Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

Embed Size (px)

Citation preview

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    1/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    2/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    3/328

    The Library of Congress Cataloged the First Issue of This Title as Follows:

    Electroanalytic chemistry: a series of advances, v. 1

    New York, M. Dekker, 1966-

    v. 23 cm.

    Editors: 19661995 A. J. Bard

    1966- A. J. Bard and I. Rubinstein

    1. Electromechanical analysisAddresses, essays, lectures

    1. Bard, Allen J., ed.

    QD115E499 545.3 66-11287

    Library of Congress

    0-8247-4719-4 (v. 22)

    This book is printed on acid-free paper.

    Headquarters

    Marcel Dekker, Inc.

    270 Madison Avenue, New York, NY 10016

    tel: 212-696-9000; fax: 212-685-4540

    Eastern Hemisphere Distribution

    Marcel Dekker AG

    Hutgasse 4, Postfach 812, CH-4001 Basel, Switzerland

    tel: 41-61-260-6300; fax: 41-61-260-6333

    World Wide Web

    http://www.dekker.com

    The publisher offers discounts on this book when ordered in bulk quantities. For

    more information, write to Special Sales/Professional Marketing at the headquar-

    ters address above.

    Copyright nnnn 2004 by Marcel Dekker, Inc. All Rights Reserved.

    Neither this book nor any part may be reproduced or transmitted in any form or by

    any means, electronic or mechanical, including photocopying, microfilming, and

    recording, or by any information storage and retrieval system, without permission

    in writing from the publisher.

    Current printing (last digit):

    10 9 8 7 6 5 4 3 2 1

    PRINTED IN THE UNITED STATES OF AMERICA

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    4/328

    This series is designed to provide authoritative reviews in the field of mod-

    ern electroanalytical chemistry defined in its broadest sense. Coverage is

    comprehensive and critical. Enough space is devoted to each chapter ofeach volume so that derivations of fundamental equations, detailed de-

    scriptions of apparatus and techniques, and complete discussions of im-portant articles can be provided, so that the chapters may be useful without

    repeated reference to the periodical literature. Chapters vary in length and

    subject area. Some are reviews of recent developments and applications of

    well-established techniques, whereas others contain discussion of the

    background and problems in areas still being investigated extensively

    and in which many statements may still be tentative. Finally, chapters on

    techniques generally outside the scope of electroanalytical chemistry, but

    which can be applied fruitfully to electrochemical problems, are included.

    Electroanalytical chemists and others are concerned not only with

    the application of new and classical techniques to analytical problems, but

    also with the fundamental theoretical principles upon which these tech-

    niques are based. Electroanalytical techniques are proving useful in such

    diverse fields as electro-organic synthesis, fuel cell studies, and radical ion

    formation, as well as with such problems as the kinetics and mechanisms of

    electrode reactions, and the effects of electrode surface phenomena,

    adsorption, and the electrical double layer on electrode reactions.

    It is hoped that the series is proving useful to the specialist and non-

    specialist alikethat it provides a background and a starting point for

    graduate students undertaking research in the areas mentioned, and that it

    also proves valuable to practicing analytical chemists interested in learning

    about and applying electroanalytical techniques. Furthermore, electro-

    chemists and industrialchemists with problems of electrosynthesis, electro-

    plating, corrosion, and fuel cells, as well as other chemists wishing to apply

    electrochemical techniques to chemical problems, may find useful material

    in these volumes. A. J. B.

    I. R.

    INTRODUCTION TO THE SERIES

    iii

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    5/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    6/328

    L. DAIKHIN Tel Aviv University, Ramat Aviv, Israel

    STEPHEN W. FELDBERG Brookhaven National Laboratory, Upton,

    New York, U.S.A.

    E. GILEADI Tel Aviv University, Ramat Aviv, Israel

    MARSHALL D. NEWTON Brookhaven National Laboratory, Upton,

    New York, U.S.A.

    JOHN F. SMALLEY Brookhaven National Laboratory, Upton, New

    York, U.S.A.

    GREG M. SWAIN Michigan State University, East Lansing, Michigan,

    U.S.A.

    V. TSIONSKY Tel Aviv University, Ramat Aviv, Israel

    M. URBAKH Tel Aviv University, Ramat, Israel

    v

    CONTRIBUTORS TO VOLUME 22

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    7/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    8/328

    CONTENTS OF VOLUME 22

    Introduction to the Series iii

    Contributors to Volume 22 v

    Contents of Other Volumes xiii

    LOOKING AT THE METAL/SOLUTION INTERFACE

    WITH THE ELECTROCHEMICAL QUARTZ-CRYSTAL

    MICROBALANCE: THEORY AND EXPERIMENT

    V. Tsionsky, L. Daikhin, M. Urbakh, and E. Gileadi

    I. Introduction 2

    A. Is It Really a Microbalance? 3B. Applications of the Quartz Crystal Microbalance 4

    C. The Impedance Spectrum of the EQCM 5

    D. Outline of This Chapter 8

    II. Theoretical Interpretation of the QCM Response 8

    A. Impedance 8

    B. The Effect of Thin Surface Films 12

    C. The Quartz Crystal Operating in Contact

    with a Liquid 16

    D. Quartz Crystals with Rough Surfaces 26

    III. Electrical Double Layer/Electrostatic Adsorption 33

    A. Introduction 33

    B. Some Typical Results 34C. The Potential Dependence of the Frequency 36

    vii

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    9/328

    IV. Adsorption Studies 43

    A. The Adsorption of Organic Substances 43

    B. The Adsorption of Inorganic Species 53

    V. Metal Deposition 60

    A. D eposition on the Same Metal Substrate 60

    B. Early Stages of Metal Deposition on a Foreign

    Substrate 64

    VI. The Influence of Roughness on the Response of the

    QCM in Liquids 70

    A. The Nonelectrochemical Case 71

    B. The Electrochemical Case 76VII. Conclusion 83

    VIII. Appendix 86

    A. Nonuniform Film on the Surface 86

    B. Experimental Remarks 86

    References 94

    THE INDIRECT LASER-INDUCED TEMPERATURE

    JUMP METHOD FOR CHARACTERIZING FAST

    INTERFACIAL ELECTRON TRANSFER: CONCEPT,

    APPLICATION, AND RESULTS

    Stephen W. Feldberg, Marshall D. Newton,and John F. Smalley

    I. Introduction 102

    A. Why Measure Fast Interfacial Electron Transfer

    Rate Constants? And How? 103

    B. Background 104C. The Underlying Principles of the ILIT Method

    The Short Version 106

    D. Definition of Terms 108

    II. The Evolution of the ILIT Method for the Study of Fast

    Interfacial Electron Transfer Kinetics 108A. The Temperature-Jump Approach for Studies of

    Homogeneous Kinetics 108

    Contents of Volume 22viii

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    10/328

    B. The Temperature-Jump Approach for Studies ofInterfacial Kinetics 108

    III. Relevant Electron Transfer Theory: Marcuss

    Description of Heterogeneous Nonadiabatic Electron

    Transfer Reactions 112

    A. Chidseys Approach 112

    B. Temperature Dependence 116

    C. How Well Does the Butler-Volmer Expression

    Approximate Marcuss Formalism? 118

    IV. Analysis of the ILIT Response 120

    A. Response of the Open-Circuit Electrode Potential toa Change in the Interfacial Temperature in the

    Presence of a Perfectly Reversible Redox CoupleAttached to the Electrode Surface 121

    B. The Relaxation of the ILIT Response When the

    Rate of Electron Transfer Is Not Infinitely Fast 126

    C. When Is the ILIT Response Purely Thermal (i.e.,

    Devoid of Kinetic Information)? 126

    D. The Shape of the Ideal ILIT Perturbation 130

    E. Nonidealities of the Shape of the ILIT Perturbation

    and ResponseExtracting the Relaxation Rate

    Constant, km 134

    F. Correlating km to Meaningful Physical Parameters 137

    V. Experimental Implementation of ILIT 143

    A. The Cell 143

    B. The Working Electrode: Preparation and Thermal

    Diffusion Properties 148

    C. Preparation of Self-Assembled Monolayers 150

    D. The Electronics 151

    E. Potential Problems 152

    F. Energetic and Timing Considerations for Single and

    Multiple Pulse Experiments 156

    G. Some Suggested Experimental Protocols 160

    VI. A Few Examples of Measurements

    of Interfacial Kinetics 161

    A. Some Typical Transients 161B. Determining the Value ofko 163

    C. Arrhenius Plots and Evaluation ofDH p and DHk 163

    Contents of Volume 22 ix

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    11/328

    VII. The Potential of the ILIT Approach 166

    VIII. Some Thoughts About Future Experiments 166

    IX. Glossary of Terms 170

    X. Appendix: One-Dimensional Thermal Diffusion into

    Two Different Phases 173

    References 175

    ELECTRICALLY CONDUCTING DIAMOND

    THIN FILMS: ADVANCED ELECTRODE MATERIALS

    FOR ELECTROCHEMICAL TECHNOLOGIESGreg M. Swain

    I. Introduction 182

    II. Diamond Thin Film Deposition, Electrode

    Architectures, Substrate Materials, and Electrochemical

    Cells 185

    III. Electrical Conductivity of Diamond Electrodes 194

    IV. Characterization of Microcrystalline and Nanocrystalline

    Diamond Thin Film Electrodes 195

    V. Basic Electrochemical Properties of Microcrystalline and

    Nanocrystalline Diamond Thin Film Electrodes 201

    VI. Factors Affecting Electron Transfer at Diamond

    Electrodes 212

    VII. Surface Modification of Diamond Materials and

    Electrodes 216

    VIII. Electroanalytical Applications 219

    A. Azide Detection 219

    B. Trace Metal Ion Analysis 221

    C. Nitrite Detection 224

    D. NADH Detection 225

    E. Uric Acid Detection 225

    F. Histamine and Serotonin Detection 226

    G. Direct Electron Transfer to Heme Peptide andPeroxidase 227

    Contents of Volume 22x

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    12/328

    H. Cytochrome c Analysis 228I. Carbamate Pesticide Detection 228

    J. Ferrocene Analysis 229

    K. Aliphatic Polyamine Detection 230

    IX. Electrosynthesis and Electrolytic Water Purification 238

    X. Optically Transparent Electrodes for

    Spectroelectrochemistry 239

    XI. Advanced Electrocatalyst Support Materials 251

    A. Composite Electrode Fabrication and

    Characterization 252

    B. Oxygen Reduction Reaction 259C. Methanol Oxidation Reaction 264

    XII. Conclusions 267

    References 268

    Author Index 279

    Subject Index 295

    Contents of Volume 22 xi

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    13/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    14/328

