contours plotted at 0.1 Co, 0.3Co, 0.5Co, 0.7Co, and 0.9Co).
Electrokinetic Sample Dispensing in Crossing Microchannels
501
Figure 8.21. The numerically determined iso-potential contours plotted in the intersection for: (a) the steady-state loading step, (b) the dynamic loading step, and (c) the dispensing step.
502
Electrokinetics in Microfluidics
where subscripts 1-4 correspond to reservoirs 1-4, respectively. The results are shown in the image sequences in Figure 8.16. A sample injection with no dynamic loading, tpi = 0, is shown in Figure 8.16(a). The first image in the sequence is of the steady- state loading step, and the second and third images are during the dispensing step (taken 3/15 s apart). Although the injected sample is diluted from the original sample concentration, it is reasonably well defined, which is due to the application of a pull-back (or equivalently 'cut-off) voltage at reservoir 1 during the dispensing step. The effect of introducing a very short dynamic loading step of tDi = 0.02s is shown in Figure 8.16(b). Images in Figure 8.16(b) were chosen to approximately match the axial position of the samples with those of Figure 8.16(a). The matching is not exact, however, because the camera was not synchronized with the dispensing signal. The images show that the dynamically loaded sample is significantly brighter than that without dynamic loading. The effect of increasing the dynamic loading period to tpi = 0.05s is shown in Figure 8.16(c). The dispensed sample is threechannel-widths (150 um) long, measuring the full width at half maximum (FWHM). The peak height in the last image is 95% of the original sample concentration. Centerline concentration profiles obtained from the images in Figures 8.16(a), (b) and (c) are plotted in Figures 8.19(a), (b) and (c), respectively. The profile of the steady-state loading step (the starting point for all the injections) is delineated in ' o ' symbols. The original sample intensity level was taken as the signal level at the top of the intersection in the sample channel, and the images were processed such that the signal intensity shown is directly proportional to dye concentration (based on a previously determined, camera response characteristic). The small jump above the background signal, visible at x = 40 um, is an artifact of the imaging process and does not represent sample concentration. This figure illustrates the increase in peak height possible with dynamic loading. As can be seen in the profiles, the steady-state loading profile contains only a small portion of sample at the original sample concentration. The peak height of the dispensed focused sample is 64% (Figure 8.19(a)). This was increased to over 95% using a dynamic loading period of only t£)L = 0.05s (Figure 8.19(c)). The corresponding length increase was from 100 um to 160 um (FWHM). It was also found that the dynamically loaded samples exhibited a greater degree of cross-stream uniformity than the focused samples. The focused pinched-valve injection resulted in a dispensed sample with a higher sample concentration on the sample channel side of the dispensing channel. This is apparent in the iso-concentration contours plotted in Figure 8.20(a) (corresponding to the dispensed sample shown in Figure 8.18(a)). Concentration contours are plotted at concentrations of 0.1Co, 0.3Co, 0.5Co, 0.7Co, and0.9Co, where Co is the original sample concentration. In focused injections, the degree
Electrokinetic Sample Dispensing in Crossing Microchannels
503
Figure 8.22. Processed Images and corresponding centerline concentration profiles for: (a) a long sample created by reducing focusing velocities; (b) a long sample created by further reducing focusing velocities (note the sample drift); and (c) a long sample created by a tDL = 0.2s dynamic loading step.
504
Electrokinetics in Microfluidics
of non-uniformity correlates with the degree of focusing potential applied. This is somewhat unfortunate because it creates larger cross-stream concentration gradients in small samples that, by nature of their size, are more susceptible to dilution. Iso-concentration contours for the dynamically loaded samples, corresponding to Figure 8.18, are given in Figure 8.20. Unlike the focused sample, both dynamically loaded samples exhibit a significant region in the top concentration range. In the case of the longer dynamic loading time (Figure 8.20(c)), this range extends across the channel, which represents a significant improvement in concentration density and cross-stream uniformity over the sample injected without dynamic loading (Figure 8.20(a)). To gain insight into the driving force behind the dynamic loading step, the potential fields for the loading, dynamic loading, and the dispensing steps were calculated using the numerical model described in Section 8-1. Iso-potential contours near the intersection are plotted in Figure 8.21. It is interesting to note that the steady-state loading steps have similar electric field patterns. In both cases, the highest potential values occur in the sample channel (at top). In the loading case (Figure 8.21 (a)), the curvature of the potential lines in the intersection shows how the sample and both buffer streams are electroosmotically pumped into the sample waste reservoir. In the dynamic loading case (Figure 8.2l(b)), the opposite curvature shows how the sample is electroosmotically pumped into both focusing channels and the sample waste channel. As expected, the electric field in the dispensing step, shown in Figure 8.2l(c), is mostly uniform throughout the dispensing channel. In addition to short concentration-dense samples, longer samples (samples of greater axial extent) are also of interest. In Figure 8.22(a), a reduced-focused sample is shown in the loading stage, and the corresponding centerline concentration profile is shown at right. This sample was produced by setting S£_2 =£L-4 =0.935. This 6.5% reduction in focusing potential resulted in a two-fold increase in sample length. The results of a further reduction to Si_2 =££-4 =0.918 are shown in Figure 8.21(b). The drift to the right, apparent in the figure, is likely due to undesired pressure disturbances originating from Laplace pressure acting on the free surfaces in the reservoirs. These effects only become significant when the electroosmotic buffer velocities are reduced. This presents a practical constraint on the length of samples that can be produced by this reduced-focusing technique. Fundamental to that technique is the use of diffusive forces to broaden the sample. Although this can create larger samples, it also results in an increased pressure sensitivity and sample dilution. Using dynamic loading, however, sample broadening is achieved by direct advection of sample into the dispensing channel. Thus, well-defined samples of high concentration density can be produced at any length. This is demonstrated in Figure 8.22(c) with a sample at the end of a dynamic loading period of tjji =
Electrokinetic Sample Dispensing in Crossing Microchannels
505
Figure 8.23. The measured centerline concentration profile sequences of the dispensing of the samples shown in Figure 8.22. In each case, the left-most profile is that of the loaded sample, delineated in ' o ' symbols. The following two profiles were obtained during the dispensing step (at 3/15 s intervals).
506
Electrokinetics in Microfluidics
0.2s (prior to dispensing). The flat-topped concentration profile indicates the concentration of the bulk of the sample is at the level of the original sample concentration. The high concentration gradients that sharply define the sample are a product of the unsteady nature of the dynamic loading process, and would not be possible with previous steady-state loading methods. Centerline concentration profiles obtained during the dispensing of these samples are given in Figure 8.23. In each case, the left-most profile is that of the loaded sample, delineated in ' o ' symbols. Regarding the dispensed sample profiles (right-most profiles), the peak height of the samples injected by the reduced-focusing technique (Figures 8.23(a) and (b)) are shown to have decreased significantly from their original values. In contrast, the dynamically loaded sample (Figure 8.23(c)) has maintained much of its volume at the original sample concentration.
Figure 8.24. The measured iso-concentration contours of the dispensed samples corresponding to injections of Figure 8.23 (contours plotted at 0.1Co, 0.3Co, 0.5Co, 0.7Co, and 0.9Co).
Electrokinetic Sample Dispensing in Crossing Microchannels
507
high percentage of the volume contains the original sample concentration has two important implications. Firstly, the net concentration density of the sample is increased. Secondly, as concentration gradients exist only at the edges of the sample, the plateau region serves to forestall diffusive effects. This is in contrast to Gaussian shaped samples in which diffusion occurs immediately throughout the sample volume. The iso-concentration contours for the dispensed samples in Figure 8.23 are shown in Figure 8.24. Although the dynamically loaded sample (Figure 8.24(c)) may be too long for some applications, the degree of cross-stream and axial concentration uniformity achieved is noteworthy. The ability to dispense samples in which a high percentage of the volume contains the original sample concentration has two important implications. Firstly, the net concentration density of the sample is increased. Secondly, as concentration gradients exist only at the edges of the sample, the plateau region serves to forestall diffusive effects. This is in contrast to Gaussian shaped samples in which diffusion occurs immediately throughout the sample volume. To demonstrate the ability to inject very long concentration-dense samples, the dynamic loading of a millimeter-sized sample was performed. Because the dynamic loading period here, tDL = 0.8 s, was relatively long, the imaging system was able to capture the symmetric growth process. An image sequence of the process, at 4/15 s intervals is given in Figure 8.25(a). Centerline concentration profiles corresponding to the dynamic loading step, at 2/15s intervals, are shown in Figure 8.25(b). Imaging artifacts at the surface of the chip caused the noise visible on the far right-hand side of the later images and the final two concentration profiles. At its final size, the sample is over 0.9 mm long, of which 88% contains the original sample concentration. This large sample was chosen to demonstrate the methods potential usefulness in applications that previously employed offset twin-T configurations with gap lengths of 250 urn and 500 um. The dynamic loading method has two key advantages over twin-T injection methods. Firstly, the size of the sample is not coupled to the geometry of the chip, and can be varied simply by varying the potentials applied or the dynamic loading period, tpi. Secondly, the use of the dynamic loading method does not incur the sample dilution inherent in the steady-state focusing step in twin-T injections (where focusing buffer and sample run adjacently in the injection length). As a final note, when large samples are injected using the dynamic loading method, the pull-back voltage(s) applied during the dispensing step (to 'cut' the sample) can lead to sample length reduction (sample loss). This loss occurs while the sample loaded into the buffer channel 2, is drawn through the intersection and into the dispensing channel. Thus, to achieve the desired sample length in the dispensing channel, the dynamic loading period (or potentials) must be set with this taken into account.
508
Electrokinetics in Microfluidics
Figure 8.25. The injection of a millimeter-sized sample: (Top) an image sequence at 4/15 s intervals showing the dynamic loading and dispensing of the sample, and (Bottom) a centerline concentration profile sequence, at 2/15s intervals, of the dynamic loading process.
Electrokinetic Sample Dispensing in Crossing Microchannels
509
In summary, the dynamic loading method for discrete sample dispensing/injection in cross microchannel chips introduces a dynamic loading step between the steady-state loading and dispensing steps. Injected samples were shown to be more concentrated and more uniform in the cross-stream direction than traditional pinched-valve injections. Short dynamic loading times were shown to increase the peak height of short discrete samples. Long dynamic loading times resulted in samples exhibiting a constant concentration plateau at the original sample concentration, extending over the bulk of the sample. With a single cross microchannel, injections of well-defined samples with lengths varying from two channel widths (100 um), to twenty channel widths (millimeter sized) can be realized. In applications such as high-speed capillary zone electrophoresis, this injection technique may provide a desirable increase in signal with tolerable increase in sample length. This technique may also be preferred in many other applications where large samples are required.
510
8-4
Electrokinetics in Microfluidics
EFFECTS OF SPATIAL GRADIENTS OF ELECTRICAL CONDUCTIVITY
The dispensing processes discussed in the previous sections all use the same solution (the same chemical composition and concentration) as the driving buffer and the sample-carrying buffer, did not consider any spatial gradient of electric conductivity in the liquid, and assumed negligible effects of samples on the bulk conductivity. In many microfluidic systems, the gradient of electric conductivity exists in the liquid. For example, when a sample solution containing different sample species is required to be transported through microchannels without sample separation [21], a lower flow velocity in the sample zone is desirable during the loading process so that the electromigration and the separation can be minimized by the slow motion. This can be achieved by using a high conductivity buffer solution as the sample-carrying buffer (to have a lower electroosmotic velocity) and a low conductivity buffer solution as the driving buffer (to have a higher electroosmotic velocity). In such a case, the spatial gradient of electric conductivity exists. In other applications where a sample is required to be stacked in the system [22-25], the use of lower conductivity buffer solution in the sample region and higher conductivity buffer in the driving liquid region is a good choice to meet this need. This is because a higher velocity will be generated in the sample region and thus the sample can be stacked by the lower velocity buffer at the front of sample region. The objective of this section is to show the effects of spatial gradients of the electrical conductivity on the onchip microfluidic dispensing processes [13].
Figure 8.26. The schematic diagram of the crossing-microchannel.
Electrokinetic Sample Dispensing in Crossing Microchannels
511
A crossing microchannel dispenser is shown in Figure 8.26. The depth and the width of all the channels are 20 um and 50 um, respectively. There are four reservoirs connected to the four ends of microchannels, in which the electrodes are inserted to set up the electrical field across the channels. The sample reservoir is Rl, where a sample solution (the sample-carrying buffer) is initially loaded. The sample waste reservoir is R3, and the buffer focusing reservoirs are R2 and R4. Reservoirs 2 and 4 and the channels are filled with the driving buffer solution initially. The lengths of the channels from intersection to reservoir are 2.5 mm, 10 mm, 2.5 mm, and 10 mm for channels 1, 2, 3, and 4 respectively. The sample-carrying buffer and the driving buffer have the same chemical composition, but different ionic concentrations. When a set of chosen electrical potentials is applied to the reservoirs of the dispenser shown in Figure 8.26, the sample solution (the buffer solution carrying the sample) in reservoir 1 will be flow toward reservoir 3 passing through the intersection of microchannels and the driving buffer solutions in channel 2 and 4 are also driven to flow into the waste channel. Consequently, the sample solution is loaded into the intersection. Once the loading process reaches the steady state, the dispensing step is initiated by adjusting the applied voltages and the loaded sample in the intersection will be dispensed downstream towards reservoir 4. In order to consider the conductivity difference, the mathematical description of the loading and dispensing processes is different from that discussed in Section 8-1. A set of 2D governing equations describing the potential field, the flow field and the concentration field during the injection process will be introduced below. Electric potential field When an electric field is applied along the microchannel, the current is setup along the channel and the local electric current vector is given by: (21)
where / is local electric current density vector, u is the local velocity vector and 0 is the local electrical potential. The jth ionic species has a valence zj, a diffusion coefficient Dj and an ionic number concentration nj. Here, e, kj, and T represent the fundamental elementary charge, Boltzmann constant and the system absolute temperature, respectively. For large KO, channel, where the electric double layer is very thin, electro-neutrality is assumed to dominate in the channel and the first term of Eq. (21) defining the current transport due to
512
Electrokinetics in Microfluidics
convection can be neglected. Here a is the hydraulic radius and K is the Debye double layer thickness. Usually, one can neglect the second term of Eq.(21) which defines the current flow due to diffusion and is small as compared with the third term. Consequently, for large K a flows, one can write the electric current in terms of molar concentration as (22)
where Na is Avogadro number and the molar concentration is given by Cj = rij jNa . For a given electrolyte solution, Eq. (22) can be rewritten as (23) where X is the electric conductivity of electrolyte solution and takes the form of (24) Eq.(24) is an expression of Ohm's law for electrically neutral dilute solutions or solutions in a microchannel having large Ka, where C,- can be determined by a set of concentration equations. The charge conservation in the liquid has to be satisfied, which is described by: (25) Substituting Eq.(23) into Eq.(25), the equation of the electrical potential field is obtained as: (26) With the given concentration field and proper boundary conditions, Eqs. (24) and (26) can provide the distribution of the applied electric field, <j), in the crossing microchannel. It should be pointed out that if there is no spatial conductivity gradient, X is a constant, and hence Eq.(26) is reduced to: (26a)
Electrokinetic Sample Dispensing in Crossing Microchannels
513
Flow filed The flow field is described by the modified Navier-Stokes equation and the continuity equation as follows: (27)
(28) where V stands for the mass average velocity vector, p is the density of liquid, P is the pressure, JJ. is the viscosity of liquid, and pe is the net charge density in the solution. For high ionic concentration solutions commonly used in on-chip microfluidic dispensing processes, the electric double layer usually is very thin, i.e., less than 0.2% of channel width. The thin electric double layer results in a plug-like electroosmotic velocity profile cross the channel except in the double layer region where velocity dramatically decreases to zero at the wall. For simplicity, we may treat the electroosmotic velocity as the boundary slip velocity and use it as the boundary condition for the momentum equation, Eq.(27). Mathematically, this slip boundary condition removes the last term in Eq. (27) and includes its effects in the boundary condition: (29) where neo is the electroosmotic mobility of the buffer solution. Concentration field We consider that the spatial gradients of buffer concentration exist and the gradients are time dependent, which have significant effects on the electrical potential field and the flow field during the dispensing process. Therefore, the buffer concentration field must be modeled simultaneously at each time step. The concentration field is governed by the mass conservation and can be described by: (30)
(31)
514
Electrokinetics in Microfluidics
where Vepj is the electrophoretic velocity of the ith species, given by Vepi = Vepi^'t'' where fxepi is the electrophoretic mobility of this species. Please note that Eq.(30) and Eq. (31) describe the concentration fields of the sample and the buffer, respectively. Numerical scheme The complete set of equations, Eqs. (24), (26), (27), (28), (30) and (31), were normalized and numerically solved using the semi-implicit method for pressure-linked equation (SIMPLE) algorithm developed by Patankar [16]. The algorithm is based on a finite control volume discretization of the governing equations on a staggered grid. In order to capture all the detailed features near the four corners of the intersection shown in Figure 8.26, the non-uniform grid system is employed. The control volume size next to the wall is minimum. The size of successive control volumes away from the walls is increased by a factor of 1.2. In this implementation, the solution to this set of equations is obtained by an iterative procedure. During each iteration procedure, the discretized equations are solved by a line-by-line iteration method. Due to the unsteady loading and dispensing processes, the numerical calculation of the potential field, flow field and concentration field must be converged at each iteration step and each time step. The conductivity gradient effects analyzed by the model simulation In the calculations, Rhodamine 110 standard is used as a sample with an electrophoretic mobility - 2 . 0 x 1 0 " m /(V-s)
and a diffusion
coefficient
^buffer = 1 -0 x 10~ m / s , respectively, which are close to the measured value . The driving buffer and the sample-carrying buffer have the same chemical composites, sodium carbonate, the same pH value (pH = 9.0), the same diffusion 22
coefficient, D^uffer =1.0 x 10
m /s, but the different ionic concentrations and
hence different electrical conductivities. Two scenarios are studied here to investigate the effects of the conductivity gradient on the transport phenomena of the sample injection processes. In the first scenario, the driving buffer has a concentration of 10 mM and an electroosmotic mobility of 5.5 x 10~ m f(V • s). The concentration of the sample-carrying buffer is chosen to be 50 mM (5 times that of the driving buffer) and its electroosmotic mobility is assumed to be 5.0x10" m /(V-s). In this case, the driving buffer has a lower conductivity than that of the sample-carrying
Electrokinetic Sample Dispensing in Crossing Microchannels
515
buffer and the relationship of the conductivity between these two solutions is approximated as {Xsampie buffer)1(^driving buffer) = 5 - I n t h e second scenario, the driving buffer has a concentration of 50 mM and the electroosmotic mobility is assumed to be 5.0x10" m /(V-s). The sample-carrying buffer has a concentration of 5 mM (1/10 times of the driving buffer) and the electroosmotic mobility is assumed to be 5.56x10" m /{V • s). The relationship of the conductivity between these two solutions is approximated as (^sample buffer ) ^driving
buffer ) = 1/10 .
Figure 8.27. The effects of the conductivity difference on the applied electrical field in the dispenser. The iso-potential contours in the intersection in the steady-state loading step (left) and the corresponding dispensing step (right) for: (a) (^sampie buffer^1^driving buffer^5' <W ^sample
buffer^1 ^driving
buffer^1-
516
Electrokinetics in Microfluidics
If the spatial gradient of the electrical conductivity in the liquid is not considered, the electric potential distribution is not affected by the conductivity (mathematically, the conductivity is not involved in the applied electrical potential equation, such as Eq.(26a)) and only need to be solved once prior to solving momentum and concentration equations. This simplicity, however, is not valid in this case where the spatial gradient of the conductivity exists. The equation of the electric potential, Eq.(26), must be solved at each time step when the concentration distribution changes and in turn the conductivity distribution changes. Figure 8.27(a) shows the normalized electric potential distribution of the first scenario {{lsampie buffer) ^driving buffer) = 5 ) i n t h e loading and the dispensing step and Figure 8.27(b), however, shows the normalized electric potential distribution without the consideration of spatial gradient of conductivity (i.e., (^sampie buffer) Wdriving buffer) = 1) i n t h e loading and dispensing steps. In all the plots, the chip was oriented with channels from Rl, R2, R3, and R4, counterclockwise from left and the same configuration is used for all the other plots in this paper, except stated otherwise. The normalized potentials applied to the reservoirs 1 - 4 are the same for all the plots in Figure 2, sy -1.0, £j = 1.0, £3 = 0.0 and £4 =1.0 in the loading step and £j = 0.2, £2 = 2.0, £3 = 0.2 and £4 = 0.0 at the dispensing step, respectively, where e indicates the normalized applied voltages and subscripts correspond to reservoir numbers. The lines shown in these plots are equal potential lines. The potential drop between two lines is the same through all the plots and the electrical field strength can be evaluated by the ratio of the potential drop to the distance between the lines. The variation of the electric field strength is clearly demonstrated between the situations with and without the conductivity gradient. During the loading process, the sample solution is pumped into channel 1 from the reservoir 1. Figures 8.28(a) and (b) show the concentration distribution of the sample near the intersection for the two scenarios, ^sample buffer)^driving
buffer) = 5and {^sample buffer)/(^driving buffer) = 1/10,
respectively. In all the plots, the concentration increases from black to white and the iso-concentration lines are 10%, 30%, 50%, 70% and 90% of the original sample concentration. The sequence of plots shows the unsteady loading of the sample at different time. One can see that the conductivity gradient has significant effects on the unsteadying loading process. The time to reach the steady state is different for the two scenarios, tstea(jy=\2s for the case of (^sample buffer)^driving buffer) =5 a n d * steady =6s f o r t h e c a s e o f (^sample buffer)^driving buffer) =ino- T h a t indicates that the average bulk velocity is affected significantly by the spatial gradient of conductivity. In addition, the concentration contour of the sample is distorted backward to the
517
Electrokinetic Sample Dispensing in Crossing Microchannels
Figure 8.28. The effects of the conductivity difference on the loading process: the sequences of the sample concentration profile of the unsteady loading process for: (a) (^-sample buffer)'(^driving
buffer)^5
and
0>)
(^-sample buffer)^driving
buffer)=Vl(> •
The
concentration decreases from white to black and the iso-contours are plotted at 0.1, 0.3, 0.5, 0.7 and 0.9 of the original sample concentration
518
Electrokinetics in Microfluidics
sample reservoir (Rl) in the case of (lsamph buffer) ^driving buffer) = 1/10, as shown in the plot of t = 5s in Figure 8.28(b). This may be understood as follows. Because the conductivity of the sample region (the sample-carrying buffer and the sample) is much lower than that in the driving buffer region (-1/10), the voltage drop or the electric field strength in this region is bigger than that in the downstream of waste channel (from the intersection to right side reservoir 3). Additionally, the sample region has a lower ionic concentration and hence a higher electroosmotic mobility. The combination of the high electrical field
Figure 8.29. The sequences of the sample concentration profile of an unsteady loading process at Is intervals at the horizontal centerline (left) and the vertical centerline (right) for: (a) ^sample buffer)1 ^driving buffer) = 5 »a n d ( b ) ^sample buffer) W driving buffer)=l/i0 • The shown in the plots indicates the direction of time increase.
arrow
519
Electrokinetic Sample Dispensing in Crossing Microchannels
Figure 8.30. The effects of the conductivity difference on the loading and the dispensing processes under the same applied voltages. The loaded sample at steady-state (la-3a) and the corresponding dispensed sample at t = 1.25s (lb-3b) for: (1) ^sample buffer) ^driving 3
l
( ) ^sample buffer)'( driving
buffer)
= 5
> @) (^sample buffer) ^driving
buffer) = • •
buffer)=l/l°
.
an
d
520
Electrokinetics in Microfluidics
strength and the high electroosmotic mobility yields a much higher local electroosmotic velocity than that in the downstream. This intends to generate a higher flow rate in the upstream (sample-carrying buffer region) than the flow rate in the downstream (driving buffer region). For an incompressible liquid, the continuity condition requires the same flow rate throughout the microchannel. In order to achieve the same flow rate in both the upstream and the downstream, a negative pressure gradient (the pressure decreases in the flow direction) is induced in the downstream to increase the local flow rate, and a positive pressure gradient (the pressure increase in the flow direction) is induced in the sample region to decrease the local flow rate. This induced positive pressure gradient (the pressure increase in the flow direction) in the sample-carrying buffer region will add a backward pressure driven flow to the electroosmotic flow (plug-like velocity profile), resulting in the distorted backward velocity profile. Consequently, the sample concentration profile is distorted backward to reservoir 1 as well, because the sample concentration profile is dependent on the convection transport (bulk velocity). One can also find that for the second scenario ((^sample buffer) ^driving buffer) =l/l())> t n e sample is loaded slowly before t = 4s (i.e. there are no sample present at the intersection before this moment); however, the sample is suddenly pumped into the intersection when t = 5s and reaches the steady state at t = 6s. This is because the sample-carrying buffer has a lower conductivity than that of the driving buffer in this case. Initially, the entire channel is filled with the driving buffer solution (high conductivity and low resistance), which results in a low average velocity due to its low electroosmotic mobility. At the earlier stage of the loading process, when a small section of the sample channel (channel 1) is filled with the sample-carrying buffer (low conductivity and high resistance), the major potential drop still occurs in the region filled with the driving buffer and the average velocity is still low due to the fact that the liquid motion in most part of the microchannel is slow. However, when more and more of the sample-carrying buffer are pumped into the sample channel, the major potential drop occurs in the region filled with the sample-carrying buffer (low conductivity and high resistance), resulting in a high local electric field strength in this region. The combination of a high electric field strength and a high electroosmotic mobility (low ionic concentration) gives rise to a much higher cross-sectional average velocity than that at the earlier time. That is the reason that the sample is pumped slowly at the first 4 seconds and is quickly pumped into the intersection after that moment. It should be noted that the moment after which the sample is pumped much faster than before depends on the conductivity gradient. When the conductivity gradient changes, this time period changes.
Electrokinetic Sample Dispensing in Crossing Microchannels
521
In order to gain the insight into the transport phenomena behind the unsteady loading process, the sample concentration distribution at the centerline of the horizontal channel (from Rl to R3) and the vertical channel (from R2 to R4) are plotted in Figure 8.29 for both scenarios. The time between each line is 1 s and the direction of the time increase is indicated in the figure. Figure 8.29(a) shows the sample concentration at the centerline of horizontal channel (left) and vertical channel (right), for the scenario of (lsampie buffer) ^driving buffer) = 5> and Figure 8.29(b) shows that for the scenario of (^sample buffer) ^driving buffer) =1/10• F r o m Figures 8.29(a) and (b), one can see that the sample is gradually pumped into channel 1 ( 0 < x < 5 0 ) until a steady state is reached. A concentration drop is found at the inlet of the waste channel ( 5 1 < J C < 1 0 1 ) , which is due to the flow of the driving buffer solutions in channel 2 and 4 into the waster channel and the focus of the sample at the intersection. Also, the sample concentration at the vertical centerline is symmetric to the center point, which is because the same voltages are applied to R2 and R4 and channel 2 and 4 have the same length. It is also clearly shown that the sample is focused by the driving buffers at the intersection of microchannels (200 < j < 2 0 1 ) . Note that the extent to which the sample is focused depends on the combination of the four applied voltages and can be adjusted. A significant increase of the sample concentration with time is also found in the case of {Xsampie buffer)'(^driving buffer) = 1/10, as shown in the right plot of Figure 8.29(b). This is due to the speed of pumping the sample is much faster than earlier after a specific moment (i.e. t = 4s). This can be understood in the same way as explained previously for the phenomena that the pumping before t = 4s is very slow and the sample is pumped much faster after that moment. As discussed above, the loading processes in the presence of the spatial conductivity gradient is different from that without the spatial conductivity gradient. It can be expected that the corresponding dispensing processes under the same applied voltages will be different. Figure 8.30 shows the comparison of the loaded sample at the steady state (top row) and the dispensed sample at t = 1.255 (bottom row) between the two cases with the conductivity difference and the case without the conductivity difference. For all the three cases, the normalized applied potentials s\ ~s^ are 1.0, 1.0, 0.0, 1.0 in the loading step and 0.2, 2.0, 0.2, 0.0 in the dispensing step. In all the plots, the white region shows the sample solution with a concentration higher than 80% of the original sample concentration. Figures 8.30(la) and (lb) show the loaded sample and the dispensed sample for the case of (hsample buffer) ^driving buffer)l = 5> Figures 8.30(2a) and (2b) show that for the case of
522
Electrokinetics in Microfluidics
{^sample buffer) ^driving buffer) = 1 / 1 ° a n d Figures 8.30(3a) and (3b) show that for the case of (Xsampie buffer) ^driving buffer) = 1 • The sensitivity of the loaded sample size (Figures 8.30(1 a)—(3a)) and the dispensed sample size (Figures 8.30(1 b)—(3b)) to the spatial conductivity gradient is clearly demonstrated. There is essentially no sample present in the dispensing channel shortly after the dispensing for the situations with the conductivity gradient. This is because the loaded sample size in the intersection is small (Figures 8.30(1 a) and (2a)) and
Figure 8.31. The variation of the loaded sample's size and shape at the intersection for different conductivity difference and different applied voltages. The loaded sample at steadystate with (Xsampie buffer)^driving buffer)=i under the applied voltages, s1 ~ e 4 , of: (a) 1.0, 1.0, 0.0, 1.0, (b) 1.0, 0.2, 0.0, 0.2. The loaded sample at steady-state with ^sample buffer^'^driving buffed = 1 / 1 0 u n d e r t h e applied voltages, el ~ £ 4 , of: (c) 1.0, 1.0, 0.0, 1.0, (d) 1.0, 0.2,0.0,0.2.
Electrokinetic Sample Dispensing in Crossing Microchannels
523
the major part of the sample in the intersection is driven to the reservoir 1 and 3 when the dispensing is initiated (see Figures 8.30 (lb) and (2b)) due to the small potentials applied to reservoir 1 and 3 in the dispensing step (i.e. (ej = £3 = 0.2 compared to £2 = 2.0). As seen from Figures 8.30 (1) and (2), with a strongly focused sample in the loading step, one cannot dispense or inject any sufficiently large sample in the dispensing channel. In order to obtain a sufficiently large dispensed sample, the size of the loaded sample at the intersection must be large at the end of the loading process. The applied voltages were found to be the most important parameter controlling the size of the loaded sample and the dispensed sample. Figures 8.31 (a) - (b) show the effects of the applied voltages on the steady-state loaded sample for the case of {Xsampie buffer)^driving buffer) =5- Figures 8.31(c)-(d) is for the case of dsample buffed ^driving buffer>XI™• T h e normalized applied voltages, s\ - £ 4 , for Figures 8.31 (a) and (c) are 1.0, 1.0, 0.0, 1.0, and for Figures 8.3 l(b) and (d) are 1.0, 0.2, 0.0, 0.2, respectively. When the same normalized voltages applied to the reservoirs 2 and 4, £2 = £4, are decreased from 1.0 (Figures 8.31 (a) and (c)) to 0.2 (Figures 8.3 l(b) and (d)), the loaded sample increases significantly for both scenarios because of the less focusing effects from channel 2 and 4 on the loaded sample in the intersection. Figure 8.32 shows the dispensed sample at different time for the case of (^sample buffer) ^driving buffer) = 5 b a s e d o n t h e different sizes of the loaded sample. The normalized applied voltages in the dispensing process are 1.1, 2.0, 1.1, 0.0 for s\ ~ £4 in all the plots. Figure 8.32(a) shows a loaded sample, and the corresponding dispensed sample at t = 0.9s and t = 1.8s, respectively. Figure 8.32(b) shows a bigger loaded sample and the corresponding dispensed sample at t = 0.9s and 7 = 1.8s, respectively. One can see that, the bigger the loaded sample, the bigger the dispensed sample at the same dispensing time and under the same applied voltages. Indeed, the loaded sample size and hence the applied voltages in the loading step are critical in determining the dispensed sample size. After the loading process reaches a steady state, the dispensing process can be initiated by changing the applied voltages to the four reservoirs. During the dispensing process, the loaded sample in the intersection is expected to be driven or dispensed into channel 4 and be surrounded by the buffer solutions from both the upstream and the downstream, providing a cut of sample for the subsequent chemical and biomedical analysis. This requires that the main flow direction in the dispensing process should be from R2 to R4 (perpendicular to the loading flow direction from Rl to R3), and the flow in channel 1 and 3 could be either from Rl (or R3) to R4 helping the major flow from R2 to R4 to provide a big sample, or from R2 to Rl (or R3) splitting the major flow from R2 to R4 to
524
Electrokinetics in Microfluidics
Figure 8.32. The dependence of the dispensed sample size on the size of the loaded sample (or the loading voltages). The loaded sample at steady-state and the corresponding dispensed sample with the normalized applied voltages, E\ - £ 4 , of: (a) 1.0, 0.5, 0.0, 0.5 in the loading step and 1.1, 2.0, 1.1, 0.0 in the dispensing step, and (b) 1.0, 0.2, 0.0, 0.2 in the loading step and 1.1, 2.0, 1.1, 0.0 in the dispensing step.
Electrokinetic Sample Dispensing in Crossing Microchannels
525
provide a clear cut of sample in the downstream of channel 4. In order to perform this process, normally the voltage applied to reservoir 2 must be bigger than that applied to reservoir 1 and 3. Otherwise, there will exist a flow from Rl to R2 (R3 to R2 as well) besides the flow from Rl to R4 (R3 to R4 as well since R4 is grounded). This will split the loaded sample in the intersection to R2 and R4 and can not provide a cut of sample (for the purpose of the dispensing) because the sample is continuously driven to R2 and R4 from the sample reservoir (Rl). However, as we discussed earlier, when the voltages applied to reservoir 1 and 3 are much smaller than that to reservoir 2 during the dispensing, there will be too much loaded sample flowing to reservoirs 1 and 3. This will results in a small dispensed sample size. In order to verify this, the same loaded sample as shown in Figure 8.32(b) is dispensed using a new set of applied voltages s\ = 0.2,£ 2 = 2.0,£3 = 0.2,£4 = 0.0, where the applied voltages to reservoir 2 and 4 are the same as that used in Figure 8.32(b), however, the applied voltages to reservoir 1 and 3 are much smaller (i.e. s\ = £3 = 0.2) than that used in Figure 8.32(b) (i.e. ^ = £ 3 =1.1). We found that there is essentially no sample dispensed in the dispensing channel at t = 0.95 when using this new set of applied voltages {s\ =£3 =0.2) in the dispensing step. Since no sample is present in the downstream of channel 4, the simulation results are not shown here. This indicates that when the applied voltages to reservoirl and 3 are too low in the dispensing process, a significant amount of sample will be driven to reservoir 1 and 3 when the dispensing process is initiated. The left small amount of the dispensed sample easily diffuses into the buffer solution in channel 4, consequently, no sample with a sufficiently high concentration appears. Therefore, in addition to the loaded sample size, the combination of the applied voltage in the dispensing process is another very important controlling parameter for the dispensed sample size. From the above discussions, it is clear that the presence of the conductivity gradient has a strong effect on the sample loading and dispensing. The applied voltages can control the shape and the size of the loaded sample and hence the shape and the size of the dispensed sample. For a given sample-buffer system, there exists an optimal set of the controlling voltages in order to obtain a desired dispensed sample size. The model presented here is able to find such a set of the optimal voltages through the numerical simulation.
526
8-5
Electrokinetics in Microfluidics
CONTROLLED ON-CHIP SAMPLE INJECTION, PUMPING, STACKING WITH LIQUID CONDUCTIVITY DIFFERENCES
The net velocity of an individual ionic species in an electroosmotic flow is a combination of the bulk fluid velocity and the specific electrophoretic velocity of the species. An initially discrete multi-component sample can therefore be separated into bands based on the charge-to-mass ratios of the individual species. The rate of separation of the sample can be changed using the electrical conductivity difference between the sample and the running buffer. Although the sample and the choice of buffer are typically determined by the application, the relative buffer concentrations can often be varied (to some extent) to enhance or inhibit separation. An understanding of these systems is also required in less ideal cases where extremely high or extremely low sample concentrations are necessitated by the application. Conductivity differences alter the shape of the electric field, which is the driving force behind both the electroosmotic bulk flow and the electrophoretic velocity of individual species. In most Lab-on-a-chip applications the channel length greatly exceeds the sample length, and thus the sample velocity is dictated primarily by the electroosmotic flowrate in the running buffer (regardless of the electroosmotic velocity developed at the sample/wall interface). Thus the bulk fluid velocity developed by electroosmosis is constrained by the conservation of mass requirement. The electrophoretic velocity of a charged species in the sample, however, is not constrained, and responds directly to the local electrical field gradient. The two cases of interest are illustrated in Figure 8.33. The pumping case, in which sample separation is reduced, is illustrated in Figure 8.33(a). In this case, the sample conductivity, ASj is higher than that of the running buffer, Ao and the electric field strength in the sample, Es, is lower than in the running buffer, Eo. The result is reduced electrophoretic velocities in the sample region, and hence a reduced rate of separation. Although the electroosmotic velocity of the sample, veo.s, is likewise reduced, pressure forces induced by the comparatively high running buffer electroosmotic velocity, veo.o, pull the sample along. The opposite is observed in the stacking case (field amplified stacking), illustrated in Figure 8.33(b). In the stacking case, separation in the sample is enhanced using a sample with a lower electrical conductivity than that of the running buffer. The increased electric field leads to high electrophoretic velocities in the sample region. The mean velocity of the sample, however, is constrained by the comparatively slow electroosmotic velocity in the running buffer. This combination results in rapid separation of species in a relatively slow moving sample. A key aspect of the stacking case is that it is possible to obtain peaks with concentrations of charged components higher than in the original sample solution. Increasing the peak
Electrokinetic Sample Dispensing in Crossing Microchannels
527
Figure 8.33. Schematic diagram of pumping (high conductivity sample transport) and stacking (low conductivity sample transport) cases. In the pumping case, the electric field strength, and electrophoretic velocities are reduced in the sample region. In the stacking case, the electric field strength and electrophoretic velocities are increased in the sample region.
528
Electrokinetics in Microfluidics
height allows for more accurate detection, and in some cases extends the analytical capabilities of microfluidic chips. To achieve significant sample concentration for either pumping or stacking, many studies have employed larger samples [26-30], in contrast to focused samples employed for direct separation by capillary zone electrophoresis [5,6]. Stacking on chips was first demonstrated by Jacobson and Ramsey [26] using a gated injection scheme. Due to the circulation generated (see Figure 8.33a), they suggested that a compromise between stacking enhancement (due to either increased conductivity difference or sample size) and separation efficiency must be reached. Haab and Mathies [28] and Vazquez et al. [29] employed offset twin-T chip configurations where the sample and buffer run adjacently in the dispensing channel for a short distance. Unfortunately, the running of buffer and sample together in this way tends to dilute the sample. To avoid this, Lichtenberg et al. [30] developed a new injector (essentially an offset twincross). Although more complicated, the new configuration was shown to successfully resolve the adverse effects of sample pinching in the offset twin-T. Maintaining sample purity is of particular importance because the degree of stacking or pumping achievable is critically dependent on the difference in conductivity of the injected sample relative to that of the running buffer. Any dilution of the sample during injection results in a reduction in the conductivity difference and a decrease in performance [30]. A new three-step technique [12] for discrete sample injection in straightcross microfluidic chips was discussed in Section 8-3. The technique has an intermediate dynamic loading step in which sample is pumped directly into the intersection and three connecting channels. A key feature of this technique, especially in the context of pumping and stacking, is the ability to inject welldefined samples of high concentration density. Another key feature of this loading technique is that samples of any axial length can be injected, and in contrast to offset twin methods [28-30] the sample size is not coupled to the chip geometry. The dynamic loading method has been demonstrated in Section 8-3 using only conductivity-matched liquids [12]. In this section, we will present applications of this dynamical loading method in on-chip injection and electrokinetic transport of samples differing in conductivity from the running buffer by fluorescence-based visualization [14]. A sample containing both a neutral dye (rhodamine-B) and negatively charged dye (fluorescein) is employed. While the charged dye responds to the field gradients induced by the conductivity difference between the sample and running buffer, the neutral dye simply tracks the location of the original sample buffer. Both electroosmotic pumping and stacking cases are studied. The coupled flow phenomena inherent in these situations are investigated and discussed in the context of microfluidic chip applications.
Electrokinetic Sample Dispensing in Crossing Microchannels
529
The glass chips used here were manufactured by Micralyne (Edmonton, Canada). The crossing microchannels are 20 |am deep (at the center) and 50 urn wide (across the top), and the reservoirs are 2 mm in diameter. The D-shaped channel cross-section is a product of the manufacturing technique. The sample was loaded into reservoir Rl, the sample waste reservoir was R3, and the buffer focusing reservoirs were R2 and R4. The lengths of the channels from the intersection to the reservoir are 5 mm, 4 mm, 80 mm, and 4 mm for channels 1, 2, 3, and 4 respectively. The chip was clamped to a precision 3-axis stage. The two horizontal positioning axes were used to position the field of view, and the vertical axis to focus the microscope. The chips were prepared by rinsing with each of the following solutions for 20 minutes (in sequence): 1M filtered Nitric acid; 1M filtered Sodium hydroxide (NaOH); and filtered buffer. To pull these solutions through each channel, a vacuum was applied to R3 using a 30 mL syringe. A tapered 1000 ul pipette tip was cut to the appropriate diameter and friction fitted to join the syringe to the chip. The fluorescent dyes employed here were rhodamine-B (478.68 MW) as supplied by Fisher, and fluorescein (332.31 MW) as supplied by Molecular Probes. The dyes were dissolved in sodium carbonate buffer of pH = 9, also used as the running buffer. Immediately before use, all solutions were filtered using 0.2 ^m pore size syringe filters. Rhodamine-B at pH = 9 was found to be neutral. Fluorescein, however, carries a valence charge of z = -2 at pH = 9.0, with an electrophoretic mobility of vph F = -3.3x10"8 mV^s" 1 . The excitation source (488 nm) is well suited to excite fluorescein with an adsorption maximum at 490 nm, and less suited to excite rhodamine-B with an adsorption maxima at 570 nm. The result is a consistent, albeit less efficient, excitation of the rhodamine. The dye concentration loadings were adjusted to compensate for that effect somewhat, although the peak height of the fluorescein was intentionally made higher than that of the rhodamine. The dye concentrations, sample and buffer ionic strengths, employed in each run are summarized in Table 1. The electrical conductivities of the 10, 25, and 50 [mM] ionic strength buffer solutions were estimated to be A = 0.04, 0.10 and 0.20 [S/m]. Liquid temperatures in the chip were maintained at the ambient temperature (25 °C). A fluorescent epi-illumination video microscope was employed to visualize the process. The dye was excited by a continuous flood of blue light provided by a single-line, 200 mW, 488 nm argon laser (American Laser Corp.), through a 32x, NA = 0.3 microscope objective (Leitz). The received signal was split by the dichroic mirror (510 nm LP/ short-reflecting) and passed through an additional filter (515 nm LP) and a camera mount with 0.63x magnification before reaching the camera. The three-step process requires independent control of three potential levels for each of the four electrodes (12 settings). Here, a high voltage power
530
Electrokinetics in Microfluidics
supply (Glassman PS/EH20R02.0) was used in conjunction with a custom-made voltage controller. The timing of each switch was controlled via a digital delay generator (Stanford Research Systems). The voltage controller was programmed for default outputs corresponding to the steady-state loading step. Rising edge triggers from the delay generator commenced dynamic loading and dispensing steps. Switch frequencies up to 200 Hz were verified, indicating a transition delay of under 2.5 ms. Since the transition delay after each trigger is expected to be consistent, the critical dynamic loading period, tDL, between triggers can be specified with considerably less uncertainty (estimated here < 1 ms).
Figure 8.34. Iso-intensity coutour plots during the dispensing of a two-analyte sample with: (a) a relatively high conductivity sample (pumping case, y = 0.2), (b) a sample with conductivity matching the running buffer (y = 1), and (c) a relatively low conductivity sample (stacking case, y = 5). No dynamic loading was employed. Contours are plotted at 5 even intervals.
Electrokinetic Sample Dispensing in Crossing Microchannels
531
Images were captured and saved on the computer at a rate of 15Hz. A progressive scan CCD camera (Pulnix, TM-9701) was used to avoid image defects due to field-field interlacing. The acquired images had a resolution of 640x484 pixels and an 8-bit dynamic range. This pixel matrix corresponded to a viewed region of 550x416 ^m. The camera orientation was carefully adjusted before each run such that the pixel grid was aligned with the coordinate directions of the intersection. Digital image processing was performed to remove any non-uniformity present in the imaging system and to relate the pixel intensity values to dye concentration. A background noise signal was subtracted, and brightfield image normalization was performed for each image.
Figure 8.35. Iso-intensity contour plots of the injection of a two-analyte sample with conductivity matched to that of the running buffer (y = 1): (a) the steady-state loading step (t = 0 s), (b) the dispensing step (t = 4/15 s), and (c) later in the dispensing step (t = 8/15 s). No dynamic loading was employed. Contours are plotted at 10 even intervals.
532
Electrokinetics in Microfluidics
These images were then smoothed with a distance-based kernel, and scaled by a single factor such that the image series filled the grayscale range. The period over which the dynamic loading voltages are applied is termed the dynamic loading period, tDL. For a given solution/chip combination and electrode potential settings, the dynamic loading period determines the axial length of the injected sample. The degree of stacking or pumping is dependent on both the initial size of the sample, and the conductivity of the running buffer relative to that of the sample buffer, y. The dye concentrations, sample and running buffer ionic strengths, and normalized electrode potentials employed in each case are summarized in Table 1. The effect of varying the conductivity difference (keeping initial sample size approximately constant) is shown in Figure 8.34. Iso-intensity profiles obtained during the dispensing of a two-dye mixture are shown for each of the three cases: pumping, uniform conductivity, and stacking (7 = 0.2, 7 = 1 , and 7 = 5 respectively). No dynamic loading was used, and the injected samples were of similar initial size. The two dyes in the relatively high conductivity sample (7 = 0.2, Figure 8.34a) do not separate, and in that sense, the sample was effectively 'pumped'. However, the sample is severely distorted by mismatch in electroosmotic velocities (see Figure 8.34a), and non-uniformity in the crossstream direction due to the nature of the injection. The effects of both these factors can be mitigated using larger initial sample lengths, as will be shown later. In contrast to the pumping case, the two dyes separate readily in the uniform conductivity case ( 7 = 1 , Figure 8.34b). The two dyes are even more rapidly separated in the relatively low conductivity sample (7= 5, Figure 8.34c). The fluorescein band in this case is thinner than that in the uniform conductivity case due to field amplified sample stacking. Figure 8.34 highlights the very significant role conductivity differences play in electrokinetic sample transport. Although the effectiveness of sample transport (pumping case) and sample separation (stacking case) are shown to improve somewhat over that of the uniform conductivity case, the injections shown in Figure 8.34 are far from ideal. The underlying phenomena and the effect of initial sample length for each case (uniform conductivity, y = 1, sample stacking, y = 5, and sample pumping, y = 0.2) will now be discussed in turn. When the conductivity of the sample matches that of the running buffer (y = 1) neither sample stacking or pumping occurs. In such cases, individual analytes in the sample move independent of the original sample-buffer solution at a velocity equal to the summation of the bulk electroosmotic velocity and their specific electrophoretic velocity. Upon dispensing, neutral rhodamine and negatively charged fluorescein are separated readily as shown in the sequence of iso-intensity contour plots in Figure 8.35. Regarding the size of the analyte bands, the fluorescein band remains in the intersection much longer than the
Electrokinetic Sample Dispensing in Crossing Microchannels
Table 1 Run parameters for each case studied. maximum applied voltage.