    CONTENTS OF OTHER VOLUMES

    VOLUME 1

    AC Polarograph and Related Techniques: Theory and Practice,

    Donald E. SmithApplications of Chronopotentiometry to Problems in Analytical

    Chemistry, Donald G. Davis

    Photoelectrochemistry and Electroluminescence, Theodore Kuwana

    The Electrical Double Layer, Part I: Elements of Double-Layer Theory,

    David M. Monhilner

    VOLUME 2

    Electrochemistry of Aromatic Hydrocarbons and Related Substances,

    Michael E. Peover

    Stripping Voltammetry, Embrecht Barendrecht

    The Anodic Film on Platinum Electrodes, S. GilamanOscillographic Polarography at Controlled Alternating Current,

    Michael Heyrovksy and Karel Micka

    VOLUME 3

    Application of Controlled-Current Coulometry to Reaction Kinetics,

    Jiri Janata and Harry B. Mark, Jr.

    Nonaqueous Solvents for Electrochemical Use, Charles K. Mann

    Use of the Radioactive-Tracer Method for the Investigation of the

    Electric Double-Layer Structure, N. A. Balashova and

    V. E. Kazarinov

    Digital Simulation: A General Method for Solving ElectrochemicalDiffusion-Kinetic Problems, Stephen W. Feldberg

    xiii

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    15/328

    VOLUME 4

    Sine Wave Methods in the Study of Electrode Processes, Margaretha

    Sluyters-Rehbach and Jan H. Sluyters

    The Theory and Practice of Electrochemistry with Thin Layer Cells,

    A. T. Hubbard and F. C. Anson

    Application of Controlled Potential Coulometry to the Study of

    Electrode Reactions, Allen J. Bard and

    K. S. V. Santhanam

    VOLUME 5

    Hydrated Electrons and Electrochemistry, Geraldine A. Kenney and

    David C. Walker

    The Fundamentals of Metal Deposition, J. A. Harrison and H. R. Thirsk

    Chemical Reactions in Polarography, Rolando Guidelli

    VOLUME 6

    Electrochemistry of Biological Compounds, A. L. Underwood and

    Robert W. BurnettElectrode Processes in Solid Electrolyte Systems, Douglas O. Raleigh

    The Fundamental Principles of Current Distribution and Mass Transport

    in Electrochemical Cells, John Newman

    VOLUME 7

    Spectroelectrochemistry at Optically Transparent Electrodes; I. Electrodes

    Under Semi-infinite Diffusion Conditions, Theodore Kuwana and

    Nicholas Winograd

    Organometallic Electrochemistry, Michael D. Morris

    Faradaic Rectification Method and Its Applications in the Study ofElectrode Processes, H. P. Agarwal

    Contents of Other Volumesxiv

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    16/328

    VOLUME 8

    Techniques, Apparatus, and Analytical Applications of Controlled-

    Potential Coulometry, Jackson E. Harrar

    Streaming Maxima in Polarography, Henry H. Bauer

    Solute Behavior in Solvents and Melts, A Study by Use of Transfer

    Activity Coefficients, Denise Bauer and Mylene Breant

    VOLUME 9

    Chemisorption at Electrodes: Hydrogen and Oxygen on Noble Metals

    and their Alloys, Ronald Woods

    Pulse Radiolysis and Polarography: Electrode Reactions of Short-lived

    Free Radicals, Armin Henglein

    VOLUME 10

    Techniques of Electrogenerated Chemiluminescence, Larry R. Faulkner

    and Allen J. Bard

    Electron Spin Resonance and Electrochemistry, Ted M. McKinney

    VOLUME 11

    Charge Transfer Processes at Semiconductor Electrodes,

    R. Memming

    Methods for Electroanalysis In Vivo, Jir Koryta, Miroslav Brezina,

    Jir Prada c , and Jarmila Prada cova

    Polarography and Related Electroanalytical Techniques in

    Pharmacy and Pharmacology, G. J. Patriarche,

    M. Chateau-Gosselin, J. L. Vandenbalck,

    and Petr Zuman

    Polarography of Antibiotics and Antibacterial Agents,Howard Siegerman

    Contents of Other Volumes xv

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    17/328

    VOLUME 12

    Flow Electrolysis with Extended-Surface Electrodes, Roman E. Sioda

    and Kenneth B. Keating

    Voltammetric Methods for the Study of AdsorbedSpecies, Etienne Laviron

    Coulostatic Pulse Techniques, Herman P. van Leeuwen

    VOLUME 13

    Spectroelectrochemistry at Optically Transparent Electrodes,II. Electrodes Under Thin-Layer and Semi-infinite Diffusion

    Conditions and Indirect Coulometric Iterations, William H.

    Heineman, Fred M. Hawkridge, and Henry N. Blount

    Polynomial Approximation Techniques for Differential Equations in

    Electrochemical Problems, Stanley Pons

    Chemically Modified Electrodes, Royce W. Murray

    VOLUME 14

    Precision in Linear Sweep and Cyclic Voltammetry, Vernon D. Parker

    Conformational Change and Isomerization Associated with Electrode

    Reactions, Dennis H. Evans and Kathleen M. OConnellSquare-Wave Voltammetry, Janet Osteryoung and John J. ODea

    Infrared Vibrational Spectroscopy of the Electron-Solution Interface,

    John K. Foley, Carol Korzeniewski, John L. Dashbach, and

    Stanley Pons

    VOLUME 15

    Electrochemistry of Liquid-Liquid Interfaces, H. H. J. Girault and

    D. J. Schiffrin

    Ellipsometry: Principles and Recent Applications in Electrochemistry,

    Shimson Gottesfeld

    Voltammetry at Ultramicroelectrodes, R. Mark Wightman andDavid O. Wipf

    Contents of Other Volumesxvi

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    18/328

    VOLUME 16

    Voltammetry Following Nonelectrolytic Preconcentration, Joseph Wang

    Hydrodynamic Voltammetry in Continous-Flow Analysis, Hari

    Gunasingham and Bernard Fleet

    Electrochemical Aspects of Low-Dimensional Molecular Solids, Michael

    D. Ward

    VOLUME 17

    Applications of the Quartz Crystal Microbalance to Electrochemistry,

    Daniel A. Buttry

    Optical Second Harmonic Generation as an In Situ Probe of

    Electrochemical Interfaces, Geraldine L. Richmond

    New Developments in Electrochemical Mass Spectroscopy,

    Barbara Bittins-Cattaneo, Eduardo Cattaneo, Peter Ko nigshoven,

    and Wolf Vielstich

    Carbon Electrodes: Structural Effects on Electron Transfer Kinetics,

    Richard L. McCreery

    VOLUME 18

    Electrochemistry in Micelles, Microemulsions, and Related

    Microheterogeneous Fluids, James F. Rusling

    Mechanism of Charge Transport in Polymer-Modified Electrodes,

    Gyo rgy Inzelt

    Scanning Electrochemical Microscopy, Allen J. Bard, Fu-Ren F. Fan,

    and Michael V. Mirkin

    VOLUME 19

    Numerical Simulation of Electroanalytical Experiments: Recent Advances

    in Methodology, Bernd Speiser

    Electrochemistry of Organized Monolayers of Thiols and RelatedMolecules on Electrodes, Harry O. Finklea