533
All potentials listed were normalized with the
Uniform Conductivity
Pumping Case
Stacking Case
Figures
4b, 5,6
4a, 7,8
4c, 9, 10
Relative Conductivity, y
1
0.2
5
Rhodamine-B Concentration [uM]
100
100
200
Fluorescein Concentration [uM]
25
25
50
Running Buffer Ionic Strength [mM]
25
10
50
Sample Buffer Ionic Strength [mM]
25
50
10
Maximum Voltage Applied, Vmax [V]
1415
1415
1415
Normalized Reservoir Potentials: Loading Step 1
1.000
1.000
1.000
2
1.000 0.911
1.000
3
0.000 0.000 0.000
4
1.000 0.911
1.000
Dynamic Loading Step: 1
0.964 0.963 0.964
2
0.707 0.708 0.710
3
0.000 0.000 0.000
4
0.707 0.708 0.710
1
0.014 0.049 0.008
2
0.124 0.127 0.067
3
0.000 0.000 0.000
4
0.000 0.000 0.000
Dispensing Step:
534
Electrokinetics in Microfluidics
rhodamine band. This positioning results in a loss of fluorescein into the sample and sample waste channels due to the application of pull-back (or equivalently 'cut-off) voltages. When a dynamic loading step was applied, the sample stream was electroosmotically pumped directly into the dispensing channel. The effect of dynamically loading the same sample mixture as employed previously (Figure 8.35), under the same conditions (y = 1) is shown in Figure 8.36. Images with samples at similar locations were chosen from the dispensing step of each run and processed identically. Dynamic loading is shown to increase the peak height of both analytes. As dynamic loading time was increased, however, there was little increase in the peak height of the rhodamine band as it approached a plateau concentration value equal to that of the original sample stream. The dynamic loading step, where neutral ions travel much faster in the dispensing channel than anions, amplifies the injection bias. This bias was evidenced by the larger
Figure 8.36. Centreline axial intensity profiles obtained during the dispensing of a twoanalyte sample with conductivity matched to that of the running buffer (y = 1). Profiles for dynamic loading times of t DL = 0 s, 0.1 s, and 0.2 s are superimposed with corresponding isointensity contour plots inset.
Electrokinetic Sample Dispensing in Crossing Microchannels
535
increase in rhodamine bandwidth (compared to that of the fluorescein) accompanying a dynamic loading period of just 0.1 seconds. In this uniform conductivity system, the benefit of increased peak height obtained with dynamic loading is at least partially offset by the cost of delayed separation due to increased band length. To facilitate the electroosmotic pumping case for reduced separation, a fluorescein/rhodamine sample mixture was prepared with buffer five times more conductive than the running buffer, y = 0.2 (run details in Table 1). The dynamic loading of the high conductivity sample is shown in Figure 8.37. In order to pull the high conductivity sample to the intersection, the focusing potentials in the steady state loading step were decreased from that of the previous run. In the
Figure 8.37. An image sequence of the three-step dynamic loading injection process of a two-analyte sample of higher conductivity than the running buffer (pumping case, y = 0.2). The dynamic loading time was t D L = 0.3 s.
536
Electrokinetics in Microfluidics
dynamic loading step, electroosmotic flows of the running buffer pulled the comparatively slow sample stream into the dispensing channel. Upon application of the dispensing voltages, the sample takes on a curved shape in the dispensing step, characteristic of the pumping case (see Figure 8.33a). No separation was observed in the limited field of view shown in Figure 8.37c. This result is in contrast to the rapid separation observed in the y = 1 case (Figure 8.35). To investigate the affect of initial sample length in the electroosmotic pumping case, the microscope was repositioned 550 |j,m (11 channel widths) downstream of the intersection in the dispensing channel. The central pixel in the CCD array was chosen as a point fluorescence detector, and the signal
Figure 8.38. The signal recorded (at a location 550 (j.m downstream of the intersection) over time through a series of seven injections. The sample contains two-analytes and has a higher conductivity than the running buffer (pumping case, y = 0.2). The first injection employed no dynamic loading (t DL = 0 ) , the following injections employed increased dynamic loading times and the final injection was a repeat of the no dynamic loading case. Images of the dynamically loaded samples at the point of detection are shown as the inset.
Electrokinetic Sample Dispensing in Crossing Microchannels
537
recorded over time through a series of seven injections is shown in Figure 8.38. The first injection employed no dynamic loading (tDL-0), the following injections employed increased dynamic loading times and the final injection was a repeat of the no dynamic loading case. Note that the although the scale of the time axis is accurate, detection was paused after each dispensing step to allow the analytes to return to the intersection and fill into the sample waste reservoir. In the first and last case, which employed the shortest samples, the sample had begun to separate into two bands, the rhodamine peak arriving at the detector before the fluorescein peak. With a dynamic loading step of tDL= 0.1 s, the peak height more than doubled, although the sample is shown to be fronting in time (or 'trailing' in space). This fronting is primarily due to cross-stream diffusion in the wake of the convex sample shape, not to be confused with fronting due to mobility differences between the analyte and that of the running buffer co-ion. Due to diffusion, the characteristic shape is only barely noticeable in the inset images of the samples as they crossed the detection point in Figure 8.38. As the initial sample length was increased (by increasing tDL), the peak height increased greatly. The band length, however, was not similarly increased. Although the tDL= 0.8 s sample was the largest sample injected, it was the sample with the smallest band length at the detector, and thus the most effectively 'pumped'. This is because in multi-component, high conductivity samples, the separation rate is coupled to the concentration density of the sample buffer. Dilution occurs at the sample ends due to an internal circulation generated by the mismatch in electroosmotic mobilities between the sample and the running buffer (Figure 8.33a). The rate at which that dilution affects the average concentration of the sample is a function of sample length. Thus, larger samples experience a reduced rate of overall change than shorter samples under similar circumstances. By maintaining the conductivity difference, larger samples maintain lower internal electrical potential gradients and hence experience less broadening due to the separation of analytes. The similarity of the first and last (tDL =0) runs verifies that the increase in pumping effectiveness observed, evidenced in an 8-fold increase in peak height and a decrease sample length, was attributable to the dynamic loading of the high conductivity sample. To facilitate the electroosmotic stacking case for enhanced sample separation, a fluorescein/rhodamine sample mixture was prepared with buffer one-fifth as conductive as the running buffer, y = 5 (run details in Table 1). In Figure 8.39, iso-intensity contour plots and centerline intensity profiles are plotted at 4/15 s intervals during the dynamic loading and separation process. In the steady state loading step (Figure 8.39a), the sample is transported to the intersection by the relatively high electroosmotic velocity in the sample channel,
538
Electrokinetics in Microfluidics
Figure 8.39. Plot sequence of the dynamic loading and dispensing of a two-analyte sample of lower conductivity than the running buffer (stacking case, y = 5). Iso-intensity contour plots at 4/15 s intervals are shown at left, and corresponding centerline axial intensity profiles are shown at right. The dynamic loading time was t DL = 0.2 s. Contours are plotted at 5 even intervals.
Electrokinetic Sample Dispensing in Crossing Microchannels
539
and experiences a relatively high electrical field strength. With application of the dynamic loading potentials, the sample pumps into the dispensing channel for tDL= 0.2 s (Figure 8.39b). The sample entering the dispensing channel is noticeably depleted in fluorescein. This depletion is due to the high electrical field strength and reduced bulk flow in the sample stream. Essentially, the sample-stacking mechanism is being applied to the whole sample channel, and fluorescein is being 'stacked' toward the sample reservoir. Also apparent is an M-shaped centerline intensity profile in Figure 8.39b at right. It is suspected that during the steady-state loading step, the fluorescein concentration was increased near the intersection as was found by Jacobson [5]. This local high concentrated portion was split during the dynamic loading step forming the M-shaped profile. With application of the dispensing potentials (Figure 8.39c), the fluorescein in the sample rapidly stacked to the rear (left) of the sample, creating a large peak in intensity. Because of the symmetry of dynamic loading, the stacked peak is formed to the left of the intersection (x < 0), partially insulated from the sample and sample waste channels. In contrast, the neutral rhodamine did not stack, marking the location of the low conductivity sample buffer. In addition, the front of the rhodamine band exhibited the concave shape characteristic of the
Figure 8.40. Centreline axial intensity profiles obtained during the dispensing of a twoanalyte sample of lower conductivity than the running buffer (stacking case, y = 5). Profiles for dynamic loading times of t D L = 0 s, 0.1 s, 0.2 s, and 0.3 s are superimposed.
540
Electrokinetics in Microfluidics
stacking case (Figure 8.33b). As the fluorescein and the rhodamine further separate, some fluorescein was consumed by the sample and sample waste channels due to the pull-back voltages applied. Although the analytes were totally separated in Figure 8.39e, the two separate bands were distinguishable almost immediately after the application of the dispensing potentials (Figure 8.39c). This rapid separation is due to the high field strength developed in the sample. To investigate the role of dynamic loading in sample stacking, injections with varying dynamic loading times were performed. Images from each run (taken at a similar time period following the application of the dispensing potentials) were processed identically. The superimposed centerline intensity profiles are shown in Figure 8.40. Being neutral, rhodamine behaved here as in the uniform conductivity case, reaching a plateau concentration value equal to that of the original sample stream. The peak height and band length of the fluorescein band, however, are shown to increase significantly with dynamic loading. A dynamic loading period of tDL= 0.3 s resulted in a three-fold increase in fluorescein peak height over that produced through sample stacking without dynamic loading. Also, as the dynamic loading time is increased the sample length increases, moving the resulting location of charged analyte stacks further from the intersection. Imposing this distance has the benefit of insulating the analyte from the sample and sample waste channels, further increasing peak height. In summary, the electrical conductivity differences between sample and running buffer streams can greatly influence the transport of individual analytes in electrokinetically driven microfluidic systems. The two situations discussed in this section are sample pumping (where bulk transport is increased and separation of charged analytes is delayed using a relatively high conductivity sample), and sample stacking (where bulk transport is decreased and separation of charged analytes is expedited using a relatively low conductivity sample). It was shown that by employing the conductivity differences alone, the effectiveness of either sample transport or sample separation was improved over the uniform conductivity case. Then it was demonstrated that increasing the sample length, through dynamic loading, further increased the effectiveness of sample pumping, evidenced in an eight-fold increase in peak height as well as a decrease in total sample length at a downstream detector. Dynamic loading the in sample stacking case was shown to increase peak height (three-fold) in rapid separations. The success of these applications of dynamic loading was attributed to the ability to inject concentration dense samples of any length. Although these processes would benefit from a formal optimization, it was demonstrated that the dynamic loading technique used in conjunction with strategic conductivity differences significantly extends the capabilities of microfluidic chips.
Electrokinetic Sample Dispensing in Crossing Microchannels
541
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30]
D. Harrison, A. Manz, Z. Fan, H. Ludi and H. Widmer, Anal. Chem., 64 (1992) 19261932. K. Seiler, D. Harrison and A. Manz, Anal. Chem., 65 (1993) 1481-1488. C. Effenhauser, A. Paulus, A. Manz and H. Widmer, Anal. Chem., 66 (1994) 2949-2953. A.J.P. Martin and F. Everaerts, Proc. Roy. Soc. London, 316 (1970) 493. S.C. Jacobson and J.M. Ramsey, Anal. Chem., 69 (1997) 3212-3217. S.C. Jacobson, C.T. Culbertson, J.E. Daler and J.M. Ramsey, Anal. Chem., 70 (1998) 3476-3480. S.V. Ermakov, S.C. Jacobson and J.M. Ramsey, Anal. Chem., 70 (1998) 4494-4504. S. Ermakov, S. Jacobson and J.M. Ramsey, Anal. Chem., 72 (2000) 3512-3517. N.A. Patankar and H. Hu, Anal. Chem, 70 (1998) 1870-1881. L. Ren and D. Li, J. Colloid Interface Sci., 254 (2002) 384-395. D. Sinton, L. Ren and D. Li, J. Colloid Interface Sci., 260 (2003) 431-439. D. Sinton, L. Ren and D. Li, J. Colloid Interface Sci., 266 (2003) 448-456. L. Ren, Ph.D. thesis, University of Toronto, 2004. D. Sinton, L. Ren, X. Xuan and D. Li, Lab on Chip, 3 (2003) 173-179. P.H. Paul, M.G. Garguilo and D.J. Rakestraw, Anal. Chem, 70 (1998) 2459-2467. S. V. Pantakar, Numerical Heat Transfer and Fluid Flow, Hemisphere Publishing Corp, New York, 1980. S.R. Wallenborg and C.G. Bailey, Anal. Chem, 72 (2000) 1872-1878. S.C. Jacobson, L.B. Koutny, R. Hergenroder, A. M. Moore Jr., J.M. Ramsey, Anal. Chem, 66 (1994) 3472-3476. S. Liu, S. Yining, W.J. William and R.A. Mathies, Anal. Chem, 71 (1991) 566-573. C. Zhang and A. Manz, Anal. Chem, 73 (2001) 2656-2662. L. Bousse, C. Cohen, T. Nikiforove, A. Chow, A.R. Kopf-Sill, R. Dubrow and J.W. Parce, Annu. Rev. Biophys. Biomol. Struct, 29 (2000) 155-181. D.S. Burgi and R. Chien, Anal. Chem, 63 (1991) 2042-2047. W. Thormann, C.X. Zhang, J. Caslavaska, P. Gebauer and R.A. Mosher, Anal. Chem, 70 (1998) 549-562. J. Olgemoller, G. Hempel, J. Boos and G. Blaschke, J. Chromatography B, 726 (1999) 261-268. J.H. Lee, O.K. Choi, H.S. Jung, K.R. Kim and D.S. Chung, Electrophoresis, 21 (2000) 930-934. S.C. Jacobson and J.M. Ramsey, Electrophoresis, 16 (1995) 481-486. J.P. Kutter, R.S. Ramsey, S.C. Jacobson and J.M. Ramsey, J. Microcolumn Separations, 10(4) (1997) 313-319. B.B. Haab and R.A. Mathies, Anal. Chem, 71 (1999) 5137-5145. M. Vazquez, G. McKinley, L. Mitnik, S. Desmarais, P. Matsudaira and D. Ehrlich, Anal. Chem, 74 (2002) 1952-1961. J. Lichtenberg, E. Verpoorte and N.F. de Rooij, Electrophoresis, 22 (2001) 258-271.
542
Electrokinetics in Microfluidics
Chapter 9
Electrophoretic motion of particles in microchannels Consider a charged particle suspended in a bulk liquid. When an electrical field is applied, the particle will move in the liquid towards either the cathode or the anode depending on the sign of the surface charge of the particle. Such a particle motion in a stationary liquid phase is called the electrophoresis. Electrophoresis is a major subject in colloid and interface science, because it is one of the most widely used separation techniques in various engineering and science applications [1,2]. Electrophoretic motion of rigid particles in unbounded aqueous electrolyte solutions has been investigated extensively, and mathematical models have been developed to describe this phenomenon in detail. Excellent summaries can be found elsewhere [3-5]. In the electroosmosis, the solid — the channel wall — is stationary, the liquid is moving under an applied electrical field. The electrophoresis is the reversed process: the liquid is stationary, the solid - the particle - is moving under the applied electrical field. The physical nature of these two processes is the same. Therefore, some analyses of electroosmosis can be applied to the electrophoresis. For example, in Chapter 4-1, we derived the electroosmotic flow velocity for the case of thin electrical double layer (i.e., KCI = CI/(1/K) is large) as, (1) where the £waU is the zeta potential of the channel wall. Applying Eq.(l) to the electrophoresis, we have the particle's electrophoretic velocity as: (2) where C,p is the zeta potential of the particle. Eq.(2) is usually referred to as the Smoluchowski equation. In the literature, the electrophoretic mobility is defined as:
Electrophoretic Motion of Particles in Microchannels
543
(2a) By the definition, it is the electrophoretic velocity per unit applied electrical field strength, characterizing how fast a particle will move in an electrical field. The mobility is proportional to the zeta potential of the particle. For a more general treatment, we can derive the electrophoretic velocity by balancing the electrical force and the flow frictional force acting on the particle. The flow frictional force is given by the Stokes equation:
where a is the radius of a spherical particle. The electrical force is given by:
where Q is the total charge of the spherical particle. By using Eq.(34) in Chapter 2,
the force balance will give us: (3) When Ka « 1, Eq. (3) is reduced to (4) Apparently, Eq.(2) and Eq.(4) are different, Eq.(2) has a constant 1 and Eq.(4) has a constant 2 / 3 . The difference reflects the consideration of the socalled electrophoretic retardation effect. Realizing the existence of the EDL around the particle, the action of the applied electrical field on the excess counterions in the EDL region will generate electroosmotic flow. For example, consider a negatively charged particle. Its electrophoretic motion will be towards the positive electrode. The excess counterions in the particle's EDL region are positive ions. Under the same applied electrical field, the electroosmotic flow of
544
Electrokinetics in Microfluidics
these positive ions will be the direction towards the negative electrode. Thus, the electroosmosis in the EDL region of the particle will cause a reduction in the velocity of the particle's electrophoretic motion. This is called the electrophoretic retardation effect. The underline assumption in the Smoluchowski equation, Eq.(2), is that the electroosmosis is the dominant force and the particle's motion is equal and opposite to the liquid motion. That is why the constant in Eq.(2) is 1. Eq.(4), however, is valid only for very small Ka values (or very thick EDL) and considers that the main retardation force to the electrophoresis is the flow frictional force. That is why the constant in Eq.(4) is - and not 1. A more general formulation, the Henry's equation,
(5)
unites the Eqs.(2) and (4). Here /(red) is the Henry's function, and it approaches 1 for small Ka and 3/2 for large Ka. The specific expressions of the Henry's function can be found elsewhere [3,4]. In addition to the above-mentioned retardation effect, the particle's electrical conductance and the particle shape will determine the local distortion of the applied electrical field due to the presence of the particle. Interested readers can find the related information from the literature [3,4]. Electrophoretic motion of rigid particles in unbounded aqueous electrolyte solutions has been studied extensively. In contrast, much less work has been done on the boundary effects on the electrophoretic motion of a particle. The boundary effects on the electrophoretic motion of particles need to be considered in cases where a particle moves in a microchannel whose size is close to the particle's size (e.g., electrophoretic motion of a protein through a porous membrane, electrokinetic transport of biological cells and bacteria in microchannels). With the emergence of bio-lab-on-chip technology, electrophoretic separation of particles in microchannels is often required. Therefore, studies of the boundary effects on the electrophoretic motion of particles are important. Keh et al. [6], Ennis et al. [7] and Shugai et al. [8] examined boundary effects on electrophoretic motion of a sphere for the following cases: a sphere near a non-conducting planar wall with an electric field parallel to the wall; a sphere near a perfectly conducting planar wall with an electric field perpendicular to the wall; and a sphere on the axis of a cylindrical pore with an electric field parallel to the axis. Ennis, et al. [9] experimentally investigated the electrophoretic mobility of proteins in membrane. They found that the protein mobility is identical to the free protein mobility when the size of the protein is
Electrophoretic Motion of Particles in Microchannels
545
small relative to that of the pore, and the mobility is significantly reduced as the pore radius approaches the protein radius. In addition, Zydney [10] and Lee, et al. [11] investigated boundary effects on the electrophoretic motion of a charged spherical particle in a spherical cavity. Their results indicate that the boundary effect is weak for thin double layers but significant for thick double layers. They also found that the charge on the boundary alters the particle motion through the development of an induced charge on the particle and through the generation of an electroosmotic recirculation flow in the spherical cavity. In this chapter, we will discuss how to model and simulate the electrophoretic motion of particles in microchannels filled with an aqueous electrolyte solution [12-14]. Two idealized particle shapes will be considered: rigid spherical particles and rigid circular cylindrical particles with hemispherical ends. By studying the electrophoretic motion of these particles, possibilities of separating the particles in aqueous solutions and manipulating the particle motion in microchannels according to their sizes will be discussed.
546
9-1
Electrokinetics in Microfluidics
SINGLE SPHERICAL PARTICLE WITH GRAVITY EFFECTS
Let's consider a spherical particle moving in a rectangular microchannel filled with an aqueous electrolyte solution. The microchannel is connected to two solution reservoirs that open to the atmosphere, i.e., there is no pressure difference along the microchannel. Figure 9.1 (a) shows the geometry of the microchannel, which has a height of 2H, a width of 2JFand a length of L. Cartesian coordinates (x,y,z) are used with the origin located at the center of the entrance cross-section. The height of the microchannel is much less than its width so that the side effect is negligible and the motion of the sphere is on the x-z plane. Figure 9.1 (b) shows the cross-section in the x-z plane, with the sphere entering at the center of the entrance cross-section (PI).
Figure 9.1. Illustration of electrophoresis of a sphere in a microchannel: (a) the geometry of the microchannel; and (b) the x-z cross-section plane of the system.
Electrophoretic Motion of Particles in Microchannels
547
Generally, the particle and the carrying electrolyte solution have different densities. For example, the density for protein is about 1350kg/m3, which is 1.35 times higher than that of the water. The gravity force tends to pull the heavier particles near the lower wall of the channel where the particles will experience different flow resistance from that in the center of the channel. This may influence the electrophoretic motion of particles in microchannels. Therefore, the gravitational effect should be considered. Here we assume that the density of the sphere is higher than that of the solution. Both the sphere surface and the microchannel surface carry uniform negative charges that are characterized by their respective zeta potentials: C, p and C,w. When an external electric field, with strength Ex, is applied uniformly along the channel, an electroosmotic flow will be developed toward the cathode. Generally, the motion of the sphere in the microchannel can be divided into Phase I ( P I - P 2 ) and Phase II (P2-P3). In Phase I, the sphere moves at a constant velocity £/(/), which has both a x-component and a z-component of velocity due to the electrostatic force in the x-direction and the gravity force in the z-direction. LI is a x-component of the distance between PI and P2. The sphere will reach an equilibrium height (a separation distance d) above the lower channel wall at the position P2. In Phase II, the net force acting on the sphere in the z-direction vanishes, so there is no particle motion in the zdirection. The sphere moves at a constant velocity £/(//) parallel to the lower channel wall. The dashed spheres in Figure 9.1(b) illustrate the possible trajectories of the sphere motion in the microchannel. To determine the velocities of the sphere in Phase I and in Phase II and the separation distance d, and to capture the main characteristics of the electrophoretic motion under the gravitational field, the following assumptions are made: • Both the sphere and the microchannel are rigid and non-electrically conducting; • The electrolyte solution is Newtonian and incompressible. And the electrolyte is symmetric, that is, n^ - n_00 = « +00 and z = z+ = -z_; • The fluid flow field is steady, and the sphere moves at constant velocities respectively in Phase I and in Phase II. The flow is slow enough so that it can be considered as a Stokes flow; • The Brownian motion and the effect of the electrokinetic lift force are negligible; • The wall effect in Phase I are negligible; the upper wall effect in Phase II are negligible; and • The sphere moves in a uniform electroosmotic flow field.
548
Electrokinetics in Microfluidics
The separation distance d When the sphere moves very close to the lower wall, the electric double layer interaction force and the van der Waals force [15] become effective, and they will balance the gravity force and the buoyancy force. In this system, we define the separation distance h (shown in Figure 9.2(c)) as the distance from the center of the sphere to the lower wall when the forces in the z-direction are balanced, and the separation distance d as d-h-a. Thus, a force balance is used to determine d. When the sphere moves at a constant velocity UX(II) parallel to and very close to the lower wall (shown in Figure 9.2(c)), there are four forces (shown in Figure 9.2(d)) acting on the sphere in the z-direction: the gravity force (FQ), the buoyancy force (Fg), the electric double layer interaction force (Fe) and the van der Waals force {FV(^W). These forces are subject to the following force balance equation: (6) The gravity and buoyancy forces can be calculated by the following equations:
Figure 9.2. The sphere's motion in Phase I and Phase II: (a) the motion of the sphere in the zdirection in Phase I; (b) the motion of the sphere in the x-direction in Phase I; (c) the motion of the sphere in Phase II; (d) the forces acting on the sphere in the z-direction in Phase II.
Electrophoretic Motion of Particles in Microchannels
549
(7)
(8) where pp is the density of the sphere, p/ is the density of the electrolyte and g is the gravitational accelerating rate. Eqs. (7) and (8) reveal that FQ and Fg are dependent on the sphere's size, the sphere's density and the solution's density. Hogg, et al. [16] derived the electrical double layer interaction potential between two spheres: (9) where VR (d) is the electrical double layer interaction potential, a\ and a^ are the radius of the two spheres respectively, q\ and q^ a r e m e zeta potentials of the two spheres respectively, d is the separation distance between the two spheres, and K: is the reciprocal Debye length. For the symmetric electrolyte, K is given by (10) where s is the dielectric constant of the electrolyte solution, e 0 is the permittivity of vacuum (8.85xlO~ 12 C7F-m), z is the valence of the ion species, e is the charge on an electron (1.602 xlO~19C), nx is the ionic number concentration (m~3), k/, is the Boltzmann constant (1.381 x 10~23 JIK), and T is the absolute temperature (K). From Eq. (9), we can derive the electrical double layer interaction potential between a sphere and a flat plate by setting a\ =a, 02=°°, q\=qp (the particle) and £2 = gw (the wall or the plate). We obtain
(11)
550
Electrokinetics in Microfluidics
The electric double layer interaction force Fe can be derived by using the following relation: (12)
As seen from the above equation, Fe is dependent on the sphere's size, the separation distance, the Debye length, the zeta potentials of the sphere and the channel wall. Generally, the van der Waals force can be attractive or repulsive depending on the properties of the sphere and the channel wall and those of the solution. Fvc{w can be approximated by the following equation [15]: (13) 1-20-
where A is the Hamaker constant. In this work, we choose A = 0.83x10 J, that is, FV(jw is an attractive force. This force is dependent on the sphere's size and the separation distance. Substituting the Eqs. (7)—(13) into Eq. (6), we have
(14)
Combining the gravity force and the buoyancy force into a net gravity force FgQ, we have (15) Thus Eq. (14) can be rewritten as:
Electrophoretic Motion of Particles in Microchannels
551
(16)
The above equation reveals that the separation distance depends on the sphere's size, the density difference between the sphere and the solution and the electrochemical properties of the solution. The motion of the particle in Phase I As mentioned before, we assume that the flow is the Stokes flow, that the sphere moves in a uniform electroosmotic flow field and that the wall effects in Phase I are negligible. The motion of the sphere can be decomposed into the motion in the z-direction and the motion in the x-direction (shown in Figures 9.2(a) and (b)), where Uz (/) and Ux (I) are the velocity components in the zdirection and in the x-direction respectively.
Figure 9.3. The separation distance between the sphere and the lower channel wall as a function of sphere's radius with Ex = 30V/cm, p« = 1200 kg/m , C, „ = -40mV and C w = -20mV.
552
Electrokinetics in Microfluidics
Figure 9.2(a) shows the forces acting on the sphere when it moves at a constant velocity Uz (I) through a bulk solution. In this direction, two forces act on the sphere: FgQ and F^Z{I). F^Q is the net gravity force given in Eq. (15); F/jZ(7) is the hydrodynamic force that can be determined by Stokes' law of resistance. The force balance on the sphere in the z direction yields (17) That is, the sedimentary velocity Uz (I) is proportional to the square of the sphere's radius and the density difference between the sphere and the solution. Figure 9.2(b) shows the forces acting on the sphere when it moves at a constant velocity Ux (I) through a uniform electroosmotic flow field (U eo ) in the x-direction. Fg is the electrostatic force, and Ffa (I) is the hydrodynamic force. For this system, we have (18)
Figure 9.4. The z-component of the sphere velocity in Phase I with respect to sphere's radius
(Ex = 30V/cm, pp=l200kg/m3,
C,p = -40mF and C w =-20mV).
Electrophoretic Motion of Particles in Microchannels
553
where Uep is the electrophoretic velocity of the sphere. By Henry's solution, Uep can be expressed as
(19) where \x is the viscosity of the electrolyte solution, and / ( K H ) is Henry's function, which was derived by Henry [17] and was simplified recently by Ohshima [18]. For a sphere, an expression for f(Ka) is given by [18]
(20)
Note that the electroosmotic flow velocity of the solution is given by: (21) Therefore, we have (22) Eq. (22) indicates that UX(I) is proportional to Ex and it is dependent on the sphere's size, the electrochemical properties of the solution and the zeta potentials of the sphere and the channel wall. When the sphere velocity components in the x-direction and the zdirection are determined, we can calculated the time tl needed for the sphere moving from PI to P2 and the distance between them in the x-direction: (23)
(24)
554
Electrokinetics in Microfluidics
Figure 9.5. The x-component of the velocity in Phase I and the velocity in Phase II as a
The motion of the particle in Phase II As mentioned earlier, in Phase II, the forces acting on the sphere in the zdirection are balanced, so there is no motion in the z-direction and the sphere moves at a constant velocity U(II) parallel to the lower wall through the uniform flow field (Ueo) (shown in Figure 9.2(c)). Ennis, et al. [7] studied the electrophoretic motion of a charged sphere parallel to a flat wall and considered the electrophoretic retardation effect. They used the method of reflections and derived the particle velocity. Their solution is valid for low zeta potentials and any value of ka. Using Ennis' solution, we have
(25)
Electrophoretic Motion of Particles in Microchannels
555
where X is the ratio of the particle radius to the distance from the boundary, and the functions of fx(x), L(x), M(x), Wj,{x), W<-(x) and W^{x) can be founded in Reference [7]. Eq. (25) reveals that £/(//) is proportional to Ex and depends on the sphere's size, the separation distance, the electrochemical properties of the solution and the zeta potentials of the sphere and the channel wall. Using Eqs. (17, 22-25), we can predict the transport distance of the sphere in the microchannel within a given time period t: (26) (27) where s is the distance (in the x-direction) that the sphere travels in time period / in the microchannel. In the above, we have presented a model to determine the sphere's separation distance d from the lower channel wall, and the motion of the sphere in Phase I and Phase II. The model indicates that both the separation distance and the sphere's motion depend on sphere's size, the density difference between the sphere and the solution, and the electrochemical properties of the system. In the following, we will discuss the effects of the sphere's size, the density difference, the zeta potentials of the sphere and the channel wall, and the applied electric field strength on the electrophoretic motion of a sphere in an aqueous solution in a microchannel. In the calculations, we choose a rectangular microchannel with H = 50/Mn, W = \mm and L = 200mm. For simplicity, we consider an aqueous, symmetric electrolyte (e.g., KC1, z + = - z ~ = l ) concentration 1.0 x 1(T3M. At a room temperature T electrochemical properties of the solution are Pl =1000kg/m3, e = 78.5. As mentioned earlier, constant to be A = 0.83 x 10~ 20 J.
solution with an ionic = 293 K, some physical and p. = 1.003 xlO~ 3 kg/(m-s), we choose the Hamaker
The effects of particle's size As discussed earlier, the sphere's size affects the separation distance and the motion of the sphere in the microchannel. In this part of the calculation, we considered that the density of the sphere is pp = 1200kg/m3, the applied electric
556
Electrokinetics in Microfluidics
field strength is Ex =30V/cm, and the zeta potentials are C,P =-40mV and C w =-20mV. Figure 9.3 shows that the separation distance d decreases with sphere's radius a. As discussed earlier, the net gravity force and the van der Waals force tend to draw the sphere toward the lower wall, while the electrical double layer force tends to push the sphere away from the lower wall. If d decreases, both the van der Waals force and the electrical double layer interaction force increase, while the latter one increases much faster. Thus, when a increases, the net gravity force increases and d decreases; at a smaller d, the electrical double layer interaction force increases to balance the increased gravity force. Figure 9.4 shows that the z-component of the sphere's velocity UZ(I) in Phase I increases with the sphere's radius a. In Phase I, the net gravity force is the driving force for the motion of the sphere in the z-direction, and it is proportional to a 3 . The Stokes hydrodynamic force (flow friction) is proportional to a. The net effect is that UZ(I) is proportional to a2. As expected, the larger particle settles down faster.
Figure 9.6. The effect of sphere's size on its transport distance (Ex= 3
pp=l200kg/m ,
Q =-40mV and £ w =-20mV).
30V/cm,
Electrophoretic Motion of Particles in Microchannels
557
Figure 9.5 shows the x-component of the sphere's velocity UX(I) in Phase I and its velocity U(II) in Phase II as a function of its radius a. Both UX(I) and £/(//) increase with a, UX{I) is larger than U(II), and the difference between UX(I) and U(II) increases with a. Since the considered ionic concentration of the solution is 1.0xl(T 3 M, the value of K: is very large. As a increases, Ka increases. For the radius range from OAjjm to \/jm, Ka ranges from 10 to 104. Within this range, the electrophoretic velocity of the sphere increases with m. So Ux (I) and U{II) increase with a. Furthermore, the electrophoretic retardation effect become stronger as Ka increases, and this explains why the curves of UX{I)~ a and U(II)~ a approach to horizontal lines with the increase of a. In addition, when the sphere moves close to the channel wall, the flow friction increases due to the wall effect. This explains that UX{I) is larger than U{11). The closer the sphere to the channel wall, the stronger the wall effects. As mentioned earlier, the separation distance decreases with a, so the wall effects become stronger for larger spheres, this is why the velocity difference (UX(I)-U(II)) increases with a. However, the velocity difference is small. As Ennis, et al. [7] found that the order of magnitude of the wall effect on the sphere's motion is O(A3), where A =
, and they found that a+d
the higher the value of Ka, the less effective the wall effect. Because both UX(I) and U(II) increase with the sphere size, it can be expected that the travel distance with time will be different for spheres of different sizes. Figure 9.6 shows that the travel distance over time for three spheres of different radius but the same zeta potential. Clearly these spheres can be separated by size in the aqueous solution in the microchannel. However, it was found that for particles with radius larger than 1 um, the velocity of the particles is essentially independent on the particle size (see Figure 9.5). In other words, all large particles will have essentially the same velocity, and hence cannot be separated in the aqueous solution in the microchannel. The effects of the density difference between the particle and the solution The net gravity is a force pulling the sphere close to the lower wall, and this force is determined by the sphere's size and the density difference (Ap = pp - pi). Therefore, the change of Ap affects the separation distance d and the sphere's velocity. To study the density effects, we kept the density of the solution as p/ = \000kglm3, and used three different densities of the sphere: pp = H00kg/m3, 1350kg/m3 and 1700kg/m3. All the following results were calculated at Ex = 30V/cm, C,P =-40mV and C,w =-20mV.
558
Electrokinetics in Microfluidics
Figure 9.7. The effect of the density difference on the separation distance between the sphere and the lower wall (£^=30^/ cm, ^p=-40mV and £ w =-20/»F).
Figure 9.8. The effect of the density difference on the transport distance of the sphere (a=0.5^im,E =30V/cm, Q =-40mV and Q
Electrophoretic Motion of Particles in Microchannels
559
Figure 9.9. Effect of the zeta potential of the particle on the x-component of the velocity in Phase I and on the velocity in Phase II (Ex = 30V/cm, pp = \200kg/m3 and Cw=-20mV).
Figure 9.10. The effect of the zeta potential of the sphere on the transport distance of the sphere (a=0.5pm, Ex=30V/cm, pp =l200kg/m3 and £ w =-20mF).
560
Electrokinetics in Microfluidics
Figure 9.7 shows the effect of the density difference on d. The separation distance d decreases with the density difference Ap. As mentioned earlier, both the van der Waals force and the electric double layer interaction force increase when d decreases, but the latter one increases faster. Therefore, when Ap increases, the net gravity increases, the combination of the electric double layer interaction force and the van der Waals force should increase to balance the net gravity force increase; the net effect is the decrease of d. As mentioned earlier, Uz (/) is proportional to Ap. In the case of a larger Ap, it takes less time for the sphere to move from Phase I to Phase II. The decrease of d with the increase of Ap will lead to a small decrease of U(II) . Consequently, the particle transport will be slower in the case of a larger Ap. However, as shown in Figure 9.8, the density difference effect on the travel distance is insignificant. The effects of the zeta potential of the particle As discussed earlier, d, UX(I) and U(II) are dependent on the zeta potential of the sphere £ „ . To investigate the effect of C, „ on UX(I) and U(II),
Figure 9.11. The effect of the zeta potential of the channel wall on the transport distance of
the sphere (a=0.5fim, E =30V/cm, p =\2QQkgln? and C, =-50mV).
Electrophoretic Motion of Particles in Microchannels
561
Figure 9.9 shows the effects of C,p on UX(I) and U(II). Both UX(I) and [/(//) increase with the absolute value of C,p. Figure 9.10 shows the effect of C,p on the transport distance J for a sphere with a radius a - 0.3pan. For a given time period, s increases with the absolute value of C,p. As discussed earlier, both UX(I) and U(II) are proportional to C,p. Therefore, the particle travel distance s is proportional and very sensitive to £ „ . This implies that the spheres with the same size but different surface charge can be separated in aqueous electrolyte solutions in the microchannel. The effects of the zeta potential of the channel wall For a given applied electrical field strength, the electroosmotic flow velocity Ueo is proportional to the zeta potential of the channel wall. If the sign of the surface charge of the wall is the same as that of the particle, the electroosmotic flow will be in the opposite direction to the particle's electrophoretic motion, and the particle's motion will be affected by the electroosmotic flow and hence the zeta potential. In our system, both the channel
Figure 9.12. The effect of the applied electric strength on the transport distance of the sphere
(pp=\200kg/m3,
Qp=-4QmV and C,w=-2QmV).
562
Electrokinetics in Microfluidics
wall and the sphere are considered carrying the surface charges of the same sign, therefore, the electroosmotic flow retards the sphere's motion. As a result, both UX(I) and £/(//) decrease with the absolute value of £ w . Figure 9.11 shows that the travel distance s decreases with the absolute value of £ w , where the following
parameters
were
used:
pp=l200kg/m
,
Ex=30V/cm,
C, p = -50mV and a = 0.5fum. The effects of the applied electric strength It is known that both the electroosmotic flow velocity and the electrophoretic motion are proportional to the applied electric strength Ex, therefore UX(I) and U(II) are proportional to Ex. As a. result, the particle's travel distance s is proportional to Ex. Figure 9.12 shows that the travel distance s increases significantly with Ex, where the following parameters were used: pp =\200kg/m3,
£ P =-40wF,and C,w = - 2 0 w F .
In summary, this section considered electrophoretic motion of a sphere in an aqueous electrolyte solution in a microchannel under the gravitational field. It is found that the separation distance d decreases with the sphere's radius a and the density difference Ap. The sphere's motion in the microchannel is affected by its size, the density difference Ap, the zeta potentials of the sphere and the channel wall and the applied electric strength. Uz (I) increases with a and Ap; both UX{I) and U(II) increase with a, Ex and the absolute value of C,p. Because the surface charges of the channel wall and the sphere are of the same sign, the electroosmotic flow retards the sphere's motion so that both UX(I) and U(II) decrease with the absolute value of t,w. The particle's traveling distance 5 increases with the particle's size a, the applied field strength Ex and the absolute value of particle's zeta potential £ „ , but decreases with the density difference Ap and the absolute value of channel wall's zeta potential £ w . The effects of Ex, £ „ and £ w on the sphere's transport distance in a time period are significant, while the effect of the density difference is very small. Within Ka range approximately from 10 to 100, we found the particles of the same surface charge can be separated by electrophoresis in aqueous solutions in a microchannel. For system with Ka larger than 100, particles of different size but with the same surface charge will have essentially the same velocity, and hence cannot be separated in aqueous solutions in a microchannel.
Electrophoretic Motion of Particles in Microchannels
9-2
563
SINGLE CYLINDRICAL PARTICLE WOTHOUT GRAVITY EFFECTS
This section considers electrophoretic motion of a rigid cylindrical particle with circular cross-section and hemispherical ends along the axis of a circular cylindrical microchannel filled with an aqueous electrolyte solution. The influences of three parameters on the electrophoretic motion of a circular cylindrical particle will be discussed: the ratio of the particle radius to the channel radius; the ratio of the axial length of the particle to its radius; and the ratio of the zeta potential of the channel to that of the particle. A simple analysis of the electrophoretic separation of circular cylindrical particles in small circular cylindrical microchannel will also be presented. Figure 9.13 (a) shows a circular cylindrical particle with hemispherical ends in a circular cylindrical microchannel filled with an aqueous electrolyte solution. The two ends of the microchannel are open to the atmosphere, i.e. this is an open system where no overall pressure gradient presents in the microchannel. The particle is suspended co-axially in the microchannel. The particle has a radius a and an axial length L, and the microchannel has a radius b and an axial length lw, where lw > 100b. Both the particle and the channel surfaces carry uniform negative surface charges, which are characterized by their respective zeta potentials: C,p and C,w. When they are brought into contact with
Figure 9.13. The schematic diagram of the particle-microchannel system: (a) the geometry of the system; and (b) the division of the liquid phase.
564
Electrokinetics in Microfluidics
an electrolyte solution, there are two electrical double layers: one is surrounding the particle and the other is next to the channel wall surface. When a uniform electric field is applied along the channel, the electrophoretic motion of the particle will be toward the anode, while the electroosmotic motion of the liquid will be toward the cathode. We consider that the particle will move at a velocity U along the axis of the channel without rotation resulting from the electrophoretic motion of the particle and hydrodynamic effects of the electroosmotic flow of the liquid. To determine the particle velocity, U, a model governing the electrical field, the flow field, and the forces acting on the particle need to be formulated and solved. Cylindrical coordinates (r, 6, z) are used with the origin located at the center of the particle; hence the particle velocity is zero, and the channel wall moves with a velocity - U. Due to the symmetry, all of the 9 -dependent terms vanish in the subsequent analysis. To further simplify the analysis, the following four assumptions have been made: (1) both the particle and the channel wall are rigid and non-conducting; (2) the aqueous electrolyte solution is Newtonian and incompressible, the continuity equation and the Navier-Stokes equations are valid. The flow is of low Reynolds number, so that the inertia terms in Navier-Stokes equations can be neglected; (3) the particle and the aqueous electrolyte solution are of the similar density, so that the gravitational effects are negligible; and (4) the electrical double layers are thin, i.e., K a —> oo. The liquid phase will be divided into two regions (as shown in Figure 9.13(b)): an inner region which is defined as the electrical double layers adjacent to the particle and the channel wall, where the characteristic length scale is the Debye length K~ ; and an outer region which is defined as the remainder of the liquid. Inner region As was mentioned above, the inner region is the electrical double layer region. Variables in this region are denoted by a hat (A) over them. At the liquidsolid interface, either on the particle or the channel wall, a local coordinate system (y, s) is used. The vector, y, is in the outward normal direction pointing into the liquid phase with y = 0 at the liquid-solid interface. The scale of y is K~ , where K~ is the Debye length. The vector, s, is the two-dimensional position vector in the tangent plane of the solid-liquid interfaces. In the inner region, the electrical potential, y/, is the sum of the electrical double layer potential, y/j, and the applied electrical potential, y/ 2 , that is (28a)
Electrophoretic Motion of Particles in Microchannels
565
Since it is assumed that both the particle surface and the channel wall are non-conducting, we have the following equation at the liquid-solid interface: (29) where n is the unit normal vector pointing into the liquid phase. Eq. (29) means that the applied electrical potential, \p2, is only dependent on s, the liquid-solid interface position vector. In the limit Ka -» oo, it can be assumed that Eq. (29) prevails over the whole inner region. The potential, \p2, must match with the electrical potential determined from the equations of the outer region. Under the thin electrical double layer condition, Keh and Anderson [6] showed that the electrical double layer potential, i/?j, is independent of s. Thus Eq. (28a) becomes (28b) where y/} is governed by Poisson equation: (30) where s is the dielectric constant of the electrolytic solution, £ 0 is the permittivity of vacuum (8.85 x 10~12C/V • m), and pe is the local volume charge density. The liquid is driven to flow in the s -direction and it is governed by the following Navier-Stokes equation: (31) where /u is the liquid (electrolyte) viscosity, and v (s) is the tangential component of the liquid flow velocity vector v. Combining Eq. (30) and (31) gives (32)
with the following boundary conditions
566
Electrokinetics in Microfluidics
(33a)
(33b) where / is the unit dyadic, and v is the velocity vector at the liquid-solid interface, v = -Uez for the particle, v = 0 for the channel wall, and £ is the zeta potential at the liquid-solid interface. Integrating Eq. (32) twice over y with boundary conditions (33) gives (34)
The normal component of the liquid flow velocity is (35)
Letting y -> oo, we can derive the tangential component of the liquid flow velocity at the outer edge of the inner region from Eq. (34). Combining the normal and tangential components gives the following liquid flow velocity at the outer edge of the inner region: (36) Outer region In this region, the ionic concentrations are uniform and the local volume charge density is zero. The electrical potential distribution is governed by (37)
with the following boundary conditions: on the particle surface and the channel surface
(38a)
far from the particle
(38b)
Electrophoretic Motion of Particles in Microchannels
567
where Ez is the applied electric strength along the channel without the suspension of the particle. Since the local net volume charge density is zero in the bulk liquid (outer region), no body force acts on the liquid. The liquid flow in the outer region is governed by (39a) (39b) where v is the liquid flow velocity vector and p is the pressure in the outer region. The boundary conditions for Eq. (39) are: on the particle surface on the channel surface
(40a) (40b)
where the term "surface" in boundary conditions (38) and (40) means the outer surface of the inner region. The outer surface of the inner region encloses a neutral body (i.e., the charged particle plus the oppositely charged inner region). Therefore, the electric field produces no force on the outer surface of the inner region. At a steady state, as the particle is freely suspended in the liquid, the net force exerted by the liquid flow on the outer surface of the inner region must be zero: (41) where S is the area of the outer surface of the inner region, and a is the stress tensor. By satisfying Eq. (41), the particle velocity, U, can be determined. From the above analysis, the liquid phase is divided into an inner region and an outer region. As Ka approaches infinity, the inner region offers the boundary conditions (38a) and (40) for the outer region. In the outer region, since the local net volume charge density (p e ) is zero, the electrical potential is contributed only by the applied electric field governed by the Laplace equation (37). The liquid flow is governed by the Stokes equations (39), with the slipping velocity boundary conditions, Eq. (40). Substituting the flow field in the force balance equation (41), the particle velocity, U, can be determined.
568
Electrokinetics in Microfluidics
However, due to the nature of the system of equations and the complex geometry, it is difficult to seek an analytical solution. A numerical method must be employed to obtain the solution. The first step is to nondimensionalize the governing equations (37) and (39) and boundary conditions (38) and (40). Let (42a)
(42b)
(42c)
(42d)
(42e)
(42f)
pp
where Uep=
-£pEzez is used as a characteristic velocity, the particle A* radius, a, is set as a characteristic length scale, and <j> = Eza is a characteristic electrical potential. The stared quantities are the dimensionless variables. Substituting these variables into Eqs. (37) and (39) and the boundary conditions (38) and (40) yields the non-dimensional equations and boundary conditions: (43) on the particle surface and the channel surface far from the particle
(44a) (44b) (45a)
Electrophoretic Motion of Particles in Microchannels
569
(45b) (46a) (46b) where U =UIU , which is the dimensionless particle velocity, and y = £ w / £ „, which is the ratio of the zeta potential of the channel wall to that of the particle. The force balance equation (41) becomes (47) The above nondimensionalized equations are solved through the following procedures: • Solve Eq. (43) with boundary conditions (44) for the electrical potential field; • Substitute the calculated V y/2 into boundary conditions (46); • Solve Eq. (45) with boundary conditions (46) for the flow field; and • Calculate the hydrodynamic force acting on the outer surface of the inner region. By satisfying Eq. (47), the particle velocity, U , can be determined. A numerical code, based on Taylor-Hood triangle element (i.e., second order interpolation for the electrical potential and the liquid flow velocity, and first order interpolation for the pressure), was developed to solve the above set of equations. An unstructured triangular mesh is used, with a fine mesh near the particle and a coarse mesh far from the particle (as shown in Figure 9.14). Since the dimensionless particle velocity, U , is also involved in the boundary conditions, an iterative solution procedure is required. To verify the code, a test case was considered first: a rigid sphere with a radius a moving in a still liquid along the axis of a circular cylindrical tube with a radius b (in the absence of electrokinetic effects). For this test case, Bohlin [19] gives the following formula (48)
570
Electrokinetics in Microfluidics
Figure 9.14. Illustration of fine mesh near a particle surface used in the numerical calculation.