    Contents of Other Volumes xvii

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    19/328

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    20/328

    LOOKING AT THE METAL/SOLUTION INTERFACEWITH THE ELECTROCHEMICAL QUARTZ CRYSTAL

    MICROBALANCE: THEORY AND EXPERIMENT

    V. Tsionsky, L. Daikhin, M. Urbakh, and E. Gileadi

    School of Chemistry

    Raymond and Beverly Sackler Faculty of Exact Sciences, Tel Aviv University

    Ramat Aviv, Israel

    I. INTRODUCTION 2

    A. Is It Really a Microbalance? 3

    B. Applications of the Quartz Crystal Microbalance 4

    C. The Impedance Spectrum of the EQCM 5

    D. Outline of This Chapter 8

    II. THEORETICAL INTERPRETATION OF THE QCM

    RESPONSE 8

    A. Impedance 8

    B. The Effect of Thin Surface Films 12

    C. The Quartz Crystal Operating in Contact

    with a Liquid 16D. Quartz Crystals with Rough Surfaces 26

    III. ELECTRICAL DOUBLE LAYER/ELECTROSTATIC

    ADSORPTION 33

    A. Introduction 33

    B. Some Typical Results 34C. The Potential Dependence of the Frequency 36

    IV. ADSORPTION STUDIES 43

    A. The Adsorption of Organic Substances 43

    B. The Adsorption of Inorganic Species 53

    V. METAL DEPOSITION 60

    A. Deposition on the Same Metal Substrate 60

    B. Early Stages of Metal Deposition on aForeign Substrate 64

    1

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    21/328

    VI. THE INFLUENCE OF ROUGHNESS ON THERESPONSE OF THE QCM IN LIQUIDS 70

    A. The Nonelectrochemical Case 71

    B. The Electrochemical Case 76

    VII. CONCLUSION 83

    VIII. APPENDIX 86

    A. Nonuniform Film on the Surface 86

    B. Experimental Remarks 86

    References 94

    I. INTRODUCTION

    The literature concerning the quartz crystal microbalance (QCM) and its

    electrochemical analogue, the electrochemical crystal microbalance

    (EQCM) is wide and diverse. Many reviews are available in the literature,

    discussing the fundamental properties of this device and its numerous

    applications, including its use in electrochemistry [15]. In this chapter we

    concentrate on electrochemical applications, specifically in studies of

    submonolayer phenomena and the interaction of the vibrating crystal with

    the electrolyte in contact with it.

    A few examples are treated in detail here, and the advantages and

    limitations of the EQCM as a tool for the study of fundamental phenom-

    ena at the metal/solution interface are discussed.When the quartz crystal microbalance was first introduced in 1959

    [6], it represented a major step forward in our ability to weigh matter. Until

    then, routine measurements allowed an accuracy of 0.1 mg, and highly

    sensitive measurements could be made with an accuracy limit of 0.3 Ag

    under well-controlled experimental conditions, (see Ref. 7). The QCM

    extended the sensitivity by two or three orders of magnitude, into the sub-

    nanogram regime.

    Even used in vacuum or in an inert gas atmosphere at ambient

    pressure, the QCM acts as a balance only under certain conditions, as

    discussed below. Then the change of mass caused by adsorption or de-

    position of a substance from the gas phase can be related directly to the

    change of frequency by the simple equation derived by Sauerbrey [6]:

    Df Cm Dm 1

    Tsionsky et al.2

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    22/328

    where Cm is a constant, representing the mass sensitivity, which is related toknown properties of quartz and the dimensions of the crystal, andDm is the

    added mass density, in units of g/cm2.

    A. Is It Really a Microbalance?

    Is the quartz crystal microbalance really a microbalance? For one thing, it

    should rightly be called a nano-balance, considering that the sensitivity of

    modern-day devices is on the order of 12 ng/cm2 and could be pushed

    further, if necessary. More importantly, calling it a balance implies that the

    Sauerbrey equation applies strictly, namely that the frequency shift is the

    sole result of mass loading. It is well known that this is not the case, andthe frequency shift observed could more appropriately be expressed by a

    sum of terms of the form

    Df Dfm Dfg DfP DfR Dfsl DfT 2where the different terms on the right-hand side (rhs) of this equation

    represent the effects of mass loading, viscosity anddensity of the medium in

    contact with the vibrating crystal, the hydrostatic pressure, the surface

    roughness, the slippage effect, and the temperature, respectively, and the

    different contributions can be interdependent. Even this equation does not

    tell the whole story, certainly not when the device is immersed in a liquid or

    in gas at high pressure. It does not account for solution occluded between

    the ridges of a rough surface or in the pores of a porous substrate. The

    nature of the interaction between the liquid and the surface, the type ofroughness, and internal stress or strain could all affect the response of the

    quartz crystal resonator. These effects become of major importance

    particularly when small changes of frequency, associated with submono-

    layer phenomena, are considered. Some of these factors will be discussed in

    this chapter.

    It should be evident from the above arguments that the term quartz

    crystal microbalance is a misnomer, which could (and indeed has) lead to

    erroneous interpretation of the results obtained by this useful device. It

    would be helpful to rename it the quartz crystal sensor (QCS), which

    describes what it really doesit is a sensor that responds to its nearest

    environment on the nano-scale. However, it may be too late to change the

    widely used name. The QCM or its analogue in electrochemistry, the

    EQCM, can each act as a nano-balance under specific conditions, but notin general.

    Electrochemical Quartz Crystal Microbalance 3

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    23/328

    B. Applications of the Quartz Crystal Microbalance

    The most common commercial use of the QCM is as a thickness gauge in

    thin-layer technology. When used to monitor the thickness of a metal film

    during physical or chemical vapor deposition, it acts very closely as a nano-

    balance, providing a real-time measurement of the thickness. Indeed,

    devices sold for this purpose are usually calibrated in units of thickness

    (having a different scale foreach metal, of course), and claim a sensitivity of

    less than 0.1 nm, which implies a sensitivity of less than a monolayer.

    The other common application of the QCM is as a nano-sensor

    proper, made sensitive to one gas or another by suitable surface treatment.

    Selecting the suitable coating on the electrodes of the QCM can determineselectivity and enhance sensitivity. It is not our purpose to discuss sensors

    in the present review. It should only be pointed out that any such sensor

    would have to be calibrated, since the Sauerbrey equation would not be

    expected to apply quantitatively.

    1. Applications for Gas-Phase Adsorption

    The high sensitivity of the QCM should make it an ideal tool for the study

    of adsorption from the gas phase. We note that the number of sites on the

    surface of a metal is typically 1.3 1015/cm2, hence a monolayer of a smalladsorbate, occupying a single site, would be about 2.2 nmol/cm2. A

    monolayer of water would therefore weigh about 40 ng/cm2, while a

    monolayer of pyridine would weigh 3060 ng/cm2, depending on its

    orientation on the surface. Comparing these numbers with the sensitivity

    of 2 ng/cm2 shows that adsorption isotherms could be measured in the gas

    phase employingthe QCM. This hasnot been done properly until relatively

    recently, mainly because the device was treated as a microbalance, i.e., it

    was assumed that the Sauerbrey equation could be applied, and several

    important terms in Eq. (2) were ignored. Obtaining adsorption isotherm

    one has to change the pressure over a wide range. Therefore, the changes of

    properties of the surrounding gas cannot be ignored. This shortcoming was

    overcome by the present authors [8], who developed the supporting gas

    method. When this method is employed, the overall pressure is kept

    constant by a large excess of an inert gas, and the frequency shift of the

    QCM is measured as a function of the partial pressure of the material being

    investigated. In this manner all terms in Eq. (2), other thanD

    fm, areessentially zero, and the device acts as a true nano-balance. One intriguing

    result that was obtained using this method came from a comparison of

    the adsorption of benzene and pyridine on a gold surface. It was found

    Tsionsky et al.4

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    24/328

    that a monolayer of pyridine weighs roughly twice as much as a monolayerof benzene. Since the two molecules have almost the same size and

    molecular weight, it must be concluded that their configuration in the

    adsorbed state is different. Benzene is probably adsorbed flat on the

    surface, while pyridine must be adsorbed perpendicular to it, occupying

    only half as many sites.

    Although the nominal resolution of 2 ng/cm2 should be enough to

    study the adsorption isotherm if the monolayer weighs around 3060 ng/

    cm2, it is somewhat marginal, and an increase of sensitivity of about one

    order of magnitude would be desirable. Part of this enhancement could be

    achieved by increasing the roughness factor on the atomic scale, without

    influencing the roughness on a scale relevant to the resonance frequency(see Sec. VI).

    2. Use of QCM in Liquids

    It was not initially obvious that the quartz crystal resonator would operate

    in liquids until this was proven experimentally [9,10]. The term associated

    with the influence of the viscosity and density of liquid in Eq. (2) can be

    written [11] as

    Dfg Cg gq 1=2 3Since the product of

    ffiffiffiffiffiffigq

    pin liquids is about two orders of magnitude

    higher than in gases at ambient pressure, the crystal is heavily loaded when

    transferred from the gas phase into a liquid.Once the door had been opened to its use in liquids, the potential of

    the QCM for interfacial electrochemistry was obvious, and the EQCMbecame popular.

    When a QCM is placed in contact with a dilute aqueous solution, the

    frequency should shift to lower values by about 0.7 kHz according to Eq.

    (3). In practice, a shift of 1.02 kH is observed, depending on the surface

    roughness. The effect of roughness is also related indirectly to viscosity and

    density, since the hydrodynamic flow regime at the surface is altered as a

    result of roughness [1214]. Roughness is a major issue in the interpreta-

    tion of the response of the QCM in liquids, and it is discussed in some detailin the following sections.

    C. The Impedance Spectrum of the EQCM

    In early studies of the QCM and the EQCM, only the resonance frequency

    was determined and conclusions were drawn based on the shift of

    Electrochemical Quartz Crystal Microbalance 5

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    25/328

    frequency. Unfortunately, in many cases this shift was attributed to massloading alone, and it was used to calculate the weight added or removed

    from the surface, disregarding other factors that affect the frequency. In

    the past decade, more and more laboratories expanded such studies to

    include measurements of the impedance spectrum of the crystal [1525].

    This provides an additional experimental variable that can obviously yield

    further information and a deeper understanding of the structure of the

    interface. For instance, a variation in the resonance width provides

    unambiguous proof that mechanisms other than mass loading are also

    involved.