Figure 9.15. Comparison between Bohlin's formula and the model in this section of wall correction factor, k, versus the ratio of a particle radius to a tube radius, a I b, for a rigid sphere moving in a stationary liquid along the axis of a circular cylindrical tube.
Electrophoretic Motion of Particles in Microchannels
571
(49)
where Fz is the hydrodynamic force acting on the sphere when it moves at a constant velocity Uz along the axis of the tube, and £ is a wall correction factor. Figure 9.15 compares the Bohlin formula for the wall correction factor, k, with the numerical solution of the above described model. As can be seen, the numerical solution matches Bohlin's formula very well. Furthermore, the electrophoretic motion of a rigid sphere with a radius a along the axis of a circular cylindrical pore of a radius b was used to test the above model and the numerical code. In the limit of thin electrical double layer, Keh et al. [6] derived the following approximate solution by a method of reflections:
(50)
where U is the sphere velocity. From the above equation, the following dimensionless sphere velocity can be derived:
(51)
Figure 9.16 compares the Keh's solution with the numerical solution of the model developed here at y - 0.6. As can be seen, the numerical solution matches the Keh's solution very well. Figure 9.17 shows the streamlines of the electrophoretic flow field near the cylindrical particle in a cylindrical microchannel and the contours of the absolute value of the flow velocity at a I b = \l 4 and y - 0.6. In this figure, the darker areas at the ends of the particle denote the lower flow velocity regions and the lighter areas on the sides of the particle denote the higher velocity regions.
572
Electrokinetics in Microfluidics
Figure 9.16. Comparison between Keh's solution and the model in this section of dimensionless particle velocity, £/*, verses the ratio of a particle radius to a channel radius, alb, with y = 0.6 for the electrophoretic motion of a rigid sphere along the axis of a circular cylindrical channel.
Figure 9.17. Streamline plot of the flow field near the particle with a contour plot of absolute values of flow velocity at alb = \/4, L/a = 4 and 7 = 0.6.
Electrophoretic Motion of Particles in Microchannels
573
Influence of the ratio of particle radius to channel radius {alb) Figure 9.18 shows the dimensionless particle velocity, Up,vs. alb, with y = 0.6. As can be seen, U„ decreases with alb. The effect becomes stronger when the ratio of particle's axial length to its radius increases. Compared with a spherical particle, the influence of alb on U for a circular cylindrical particle is stronger. When the channel radius is far larger than the particle radius, i.e., (a/b)< (1/20), the boundary effect is negligible. However, for the cases of (a/b)>(l/20), the boundary effect is relatively significant. This result implies that, for a given channel size, the cylindrical particles with a smaller radius move faster, where the particles are assumed to have the same zeta potential and the same axial length but different radii. In other words, particles of the same zeta potential and length can be electrophoretically separated by their radii in an aqueous solution in a microchannel. Influence of the ratio of particle's axial length to its radius (LI a) Figure 9.19 shows the dimensionless particle velocity, U , vs. LI a, with y - 0.6. As can be seen, the influence of LI a on Up becomes stronger as a I b increases. For a given alb, Up decreases with LI a and gradually approaches to a constant. The slope of U*p ~ LIa curves is steeper in the range of LI a < 10, that is, the particle's motion is more sensitive to LI a in this range. As shown in Table 1, to move through a 50mm long cylindrical microchannel, when particles' radii are set the same, the particle with a shorter axial length moves faster and spends less time. That is, circular cylindrical particles of the same zeta potential and the same radius can be electrophoretically separated by their axial lengths in an aqueous solution in a microchannel. Table 1. Time required for a cylindrical particle moving through a 50mm long circular cylindrical microchannel
with
a = l^m,
£ =80.1,
Ciyl=-A%mV, £ p =-80mF,
£ z =300F/cm
3
/i = 1.003* 10" kg/(m •s).
a/b
z(um)
1/4
1/3 4
6
8
4
6
1/8 8
4
6
8
0.371 0.365 0.362
0.386 0.382 0.380
0.398 0.397 0.396
U(mmls)
0.628 0.619 0.614
0.654 0.648 0.644
0.674 0.673 0.671
<(s)
79.56 80.72 81.46
76.50 77.16 77.65
74.19 74.34 74.46
and
574
Electrokinetics in Microfluidics
Figure 9.18. Calculated dimensionless particle velocity, U*p, verses the ratio of a particle radius to a channel radius, alb, with y = 0.6 .
Figure 9.19. Calculated dimensionless particle velocity, u'p, verses the ratio of a particle axial length to its radius, LI a, with y = 0.6 .
Electrophoretic Motion of Particles in Microchannels
575
Influence of the zeta potential ratio (y = £ w / £ „ ) Figure 9.20 shows the dimensionless particle velocity, U , vs. y, with a I b = 1 / 5, and LI a = 6. Up is shown to decrease linearly with y. It is known that the electrophoretic velocity of a particle is proportional to its zeta potential, C,p, and that the electroosmotic flow velocity is proportional to the zeta potential of the boundary surface, £ w . In this system, the particle velocity is a result of the electrophoretic motion of the particle and the hydrodynamic effects of the surrounding electroosmotic flow of the liquid in the microchannel. As we consider that the particle and the channel wall carry the same type of charge (i.e., both positive or both negative), the electrophoretic motion of the particle is in the opposite direction to the electroosmostic flow of the liquid. In the limit of KQ —» oo, the particle velocity is proportional to (£ - C,w). As shown in Figure 9.20, particle velocity is zero when the zeta potentials are equal (i.e., y = 1). Electrophoretic separation of cylindrical particles in a microchannel Electrophoresis is one of the most widely used separation techniques. It is known that the electrophoretic mobility of a particle is proportional to its zeta potential. Therefore, particles with the same size but different zeta potentials can be electrophoretically separated. In the limit of thin electrical double layer, the electrophoretic motion of a particle (Uo) in an unbounded aqueous electrolyte
Figure 9.20. Calculated dimensionless particle velocity, U*p, verses the ratio of the zeta potential of the channel to that of the particle, y, with alb = 1/5 and LI a = 6.
576
Electrokinetics in Microfluidics
solution can be expressed by Smoluchowski's formula, U0=ss0£pEao//j,, where Ex is an applied electric strength. It implies that the particle velocity is independent on its size or shape so that particles with the same zeta potential cannot be separated electrophoretically by their size or shape in unbounded aqueous solutions. However, for spherical particles moving in small channels, it is found that the particle velocity is dependent on its size (Figure 9.16) and hence the spherical particles with the same zeta potential can be separated by size in microchannels. This is possible even in an unbounded liquid, if the double layer is not thin or if ka is finite as the mobility depends on ka. As for circular cylindrical particles in small circular cylindrical microchannels, our model analysis and numerical results indicate that the particle velocity is dependent on its size, due to the boundary (channel wall) effects. An analysis of electrophoretic separation of circular cylindrical particles in a circular cylindrical microchannel filled with an aqueous solution is presented below. In the calculations, the following conditions were used: e = 80.1, C w =-48mV, £ p = - 8 0 m V , £ z =300V/cm, and n = 1.003 * 10"3 kg/(m • s). First, we considered three circular cylindrical particles with different radii,
Figure 9.21. Illustration of circular cylindrical particles separation by their radii in a circular cylindrical microchannel with L = 6pim, b = l2/jm, £ = 8 0 . 1 , c w =-48m7, ;p=-mmv, Ez=3oov/cm and ti=\.oo3xicr3kgUm*).
Electrophoretic Motion of Particles in Microchannels
577
0.5/iw, l/jm and 1.5/mi, respectively and with the same axial length, L = 6\im, in a circular cylindrical microchannel of radius b-12pan. The numerical analysis shows that these three particles have velocities of 0.677mm/s, 0.676mm/ s and 0.674mm I s, respectively. As shown in Figure 9.21, the three particles are well separated after 500 seconds, the particle of the smallest radius moves the fastest. Here we assumed that each particles moves at a velocity as it moves along the axis of the channel in the absence of other particles. In Figure 9.22, another three particles with the same radius, a = \.0/jm, but different axial lengths, 4/urn, 6pan and S/jm, respectively, in a circular cylindrical microchannel of a radius b = 12fj.m were considered. Under the same conditions, these particles move at 0.67644mm/s, 0.67602mmIs and 0.6756\mm/s respectively. As shown, these particles can be well separated after 1000 seconds, the shortest particle moves the fastest. In summary, this section considered the electrophoretic motion of a circular cylindrical particle with hemispherical ends in a circular cylindrical microchannel filled with an aqueous electrolyte solution. The purpose is to investigate the boundary effects on the electrophoretic motion of the particle. As predicted by the numerical simulations, in a circular cylindrical microchannel, the boundary effects on the electrophoretic motion of a circular cylindrical
Figure 9.22. Illustration of circular cylindrical particles separation by their axial lengths in a circular cylindrical microchannel with a = 0.5pm, 6 = l0pm, £=80.1, f w =-48mK, C, =-80mF, Ez =300V/cm and fi = \.003 *\0~3 kg/(m-s) .
578
Electrokinetics in Microfluidics
particle are appreciable when the ratio of the particle radius to the channel radius, alb, is large and when the ratio of the axial length of the particle to its radius, LI a, is small. The dimensionless particle velocity, U , decreases with alb for a fixed LI a and U decreases with LI a for a fixed a I b. It was found also that U decreases as the ratio of the zeta potential of the channel to that of the particle increases. Based on these results, in a small circular cylindrical microchannel filled with an aqueous electrolyte solution, circular cylindrical particles of the same zeta potential may be separated electrophoretically by their radii or by their axial lengths. The numerical calculation method presented in this section is suitable to the 2D steady state particle-liquid system. However, it is not advisable to expand this technique to more complicated cases such as 3D unsteady state particleliquid systems. In this method, hydrodynamic force needs to be calculated explicitly. How to calculate this force accurately is a big challenge. Furthermore, if the rotational effects of the particle cannot be neglected in some more complicated cases, the interacting moments between the particle and the surrounding liquid need to be determined explicitly too. In addition, extra governing equations of rotational equilibrium/dynamics will complicate the iterative solving procedures. For these more complicated situations, direct numerical formulation [20] should be used. Furthermore, in this numerical simulation, the ratio of the particle's radius to the channel's radius, alb, was chosen to be less than 0.5. There are two basic reasons: First, if the ratio, alb,is close to unity, the gap between the particle and the channel wall is very small and the continuum and Stokes equations might not be valid in some cases. Second, for large alb values, or the very small gaps between the particle and the channel wall, the flow velocity changes more abruptly in this gap and a finer mesh system is require describing the flow field. This would cause more complication to the numerical calculation and decrease the accuracy.
Electrophoretic Motion of Particles in Microchannels
9-3
579
SPHERICAL PARTICLE IN A T-SHAPED MICROCHANNEL
In many applications in lab-on-a-chip devices, it requires manipulating the motion of bio-particles through relatively complicated microchannels such as Tshaped microchannels or Y-shaped microchannels. When a particle undergoes electrophoretic motion through the T-shaped or Y-shaped junction region of the channels, transient effects have to be considered. Because of the dimension change of the channel and the presence of the moving particle in this region, the local electric field and hence the flow field varies with time simultaneously. This section discusses three-dimensional transient electrophoretic motion of a spherical particle in a T-shaped rectangular microchannel [14]. The objective is to provide a fundamental understanding of the characteristics of the particle motion in the complicated T-shaped junction region. The influences of the applied electric potentials, the zeta potentials of the channel walls and the particle as well as the size of the particle on the electric field and on the particle motion will be examined. Theoretical model We consider a spherical particle suspended in an aqueous electrolyte solution in a T-shaped rectangular microchannel, as shown in Figure 9.23. Cartesian coordinates (x,y,z) are used in the analysis, with the axis z in the vertical direction. For simplicity, we assume that the particle has the same density as that of the solution so that the gravitational effects are negligible. The three ends (i.e., the inlet, the outlet 1 and the outlet 2) of the channel are connected to three reservoirs that are open to the atmosphere, i.e., this is an open system where no overall pressure gradient presents in the system. Both the particle and the channel surface carry uniform negative surface charges, which are characterized by their respective zeta potentials: C,p and C,w. When they are brought into contact with an electrolyte solution, there are two electrical double layer fields: one is surrounding the particle and the other is next to the channel wall. In order to reduce the computational time in the numerical simulation, we assume that the particle is initially suspended in the T-shaped junction region with zero velocity. When the electric potentials are applied at the three ends of the channel, the electrophoretic motion of the particle will be in the direction against the electric field (i.e., the gradient of the electric potential), while the electroosmotic flow of the liquid will be along the electric field. In this system, due to the complicated geometry, the electric field changes with the movement of the particle, which in return influences the flow field and the particle motion. Thus, the electric field, the flow field and the particle motion are coupled
580
Electrokinetics in Microfluidics
Figure 9.23. The schematic diagram of a spherical particle in a T-shaped microchannel and the non-dimensional coordinate system;
together and this is a transient process. To capture the characteristics of the particle motion in the junction region, a model governing the electric field, the flow field and the particle motion needs to be developed and solved. To further simplify the analysis, the following four assumptions have been made: (1) both the particle and the channel wall are rigid and non-conducting; (2) the aqueous electrolyte solution is Newtonian and incompressible, and the flow is a Stokes flow; (3) the electric field is developed instantly; and (4) the electrical double layers are thin. In the theoretical analysis, we divide the liquid phase into two regions: An inner region that is defined as the electrical double layers adjacent to the particle and the channel wall. An outer region that is defined as the remainder of the liquid. For the inner region, the characteristic length scale is the Debye length K:~ . Under the assumption of thin electrical double layers, the electroosmotic
Electrophoretic Motion of Particles in Microchannels
581
Figure 9.24. Illustration of the outer region.
flow velocity is used to describe the flow in the inner regions adjacent to the particle and the channel wall [6,13]. The outer region is shown in Figure 9.24, where Q denotes the domain of the outer region, Tw denotes the outer edge of the electrical double layer adjacent to the channel wall, r ^ denotes the outer edge of the electrical double layer adjacent to the particle, and Tin, Tout\ and Tout2 denote the inlet, the outlet 1 and the outlet 2 of the channel respectively. In the outer region, the ionic concentrations are uniform and the local volume charge density is zero. The model governing the distribution of electric potential and the flow field in the outer region and the particle motion are described in the following sections. Distribution of the Electric Potential In the outer region, the ionic concentrations are uniform and the local volume charge density is zero. The distribution of the electric potential is governed by (52)
582
Electrokinetics in Microfluidics
with the following boundary conditions: (53) (54)
where y/ is the applied electric potential and n is the unit normal vector pointing into the liquid phase. Flow fields around the particle Under the consideration of thin electrical double layers, the electroosmotic flow velocities are employed to describe the flow in the inner regions adjacent to the particle and the channel wall. In the outer region, since the local net charge density is zero in the bulk liquid (the outer region), no electrostatic force acts on the liquid. The liquid flow in the outer region is governed by (55) (56) where v is the liquid flow velocity vector, p is the density of the electrolyte solution, 77 is the viscosity of the liquid (i.e., the electrolyte solution) and p is the pressure in the outer region. Under an applied electrical field, the charged particle will move following the electrical field force (i.e., the electrophoresis), while the bulk liquid will undergo an electroosmotic flow due to the presence of the EDL field near the microchannel wall. To the liquid flow, the microchannel wall is a fixed boundary, and the particle surface is a moving boundary. In addition to the particle's absolute velocity, the liquid immediately next to the particle surface also experiences an electroosmotic flow (flow slip at the particle surface) due to the presence of the EDL around the charged particle. Therefore, the liquid flow boundary condition at the particle surface is the sum of the particle's velocity and the liquid electroosmotic flow velocity. Under the thin EDL or thin inner region consideration, the electroosmotic flow velocities at the particle surface and at the channel wall surface are considered as the slipping flow boundary velocities for the outer region flow. Therefore, the slipping flow boundary conditions for the outer region flow field are
Electrophoretic Motion of Particles in Microchannels
583
(57) (58) where e is the dielectric constant of the electrolytic solution, £Q is the permittivity of vacuum (8.85xl(T12C/V-m), I is the unit dyadic, Vp and a>p are the translational velocity and the angular velocity of the particle respectively, xp is the position vector on the particle surface, and X „ is the position vector of the particle center. Eq.(57) is the slipping flow boundary condition, i.e., the electroosmotic flow velocity, at the microchannel wall. Eq. (58) is the velocity boundary condition at the particle surface. In Eq. (58), the 1st term is the particle's translational velocity, the second term represents the particle's angular velocity and the last term is the electroosmotic velocity of the liquid around the particle. Particle motion Generally, the Newton's second law governs the particle's motion. The translational motion is described by:
where mp is the mass of the particle, Vp is the translational velocity, and Fnet is the net force acting on the particle. Since the particle carries uniform surface charge, there is electrostatic force acting on the particle by the applied electric field. At the same time, the flow field around the particle exerts a hydrodynamic force on the particle surface. The net force acting on the particle is:
where Fg is the electrostatic force acting on the particle and Ff, is the total hydrodynamic force acting on the particle. The hydrodynamic force can be divided into two components:
584
Electrokinetics in Microfluidics
where Ff,jn is the hydrodynamic force acting on the particle surface due to the flow field in the inner region (i.e., electroosmotic flow in the EDL region around the particle), and FfjO is the hydrodynamic force acting on the particle surface due to the flow field originated in the outer region. In this model, we assume the EDL around the particle is so thin that its thickness can be neglected in comparison with the size of the particle (1-30 |j.m) and the size of the channel (~100 um). For example, in a solution with a high electrolyte concentration such as 10~2M, the EDL thickness is approximately 3 nm. Thus, we don't consider the detailed flow field in the EDL—the inner region, and simply replace the flow field in the thin EDL by the electroosmotic velocity as a slipping flow boundary condition for the outer region flow field. In this way the flow field around the particle is the flow field originated in the outer region and subject to the slipping flow boundary condition at the particle surface. Consequently, F^o is a hydrodynamic force acting on the particle surface. It can be shown that Fg and F^in have the same value but operate in the opposite directions, therefore, the net force acting on the particle becomes:
Consequently, the equation governing the particle's translational motion becomes:
(59)
(60) The rotational motion of the particle is governed by: (61) where J is the moment of inertia of the particle, and T), is the torque (about the particle centre) on the particle by the flow field in the outer region. The hydrodynamic force and the torque are given by
585
Electrophoretic Motion of Particles in Microchannels
(62) (63) where a is the stress tensor that is given by (64)
The initial conditions are given by: (65) In order to non-dimensionalize the equations, we chose / = a {a is the particle as a radius) as a characteristic length, the electric potential 0 = y/ r •MB
(66) (67) (68)
(69) (70)
586
Electrokinetics in Microfluidics
(71) (72) where y is the ratio of the zeta potential of the channel wall to that of the particle. The quantities with a star in the above equations are the dimensionless variables.
is the density of the particle, we can write the dimensionless governing equations for the particle motion as follows:
(73)
(74)
(75)
where
which is a coefficient.
Finally, the initial conditions
become (76) Numerical method Due to the complicated geometry in the T-shaped junction region, temporal effects have to be considered. Therefore, the particle motion, the flow field and the electric field are coupled in the following way: the particle motion provides the moving boundary for the flow field and the electric field; the flow field exerts the hydrodynamic force and toque on the particle; and the electric field provides the slipping flow boundary conditions for the flow field. In order to solve the complicated system of coupled equations, numerical simulation is required.
Electrophoretic Motion of Particles in Microchannels
587
Generally, two challenges exist for simulating the liquid-particle flow: (1) accurately calculating the hydrodynamic interactions (i.e., hydrodynamic force and torque); and (2) tracking the position of the particle. For a complicated geometry, it is extremely difficult to calculate accurately the hydrodynamic interactions. A direct numerical simulation method is preferable since it evaluates the hydrodynamic interactions with no averaging or approximation. For example, Hu [20] and Glowinski et al. [21] employed a generalized Galerkin finite element method. This method incorporates both equations of the fluid flow and equations of the particle motion into a single variational equation where the hydrodynamic interactions are eliminated and the explicit calculation of the hydrodynamic interactions is not required. Ritz et al. [22] assumed solid particles to have a very high viscosity and developed a continuous hydrodynamic model that leads to a unique system of equations for both phases. By this method, no explicit calculation of the hydrodynamic interactions is required. To track the position of the moving particle, two strategies have been developed. One is the moving mesh method that considers the relative position of the particle as a boundary of the fluid domain. Re-eshing is then needed at each time step. The other method is the fixed mesh method that tracks the position of the moving particle over time. As an example of the moving mesh method, Hu [20] adopted an arbitrary Lagrangian-Eulerian (ALE) technique to deal with the motion of the particles. The advantage of this method is that the interface of the liquid-particle is clearly defined at each time step. As an example of the fixed mesh method, Glowinski et al. [21] used a finite element based fictitious domain approach where the solid-liquid interface is defined by control points. At each of these points, a kinematic condition is imposed using Lagrange multipliers. The motion of these control points is governed by the action of the fluid on the surface of the particles. The advantage of this method is that no remeshing is needed. However, the interface of the liquid-particle is not clearly defined at each time step. In the numerical simulations presented in this section, the method of a generalized Galerkin finite element formulation [20,21] was used to incorporate both equations of the fluid flow and equations of the particle motion into a single variational equation where the hydrodynamic interactions are eliminated. In addition, for the system considered here, both the electric field and the slipping flow boundary conditions for the liquid phase require that the interface of the liquid and the particle be defined clearly. In this work, an ALE method [23] is used to track the position of the particle at each time step. We used the same procedures as used by Hu [20]: The nodes on the particle surface and in the interior of the particle are considered to move with the particle. The moving velocity of the nodes in the interior of the liquid is computed using Laplace's equation to guarantee a smoothly varying distribution of the nodes. At each time step, the grid is updated according to the motion of the particles and is checked for element degeneration. If unacceptable element distortion is detected, a new
588
Electrokinetics in Microfluidics
finite element grid is generated and the flow fields are projected from the old grid to the new grid. The first order Euler backward method is used to discretize the temporal term. We use GAMBIT (a mesh generator developed by Fluent Inc.) to generate unstructured tetrahedral meshes. A program based on Taylor-Hood tetrahedral element [24,25] has been developed to solve the above set of dimensionless equations. To verify the numerical simulation programs, we considered a test case: a non-conducting spherical particle carrying uniform surface charge is freely suspended in the center of a large circular cylindrical channel. An external electric field is applied along the axis of the channel. The channel wall carries no surface charge and the radius of the channel, b, is 30 times larger than the radius of the particle, a. Therefore, the channel wall has a negligible boundary effect on the particle motion. Furthermore, we assume that the electrical double layer surrounding the particle is very thin, i.e., Ka -> oo, and that the double layer is not distorted by the applied electric field. The conditions of this test case match that of Henry's solution for the electrophoresis of a spherical particle in unbounded solution. For KM—»oo, according to Henry's formula, the electrophoretic velocity of the particle is given by: (77) Using this Uf, as a characteristic velocity, the calculated dimensionless velocity for the spherical particle in the large channel is Vp = —^- = 0.9988664. The discrepancy between the calculated result and Henry's solution is approximately 0.113%. Furthermore, the numerical simulation program was tested against Keh's solution. Keh et al [6] considered the electrophoretic motion of a rigid sphere with a radius a along the axis of a circular cylindrical pore of a radius b. Considering a thin electrical double layer, they derived the following approximate solution by a method of reflections:
(78)
Electrophoretic Motion of Particles in Microchannels
589
Figure 9.25. Effect of the applied electric potentials on the electric field: (a) y/* = C* = 0.0 and (b) y/* =C* =1.0, with y/" = A' =1.0, y/* = B* =0.0 and a =1.0.