    A series of typical admittance spectra are presented here. In Fig. 1a

    we show a simple case of metal deposition (gold on a gold substrate). TheEQCM acts as a true microbalance in this case. The resonance frequency is

    shifted to lower values with increasing load, but the shape of the spectrum

    remains unaltered.

    In Fig. 1b the effect of viscosity on the admittance spectrum is shown.

    Here again the resonance frequency is shifted to lower values with

    increasing viscosity, but this has nothing to do with mass loading.

    However, the shape of the spectrum is quite different, and the width at

    half-height (see below) increases dramatically with increasing viscosity and

    density of the liquid. Line 1 and the inset in this figure show the response of

    theQCMinH2 at ambient pressure. The product of viscosity and density is

    about four orders of magnitude smaller than in any of the liquids.

    Correspondingly, the width of the resonance is only about 20 Hz,

    compared to about 2.5 kHz in the liquid corresponding to line 2.

    Another aspect of the admittance spectrum is shown in Fig. 1c. Here

    the same metal deposition was conducted as in Fig. 1a, but the conditions

    were purposely chosen to produce a very rough surface (by plating at a

    current density close to the mass-transport limited value). The width of the

    resonance is increased and the frequency is shifted to lower values with

    increasing roughness.

    We chose rather extreme cases of viscosity and roughness in Fig. 1b

    and 1c, for the purpose of illustration. The corresponding shift in fre-

    quency is very high, in the range of 515 kHz, as compared to changes of

    frequency of 540 Hz typically observed in the studies of electrosorption,

    double layer, upd, and other submonolayer phenomena. The important

    conclusion is that even very small changes of viscosity and/or surfaceroughness (produced inadvertently) could lead to a shift of frequency

    comparable to that expected for such submonolayer phenomena, and the

    Tsionsky et al.6

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    26/328

    FIG. 1. (a) The real part of the admittance versus frequency: during depositionof gold on a gold-covered EQCM at a current density of 20 AA/cm2. (c) The same

    at 500 AA/cm2. (b) The response of the QCM immersed in different media: 1,

    hydrogen, 1 atm; 2, dimethyl ether; 3, water; 4 and 5, 40% and 50% aqueoussolutions of sucrose, respectively. (Inset) Admittance for H2, on an expanded scale.

    Arrow gq shows the increase of product gq. (From Ref. 24.)

    Electrochemical Quartz Crystal Microbalance 7

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    27/328

    change of frequency cannot be generally interpreted to be a result of mass

    loading alone.

    D. Outline of This Chapter

    This chapter contains theoretical and experimental sections. In the theo-

    retical section we consider different aspects of the behavior of the vibrating

    resonator: when it was loaded by additional mass, immersed in viscous

    media, has undergone changes in surface roughness, etc. We discuss the

    universal perturbation theory of the influence of slightly rough surfaces on

    the QCM response and consider the special model for strong roughness,

    noting that a general model does not exist for such surfaces. Specialattention was paid to consideration of the influence of slippage on the

    QCM at the solid/electrolyte interface.

    The QCM is now so widely and extensively used that, in the frame-

    work of this chapter, it is not possible to review all the available litera-

    ture. Hence we limited ourselves here to a review of the experimental data

    and ideas concerning the studies of submonolayer adsorption and inter-

    actions taking place at the metal/solution interface. In other words, this

    review is restricted to the use of the QCM in fundamental electrochem-

    istry. Furthermore, we did not include studies of electrochemical kineticswith the help of the EQCM, which merits a separate review. The problems

    of the interpretaion of the EQCM response caused by changes taking

    place at the metal/solution interface are obviously of first priority.

    We did not present here a full description of the operation of anEQCM. This topic is well described in previous reviews (see Refs. 1,2) and

    in many articles published in readily accessible electrochemical journals.

    However, a few aspects of the experiments with the EQCM are covered in

    the Appendix (Sec. VIII.B).

    II. THEORETICAL INTERPRETATION OF THE QCM

    RESPONSE

    A. Impedance

    The shear mode resonator consists of a thin disk of AT-cut quartz crystal

    with electrodes coated on both sides. The application of a voltage between

    these electrodes results in a shear deformation of the crystal due to itspiezoelectric properties. The crystal can be electrically excited into a

    Tsionsky et al.8

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    28/328

    number of resonance modes, each corresponding to a unique standingshear wave across the thickness of the crystal. If a quartz resonator

    operates in contact with an outer medium, the oscillating surface interacts

    mechanically with the medium and excites motion in it. The mechanical

    properties of the medium in contact are reflected in the response of the

    resonator.

    The geometry of the system consisting of a quartz crystal in contact

    with the outer medium is schematically shown in Fig. 2. The z-axis is

    plotted perpendicular to the plane of contact - the plane z =0 coinciding

    with the unconstrained face of the quartz resonator, and the plane z = dis

    its constrained face. The thickness of the quartz crystal is d.

    When an ac voltage is applied between the electrodes, the motion ofthe AT cut quartz crystal can be described by a system of two coupled

    differential equations, which constitute the wave equation for elastic

    displacements, u(z,t) = u(z,x) exp(ixt), and the equations that establish

    FIG. 2. Schematic sketch of the quartz crystal resonator in contact with a liquid.The contacting medium is a thin film rigidly attached to the crystal surface from

    one side, at z = d. The opposite surface of the crystal (z = 0) is unconstrained. disthe thickness of the quartz crystal.

    Electrochemical Quartz Crystal Microbalance 9

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    29/328

    the relationship between displacements and the electrostatic poten-tial,u(z,t) =u(z,x) exp(ixt), [26] are

    x2qqu z;x c66d2

    dz2u z;x 4

    e22d2

    dz2u z; x e26 d

    2

    dz2u z; x 5

    Here, c66 lq e226=e22 ixgq; qq;lq are the density and shear modu-

    lus of quartz,e22,e26 are the dielectric constant, and the piezoelectric stress

    coefficient of quartz, gq, is its fictitious viscosity, x = 2p f is the angularfrequency, and f is the frequency. Equations (4) and (5) are solved under

    the following boundary conditions:

    1. At the plane z =0, the potential equals u0 and the stress is zero.

    2. At the plane z = d, the potential equals u0 and the ratio of theshear stress, c66du(z,x)/dz, acting on the contacting medium to

    the surface velocity, ixu(d,x), equals Zout.Here Zout is the mechanical impedance of the medium contacting the

    quartz surface. Solution of Eqs. (4) and (5) yields the following expression

    for the admittance of the quartz resonator [27,28]:

    Y ixC0 Z1m 6where C0=e22/d is the static capacitance and Zm is the motional imped-ance:

    Zm 1ixC0

    /q

    K2q 2tan /q=2 1

    " # /q

    4K2q xC0

    Zout=Zq

    1 iZout=Zq2tan /q=2

    26664

    37775 7

    and K2q e226=e22c66;/q kqd; Zq kqc66=x , and kq xffiffiffiffiffiffiffiffiffiffiffiffiffi

    qq=c66p

    is the

    wave number of the shear wave in quartz. The first term in Eq. (7) de-

    scribes the motional resistance of an unloaded quartz resonator. The sec-

    ond term arises from surface loading and includes the properties of the

    electrode surfaces and the contacting medium through Zout.In QCM experiments the surface loading is relatively small [27], that

    is, |Zout/Zq|

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    30/328

    nant frequency of the quartz-crystal resonator with respect to the resonantfrequency of the unloaded quartz crystal, f0, can be written as [12,29]

    D fu Df iG2

    if0p

    Zout

    Zq8

    It should be noted that the frequency shift Dfcan be a complex number,

    and its imaginary part, G, reflects the width of the resonance. Equation (8)

    shows that the complex frequency shift Dfcontains the same information

    as the mechanical impedance Zout.

    The admittance of the quartz resonator can be presented in terms of

    an electrical equivalent circuit [24,15,3036]. The equivalent circuit forthe unloaded quartz crystal consists of a motional branch, which reflects

    the vibration of the quartz, and a static capacitance, which is in parallel

    with the motional branch. The motional branch includes a resistance,

    capacitance, and inductance connected in series. The relationships between

    the electrical elements and the mechanical parameters describing the

    crystal motion (mass, compliance, and damping coefficient) were consid-

    ered in Refs. 3739. When the surface loading is small, |Zout/Zq|

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    31/328

    example, it is not obvious how one would account for a surface roughness

    of unspecified nature (slight roughness, strong roughness, or some combi-

    nation of both) in terms of an equivalent circuit.

    In order to analyze the influence of the different loading mechanisms

    on the QCM response, one has to model a dependence of the mechanical

    impedance Zout or the complex resonance frequency shift on the chemical

    and physical properties of the contacting medium. Various models for the

    mechanical contact between the oscillating quartz crystal and the outermedium are discussed below.

    B. The Effect of Thin Surface Films

    1. Uniform Film Rigidly Attached to the Surface

    First we consider the effect of a thin film, rigidly attached to an ideally flat

    crystal surface, on the response of the quartz crystal resonator (see Fig. 2).

    FIG. 3. The Butterworthvan Dyke equivalent circuit of the loaded quartzcrystal resonator. The parameters R, C, and L describe the behavior of the

    unloaded quartz resonator; Zout is the impedance of the contacting medium.

    Tsionsky et al.12

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    32/328

    For a homogeneous thin film with a thickness smaller than thewavelength of the shear oscillations, the shift of the resonance frequency

    can be expressed in terms of the change in surface mass density of the film,

    Dmf (in g/cm2). This was given by Sauerbrey [6] as

    Df 2f2

    0 Dmf

    lqqq 1=2 9

    This equation coincides with Eq. (1) with Cm 2f20 = lqqq 1=2

    . Equation

    (9) can be derived by supplementing the wave equation [Eq. (4)] with the

    Newtonian equation of motion for the surface film:

    Dmfx2uf x lq

    d

    dzu z; w at z d 10

    where uf(x) is the displacement of the film. Here the shear stress,lqdu(z,x)/dz, plays the rule of the external force acting on the film. Hereand everywhere below we use in Eq. (4) lq instead ofc66, neglecting small

    values xgq and e226=e22 . Solving Eqs. (4) and (10) under the standard

    boundary condition, namely that (1) at the unrestricted surface, z = 0, the

    shear stress equals zero and (2) at the surface z = d the quartz surface

    displacement is equal to the surface film displacement, one obtains the

    Sauerbrey equation. From Eq. (10) one can see that the frequency shift is

    determined by the inertial force of the film acting on the quartz surface.

    Equation (9) shows that the addition of mass rigidly attached to the

    surface of the quartz-crystal resonator leads to a decrease of the resonant

    frequency, but it does not influence the width of the resonance.

    2. Nonuniform Film Rigidly Attached to the Surface

    A natural question arises whether a nonhomogeneous mass distribution

    can lead to an additional shift of frequency and/or to a broadening of theresonance, compared to the result given by the Sauerbrey equation?

    In order to answer this question, we consider here the effect of

    nonuniform mass loading on the response of the QCM. Lateral displace-

    ment of the nonuniform film can be described by the equation of motion:

    DmfRx2ufR lqd

    dzuz;R for z d 11

    where uf(R) and u(z,R) represent the displacements of the surface film and

    quartz crystal andDmf(R) is the surface film density, as a functionof lateral

    Electrochemical Quartz Crystal Microbalance 13

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    33/328

    coordinates, R. This equation plays the role of a boundary condition forthe wave equation describing the shear-mode oscillations in the quartz

    crystal. Equations (4) and (11), with the standard boundary condition, lead

    to the following equation for determination of the resonant frequency (see

    Appendix):

    lqk tankd x2Dmf x4Dm2lZ

    dK

    2p2gK

    lqpKtanpKd x2Dmf12

    Here, k = x(qq/lq)1/2

    is the wave number of the shear wave in the quartz,p(K) = k2 K2, and K is the two-dimensional tangential wave vector.Considering mechanical response of the quartz crystal we use k instead

    of kq [see text following Eqs. (4) and (5)] because xgq and e226=e22 are

    small compared to lq. Also, Dml is the root mean square deviation of the

    mass density from the average value Dmf, and g(K) is the correlation

    function, which describes a nonuniform mass distribution along the

    surface. Equation (12) is a general form of the Sauerbrey equation,

    applicable for the case of an inhomogeneous surface film. For uniform

    mass distribution (corresponds to Dml= 0), it yields the Sauerbrey

    equation in its usual form.

    Assuming that the correlation function g(K) has a Gaussian form,

    with a lateral correlation length, l, Eq. (12) can be solved analytically for

    two limiting cases, kl>>1 and kl1,

    splitting of the resonant frequency occurs, and the frequency shift can be

    estimate as

    Df 2f20ffiffiffiffiffiffiffiffiffiffi

    qqlqp DmfFDml

    13In contrast to the case of uniform mass loading, Dm

    1f R 0, two

    values of the resonance frequency appear. This effect can be simulated by

    a simple equivalent circuit consisting of two Butterworthvan Dyke [33

    35] circuits in series with the inductances corresponding to the two dif-

    ferent values of the surface mass densities,D

    mf D

    mlandD

    mf D

    ml. Dueto overlap of these two resonance states, splitting can manifest itself as a

    broadening of the resonance, which will have an effective width of the

    Tsionsky et al.14

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    34/328

    order of 2f20Dml=p lqqq 1=2

    . For the 6 MHz quartz resonator thisbroadening effect becomes important when the correlation length l is

    larger than 0.02 cm.

    In thesecond limiting case, kl

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    35/328

    film yields the following expressions for the changes of the frequency, Df,and the width of the resonance, G:

    Df 2f20Dmaffiffiffiffiffiffiffiffiffiffiqqlq

    p v2

    v2 2pf0Dma 2" #

    15

    G 4f20Dmaffiffiffiffiffiffiffiffiffiffiqqlq

    p 2pf0Dmavv2 2pf0Dma 2

    " #16

    Note that

    G

    Df 4pf0

    Dma

    v 17

    Thus, the interfacial friction can be evaluated from measurement ofG and

    Df. This procedure has been applied to a number of systems in which weak

    physical adsorption occurs, such as the adsorption of Xe, Kr, N2 on Au

    and of H2O and C6H12 on Ag [4852]. In all above cases, slippage was

    observed and the ratio of the coefficient of sliding friction to the mass

    density was of the order v/Dma = (108 109)s1. As an example, the

    frictional stress acting on the monolayer Xe film sliding on a Ag (111)

    surface at a velocity v =1 nm / s, F=vv, equals about 10 N/m2 [54]. It is

    much smaller than typical shear stresses involved in sliding of a steel block

    on a steel surface under boundary lubrication condition. The shear stress in

    the latter case is of the order c108 N/m2 [53].

    In a recent paper [55] the dependence of the slip time, ss, on the

    amplitude of the crystal surface oscillations, A, and the surface coverages

    was investigated. The results refer to the absorption of krypton atoms on

    gold at 85jK. The slip time is related to the interfacial friction coefficient, v,

    as ss = Dma /v. It was found that there is a step-like transition between a

    low-coverage region, where slippage exists at the solid/film interface, and a

    high-coverage region where the film is locked to the surface. The transition

    occurs at different coverages depending on the amplitude, A. Independent

    of coverage, the film is attached rigidly to the surface for AV 0.18 nm and

    slides for A > 0.4 nm. In the region of sliding at small coverages, the values

    of the slip time are in the interval 210 nsec, for 0.18 nm < A < 0.4 nm.

    C. The Quartz Crystal Operating in Contact with a Liquid

    1. General Considerations

    When a quartz crystal resonator operates in contact with a liquid, the shear

    motion of the surface generates motion in the liquid near the interface. The

    Tsionsky et al.16

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    36/328

    velocity field, v(r,x) related to this motion in a semi-infinite Newtonianliquid is described by the linearized Navier-Stokes equation:

    ixqvr;x jPr;x gDvr;x 18where P(r,x), g, and q are pressure, viscosity, and density of the liquid,

    respectively. Under the conditions of the QCM experiments, where the

    shear velocities are much smaller than the sound velocity in the liquid, the

    displacement of the crystal does not generate compressional waves and a

    liquid can be considered as an incompressible one. If the surface is

    sufficiently smooth, the quartz oscillations generate plane-parallel laminar

    flow in the liquid, as shown in Fig. 4. The velocity field obtained as the

    solution of Eq. (18) for a flat surface has the form

    vxz vq0xexp1 iz=d 19where vq0(x) is the velocity of the liquid at the surface and d

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2g=x0q

    p.

    Equation (19) represents a damped shear wave radiating into the liquid

    from the surface of the oscillating resonator. d is the velocity decay length

    of this shear wave, which lies between 250 and 177 nm, for dilute aqueous

    solutions at room temperature, for crystals having a fundamental frequen-

    cy in the range of 510 MHz. Damping of the shear wave has a number of

    important consequences. First, it ensures that the quartz crystal can

    operate in liquids, the losses in the liquid being limited by the finite depth

    of penetration. Second, a small portion of the liquid is coupled to the

    crystal motion and a frequency decrease is observed. Third, the viscousnature of motion gives rise to energy losses, which are sensed by the

    resonator, both as a decrease in frequency and as anincrease in the width of

    the resonance.

    2. The Nonslip Boundary Condition

    The response of the QCM at the solid/liquid interface can be found by

    matching the stress and the velocity fields in the media in contact. It is

    usually assumed that the relative velocity at the boundary between the

    liquid and the solid is zero. This corresponds to the nonslip boundary

    condition. Strong experimental evidence supports this assumption on the

    macroscopic scales [56,57]. In this case the frequency shift, Dfl, and the

    width of the resonance, Gl, can be written as follows [10,11]:

    Dfl f

    3=20

    ffiffiffiffiffiffiqg

    pffiffiffiffiffiffiffiffiffiffiffiffipqqlq

    p 20

    Electrochemical Quartz Crystal Microbalance 17

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    37/328

    Gl 2f

    3=20

    ffiffiffiffiffiffiqg

    pffiffiffiffiffiffiffiffiffiffiffiffipqqlq

    p 21

    Equations (20) and (21) show that the generation of a damped

    laminar flow in the liquid causes a decrease in the resonance frequency

    and an increase in the resonance width, which are both proportional toffiffiffiffiffiffiqgp . In contrast to the case of the mass loading, where Dfis proportionalto f0

    2, the liquid induced response of the QCM is proportional to f03/2.

    FIG. 4. The system geometry and the velocity distribution. Curves 1 and 2represent the velocity distributions at the liquid/adsorbate interface without and

    with slippage, respectively. Curve 3 is the velocity distribution in the quartz. The

    thickness of various layers is not drawn to scale.