590
Electrokinetics in Microfluidics
where U^ is the velocity of the spherical particle. From the above equation, the following dimensionless particle velocity can be derived:
(79)
It can shown that the discrepancy between the numerical solution results and Keh's solution is less than 1.0% when — is within the range from — to — . b 4 30 In the system studied here, the electric field (i.e., the gradient of the electric potential) is the driving force for both the particle motion and the flow field in the channel. In addition, since the size of the channel is close to that of the particle, the boundary effects are significant and the particle motion is dependent on its size relative to that of the channel. In order to study the particle size effects, we considered two particles of different sizes: one has a radius a and the other has a radius 2a. For both cases, we used the same characteristic length scale l~a, thus the dimensionless particles' radii become a -\ and a -2, respectively, for these two particles. The dimensionless sizes of a T-shaped rectangular channel used in the numerical calculations are shown in Figure 9.23. As seen in the figure, the channel is composed of three sub-channels with a cross-section of 10x10. The initial position of the particle center is set at (19,0,0). This particle-channel system is symmetrical to both the z = 0 plane and to the y = 0 plane. When three electrical potentials are applied at the inlet and two outlets respectively, the electric field (i.e., the gradient of the electric potential) will be in x and y directions, and so will the particle motion. Therefore, in the following discussion, we present the electric field and the flow field on the z = 0 plane. Effects of the applied Electric field In the numerical calculations, the applied electric potentials at the inlet and the outlet 1 are set at i// -A
= 1.0 and y/ = B = 0.0 respectively, while the
applied electric potential at the outlet 2 (i// = C ) changes from 0.0 to 1.0. A particle with dimensionless radius a' = 1 is suspended in the channel. The electric field on the z = 0 plane is shown in Figure 9.25, where the lines with arrows denote the direction of the electric field, and the gray levels denote the
Electrophoretic Motion of Particles in Microchannels
591
Figure 9.26. Effect of the suspended particle on the electric field: (a) without a particle and (b) with a particle (whose dimensionless radius a' = 2.0) suspension, with y/* = A' =1.0, V/* = B' = 0.0 and y/* = C* = 0.5.
592
Electrokinetics in Microfluidics
magnitude of the electric field with the darker areas representing the weak electric field and the lighter areas representing the strong electric field. It is shown that the applied electric potentials have a great influence on the electric field in the T-shaped junction region. The electric field is symmetrical in Figure 9.25(a) while it is unsymmetrical in Figure 9.25(b). As mentioned before, the geometry of the particle-channel system is symmetrical to the y - 0 plane. When the electric potentials applied at two outlets are of the same value, the electric potential distribution is symmetrical to the y - 0 plane. On the other hand, when the electric potentials applied at two outlets are of the different values, the electric potential distribution is not symmetrical to the y - 0 plane. Effect of the suspended particle The suspended particle provides a non-conducting boundary for the electric field. Figure 9.26 shows the electric fields on the z = 0 plane for two different cases: (a) without a particle, and (b) with a particle of a dimensionless radius a = 2. In these two cases, the same electric potentials are applied at the inlet and the two outlets: y/ -A =1.0, y/ =B = 0.0 and y/ =C =0.5, respectively. Comparing Figure 9.26(a) with Figure 9.26(b), it is seen that the presence of a particle distorts the electric field in the region adjacent to the particle, which will have an effect on the particle motion. Particle motion As mentioned before, the particle motion is a combined result of the particle electrophoresis and the electroosmotic flow field in the channel. Both the electrophoresis of the particle and the flow field depend on the applied electric field. In addition, since the size of the channel is close to that of the particle, the effects of the non-conducting boundaries, i.e., the channel wall and the particle surface, on the electric field and the flow field will affect the particle motion in the T-shaped microchannel. The initial position of the particle center is at (19,0,0) and the particle initial velocity is zero. When the electric potentials are applied at the inlet and two outlets of the channel, the particle will begin to move. As mentioned before, in this particle-channel system, the geometry is symmetrical to the z - 0 plane. When three electrical potentials are applied at the inlet and two outlets, the electric field will be in x and y directions, and so will the particle motion. Therefore, the particle translates on the z = 0 plane. In the following sections, we will discuss the track of the particle motion in the channel on the z = 0 plane and the decomposed translational velocity components of the particle in x and y directions. In the calculations, the following conditions are used: s = 80.1,
Electrophoretic Motion of Particles in Microchannels
593
Figure 9.27. Flow velocity vector plot on the z = 0 plane at t* = 2 4 with y/' = A' =1.0, yz* =B' = 0 . 0 , i//' =C* = 0 . 5 , y = 5.
594
Electrokinetics in Microfluidics
Thus the coefficient in Eq. (75) is a = 1.0. It should be noted that, in the calculations, the total nodal number ranges from 40,000 to 50,000 and the number of the tetrahedral elements ranges from 28,000 to 33,000. It is very time-consuming to calculate the elemental matrices, solve the electric field and the flow field, calculate the mesh moving velocity and update the mesh, etc., at each time step. Since our goal here is to examine the influence of the different parameters on the trend of the particle motion, we will not calculate the particle motion from its initial position to the exit of the channel. We only calculate the particle motion with limited time steps as far as the trend of the particle motion under different influences can be shown. Figure 9.27 shows the flow velocity vector plot on the z = 0 plane at t =24 with
Figure 9.28. Effect of the applied electric potential on the particle motion. The lines carrying different symbols denote the path of the particle moving in the junction region during a period ofdimensionlesstime t* = 44, 1//* = A' = 1.0, y* = B' = 0.0, a" = 1.0, y = 5.
Electrophoretic Motion of Particles in Microchannels
595
Effect of the applied electric potentials The applied electric field is the driving force for the electroosmosis and the electrophoresis. Without the electric field, there is no liquid flow and particle motion in the channel. To calculate the influence of the applied electric potentials on the particle motion, we chose a particle with a dimensionless radius a = 1, and we set the ratio of the zeta potential of the channel to that of the particle as y =gw/gp
=5. The time step (Vt ) is Vt =2.0. The applied
electric potentials at the inlet and the outlet 1 are chosen as y/ -A
=1.0 and
y/ = B =0.0 respectively, while the applied electric potential at the outlet 2 (y/ =C ) changes from 0.0 to 0.25, 0.5 and 1.0. The trace of the particle motion under the different applied electric potentials are shown in Figure 9.28,
Figure 9.29. Effect of the zeta potential ratio on the particle motion. The lines carrying different symbols denote the path of the particle moving in the junction region during a period of dimensionless time t* = 66, y/* =A* =\.0, y/* = B' = 0 . 0 , y>* =C* = 0.5 , a =1.0.
596
Electrokinetics in Microfluidics
where the lines carrying different symbols denote the path of the particle moving in the junction region during a period of dimensionless time t - 44. As shown in the figure, the particle motion is greatly influenced by the applied electric potentials. It can be concluded that the direction of the particle motion in the Tjunction can be controlled by the direction of the local electric field and that the particle moves faster when the local electric field is stronger. Effect of the zeta potentials of the channel and the particle The zeta potentials of the channel and the particle are critical for the electroosmosis and the electrophoresis. In the extreme case, if both the channel and the particle carry no surface charges (i.e., £ w = 0 and C,p =0), there is no liquid flow and the particle motion in the channel even if an electric field is applied along the channel. The effect of the zeta potentials of the channel and the particle on the particle motion can be seen from the effect of the zeta potential ratio, 7 = Cw ^p • Three cases are considered here: (a) y = 5.0; (b) y = 3.0; and (c) y = 0.5. The applied electric potentials at the inlet and the two outlets are the same for all three cases, i.e., y/ = A =1.0, y/ = B - 0.0 and y/ = C =0.5 respectively. A particle with a dimensionless radius a = 1 is chosen. The time step (Vt ) is set as Vt =2.0. The trace of the particle motion under the different ratios are shown in Figure 9.29, where the lines carrying different symbols denote the trace of the particle moving in the junction region during a period of dimensionless time t = 66. As shown in the figure, the ratio, y, has a great influences on the particle motion. For the first two cases (i.e., y - 5.0 and y - 3.0) where the value of the zeta potential of the channel is bigger than that of the particle, the particle moves in the same direction as the electroosmotic flow, and the bigger the ratio, the faster the particle moves. However, for the case of y = 0.5 where the value of the zeta potential of the channel is smaller than that of the particle, the particle moves in the opposite direction to the electroosmotic flow. This can be understood as follows: the particle motion is the result of the electrophoresis coupled hydrodynamically with the electroosmosis. When both the channel and the particle carry negative surface charges, electrophoresis will be in the direction against the electric field while the electroosmosis will be along the electric field. For thin EDL, the electrophoresis mobility of a particle in an unbounded solution is given by \xeph = sC,p //J., and the electroosmotic mobility of the liquid in the channel is given by pieof = sC,w /JX. Therefore the ratio y =C,JC,P is also the ratio of mobility neof/fJ-eph- By definition, the mobility is the velocity of electrophoretic or electroosmotic motion under a unit applied electrical field strength. When y >1.0, i.e., the magnitude of the electrophoretic mobility is less
Electrophoretic Motion of Particles in Microchannels
597
than that of the electroosmotic mobility. The electroosmotic flow velocity is larger than the particle's electrophoretic velocity. Therefore the particle will by carried by the liquid electroosmotic flow to move in the electroosmotic flow direction but with lower speed than the electroosmotic flow. When y <1.0, the magnitude of the electrophoretic mobility is larger than that of the electroosmotic mobility. The particle's electrophoretic velocity is larger than the liquid's electroosmotic flow velocity. Therefore, the particle will overcome the resistance of the electroosmotic flow and move in the opposite direction to that of the electroosmotic flow but with a speed lower than that of electrophoretic velocity. When y = 1.0, the mobility of the electrophoresis has the same magnitude as that of the electroosmosis. However, because these two motions are in the opposite directions and have the same speed, the particle will be stationary when both electroosmotic flow and electrophoretic motion are fully developed. It is easy to understand what will happen if the particle carries the same amount of charge with opposite sign to that of the channel walls, for example, y = -1.0. In such a
Figure 9.30. Effect of the particle size on the particle motion in the same channel. The lines carrying different symbols denote the dimensionless particle translational velocities at different time, with y/* = A' = 1.0, y/' = B' = 0.0, y/' = C* = 0.5, y = 5.
598
Electrokinetics in Microfluidics
case, the electrophoresis is in the same direction as the electroosmosis, the particle will move with the electroosmotic flow at a higher speed. From Figure 9.30, it is clear that the direction and the speed of the particle motion are also controlled by the zeta potential ratio, y. Effect of the particle size As mentioned before, the presence of the non-conducting boundaries (i.e., the channel walls and the particle surface) directly influences the electric field and the flow field, and the electric field and the flow field will transfer the boundary influences to the particle motion. To calculate the influence of the particle size on the particle motion, two cases are considered here: (a) a particle with a dimensionless radius a -1; and (b) a particle with a dimensionless radius a - 2. In these two cases, the same applied electric potentials at the inlet and the two outlets are used: y/ -A =1.0,y/ =B = 0.0 and y/ =C =0.5 respectively, and the ratio of the zeta potential of the channel to that of the particle is chosen as y = 5. The time step is Vt = 2.0. The numerical results of the dimensionless translational velocity of the particle are shown in Figure 9.30, where Vpx is the velocity component in x direction and Vpy is the velocity component in y direction. As shown in the figure, the particle motion depends on its size slightly, the smaller particle moves faster under the same conditions. This is due to the boundary effects.
Electrophoretic Motion of Particles in Microchannels
9-4
599
TWO PARTICLES IN A RECTANGULAR MICROCHANNEL
In practical applications of many lab-on-a-chip devices, it requires manipulating the motion of multiple particles through relatively small microchannels. In such a situation, the electric field and the flow field surrounding one particle will be influenced by the presence of other particles nearby. Consequently, the velocity and the path of the particles will be influenced. As a first step to study multiple particle systems, this section considers the three-dimensional transient electrophoretic motion of two particles in a straight rectangular microchannel [26]. The objective is to examine how the electric field, the flow field surrounding one particle and the particle velocity are influenced by the presence of a neighbouring particle. Theoretical model We considered two spherical particles moving in a straight rectangular microchannel. Both ends of the channel are connected respectively to two reservoirs which are open to the atmosphere. The channel and the two reservoirs are filled with an electrolyte solution. Therefore, this is an open system with no overall pressure gradient. An external electric field is applied along the channel. In order to focus on the multiple-particle effects, we assume that the particles have the same density as that of the solution so that the gravitational effects are negligible. Both the particles and the channel surface carry uniform negative surface charges, which are characterized by their respective zeta potentials: £ „ and £ w . When the particles are brought into contact with an electrolyte solution, there are electrical double layers surrounding the particles and next to the channel wall. Under an externally applied electric field, the electrophoretic motion of the particles will be in the direction against the electric field (i.e., the gradient of the electric potential), while the electroosmotic flow of the liquid will be along the electric field. To save computational time, instead of the whole channel, we consider only a region where two particles are suspended. The geometry of the channel is shown in Figure 9.31. Cartesian coordinates (x, y,z) are used in the analysis, with the axis x in the horizontal direction and the axis y in the vertical direction. In this system, two particles generally will move at different velocities since they have different sizes or different electrokinetic properties, the distance between the two particles will change with time. The change of the distance will result in the change of the electric field, the flow field and particle motion unless the distance is sufficiently large. Therefore this is a transient process. To simplify the model developments and analysis, the following four assumptions have been made: (1) the particles and the channel wall are rigid and non-conducting; (2) the aqueous electrolyte solution is
600
Electrokinetics in Microfluidics
Figure 9.31. The schematic diagram of the geometry and the dimensionless parameters of the channel.
Newtonian and incompressible, and the flow is a Stokes flow; (3) the electric field is developed instantly; (3) the distances between two particles or between the particles and channel wall are not too small (larger than 10 nm) so that the van der Waals force and the electric double layer repulsion force can be neglected; and (4) the electrical double layers are thin. To develop the model, we divide the liquid phase into two regions: an inner region that is defined as the electrical double layers adjacent to the particle and the channel wall and an outer region that is defined as the remainder of the liquid. For the inner region, the characteristic length scale is the Debye length, K~ . Under the assumption of thin electrical double layers, the electroosmotic flow velocity is used to describe the flow in the inner regions adjacent to the particles and the channel wall. In this model, Q denotes the domain of the outer region, Tw denotes the outer edge of the electrical double layer adjacent to the channel wall, F p \ and F p 2 denote the outer edge of the electrical double layers adjacent to two particles respectively, and Tin and Tout denote the inlet and the outlet of the channel respectively. In the outer region, the ionic concentrations
Electrophoretic Motion of Particles in Microchannels
601
are uniform and the local volume charge density is zero. The model governing the distribution of electric potential and the flow field in the outer region and the particle motion are described below. Distribution of the electric potential In the outer region, the ionic concentrations are uniform and the local volume net charge density is zero. The distribution of the electric potential is governed by (80) with the following boundary conditions:
where y/ is the applied electric potential and n is the unit normal vector pointing into the liquid phase. Flow field Under the consideration of thin electrical double layers, the electroosmotic flow velocities are employed to describe the flow in the inner region adjacent to the particles and the channel wall. In the outer region, since the local net charge density is zero, no electrostatic force acts on the liquid. The liquid flow in the outer region is governed by the continuity equation and Navier-Stokes equation. (81) (82) where v is the liquid flow velocity vector, p and r\ are respectively the density and the viscosity of the electrolyte solution, and p is the pressure in the outer region. The flow boundary conditions for flow field in the outer region are
602
Electrokinetics in Microfluidics
where e is the dielectric constant of the electrolytic solution, SQ is the permittivity of vacuum (8.85xl(T12 C/Vm), I is the unit dyadic, Vp\, Vp2, a>pi, a>p2, Xpi, xp2, Xpi, Xp2, Cpi and C,p2 are the translational velocity, the angular velocity, the position vector on the particle surface, position vector of the particle center and zeta potential for particle 1 and particle 2, respectively. Particle motion Similar to the analysis presented in the last section, we can show that the equations governing the particles' translational motion are given by:
(83)
(84)
(85)
(86) The rotational motion of the particle is governed by: (87)
(88)
Electrophoretic Motion of Particles in Microchannels
603
Figure 9.32. Different particle positions on the j*=4 plane: (a) Case 1; (b) Case 2; (c) Case 3; (d) Case 4; (e) Case 5; and (f) Case 6.
604
Electrokinetics in Microfluidics
where J 1 ; F/JOJ and T/,j are respectively the moments of inertia, the hydrodynamic force and the torque (about the particle centre) on particle 1 by the flow field in the outer region, and J2, Fhol anc^ ^hl a r e respectively the moments of inertia, the hydrodynamic force and the torque (about the particle centre) on particle 2 by the flow field in the outer region. The hydrodynamic force and the torque are given by (89) (90) (91) (92) where a is the stress tensor that is given by (93)
In order to non-dimensionalize the equations, we chose l-a\ {a\ is the radius of particle 1) as a characteristic length, the electric potential <j> = y/ r as a '•in BE A
(h
characteristic electric potential, and U = — - £ w — as a characteristic velocity. l 1 Letting x = lx , v = Uv , p = ^—p , I/A = 0y/ and t = ——t , we can derive /
77
the following dimensionless equations for the electrical field and the flow field: (94) (95) (96)
Electrophoretic Motion of Particles in Microchannels
605
The dimensionless boundary conditions becomes
where / j and y 2 are respectively the ratio of the zeta potential of the particles to that of the channel, and A is the ratio of radius of particle 2 to that of particle 1, i.e., A = alla\.
mpl = ppl3mp*,
Defining J2 - A ppl
J
(pp
is
the
mp2 = A3 ppl3mp*, density
of
the
particles),
Jx=ppl5J*, a = -—a ,
ho\ =rlUlFhoX, Fho2 =r)UlFlo2, fhx =t]Ul2f^ and fh2 =rjUl2fh2,we can write the dimensionless governing equations for the particle motion as follows:
(97)
(98)
(99)
(100)
606
Electrokinetics in Microfluidics
(101)
(102)
The quantities with a star in the above equations are the dimensionless variables. Numerical method For such a two-particle system, the particle motion, the flow field and the electric field are coupled in the following way: the particle motion provides the moving boundaries for the flow field and the electric field; the flow field exerts the hydrodynamic force and toque on the particles; and the electric field provides the slipping flow boundary conditions for the flow field. In order to solve the complicated system of coupled equations, numerical simulation is required. To solve the above couple equations, the method of a generalized Galerkin finite element formulation [20,21] was used to incorporate both equations of the fluid flow and equations of the particle motion into a single variational equation where the hydrodynamic interactions are eliminated. For the system considered here, both the electric field and the slipping flow boundary conditions for the liquid phase require that the interface of the liquid and the particles be defined clearly. An ALE method [23] is used to track the position of the particles at each time step. The following procedures were used: The nodes on each particle surface are considered to move with the particles respectively. The moving velocity of the nodes in the interior of the liquid is computed using Laplace's equation to guarantee a smoothly varying distribution of the nodes. At each time step, the grid is updated according to the motion of the particles and is checked for element degeneration. If unacceptable element distortion is detected, a new finite element grid is generated and the flow fields are projected from the old grid to the new one. The first order Euler backward method is used to discretize the temporal term. GAMBIT (a mesh generator developed by Fluent Inc.) was used to generate unstructured tetrahedral meshes. A program based on Taylor-
Electrophoretic Motion of Particles in Microchannels
607
Hood tetrahedral element [24,25] has been developed to solve the above set of dimensionless equations. Here we will consider two kinds of initial particle velocity conditions: one is zero velocity and the other is the steady-state velocity. The steady-state velocity is determined by solving Eqs. (80)~(82) without the temporal term in Eq. (82). In the system studied here, the electric field (i.e., the gradient of the electric potential) is the driving force for both the particle motion and the flow field in the channel. In addition, since the size of the channel is close to that of the particle, the boundary effects may be significant and the particle motion may depend on its size relative to that of the channel. The dimensionless parameters of a straight rectangular channel (16x8x11) used in the numerical calculations are shown in Figure 9.31. In the calculations, the following conditions are used:
the dimensionless time via / = 2.5x10" t s. The centers of the particles are initially located on the j*=4 plane. Since the electric field is applied along the channel (in x direction), and the particle-channel system is symmetrical to the y*=4 plane, both centers of the particles will move along the y*=4 plane. Therefore, in the following discussion, we present the electric field and the flow field on the y*=4 plane. Figure 9.32 shows the initial positions of the particles for different cases. Electric field The suspended particles provide non-conducting boundaries for the electric field. Figure 9.33 shows the electric fields on the y*=4 plane for three different cases: (a) one particle, (b) two particles with the separation d*=1.0 (d* = d/al, where d is the separation distance between two particles, and al is the radius of particle 1) and (c) two particles with the separation d*=2.0. In this figure, the lines with arrows denote the direction of the electric field and the grey levels denote the magnitude of the electric field. Comparing Figure 9.33(a) with Figure 9.33(b-c), it can be seen that the electric field in the region adjacent to the bigger particle is distorted by the presence of the smaller particle. In addition, by comparing Figure 9.33(b) and Figure 9.33(c), one can see that the influence of the smaller particle on the electric field adjacent to the bigger particle is weaker when the separation distance between two particles is larger.
608
Electrokinetics in Microfluidics
Figure 9.33. The electric field of (a) one particle, (b) two particles with the separation d*=l.O, and (c) two particles with the separation d*=2.0. The lines with arrows denote the direction of the electric field and the grey levels denote the magnitude of the electric field. The other parameters used in this figure are A = 2, yx = y2 = 0.9 and Ex = 62.SWIm.
Electrophoretic Motion of Particles in Microchannels
609
Figure 9.34. The flow field of (a) one particle, (b) two particles with the separation d*=\.O, and (c) two particles with the separation d*=2.0 where the vectors denote the flow velocity and the lines with arrows denote the streamlines. The other parameters used in this figure are A = 2, y, =y 2 =0.9 and Em =62.5kV/m.
610
Electrokinetics in Microfluidics
Flow field It is well known that the electroosmotic flow is directly dependent on the electric field. In this model, the electroosmotic flow is described by the slipping flow boundary velocity. The slipping flow boundary velocity for the flow field will change with the change of the electric field. In addition, the presence of moving particles provides moving boundaries for the flow field. Therefore the flow field depends on the electric field, the particles' positions and the particles' motion. Figure 9.34 shows the flow field on the y*=4 plane for three different cases: (a) one particle, (b) two particles with the separation d*=1.0 and (c) two particles with the separation d*=2.0, where the zeta potential ratios (the zeta potential of the particle to that of the channel wall) were chosen as 7j = Y2 ~ 0-9 • In the figure, the vectors denote the flow velocity and the continuous lines with arrows denote the streamlines. As is shown, the streamlines and velocity vectors surrounding a single particle are symmetric in Figure 9.34(a) but unsymmetrical in Figure 9.34(b-c). The flow field on the right side of the bigger particle is distorted by the presence of the smaller particle. In addition, by comparing Figure 9.34(b) and Figure 9.34(c), it can be seen that the influence of the presence of the smaller particle on the flow field adjacent to the bigger particle is weaker as the separation distance between two particles increases. Particle motion As discussed above, the presence of one particle exerts influences on the electric field and the flow field adjacent to the other particle, which will influence the particle motion. The effects of particle size, particle zeta potential, particle position and the separation distance on the particle motion are discussed below. Effect of particle size To discuss the effects of particle size on particle motion, Case 2 (Figure 9.32(b)) is considered: the centers of Particle 1 and Particle 2 are initially set at (7, 4, 3) and (7, 4, 7) respectively, and The radius of Particle 2 is twice of Particle 1. We assume the particles' initial velocity is zero, and the zeta potential ratios yl=y2= 0.2. After the electrical field (62.5KV/m) is applied, the particles start moving. The dimensionless translation velocities for Particle 1 and Particle 2 in the x direction reach 0.04831 and 0.04803 respectively at f=80. In dimensional terms, the velocities are 2.400mm/s for the small particle (al = 5\im) and 2.386mm/s for the larger particle {a.2 - lOum) at 0.002s. The smaller particle (Particle 1) moves a little faster than the bigger one (Particle 2). Though the difference between these two velocities is small, after a certain time, the two
Electrophoretic Motion of Particles in Microchannels
611
particles can be separated in the x direction. For example, at /*=8000, i.e., t=0.2s, these two particles can be separated by d*=2.6 (ord= 13 |^m) in the x direction. Effect of particle zeta potential The zeta potentials of the channel wall and the particles are critical for the electroosmosis and the electrophoresis. Under the same electric field, the electroosmosis and the electrophoresis will be in the opposite direction if both the channel wall and the particles are negatively or positively charged. The particle motion is the result of the electrophoretic motion coupled with the electroosmotic flow field in the channel. Case 3 (Figure 9.32(c)) is considered: the centers of Particle 1 and Particle 2 are initially set at (7, 4, 3) and (7, 4, 8) respectively, Particle 2 has the same size as Particle 1. The zero velocity is used as the initial condition, and we set the zeta potential ratios as /j = 0.2 and Y2 = 0.4. At t*=64, the dimensionless translation velocities for Particle 1 and Particle 2 in the x direction reach 0.04960 and 0.03706 respectively. In dimensional terms, the velocities are 2.464mm/s for Particle 1 and 1.841mm/s for Particle 2 after 0.0016s. Particle 1, with the lower zeta potential, moves faster than Particle 2 in the electroosmotic flow direction. When C,p < £w (i.e., y < 1),
Figure 9.35. Paths of Particle 1 and Particle 2 for Case 3 from /*=0 to /*=288, where the dash circles and the solid circles respectively show the initial and the final positions of the particles. The parameters used in this figure are X = 1, yx = 0.2, y 2 = 0.4, and Em = 62.5kV/m.