    Tsionsky et al.18

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    38/328

    It is interesting to note that for both a surface film rigidly attached tothe resonator and a liquid in contact with the surface of the quartz crystal,

    the shift of the resonant frequency can be written in the same form, as

    Df f0 qqq

    kheff 22

    where k x0 ffiffiffiffiffiffiffiffiffiffiffiqq=lqp , q is the bulk density of the medium in contact withthe vibrating surface of the solid, film, or liquid, and heffis the thickness of

    the layer that responds to the quartz oscillations. In the case of a film, heffcoincides with the thickness. For a semi-infinite liquid, heff presents a

    thickness of liquid involved in the motion, and it should be taken equal to

    d/2. The difference in the frequency dependence of the QCM response in

    the two cases is a result of the frequency dependency of d. However, in

    contrast to the case of pure mass loading, the effect of a liquid results not

    only in a frequency shift, but also in a broadening of the resonance.

    a. Effect of a Thin Liquid Film at the Interface

    The properties (the effective viscosity and density) of the liquid layer in

    close vicinity to the interface can differ from their bulk values. There are

    various reasons for these phenomena. For example, the properties of a

    thin liquid layer confined between solid walls are determined by interac-

    tions with the solid walls [58,59]. In electrochemical system the structuring

    of a solvent induced by the substrate and a nonuniform ion distribution in

    the diffuse double layer can significantly influence the properties of the

    solution at the interface. The nonuniform distribution of species, which

    influences the properties of the liquid near the electrode, also occurs in the

    case of diffusion kinetics. The latter was considered in Ref. 60, where the

    ferro/ferri redox system was studied by the EQCM. This was the case

    where the velocity decay length (>25 Am) was much less than the thickness

    of the diffusion layer (>100 Am), in which the composition of the solution

    is different from the bulk composition.

    Nonuniform distribution of species results in nonuniform distribu-

    tion of the properties of liquid near the vibrating surface of the resonator.

    The properties change with distance from the interface, until the values

    corresponding to the bulk of solution have been reached. In order to

    simplify the description of this nonuniformity on the QCM, it is assumed

    that a thin film of liquid, having different values of gfand qf, exists at theinterface [61]. To calculate the effect of this film on the frequency shift, one

    has to solve the wave equation for the elastic displacements in the quartz

    Electrochemical Quartz Crystal Microbalance 19

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    39/328

    crystal [see Eq. (4)] simultaneously with the linearized Navier-Stokesequation for the velocities in the film and in the bulk liquid under standard

    nonslip boundary conditions.

    The shift of the resonant frequency and the width of the resonance

    can be written as

    Df f3=20

    ffiffiffiffiffiffiqg

    pffiffiffiffiffiffiffiffiffiffiffiffiplqqq

    p 2f20ffiffiffiffiffiffiffiffiffiffi

    lqqqp q

    1 g

    gf

    qf q " #

    Lf 23

    G 2f

    3=20 ffiffiffiffiffiffiqgpffiffiffiffiffiffiffiffiffiffiffiffiplqqqp

    4f20ffiffiffiffiffiffiffiffiffiffilqqqp q

    1

    g

    gf qf q " # L

    2f

    d 24

    where Lfand qfare the thickness and the density of the film, respectively.

    These equations are valid in a particular case, when Lf g the film acts as though it were rigidly attached

    to the surface: it causes a shift in frequency equal to that caused by its mass.

    3. Slip Boundary Conditions

    a. Slippage at Solid/Liquid Interface

    Although the nonslip boundary condition has been remarkably successful

    in reproducing the characteristics of liquid flow on the macroscopic scale,

    its application for a description of liquid dynamics in microscopic liquid

    layers is questionable. A number of experimental [6264] and theoretical

    [65,66] studies suggest the possibiility of slippage at solid/liquid interfaces.

    The boundary condition is controlled by the extent to which the

    liquid feels a spatial corrugation in the surface energy of the solid. This

    depends on a number of interfacial parameters, including the strength of

    the liquid-liquid and liquid-solid interactions, the commensurability of thesubstrate and the liquid densities, characteristic sizes, and also the rough-

    ness of the interface. In order to quantify the slippage effect, the slip length,

    Tsionsky et al.20

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    40/328

    k, is usually introduced [65,67,68]. The traditional nonslip boundarycondition is replaced by

    dvz;xdz

    zd

    1k

    vd;x vq0x 25

    where v(z,x) is the velocity in the liquid and vq0(x) is the velocity of thequartz crystal surface. Equation (25) expresses the discontinuity of the

    velocity across the interface. For k = 0, Eq. (25) is reduced to the usual

    nonslip boundary condition: v(d,x) = vq0 (x). The physical meaning of the

    slip length can be clarified by comparing velocity profiles for the nonslip

    and slip boundary conditions. These two profiles coincide when the nonslip

    boundary condition is imposed at the surface shifted inside the solid on the

    distance k with respect to the actual interface.

    The slip boundary condition (25) results in the following equations

    for the resonant frequency shift and the width of the resonance:

    Df f20 qdffiffiffiffiffiffiffiffiffiffiqqlq

    p 11 k=d2 k=d2" #

    26

    G 2f20 qdffiffiffiffiffiffiffiffiffiffi

    qqlqp 1 2k=d

    1 k=d 2 k=d 2" #

    27

    Equations (26) and (27) show that the influence of the slippage on the

    response of the QCM in liquid is determined by the ratio of the slip length

    k to the velocity decay length, d. Even for a small value ofk c 1 nm, the

    slippage-induced correction to the frequency shift, Dfsl, will be of the order

    of 6.5 Hz for the fundamental frequency of f0 = 5 MHz. This value far

    exceeds the resolution of the QCM, but it is difficult to separate it from the

    overall QCM signal.

    There have been attempts [53] to estimate the slip length at the solid/

    liquid interface on the basis of QCM experiments for adsorbed liquid

    layers. The slip length can be expressed in terms of the coefficient of sliding

    friction, v, at the interface

    k gv

    28

    Using the sliding friction coefficient v = 3 g/cm

    2

    s, which is obtained for amonolayer of water on Ag [49] and on Au [69], a surprisingly high slip

    length of k = 6 104 nm is obtained. Using this value for the interface

    Electrochemical Quartz Crystal Microbalance 21

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    41/328

    between Au and bulk water, Eq. (26) yields for f0= 5 MHz a value ofDfc7103 Hz, which turns out to be smaller than that observed experimen-tally by a factor of 105. This inconsistency is most likely caused by a

    roughness of the electrode surface that reduces the effective slip length.

    Another reason could be the difference between friction at the solid/

    adsorbed layer and the solid/liquid interfaces. For example, a decrease in

    the slip length with increasing film thickness has been observed recently in

    QCM studies of Kr films on gold electrodes [55].

    Recent molecular dynamics simulations [65,70] demonstrated that

    the slip length is determined by the ratio of characteristic energies of liquid-

    substrate, els, and liquid-liquid, ell, interactions, k = f(els/ell). The slip

    length is negligible for els/ell z 1 and grows with the decrease of theparameter els/ell. The slip length k may be as large as 15 diameters of liquid

    molecules for els/ellc 0.5. It should also be noted that, for a given value of

    els/ell, the slip length is minimal when substrate and liquid molecules are of

    the same size and increases with the increase of incommensurability of the

    sizes. For smaller coupling between the liquid and the substrate or

    incommensurability of their sizes, the spatial corrugation in the interfacial

    energy is weaker and interfacial slip can develop.

    The latter conditions are satisfied for partially wetting liquid/solid

    interfaces. Wetting is characterized by a contact angle, which can be esti-

    mated as [68]

    cos h

    1

    2

    qs

    q

    els

    ell 29

    where qs and q are the density of the solid and the liquid, respectively.

    Thus, the contact angle may be interpreted as a measure of the strength

    of interaction between the liquid and the solid, els. One expects a large

    value of the slip length for a nonwetting situation (cos(h) ! 1), when elsbecomes much smaller than ell. This conclusion is in agreement with several

    experimental observations [62,71] reporting large slip lengths for partially

    wetting liquids.

    The authors of Refs. 14,72,73 showed that surface treatments

    affecting liquid contact angle influence the response of quartz crystal

    resonator: resonant frequency changes caused by liquid loading were

    consistently smaller for surfaces having large liquid contact angles. These

    results were interpreted as arising from the onset of slippage at the solid/liquid interface: the solid-liquid interaction becomes sufficiently weak on a

    hydrophobic surface, and shear displacement becomes discontinuous at

    Tsionsky et al.22

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    42/328

    the interface. However, this interpretation was called into question by aseries of experiments in which the effect of a hydrophobic monolayer was

    examined on devices with various surface roughness [12].

    Correlating the wetting properties with the response of the QCM in

    contact with liquids seems to be a promising area for future research.

    Unfortunately, studies of wetting behavior require ex situ measurements of

    the contact angle, which change drastically the properties of the electro-

    chemical system at the electrode/adsorbed layer/electrolyte interfaces.

    b. Slippage at the Adsorbate/Electrolyte Interface

    Slippage is very sensitive to the molecular structure of the interface, as we

    have already discussed above. Thus, adsorption can strongly influence thisphenomenon. In order to describe the effect of adsorption, let it be assumed

    that the adsorbed layer is rigidly attached to the surface and slippageoccurs at the adsorbate/liquid interface (see Fig. 4). Then the equation of

    motion of the adsorbed layer can be written as [74]

    ixDmava x lqdu z

    dz v va x v1 x at z d 30

    where va(x) is the velocity of the adsorbed layer and Dma is its two-

    dimensional density, while vl(x) is the velocity of the liquid at the interface.