612
Electrokinetics in Microfluidics
the electroosmotic mobility is larger than the electrophoretic mobility, thus the particle motion will be in the same direction as the electroosmotic flow. It can be concluded that, within the range of 0 < y < 1, the smaller the value of y, the bigger the particle moving velocity. Figure 9.35 shows the tracks of Particle 1 and Particle 2 from /*=0 to f=288. It can be seen that at t*=288, the position of Particle 1 is ahead of Particle 2, and the separation distance between two particle's centers is 4.5um. Effect of the particle position To discuss the effects of the particle position on the particle motion, Case 4 and Case 5 (Figure 9.33(d-e)) are considered. The larger Particle 2 is ahead of the smaller Particle 1 in Case 4 and the larger Particle 2 is behind of the smaller Particle 1 in Case 5. The centers of two particles are initially set at (5, 4, 5.5) and (9, 4, 5.5) respectively. The radius of Particle 2 is twice of Particle 1. The separation distance for both cases is d*=l.O. We used the following zeta potential ratios: y\ = 0.2 (fixed) and Y2 changes from 0.1, 0.2 to 0.8, respectively. The velocities for Particle 1 and Particle 2 are shown in Table 2. It can be seen that the difference between particle velocities is very small (within 0.1%). It can be concluded the effect of the particle position, i.e., the bigger particle ahead or behind of smaller one, has a negligible effects on the particle moving velocity. It can also be seen from Table 2 that Particle l's velocity is influenced by the change of the velocity of Particle 2: Particle l's velocity increases with Particle 2's velocity. Table 2 A list of particle velocities for different particle positions with y, = 0.2. Case 4 (Figure 3d)
Case 5 (Figure 3e)
0.1
0.05092
0.05446
0.05092
0.05442
0.2
0.04995
0.04852
0.04995
0.04847
0.8
0.04410
0.01291
0.04412
0.01290
Electrophoretic Motion of Particles in Microchannels
613
Effect of the separation distance To examine the effects of the separation distance, Case 1 and Case 4 (Figure 9.33(a, d)) are considered, where The radius of Particle 2 is twice of the Particle 1 and the separation distances are respectively d*=0.3 and d*=\.O. We let 7] = 0.2 and y2 change from 0.1, 0.2 and 0.8 respectively. Table 3 shows the calculated particle velocities for different separation distance. In order to compare the effects, we calculated the velocity for Particle 1 in the channel without the presence of Particle 2, and the velocity is 0.04993. Comparing this velocity with the velocities listed in Table 3, one can see that the separation distance does affect the particle motion. With the increase of the separation distance, the effect of the separation distance weakens. In addition, this effect on Particle 1 's velocity is dependent on Particle 2's velocity. For example, when the difference between Vp2 and Vp\ is small as in the case of y 2 = 0.2, the presence of Particle 2 has a negligible effect on Particle l's velocity. However, when the difference between Vp2 and fp\ is big as in the case of y2 = 0.8, the velocity of Particle 1 is significantly influenced by the presence of Particle 2. A faster particle climbing and passing a slower particle In this part, we want to investigate and demonstrate how one particle passing the other particle. Case 6 (Figure 9.33(f)) is considered: the initial positions of the two particles' centers are set at (7, 4, 4.8) and (3, 4, 7.2) respectively and the radius of Particle 2 is twice larger than that of Particle 1. Table 3 A list of particle velocities for different separation distances with yx = 0.2. Case 1 (d*=0.3)
Case 4 (d*= 1.0)
72
VpX
Vp2
Vpl
Vp2
0.1
0.05159
0.05429
0.05092
0.05446
0.2
0.04924
0.04854
0.04995
0.04852
0.8
0.03512
0.01404
0.04410
0.01291
614
Electrokinetics in Microfluidics
The steady state velocity is used as the initial condition and we set j \ = 0.2 and Y2 = 0-8. The results show that Particle 1 moves faster than Particle 2. The paths of two particles during a climbing and passing process from t*=0 to t*=880 are shown in Figure 9.36. It can be seen that Particle 1 climbs up and passes Particle 2, and then climbs down, and Particle 2 climbs down a little during the first half process and then climbs up a little during the second half process. Here the phenomena of climbing up and climbing down would be explained by the unsymmetrical flow field surrounding the particle. Figure 9.37(a) shows the flow field at the state of t*=272 during the first half process. It can be seen that the flow field surrounding the particles is not symmetrical in the z-direction, and the unsymmetrical flow field pushes Particle 1 up while pushes the Particle 2 a little down in the z-direction. Figure 9.3 7(b) shows the flow field at the state of t*=600 during the second half process. As is shown, the flow field surrounding the particles is not symmetrical in the z-direction, and it draws Particle 1 down while draws the Particle 2 a little up in the z-direction.
Figure 9.36. The tracks of particles during a climbing and passing process from t*=0 to t*=880. The dash circles and the solid circles respectively show the initial and the final positions of a small particle and a large particle. The curves denote the tracks of the particles' center positions during the process. The other parameters used in this figure are X = 2 , =62.5kV/m. ft =0.2, n =0.8, and £ „
Electrophoretic Motion of Particles in Microchannels
615
Figure 9.37. The flow field: (a) t*=272, and (b) t*=600, where the vectors denote the flow velocity and the lines with arrows denote the streamlines. The other parameters used in this figure are A = 2, yl = 0.2, y2 = 0.8, and Em = 62.5W I m.
616
Electrokinetics in Microfluidics
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] II1] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26]
H. Shintani and J. Polonsky, "Handbook of Capillary Electrophoresis Applications", Blackie Academic & Professional, London, 1997. P.D. Grossman and J.C. Colburn, "Capillary Electrophoresis: Theory and Practice", Academic Press, San Diego, 1992. R.J. Hunter, "Zeta Potential in Colloid Science Principal and Applications", Academic Press, New York, 1981. J. Lyklema, "Fundamentals of Interface and Colloid Science", Academic Press, San Diego, 1991. T.G.M. van de Ven, "Colloidal hydrodynamics", Academic Press, San Diego, 1988. H.J. Keh and J.L. Anderson, J. Fluid Mech., 153 (1985) 417. J. Ennis and J.L. Anderson, J. Colloid Interface Sci., 185 (1997) 497. A.A. Shugai and S.L. Carnie, J. Colloid Interface Sci., 213 (1999) 298. J. Ennis, H. Zhang, G. Stevens, J. Perera, P. Scales and S. Carnie, J. Membrane Sci. 119, (1996)47. A.L. Zydney, J. Colloid Interface Sci., 169 (1995) 476. E. Lee, J.-W. Chu and J.-P. Hsu, J. Colloid Interface Sci., 196 (1997) 316. C. Ye and D. Li, J. Colloid Interface Sci, 251 (2002) 331-338. C. Ye, D. Sinton, D. Erickson and D. Li, Langmuir, 18 (2002) 9095-9101. C. Ye and D. Li, J. Colloid Interface Sci, (in press). J.N. Israelachvili, "Intermolecular and Surface Forces", Academic Press, San Diego, 1991. R. Hogg, T.W. Healy and D.W. Fuerstenau, Trans. Faraday Soc, 62 (1966) 1638. D.C. Henry, Proc. Roy. Soc, London, 133 (1931) 106. H. Ohshima, J. Colloid Interface Sci, 168 (1994) 269. J. Happel, Chapter 7 in Low Reynolds Number Hydrodynamics (2nd ed.), Noordhoff International Publishing, Leyden, 1973. H. Hu, Int. J. Multiphase Flow, 22 (1996) 335-352. R. Glowinski, T.-W. Pan, T.I. Hesla and D.D. Joseph, Comput. Methods Appl. Mech. Engrg.,25(1999)755. J.B. Ritz and J.P. Caltagirone, Int. J. Numer. Mech. Fluids, 33 (1999) 1067. T. Nomura and T.J.P. Hughes, Comput. Methods Appl. Mech. Engrg, 95 (1992)115. C. Johnson, "Numerical Solution of Partial Differential Equations by the Finite Element Method", Cambridge University Press, New York, 1987. O.C. Zienkiewicz and R.L. Taylor, "The Finite Element Method", ButterworthHeinemann, Oxford, Boston, 2000. C. Ye and D. Li, Proceeding of 2nd International Conference on Microchannels and Minichannels, Rochester, New York, June 17-19, 2004.
Microfluidic Methods for Measuring Zeta Potential
617
Chapter 10
Microfluidic methods for measuring zeta potential Zeta potential is an electrokinetic potential at the shear plane (the boundary between the compact layer and the diffuse layer) near a solid-liquid interface where the liquid velocity is zero. Zeta potential is a very important interfacial electrokinetic property to a huge number of natural phenomena, such as electrode kinetics, electrocatalysis, corrosion, adsorption, crystal growth, colloid stability and flow characteristics of colloidal suspensions and electrolyte solutions through porous media and microchannels. For example, zeta potential is a key parameter in determining the interaction energy between particles and hence the stability of colloid suspension systems. In microfluidics applications, as we have demonstrated in the previous chapters, particularly for electroosmotic flows through microchannels, the zeta potential will critically influence the velocity. From the modeling and simulation point of view, we must know the zeta potential values used in the boundary conditions. The surface conductance, another important parameter, usually is referred to the electrical conductivity through the electrical double layer region, i.e., a thin liquid layer near the solid-liquid interface where there is a net charge accumulation due to the charged solid-liquid interface. It is not difficult to understand that surface conductance may have a significant effect on streaming potential (e.g., lOOmV/cm) in pressure-driven flow of dilute aqueous electrolyte solutions in small microchannels. However, for electroosmotic flow, the effect of surface conductance on the electrical field in the microchannel is generally negligible due to the fact that the applied electrical field is usually high (e.g., lOOV/cm). In many cases, knowing the surface conductance is a must in order to evaluate the zeta potential and other electrokinetic properties correctly. Measurement of the surface conductance, therefore, is important in the studies of electrokinetic phenomena. Most techniques for measuring zeta potentials are based on electrophoresis and streaming potential measurements [1-3]. Within the scope of this book, two methods based on electrokinetic flows in microchannels to measure the zeta potential will be reviewed in this chapter. One of these methods involves measurement of the streaming potential in pressure-driven flow, and the other involves electroosmotic flow.
618
Electrokinetics in Microfluidics
10-1 STREAMING POTENTIAL METHOD As discussed in Chapter 2, the rearrangement of the charges on the solid surface and the balancing charges in the liquid is called the electrical double layer (EDL). Immediately next to the solid surface, there is a layer of ions that are strongly attracted to the solid surface and are immobile. This layer is called the compact layer, normally about several Angstroms thick. Because of the electrostatic attraction, the counterions concentration near the solid surface is higher than that in the bulk liquid far away from the solid surface. The coions' concentration near the surface, however, is lower than that in the bulk liquid far away from the solid surface, due to the electrical repulsion. Therefore, there is a net charge in the region close to the surface. From the compact layer to the uniform bulk liquid, the net charge density gradually reduces to zero. Ions in this region are affected less by the electrostatic interaction and are mobile. This region is called the diffuse layer of the EDL. The thickness of the diffuse layer is dependent on the bulk ionic concentration and electrical properties of the liquid, usually ranging from several nanometers for high ionic concentration solutions up to several microns for pure water and pure organic liquids. The boundary between the compact layer and the diffuse layer is usually referred to as the shear plane. The electrical potential at the solid-liquid surface is difficult to measure directly. The electrical potential at the shear plane is called the zeta potential, <;, and can be measured experimentally. Due to the adsorption of the counterions on the surface, i.e., the compact layer, one cannot measure the true electrical potential of the solid surface. In practice, the zeta potential is used as an approximation to the potential at the solid-liquid interface. The ion and electrical potential distributions in the electrical double layer can be determined by solving the Poisson-Boltzmann equation. As shown in Chapter 2, for symmetric electrolytes, the local net charge density per unit volume pe at any point in the solution is proportional to the concentration difference between cations and anions, and is given by the Boltzmann equation: (1) Correspondingly, the EDL field is described by the Poisson-Boltzmann equation: (2)
Microfluidic Methods for Measuring Zeta Potential
619
where s is the dielectric constant of the solution, nx and z are the bulk ionic concentration and the valence, respectively, e is the charge of a proton, Kb is the Boltzmann constant, and T is the absolute temperature. When a liquid is forced through a capillary under an applied hydrostatic pressure, the counterions in the diffuse layer (mobile part) of the EDL are carried towards the downstream end, resulting in an electrical current in the pressuredriven flow direction. This current due to the transport of charges by the liquid flow is called the streaming current. Corresponding to this streaming current, there is an electrokinetic potential called the streaming potential. This flow induced streaming potential is a potential difference that builds up along the capillary. This streaming potential acts to drive the counterions in the diffuse layer of the EDL to move in the direction opposite to the streaming current, i.e., opposite to the pressure-driven flow direction. The action of the streaming potential will generate an electrical current called the conduction current. It is obvious that when ions move in a liquid, they will pull the liquid molecules to move with them. Therefore, the conduction current will produce a liquid flow in the opposite direction to the pressure driven flow. The overall result is a reduced flow rate in the pressure drop direction. If the reduced flow rate is compared with the flow rate predicted by the conventional fluid mechanics theory without considering the presence of the EDL, it seems that the liquid would have a higher viscosity. This is usually referred to as the electro-viscous effect. Generally, the zeta potential can be determined from the measured pressure drop and the measured streaming potential along a single capillary tube of circular cross-section. In absence of an applied electric field, when a liquid is forced through a channel under hydrostatic pressure, the excess counter-ions in the diffuse layer of the EDL are carried by the liquid to flow to the downstream, forming an electrical current, i.e., the streaming current. The streaming current is given by (3)
where Ac is the cross-section of the capillary (Ac = nR2 in this case), v is the local velocity of the liquid, and pe is the local net charge density. For a liquid flowing through a cylindrical capillary tube of radius R at a steady state, the liquid local velocity v(r) is given by the solution of the Poiseuille equation: (4)
620
Electrokinetics in Microfluidics
where r is the radial variable measured from the central of the capillary, AP is the pressure drop along the capillary, t] is the viscosity of the liquid, and L is the length of the capillary. The local net charge density pe(r) in Eq. (3) can be replaced by Eq.(l). Realizing that local net charge density pe(r) is not zero only in the EDL region, i.e., only in a thin region near the channel wall, and that the EDL potential at the channel wall is approximated to be the zeta potential C,, one can show that the streaming current is given by: (5)
As explained above, the streaming potential generates a conduction current in the reversed direction. For a cylindrical capillary, the conduction current is given by:
(6) where Ab is the bulk liquid electrical conductivity, and Es is the streaming potential. At a steady state, the net electrical current should be zero, i.e., Is + Ic = 0. Using this condition, one obtains the following: (7)
As seen from Eq. (7), by measuring the streaming potential and the pressure drop, and knowing the liquid properties, the zeta potential can be determined. Generally, for a given solid-liquid system at a given temperature, the zeta potential is considered as a constant, this is because the nature of the material, its surface charge, and the nature and the concentration of the electrolyte determine the electrokinetic potential. However, early studies found that zeta potential determined from Eq.(7) depends on the size of the capillary tubes for the same solid material (the capillary wall) and the same liquid (with the same electrolyte concentration). This problem was solved later by considering the surface conduction, i.e., the electrical conduction through a thin layer at solid-liquid interface. Generally, the surface conduction is the excess conduction tangential to a charged surface, and originates from the excess counterions' concentrations in the EDL region near the solid-liquid interface [3]. Particularly in the cases of low bulk ionic concentrations, the surface conduction will have a significant contribution to the total conduction current through the capillary tube. Usually,
Microfluidic Methods for Measuring Zeta Potential
621
the surface conductivity, Xs, is considered as the conductivity of a sheet of material of negligible thickness, with a unit ohrrT'nT1. Specific surface conductivity values are of the order 10~9 ~ 10~8 for water in glass capillaries. The surface conductivity is expected to have a significant effect on the zeta potential in capillaries smaller than 1 mm in diameter at concentrations below 1(T3-5M. Briggs [4] suggested a simple procedure to correct Eq. (7). This method requires the measurement of the actual electrical resistance of the liquid in the capillary, Rexp, and the comparison of this value with calculated electrical resistance at high concentration, Rcai, where the surface conduction is expected to be negligible. The modified Eq. (7) becomes: (8)
Briggs' method does not give explicit information on surface conductance. Rutgers [5] was first to rigorously consider the surface conductance effect on the zeta potential and solved the problem of zeta potential dependence on capillary size. By considering the surface conductance, the conduction current for a cylindrical capillary of radius R and length L is given by: (9) In the above equation, the first term accounts for the conduction through bulk liquid, and the second term accounts for the surface conduction. Using Eq.(5) and Eq.(9), and the steady state condition, Is + Jc = 0, we have: (10) Realize s -ers0, where sr and £0 are the dielectric constant of the liquid and the permittivity of vacuum, respectively, the above equation can be rearranged into the following form: (11)
622
Electrokinetics in Microfluidics
r — versus — for a slit channel (or —for a r]XbEs h R cylindrical capillary) by using Eq. (12) (or Eq.(l 1)) to determine the zeta potential and the surface conductivity by the streaming potential method.
Figure 10.1. The linear relationship of
For a given solid-liquid system, sr, r\, Xb, and £ are constant. Equation (11) clearly shows that the term (srs0AP)l{r]XbEs) is a linear function of 1/R. If we plot (ers0AP)/(r]XbEs)\s. 1/R, as shown in Figure 10.1, the intersection of the line given by Eq.(l 1) with the vertical axis is \lq, and the slope of this line is (2Xs)l{g A.b). This implies that if one measures the pressure drop and the streaming potential of the same solid (capillary wall) - liquid system for different capillary radii, and plot the data according to Eq.(l 1), the zeta potential £and the surface conductivity As can be determined from the intercept and the slope of the line. In additional to cylindrical capillary tubes, another type of capillary with a simple cross-section shape is the slit microchannel, i.e., a channel formed between two parallel plates. For an electrolyte solution flowing through a slit channel, it can be shown that, in analogy to Eq.(l 1), (12)
Microfluidic Methods for Measuring Zeta Potential
623
where h is the height of the slit channel. As discussed above, the key in the streaming potential method is to measure the streaming potential and pressure drop of the same solid-liquid system for different capillary sizes. Figure 3.14 illustrates an experimental setup for such measurements, developed in the Laboratory of Microfluidics and Labon-a-Chip, Department of Mechanical & Industrial Engineering, University of Toronto. The capillary used in the test may be a single capillary tube or a bundle of capillary tubes of the same dimensions, or a slit channel formed between two parallel plates. The capillary tube is placed in a two-part symmetrical Plexiglas assembly to form a test cell as shown in Figure 3.14. Epoxy resin is applied to bond the capillary tube and the assembly together to avoid leaking. Two pairs of sumps are machined in the assembly for the pressure drop and the streaming potential measurements. For a given capillary tube and a given electrolyte solution, the precision pump is set to maintain a constant flow rate. The pressure drop across the capillary tube is monitored and recorded by a differential pressure transducer (±0.5% FS accuracy, Validyne Eng. Corp.) and a computer data acquisition system. The flow was considered to have reached a steady state when the readings of the pressure drop do not change any more. Such a constant pressure drop value corresponds to the AP in Eq.(l 1) or (12). At this steady state, two electrodes (e.g., Ag/AgCl electrodes, Sensortechnik Meinsberg GmbH) and a high resistance electrometer (Keithley Instruments Inc.) are used to measure the streaming potential. The streaming potential measured at this steady state corresponds to the Es in Eq.(ll) or (12). It should be emphasized that such a streaming potential measurement should not last more than 60 or 100 seconds (depending on the electrolyte concentration and the flow rate) to avoid the polarization of the electrodes. In addition to the problem of electrode polarization, it should also be noted that there are no two identical electrodes due to the material and the manufacturing process. Therefore, in order to obtain more reliable and repeatable results, under the same pump setting (i.e., the same flow rate), the flow into the capillary is switched in an opposite direction after the previous streaming potential measurement. When the flow reaches the steady state again, the above described pressure drop and streaming potential measurement are repeated. Usually, for one flow rate, the operation described above, switching the flow direction and measuring the pressure drop and the streaming potential, should be repeated several times to ensure the repeatability and good average values of the pressure drop and the streaming potential. It is recommended to repeat the above measurements for several different flow rates. As explained above, in order to determine the zeta potential and the surface conductivity by using Eq.(ll) or (12), one must measure the pressure drop and the streaming potential of the same solid-liquid pair for different capillary sizes. Therefore, the above-described measurements should be
624
Electrokinetics in Microfluidics
Figure 10.2. Experimental data for a glass surface with (a) De-ionized Ultra-filtered water, (b) 10~3 M aqueous NaCl solution.
625
Microfluidic Methods for Measuring Zeta Potential
conducted for cylindrical capillary tubes with different radii or slit channels with £ £ AP
different heights. Finally, by plotting —— r\XhEs
1
1
vs. —(or—) as illustrated in R
h
Figure 1, one can determine the zeta potential and the surface conductivity from the intercept and the slope of the best-fitted line to the experimental data points. As an example, the results of glass surface-NaCl solution systems are replotted here [6] in Figures 10.2 and 10.3. The measurements were done by using a slit channel formed by two parallel glass plates separated by two strips of a plastic shim of uniform thickness. Using plastic shims of different thickness will change the slit channel height. The channel height can be measured accurately by using a microscope. In Figures 10.2, each data point represents an average value of several measurements as described above for the same channel height. For a given channel height the three data points represent the measurements for three different flow rates. Figure 10.3 shows the measured zeta potential and the surface conductivity as functions of the NaCl concentration. This method can also be used to study the effects of temperature, pH, different electrolytes and surfactants on zeta potential and surface conductivity.
Figure 10.3. Variation of the bulk conductivity and the surface conductivity with NaCl concentration.
626
Electrokinetics in Microfluidics
It should be pointed out that Eq.(ll) or (12) is derived on the basis of Eq.(4) (or the equivalent form for a slit channel), the solution of the classical Poiseuille equation. The Poiseuille equation, however, does not consider any EDL or electrokinetic (e.g., electro-viscous) effect on the liquid flow. For small capillaries, the electro-viscous effects on flow may be significant and have to be considered, as discussed in Chapter 3. The Eq. (11) or (12) is no longer valid for small capillaries and may cause a significant error when used to evaluate the zeta potential and the surface conductivity. Since the electro-viscous effect on flow is negligible for flow through a capillary with a hydraulic diameter larger than 50 microns, it is recommended that above-described method should be used with capillary tubes of a hydraulic diameter larger than 50 microns.