    The first term on the right-hand side of Eq. (30) describes the driving force

    acting on the adsorbed layer from the quartz crystal, while the second term

    accounts for the friction at the adsorbate/liquid interface.

    The velocity fields in the crystal and the liquid are given by thesolutions of the wave equation [Eq. (4)] and the linearized Navier-Stokes

    equation [Eq. (18)], respectively. The solution of Eqs. (4), (18), and (30)

    with the boundary conditions for shear stresses and velocities leads to the

    following equation for the shift of the resonant frequency, Df, and the

    change of the width of the resonance, G:

    Df 2f20Dma

    qqlq1=2 f

    3=20 qg1=2

    pqqlq1=21

    1 a 2a2

    " #31

    G 2f3=20 qg1=2

    pqqlq1=21 2a

    1 a2 a2

    " #32

    Writing Eqs. (31) and (32), we introduced a dimensionless parameter a =

    g/ vd = k/d , which is the ratio of the slip length, k = g/v, and the velocity

    Electrochemical Quartz Crystal Microbalance 23

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    43/328

    decay length in the liquid, d. Equations (31) and (32) include both theinterfacial (adsorption) and the bulk solution contributions to the response

    of the QCM, given by Eqs. (20) and (21). The latter remains constant in

    adsorption studies and can be subtracted from the overall change given by

    Eqs. (31) and (32). As a result, the shift of the resonant frequency and the

    change of the width due to adsorption, which are measured experimentally,

    are given by the equations:

    Df DfluDfm Dfsl 2f20Dma

    qqlq1=2 f

    3=20 qg1=2pqqlq1=2

    aa 11 a2 a2

    " #

    33G Gl f

    3=20 qg1=2pqqlq1=2

    4a2

    1 a2 a2

    " #34

    Equation (33) shows that there are two different contributions to the

    frequency shift,Dfm andDfsl, which originate from (1) a change of the mass

    of the adsorbed layer rigidly coupled to the surface [first term on the rhs of

    Eq. (33)], and (2) partial decoupling between the quartz crystal oscillations

    and the solution, caused by slippage at the adsorbate/liquid interface

    [second term on the rhs of Eq. (33)]. It should be stressed here that, in

    contrast to adsorption from the gas phase, electrosorption can result ineither a decrease or an increase of the resonant frequency, depending on its

    effect on the mass of the layer rigidly coupled to the surface and on changeof the coefficient of sliding friction, which determines the slip length,

    according to Eq. (28).

    Consider the effect of adsorption on the parameters Dma and v. The

    layer adsorbed at the electrode/electrolyte interface contains two types of

    molecules: adsorbate and solvent. In the framework of mean field approx-

    imation, the effective interaction between the liquid and the adsorbed layer

    can be characterized by the energy elscela&a/&m + ell(1&a/&m), where elais the characteristic energy of the adsorbate/liquid interaction and&m is the

    maximum surface excess of the adsorbate. As a result, the slip length at the

    adsorbed layerliquid interface can be expressed as

    k fela=ell&a=&m 1 &a=&mcfela=ell&a=&m 35showing an increase of k with &a for ela/ell < 1. Equation (35) is the

    interpolation formula that describes correctly the behavior of k for small

    Tsionsky et al.24

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    44/328

    &a/&m and for &a/&m =1. We note that when the liquid and adsorbatemolecules are of significantly different size, the incommensurability be-

    tween the structures of the adsorbed layer and liquid grows with &a, which

    may lead to an additional enhancement of the slip length. What is

    important here is a relation between scales of corrugations of the poten-

    tial energy in the solvent and the adsorbed layer, rather than their physical

    sizes of solvent and adsorbed molecules.

    The foregoing discussion shows that for ela/ell< 1 the parameter a =

    k/d in Eqs. (31) and (32), characterizing the effect of slippage on the

    response of the QCM, increases with &a. For instance, for ela/elc 0.5, it

    may reach values as high as a c 102 for &a c &m. Correspondingly, the

    adsorption-induced slippage leads to a positive frequency shift, whichgrows with &a. This contribution can be larger than the effect of added

    weight. As a result, the overall frequency shift due to electrosorption can

    be positive and increases with &a [74]. It should be noted that for small

    values of the parameter a, the effect of slippage on the resonance frequency

    shift is much larger than its effect on the width of the resonance [see Eqs.

    (33) and (34)]. Also, slippage will always cause a decrease in the width of

    the resonance. Thus, if a positive shift of frequency with adsorption is to be

    associated with enhanced slippage, it should also be exhibited as a

    reduction of the width of the resonance, although the latter may be hard

    to detect experimentally.

    Above we discussed the situation where the adsorbed layer is rigidly

    attached to the oscillating crystal surface, and there is finite slippage at the

    adsorbate/liquid interface. An alternative model based on the assumption

    that slippage occurs at the crystal/adsorbed layer interface and nonslip

    boundary conditions apply to the adsorbate/liquid interface can also be

    considered. For a small slip length, E

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    45/328

    slip length would be equivalent to a decrease of viscosity. Unfortunately,there is at present no suitable theory to describe the effect of the excess

    surface charge density (or the corresponding high electrostatic field) on the

    viscosity of the electrolyte in the double layer. The derivation of such a

    model is complicated by the fact that electro-neutrality does not exist on

    the solution side of the interface (except at the potential of zero charge),

    although the electrostatic energy is reduced by interaction with the image

    charges on the metal side of the interface.

    D. Quartz Crystals with Rough Surfaces

    1. Quartz Crystals with Rough Surfaces Operating in LiquidsWhen the surface of quartz crystal resonator is rough, the liquid motion

    generated by the oscillating surface becomes much more complicated than

    for the smooth surface. A variety of additional mechanisms of coupling

    between the acoustic waves in the solid and the motion in the liquid can

    arise. These may include generation of nonlaminar motion, the conversion

    of in-plane surface motion to motion normal to the surface, and trapping

    of liquid by cavities and pores. It has been experimentally demonstrated

    [12,15,7579] that the roughness-induced response of the QCM includes

    both the inertial and viscous contributions. Measurements of the complex

    shear mechanical impedance [12] were used to analyze different contribu-

    tions to the roughness-induced response of the quartz resonator and to

    correlate the experimental results with the surface roughness of the quartz

    resonator. Nevertheless, this subject is poorly developed, and the inter-

    pretation of experimental results can often be ambiguous.

    The dependence of the QCM response on the morphology of the

    interface is determined by the relation between the characteristic sizes

    of roughness and the length scales of the shear modes in the liquid and

    the quartz resonator. The length scales in the liquid (the velocity decay

    length, d) and in the crystal (wave length of the shear-mode oscillations,

    kq) are defined by the Navier-Stokes equation and by the wave equation

    for elastic displacement, respectively. For typical frequencies used in

    QCM experiments, f0=510 MHz and the lengths d = (g/pf0q)1/2 and

    kq = (lq/qq)1/2f0

    1 are of the order 0.1770.25 Am and 0.030.1 cm, re-

    spectively.

    The surface profile may be specified by a single valued function z =n(R) of the lateral coordinates R that defines a local height of the surface

    with respect to a reference plane (z = 0). The latter is chosen so that the

    Tsionsky et al.26

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    46/328

    average value ofn(R) will equal zero. Surfaces used in QCM experimentsmay have various scales of roughness. In order to clarify this point, let us

    consider the two limiting cases: slight and strong roughness structures,

    which are schematically shown in Fig. 5. For the slight roughness (Fig. 5a)

    the amplitude of deviation from the reference plane z = 0 is much less

    than the lateral characteristic length. In the case of strong roughness (Fig.

    5b), the amplitude and period of repetitions are of the same order of

    magnitude.

    In order to stress the multiscale nature of roughness, the profile

    function can be written as the sum of the functions that characterize the

    profile of the specific scale i:

    nR X

    i

    niR 36

    For the calculation of the response of the QCM, the height-height pair

    correlation function is needed [80]. When rough structures having different

    FIG. 5. Schematic representation of a slight (a) and a strong (b) roughness. Theprofile of slight roughness is described by the function z =n(R). L is the effective

    thickness of the porous film that represents strong roughness. (From Ref. 24.)

    Electrochemical Quartz Crystal Microbalance 27

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    47/328

    scales do not correlate the total correlation function can be written in theform

    < nRVnRVR >X

    i

    < niRVniRVR > 37

    where is the correlation function for the scale iand means averaging over the lateral coordinates.Usually one assumes that the correlation function has a Gaussian form = h2i exp(-|R|2/li2), where hi is theroot mean square height of the roughness and li is the lateral correlation

    length, which represents the lateral scale. Thus, the morphology of the

    rough surface can be characterized by a set of lengths {hi, li}.

    It is impossible at the present time to provide a unified description of

    the response of the QCM for nonuniform solid/liquid interfaces with

    arbitrary geometrical structure. Below we summarize results obtained

    for the limiting cases of slight and strong roughness.

    a. Slight Roughness

    For slightly rough surfaces, the problem was solved in the framework of

    perturbation theory with respect to the parameters |jn(R)|

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    48/328

    integral parameter, the roughness factor, R, which is the ratio between thetrue and the apparent (geometrical) surface area. For slight roughness, the

    roughness factor is expressed through the correlation function [81] as

    R 1 h2

    2

    ZdK

    2p2 gKK2 41

    For the Gaussian random roughness g(K) = pl2 exp(l2K2/4) and Eq. (41)yields R =1+2h2/l2.

    It should be noted that the roughness factor, R, relevant to the

    operation of the EQCM is not the same as the roughness factor commonly

    referred to in interfacial electrochemistry, because of the difference in

    corresponding length scales. The EQCM roughness factor is mostly

    determined by the roughness on the scale of the velocity decay length in

    the liquid, d, which assumes values of hundreds of nm, depending on the

    frequency of the crystal and the viscosity and density of the liquid. The

    interfacial roughness factor is related to charge transfer at the interface

    and the double layer structure, and therefore its characteristic scale is

    about 1 nm.

    The first terms in braces in Eqs. (38) and (39) define the shift and the

    broadening of the resonance at the interface between an ideally smooth

    crystal and the liquid [11]. The surface roughness leads to an additional

    decrease of the resonant frequency and a broadening of the width of the

    resonance, expressed by the second terms in this equation.

    The particular form of the scaling functions F(l/d) and A(l/d) isdetermined by the morphology of the surface. However, the asymptotic

    behavior of these functions for l/d >> 1 and l/d 1 42Fl=d l=db1 b2d=l at l=d > 1 44Al=d l=d2c2 at l=d > 1 the roughness-induced frequencyshift includes a term that does not depend on the viscosity of the liquid,

    Electrochemical Quartz Crystal Microbalance 29

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    49/328

    the first term in Eq. (42) and Eq. (38). It reflects the effect of the non-uniform pressure distribution, which is developed in the liquid under the

    influence of a rough oscillating surface [80]. The corresponding contri-

    bution has the form of the Sauerbrey equation. This effect does not exist

    for smooth interfaces. The second term in Eq. (42) and Eq. (44) describes

    a viscous contribution to the QCM response. Its contribution to Df has

    the form of the QCM response at a smooth liquid/solid interface, but

    includes an additional factor R that is a roughness factor of the surface.

    The latter is a consequence of the fact that for l/d >> 1 the liquid seesthe interface as being locally flat, but with R time its apparent surface

    area.

    Results obtained in Refs. 80, 81 show that the influence of slightsurface roughness on the frequency shift cannot be explained in terms of

    the mass of liquid trapped by surface cavities, as proposed in Refs. 76,

    77. This statement can be illustrated by consideration of the sinusoidal

    roughness profile. The mass of the liquid trapped by sinusoidal grooves

    does not depend on the slope of the roughness, h/l, and is equal to S-h/p,

    where Sis the apparent area of the crystal. However, Eq. (38) demonstrates

    that the roughness-induced frequency shift increases with increasing slope.

    Equation (39) and the asymptotic behavior of the scaling functions

    show that in the regions where l/d >> 1 and l/d l2pqf0, the roughness-induced frequency

    shift approaches a constant value and the roughness-induced width tends

    to zero.

    The results obtained make it possible to estimate the effect of

    roughness on the response of the QCM if the surface profiles function

    n(R) can be found from independent measurements.

    b. Strong Roughness

    Perturbation theory cannot be applied to describe the effect of the strong

    roughness. An approach based on Brinkmans equation has been used

    instead to describe the hydrodynamics in the interfacial region [82]. The

    flow of a liquid through a nonuniform surface layer has been treated as the

    flow of a liquid through a porous medium [8385]. The morphology of

    the interfacial layer of thickness, L, has been characterized by a local

    permeability, nH, that depends on the effective porosity of the layer, /. Anumber of equations for the permeability have been suggested. For

    instance, the empirical Kozeny-Carman equation [83] yields a relationship

    Tsionsky et al.30

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    50/328

    between n2H and the effective porosity n2Hfr2/3= 1 / 2 , where r is the

    characteristic size of inhomogeneities.

    The flow of liquid through the interfacial layer can be described by

    the following equation [82]:

    ixqvz; x g d2

    dz2vz;x gn2H vq0 vz;x 47

    where vq0 is the amplitude of the quartz surface velocity and v(z,t) = v(z,x)

    exp(ixt) is the velocity of the liquid in the layer. In this equation the effect

    of the solid phase on the flow of liquid is given by the resistive force, which

    has a Darcy-like form, gn2

    H vq0 vz;x . In the case of high effectiveporosity, the resistive force is small and Eq. (47) is reduced to the Navier-Stokes equation, describing the motion of the liquid in contact with a

    smooth quartz surface. For a given viscosity, the resistive force increases

    with decreasing effective porosity and strongly influences the liquid

    motion. At very low effective porosity, all the liquid located in the layer

    is trapped by the roughness and moves with a velocity equal to the velocity

    of the crystal surface itself.

    Brinkmans equation represents a variant of the effective medium

    approximation, which does not describe explicitly the generation of non-

    laminar liquid motion and conversion of the in-plane surface motion into

    the normal-to-interface liquid motion. These effects result in additional

    channels of energy dissipation, which are effectively included in the model

    by introduction of the Darcy-like resistive force.The liquid-induced frequency shift and the width of the resonance

    have the following form [82]:

    Df 2f20 q

    lqqq1=2Re

    &1

    q0 L

    n2Hq21

    1W

    1

    n2Hq21

    2q0q1

    coshq1L 1 sinhq1L !'

    48

    G 4f20 q

    lqqq1=2Im

    &1

    q0 L

    n2Hq21

    1W

    1

    n2Hq21

    2q0q1

    coshq1L 1 sinhq1L !'

    49

    Electrochemical Quartz Crystal Microbalance 31

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    51/328

    where q0= (i2p f0q/g)1/2, q21 = q

    20 +n

    2H , and W = q1 cosh( q1L)+q0

    sinh( q1L). The first terms on the right-hand sides of Eqs. (48) and (49)

    describe the response of the QCM for the smooth quartz crystal/liquid

    interface [11]. The additional terms present the shift and the width of the

    QCM response caused by the interaction of the liquid with a non-uniform

    interfacial layer.

    When the permeability length scale is the shortest length of the

    problem,nH

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    52/328

    non. In this manner, liquid flow at a rough surface has been simulated as aflow at a smooth surface with an effective slip length.

    Application of this approach to the QCM problem yields the fol-

    lowing equation for the effective slip length:

    keff k 1 h0K02

    23 4kk01 2kk0

    !( ) k0h

    20

    2

    2 3kk01 kk01 2kk0 !

    52Equation (52) was derived for a sinusoidal profile of roughness, z(x) =

    d+ h0 sin(k0x), with an amplitude h0 and a period of 2p/k0, assuming that

    the decay length, d, is the largest characteristic length of the problem,d/k>>1 and dk0 >>1. Beyond these conditions the effective slip length is a

    complex function. Equation (52) shows that roughness diminishes the

    influence of slippage on the QCM response, namely the effective slip length

    becomes smaller than the corresponding length for the smooth interface.

    At rough interfaces, the effective slip length decreases with an increase of

    the amplitude of the surface corrugation and with a decrease of its period.

    It should be noted that an effective slip length is not an intrinsic

    property of the surface. Its value depends also on the experimental

    configuration, for instance, kefffound for the Poiseuille flow between rough

    surfaces [86] differs from the corresponding value obtained for QCM

    experiments [Eq. (52)].

    So far only a few studies [8689] have been devoted to the effect of

    roughness on slippage, and this subject requires additional investigation.

    III. ELECTRICAL DOUBLE LAYER/ELECTROSTATIC

    ADSORPTION

    A. Introduction

    We shall restrict our consideration here to the simplest electrochemical

    case: the electrical double layer, which is not complicated by charge trans-

    fer or by specific adsorption. At first glance it would seem that, if there is no

    change in massand nothing happens in the bulk ofthe solution in which the

    electrode is immersed, the EQCM response should be zero. However,

    essentially all measurements show that in the double-layer region the

    frequency of the EQCM depends on potential. The effect is rathersmalla few Hz for crystals with fundamental frequencies of 510 MHz.

    In most cases reported in the literature [9096], experiments were

    performed employing cyclic voltammetry, with the potential extending to

    Electrochemical Quartz Crystal Microbalance 33

  • 8/2/2019 Electroanalytical Chemistry a Series of Advances Volume 22 Electroanalytical Chemistry

    53/328

    the region of oxide formation. All data obtained in this manner exhibitsome degree of hysteresis in the double layer region, which increases with

    sweep rate. It also increases when the anodic limit of potential is extended,

    i.e., when the potential is swept deeper into the oxide formation region.

    There is good reason to believe that this hysteresis is due to a memory

    effect of the interface, related to residual adsorbed oxygen or to some

    traces of dissolved gold remaining near the surface. In only a few papers

    [61,9799] has the response of the EQCM to changes in potential in the

    double-layer region been studied under experimental conditions, which

    excluded hysteresis. In these studies performed on gold andsilver, thelimits

    of potential during cycling were restricted to the double-layer region. Both

    metals have sufficiently extensive potential region where only electrostaticadsorption can take place. Moreover, it is well known that the surfaces of

    these metals do not undergo any changes in their morphology in the course

    of cycling in this restrictedpotential region. Measurements were conducted

    in electrolytes that are not specifically adsorbed to a significant extent.

    B. Some Typical Results

    All attempts to find rigorous quantitative correlation between data

    obtained in different laboratories have failed. This is not surprising,considering that the effects are rather small and the surfaces studied have

    different histories (technique of producing the EQCM, electrode pretreat-

    ment, etc.) and different morphologies. The surfaces employed in EQCM

    studies are always polycrystalline, even though they often have a preferredcrystal orientation. Ne