Microfluidic Methods for Measuring Zeta Potential
627
10-2 ELECTROOSMOTIC FLOW METHOD This section will introduce a method for measuring the ^"-potential by combining the Smoluchowski equation with the measured slope of current-time relationship in electroosmotic flow [7]. It is obvious that the measurement of the total electrical current in electroosmotic flow is relatively simple and easy in comparison with other zeta potential measurement techniques based on electrophoresis and streaming potential methods. Because this approach uses electroosmotic flow, as opposed to pressure driven flow in the streaming potential case, it is more representative of the true operating conditions of a microfluidic device. This approach also allows the use of flat surfaces, which are more flexible in terms of surface treatment (for example micro-contact printing) than capillary tubes.
Figure 10.4.
Illustration of a slit microchannel.
Consider an electroosmotic flow of an aqueous solution in a slit microchannel (formed between two parallel plates). Generally most electroosmotic flows, owing to the inherently low Reynolds number, reach a steady state very quickly and the entrance length tends to be extremely short. For the case examined here the time to reach a steady state is on the order of 10 microseconds and the entrance length is approximately lOum. In addition the slit channel geometry, where the width is much larger than the height, w » h, minimizes the edge effects and makes the velocity profile essentially uniform along the x-axis. Therefore, the flow could be assumed as steady, onedimensional and fully developed. If the electrical double layer is thin or the channel height is large in comparison with the double layer thickness, it can be shown from the Smoluchowski equation that the average electroosmotic flow velocity is given by [1]:
628
Electrokinetics in Microfluidics
(13) where Ez is the applied electric field strength, so and sr are the relative dielectric constant and the electrical permittivity of a vacuum respectively, /j is the solution viscosity, and C, is the zeta potential. If the average velocity of an electroosmotic flow can be determined, the zeta potential can then be determined from the above equation. In Chapter 7, we have introduced the slope method to determine the average velocity of electroosmotic flow. During the electroosmotic flow of one solution displacing another similar solution in a microchannel, the measured current-time relationship (except the beginning and the ending) is linear as long as the concentration difference between these two solutions is small. The slope of this current-time relationship is given by: (14) where AI and At are the change in current and time taken from the linear range. In electro-osmotic flow, the total current consists of three components: the bulk conductivity current, Icond,buik, the surface conduction current, Icond,surf and the convection current, Icom. Generally, the convection current can be neglected when determining the total current, since it is several orders of magnitude smaller than the other two current components. Then, under an applied electrical field, Ez, the total current can be shown as: (15) Substituting Eq. (15) for the difference of the current (note AXS = 0), Eq. (14) can be rewritten as: (16) where (Xb2 - Xbl) is the difference of the bulk conductivity between the high concentration solution and the low concentration solution. Rearranging Eq. (16) yields:
Microfluidic Methods for Measuring Zeta Potential
629
Figure 10.5. Schematics of (a) the microchannel cell (cross-section) (b) experimental set-up used for liquid handing (c) the electro-osmotic flow measuring system.
630
Electrokinetics in Microfluidics
(17) The above expression shows that the average electroosmotic flow velocity can be determined by measuring the slope of the current-time relationship. Once the average velocity is determined, we can use the vav value and Eq. (13) to determine the ^-potential. In other words, combining Eq.(13) and Eq. (17) gives: (18)
Figure 10.5 shows an experimental setup to implement this method. There are three major parts: a microchannel cell (Figure 10.5a), a liquid handling system (Figure 10.5b), and an electroosmotic flow monitoring system (Figure 10.5c). The microchannel cell itself consists primarily of a parallel-plate microchannel confined between an inlet reservoir block and an outlet reservoir block. The parallel-plate microchannel is formed by two flat sample surfaces. A positioning clamp is placed to keep the surfaces parallel and to limit their deflection when exposed to pressure (during the flushing step as explained below). The two equal sized reservoirs form the inlet and the outlet to the microchannel. In a recent work [7], the parallel-plate microchannels were formed by using two identical testing surfaces (15.5 mm x 37 mm) separated by two thin plastic shims (Small Parts, Inc., Florida, USA) to form the channel sidewalls. The shims were placed along the lateral edges of the cleaned surfaces and fixed by applying a small amount of epoxy (Devcon, USA), leaving a flow passage 3.5mm wide. Once formed the channel was allowed to dry in room temperature air for 16 hours. After that, the width and the length of the microchannel will be measured using a precision gauge (Model CD-6"B, Mitutoyo Co., Japan) with an accuracy of ± 1 urn. Finally, the microchannel was mounted in the supporting blocks, as illustrated in Figure 10.5a, to form the microchannel cell. A positioning clamp was then used to maintain the channel position. As it is the most sensitive dimension in terms of introducing systematic error into the experimental results, the channel height must be measured accurately. In the reported work [7], the channel height was measured by two independent techniques, direct digital imaging analysis and volume flow rate calibration. Using a microscope and video camera system (Model 4815-5000,
Microfluidic Methods for Measuring Zeta Potential
631
Figure 10.6. Current versus time for lO^M KCl solution in a glass channel of 217 |im under an applied voltage of (a) 400 V, (b) 600 V, and (c) 800 V over the 37mm length of the channel.
632
Electrokinetics in Microfluidics
Cohu; Wild Heerbrugg M7-5, Switzerland) with a 40x objective the channel height and parallel could be measured to an accuracy of ± 0.8 |am prior to being placed in the experimental system. The channel heights used in that study range from 200 to 300 um. The height measurements at the channel inlet and outlet were found to be consistent within ± 2 %. Once placed in the experimental system the channel height was also calibrated by using pressure driven flow of a highly concentrated electrolyte solution (e.g., 10"2 M) to minimize any electrical double layer effects. For a steady fully-developed laminar flow, the channel height h can be related the measured pressure drop, AP, and volume flow rate, Q, via:
The pressure drop was measured using a differential pressure transducer (Validyne, Model CD23, Engineering Co., CA, USA) and volume flow rate through mass accumulation at the channel exit by using an electronic balance (Mettler Instrument AG, Model BB240) with an accuracy of 1 mg. The accuracy of the flow rate measurement was estimated to be ± 2 %. The discrepancy between the results of the channel height from the above two measuring methods was found to be within 2%, suggesting that an accurate estimate of the true channel height had been obtained. A liquid handling system has been designed (Figure 10.5b) to clean the system after each run and introduce new solutions. Prior to the experiment, pure water is pumped from a solution reservoir to the flow loop and though the microchannel for at least half an hour by a high precision pump (Masterflex, Model 7550-60, Barnant Co., IL, USA), which has a flow rate range of 0.6 to 2900 ml/min. Then the flushing process was repeated by using the test solution until the electrical conductivity of the test solution existing the microchannel reaches and remains the standard value for the specific electrolyte solution at the given concentration. The bulk electrical conductivity is measured using a conductivity sensor (Model Inpro 7001/120, Mettler Toledo Process Analytical Inc., MA, USA), which is connected to a conductivity/resistivity transmitter (CR7300, Mettler Toledo Process Analytical Inc., MA, USA). The electroosmotic flow monitoring system, shown in Figure 10.5c, consists primarily of the measuring cell with electrodes, a high voltage power source (CZE 1000R, Spellman High Voltage Electronics Co., NY, USA), and a PGA-DAS 08 data acquisition card (OMEGA Engineering, Quebec, Canada). During the experiments, the high voltage power source was used to apply a potential difference between the two reservoirs via platinum electrodes. The
Microfluidic Methods for Measuring Zeta Potential
633
Figure 10.7. Current versus time for 10"3 M KCl solution in a glass channel of 217 um under an applied voltage of (a) 400 V, (b) 600 V, and (c) 800 V over the 37mm length of the channel.
634
Electrokinetics in Microfluidics
output voltage and current were recorded on a personal computer, via the data acquisition card, using a 0~10 V output signal from the power supply. Glass surfaces to be tested were prepared from microscope slides, which had been cut and carefully polished, to a dimension of 15.5 mm by 37 mm with an accuracy of ± 0.5 mm. The glass plates were soaked in acetone (Caledon Laboratories Ltd., ON, Canada) for 12 hours and then vigorously washed with acetone several times. After that, the surfaces were submerged in deionized ultra-filtered water (DIUF) (Fisher Scientific, Canada) for 12 hours. The surfaces were then dried under a heat lamp before use. Many microfluidic devices have been fabricated by using poly(dimethylsiloxane) (PDMS) [8-10]. PDMS is a bulk polymer consisting of repeated units of -OSi(CH 3 ) 2 O-. It is a popular material for building microfluidic devices for a number of reasons: (1) It is much less expensive then glass, and not as fragile as glass. (2) Channels can be formed by molding or embossing rather than etching. (3) It is optically transparent down to 280 nm so it can be used for a number of detection schemes. (4) It cures at low temperatures. (5) It can be sealed reversibly to itself and a range of other materials, such as glass, by making molecules contact with the surface. (6) It is nontoxic - mammalian cells can be cultured directly on it. For the PDMS-coated glass surfaces, each glass plate was fixed to a glass dish lined with parchment paper. PDMS base and curing agent (Sylgard 184, Dow Corning, Midland, MI, USA) were thoroughly mixed in a ratio of 10:1 and then poured onto the cleaned surfaces. The PDMS was then degassed at room temperature for one hour to permit the removal of air bubbles. The glass dishes were then put in an oven at 65°C for 30 minutes. After cooling, the PDMScoated glass surfaces were removed from the dishes, and cleaned with acetone and DIUF water. Aqueous KC1 and aqueous LaCl3 solutions were chosen for these experiments for two reasons. Firstly, the electrokinetic characteristics of these solutions in contact with various solid surfaces have been well studied using a variety of other techniques, allowing for a direct comparison with the results obtained here. Secondly, the effect of the valence and the symmetry of the ions on the flow characteristics in a microchannel could be examined, since the KC1 solution has ion distribution, and the valence of the La+ ion is 3. The electrolyte solutions were prepared by dissolving KC1 (Bioshop Canada Inc., ON, Canada) or LaCl3-H2O (Fisher Scientific Co., NJ, USA) in DIUF water, which has a pH = 6.5 and Xb = 1.16 u-Scm"1 at 25°C. The bulk liquid conductivity of the solutions were measured using a high-precision symmetrical 1:1 ions (the valence is 1), while the LaCl3 has an unsymmetrical conductivity meter with an accuracy of ± 0.5 % of its reading. The measured conductivity of the aqueous solutions used here were 10~4M KC1 with Xb = 16.3
Microfluidic Methods for Measuring Zeta Potential
635
Figure 10.8. ^-Potential versus applied voltage and channel height, for glass surfaces in (a) 10~4 M KC1, (b) 10~3 M KC1 (c) 10~4 MLaCl3 (d) 1(T3 M LaCl3 aqueous solutions.
636
Electrokinetics in Microfluidics
Figure 10.9. Current versus time for 10 4 M KCl solution in a PDMS-coated glass channel of 229 um under an applied voltage of (a) 200 V, (b) 400 V, and (c) 600 V over the 37mm length of the channel.
Microfluidic Methods for Measuring Zeta Potential
637
uScnf1, 10~3 M KC1 with Xb = 146.0 uScm"1, 10"4M LaCl3 with lb = 41.1 uScm"1, and 10"3M LaCl3 with Xb = 410.8 uScm"1. Prior to introducing a testing solution into the apparatus the microchannel cell and the entire piping system was bi-directionally rinsed with DIUF water at for one hour in order to remove any contaminants (such as electrolyte remaining from previous experiments). Following that a second 15-minute rinsing cycle was initiated using the desired test solution. The transparent nature of the microchannel surfaces allowed for visual inspection of the channel to ensure that all bubbles had been removed. Once the flushing was complete, the microchannel cell was detached from the liquid handling system and the temperature, pH level, and bulk conductivity of the solution in the reservoirs were measured while any height difference between the fluid in the two reservoirs was allowed to equilibrate. The content of reservoir 1 was then diluted to 95 % of the test solution by pipetting a predetermined amount of DIUF water (an equivalent volume was also removed from the reservoir to ensure the system remained at mechanical equilibrium). Once both reservoirs and the microchannel were filled with electrolyte, the electroosmotic flow measurement began. Immediately after placing the platinum electrodes into each reservoir and ensuring that they were aligned perpendicular to the channel, a voltage difference was applied. The lower concentrated electrolyte from reservoir 1 thus gradually displaced the higher concentrated electrolyte in the microchannel increasing the overall channel resistance and thus decreasing the current draw, which was monitored via the data acquisition system. The voltage was applied for no longer than one minute, after which the power supply was switched off and the electrodes were removed from the microchannel cell. All experiments were repeated in triplicate and conducted at room temperature, 25°C. As examples, Figures 10.6 and 10.7 show the current change with time for 10"4 M and 10"3 M KC1 solutions with glass surfaces at three different applied voltages. In all cases measurements were done 3 times to confirm the repeatability. As seen from these figures, a linearly decreasing relationship between current and time is evident for the first 20 seconds of each set. In the low applied voltage cases the current ultimately reaches a constant value when the high concentration solution in the microchannel has been completely replaced by the lower concentration solution, such as Figure 10.6a. In the high concentration solution and high electrical field strength cases, such as Figure 10.7c, the current draw became unstable at the end of the experiment. It is believed that the fluctuation of the current value is due to the joule-heating effect. For these experiments then the slope current vs. time was taken over the first 20 seconds of the displacement process in order to avoid this complication.
638
Electrokinetics in Microfluidics
Figure 10.10. ^-Potential versus applied voltage and channel height, for PDMS surfaces in (a) 10"4 M KC1, (b) 10"3 M KC1 (c) 10"4 MLaCl3 (d) 10"3 M LaCl3 aqueous solutions.
Microfluidic Methods for Measuring Zeta Potential
639
The slopes of the current-time relationship shown in the figures were used to determine the ^-potential of the glass plate surface using Eq. (18), and the following physical properties of DIUF water were used: £r = 80, s0 = 8.854 x 10~12 CV'm" 1 , and n = 0.001 Nsm"2 [11]. Figure 10.8 shows calculated C,potential as a function of applied voltage and channel height for (a) 10~4 M KC1, (b) 10"3 M KC1, (c) 10~4 M LaCl3, and (d) 10"3 M LaCl3 solutions. As expected the ^-potential shows no dependence on channel height or the magnitude of the applied field and the results were found to be repeatable within ± 5 %. From these results the ^-potential was found to be -88 mV for 10"4 M KC1, -78 mV for 10"3 M KC1, -77 mV for 10"4 M LaCl3, and -66 mV for 10~3 M LaCl3. These values are similar to those obtained for glass surfaces using the electrophoresis technique [12] and the streaming potential technique [6]. For the PDMS-coated surfaces, a sample experimental result of current vs. time for the 10~4 M solution in a microchannel with height of 229 um is plotted in Figure 10.9. As before, measurements under the same conditions were repeated three times to confirm the reproducibility, which was found to be within ± 3 %. For the PDMS coated surfaces the input voltage was reduced to a maximum of 600V and thus the current fluctuation could be eliminated in most cases. As before the slopes of the current-time relationship were measured over the initial 20-second time period. Figure 10.10 shows the ^-potential calculated by Eq.(18) against the applied voltage for (a) 10"4M KC1, (b) 10"3 M KC1, (c) 10"4M LaCl3 (d) 10 3 M LaCl3 solutions. Similar to the glass surface, the ^-potential shows no dependence on channel height or the magnitude of the applied field and the measured results are reproducible within ± 6 %. Using these results the measured ^-potential values were -110 mV, -87 mV, -82 mV, and -68 mV for 10"4 M KC1, 10"3 M KC1, 10"4 M LaCl3, and 10"3 M LaCl3 solutions, respectively. Examination of Figures 10.8 and 10.10 indicates that the magnitude of the C,potential for a PDMS-coated glass surface is higher than that for a glass surface under same conditions. This is in agreement with the findings of other authors [13,14] for similar polymer surfaces.
640
Electrokinetics in Microfluidics
REFERENCES [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14]
RJ. Hunter, Zeta Potential in Colloid Science: Principle and Applications. Academic Press, London, 1981. J. Lyklema, Fundamentals of Interface and Colloid Science, Vol. II, Solid-Liquid Interfaces, Academic Press, London, 1995. J. Lyklema, Fundamentals of Interface and Colloid Science, Vol. II, Solid-Liquid Interfaces, Academic Press, London, 1995. D. K. Briggs, J. Phys. Chem., 32 (1928) 641. A.J. Rutgers, Trans. Faraday Soc, 36 (1940) 69. Y. Gu and D. Li, J. Colloid Interface Sci., 226 (2000) 328-339. A. Sze, D. Erickson, L. Ren and D. Li, J. Colloid Interface Sci., 261 (2003) 402-410. A. Kumar and G.M. Whitesides, App. Phys. Lett, 63 (1993) 2002. C.S. Effenhauser, G.J.M. Bruin, A. Paulus and M. Ehrat, Anal. Chem., 69 (1997) 3451. R.S. Martin, A.J. Gawron, S.M. Lunte and C.S. Henry, Anal. Chem., 72 (2000) 3196. R. Weast, M.J. Astle and W.H. Beyer, CRC Handbook of Chemistry and Physics, CRC Press, Florida, 1986. R.S. Sanders, R.S. Chow and J.H. Masliyah, J. Colloid Interface Sci., 174 (1995) 230. A. Voigt, H. Wolf, S. Lauckner, G. Neumann, R. Becker and L. Richter, Biomaterials, 4(1983)299. C. Werner, H. Korber, R. Zimmermann, S. Dukhin and H.J. Jacobasch, J. Colloid Interface Sci., 208 (1998) 329.
Subject Index
641
Subject Index A AC electroosmotic flow, 119 -150 asymmetric electrolytes/ions, 15, 16, 102, 153 asymmetric roughness, 323, 324, 333, 334, 336, 337, 339, 344, 350 - 352
B Boltzmann distribution, 12 -15, 45, 74 - 77, 79, 87, 89, 98, 105, 119, 124, 133, 152, 208, 234, 239, 299 Boltzmann equation, 13, 33, 75, 76, 95, 360 Bubble lensing, 411 - 426
C caged and uncaged fluorescent dye, 376, 378 - 381, 383, 385, 386, 389, 390, 392, 394, 402 charge origin, 8 - 1 0 chemical potential, 12, 13 compact layer, 10, 11 concentration distribution, 14, 20, 21, 25, 84, 85, 215, 352, 470, 471, 475, 477, 479,489,516,521 conduction current, 31. 32, 38, 39, 52, 53, 56,58,82,210,619-621,628 current monitoring method, 355 - 375, 403 -410
D Debye-Hiickel linear approximation, 19, 23, 35, 36,46, 47, 92, 96, 127, 306 diffuse layer of electric double layer, 10, 11 double layer thickness, 10, 15, 17 - 19, 34, 35, 53, 55, 56, 62, 64, 67, 70, 72, 73, 83, 87 dynamic sample loading, 496 - 509
E electric double layer, 7 - 2 8 electrokinetic phenomena, 28 electroosmosis, 28, 92 - 201, 23 8 - 317, 323 -352, 356-459,627-639 electroosmotic mobility, 188, 268, 274, 276, 284, 286, 290, 293, 298, 299, 327, 340, 342 -346, 387, 395, 410, 418, 425, 434, 444, 468, 475, 477, 483, 489, 513, 514, 515,518,520,596,612 electroosmotic mobility ratio, 346, 348 350, 352 electroosmotic flow, 28, 92 - 201, 238 317,323-352, 3 5 6 - 4 5 9 , 6 2 7 - 6 3 9 electroosmotic pumping, 94 - 118 electroosmotic velocity, 28, 92 - 201, 238 317,323-352, 3 5 6 - 4 5 9 , 6 2 7 - 6 3 9 electrophoresis, 28, 542 - 615 electrophoretic mobility, 291, 329, 340, 342, 343 - 345, 377, 386, 390, 391, 394, 399 - 402, 407, 423, 453, 470, 477, 483, 486, 489, 514, 529, 537, 543, 544, 575, 576, 596, 597, 612 electrophoretic mobility ratio, 330, 345 electrophoretic motion, 542 — 615 electro-viscous effect, 30 - 90 enhanced mixing, 268 - 317
F flow visualization, 376 - 426, 446 - 459,
G gravity effects, 547, 548, 550, 552, 556, 557, 560
H Henry function, 544, 553 heterogeneous microchannels, 204 - 317, 344-352
642
Electrokinetics in Microfluidics
I immobile ions, 10, 11
J Joule heating effects, 117, 118, 1 8 4 - 2 0 1 , 427 - 459
L Lab-on-a-Chip, 1 - 4 laser-induced dye injection, 376 - 410
M measurement techniques, 354 - 460, 617 — 639 micro bubble lensing method, 411 - 426 microchannel network, 298 — 317 microfluidic visualization, 376 - 426 mixing, 268 - 297 mobile ions, 10, 11 molarity, 17 multiple particles, 599 - 615
N Nernst equation, 76, 79 Nernst-Planck equation, 14, 218, 236, 252 net charge density, 8, 10, 15, 20, 22, 33, 37, 43 - 48, 62, 72, 77, 78, 87, 92, 94, 95, 98 - 108, 111. 119. 133, 140, 144, 152 - 154, 157, 208, 210, 215, 235, 238, 240 - 242, 246, 260, 266, 299, 300, 326, 327,467,468,618-620 non-symmetric electrolyte, 153
P parallel plates or slit microchannels, 18, 32 — 43, 63 - 90, 94 - 97, 215 - 237, 251 267,323-352 particle motion, 546 — 615 Peclet number, 218, 252
pH, 9, 26, 27, 80, 84, 381, 387, 394, 407, 410, 412, 436, 449, 485, 486, 489, 514, 529, 629, 634, 637 photobleaching method, 411 - 426 point of zero charge, 26 Poisson equation, 8, 15, 16, 22, 23, 44, 77, 78, 94, 98, 105, 133, 152, 186, 208, 224, 239, 254, 299, 325, 326, 467, 565 Poisson-Boltzmann equation, 10, 16, 33, 34, 45, 46, 76, 96, 102, 105, 106, 124, 125, 134, 145, 153,186,208,618 potential determining ions, 77, 78 pressure-driven flow, 30 - 90, 206 - 237, 309-311,315,316
R Reynolds number, 48, 67 - 69, 89, 222, 228, 232, 236, 253, 266, 270, 324, 325, 327, 329, 468 rectangular microchannel, 44 - 62, 104 118,206-214,599-615 rough microchannel, 321 - 352
S sample dispensing/injection, 463 - 540 slip boundary velocity, 135, 188, 189, 193, 272, 274, 327, 328, 331 - 333, 336, 343 -346,349,468,470,513 slit microchannel, 32 - 43, 63 - 90, 94 - 97, 215-237,251-267,323-352 slope method, 369 - 375, 627 - 639 shear plane, 10, 11, 18 - 20, 23 Smoluchowski equation, 543 solution displacing process, 151 — 183 spatial gradients of electrical conductivity, 510-540 species transfer, 151 - 183, 268 - 297, 323 343, 466 - 540 streaming potential, 28, 32, 36 -40, 42, 52, 82, 83, 88, 208, 210, 224, 226, 617-621 streaming current, 28, 32, 37, 38, 52, 53, 82, 210,619,620
Subject Index
surface charge, 9,10, 18, 22, 25, 27, 80, 81, 84, 254, 256, 258, 260 - 266, 288-297, 346, 348 surface conductance, 53, 56, 70, 617, 620 626 symmetric electrolyte, 12, 15 - 17, 24, 33, 45, 68, 69, 77, 95, 99, 101, 133, 153, 187, 239, 252, 271, 298, 360, 618, 634 T T-shaped microchannels, 268 - 297, 427 445,579-598 two particle effects, 599-615 transient electroosmotic flow, 119 - 150 V van der Waals force, 548, 550 velocity profile, 376 - 402, 411 - 426, 446 459 Z zeta potential, 12, 15, 24 - 27, 617 - 639 zeta potential measurement, 617-639
643
This Page Intentionally Left Blank