Download as pdf or txt
Download as pdf or txt
You are on page 1of 348

Han-Ill 

Yoo

Lectures
on Kinetic
Processes in
Materials
Lectures on Kinetic Processes in Materials
1.40

1.35
¬
log aH O= –3.29 –2.75
σ/10–2Ω–1cm–1

2
1.30

1.25

¬
1.20 log a H O= –2.75 –3.28
2

1.15

–2 0 2 4 6 8 10
t × 10–3/s

Twofold electrical conductivity relaxation upon hydration and reverse dehydration of


BaCe0.95Yb0.05O2.975 at 700  C in a fixed oxygen activity atmosphere. (From H.-I. Yoo et al.,
J. Electrochem. Soc. 156, B66 (2009))
Han-Ill Yoo

Lectures on Kinetic Processes


in Materials
Han-Ill Yoo
Seoul National University
Seoul, South Korea
Daegu Gyeongbuk Institute of Science & Technology
Daegu, South Korea

ISBN 978-3-030-25949-5 ISBN 978-3-030-25950-1 (eBook)


https://doi.org/10.1007/978-3-030-25950-1

© Springer Nature Switzerland AG 2020


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Dedicated to my mother who passed away on
March 1, 2018, and my father
Preface

Thermodynamics teaches us how to describe the state of existence of a system


material and in which direction it should change when its state of existence is
disturbed by “heat and beat” or “bake and shake.” But it never tells how fast it
does; it is kinetics that tells how to describe and evaluate this rate of change. The
former may, thus, be regarded as the necessity and the latter as the sufficiency for
materials engineering in general, and hence, these two subjects are taught as
mandatory courseworks in materials engineering since this discipline was born.
Kinetics is, however, notoriously known to be a not-so-easy subject to learn, as
much as thermodynamics does, to the students in materials disciplines including the
author himself in his student days far back 40 odd years ago. It is often said that
thermodynamics is the first activation barrier and kinetics the second, and the highest
as well, in materials courseworks.
Kinetic processes in materials typically involve chemical reactions and solid state
diffusion in parallel or in tandem. There are some excellent monographs, rather than
textbooks, on materials kinetics, but they are mostly somewhat too deep or too
extensive for the beginning students in materials engineering to learn as a semester
coursework. There has, thus, long been a need for an introductory textbook, rather
than another monograph, which is easy enough for those to learn.
This textbook is a precipitation of the author’s own 30-odd-year-long teaching
and learning in the Department of Materials Science and Engineering, Seoul
National University. It aims to provide senior-level undergraduates or beginning
graduates in materials disciplines with very fundamental and quantitative ideas on
kinetic processes in materials in general, which, no matter how diversified or
sophisticated materials applications grow, will forever continue to work. Further-
more, it is written in the easiest way possible based on the classic kinetics wisdoms
and almost verbatim as being lectured in class so that students, only with reasonable
knowledge on physical chemistry and calculus, can learn just by reading through
even without having to consult references at all.
This textbook is designed as a semester coursework for 24–25 lectures of 75 min
long each or 36–37 lectures of 50 min long each. At the end of each chapter are

vii
viii Preface

furnished a number of heuristic classic exercise problems dating back to C. Wagner


in 1950s, which were inherited to the author mostly from his life-long mentors and
would otherwise highly likely stop to be inherited.
There are two things missing in this textbook which often frustrate the novices in
particular: one is “It can be easily shown that. . .” and the other a long list of “further
readings” at the end of every each chapter. Instead are provided two appendices
particularly for immediate reference on transport and reaction kinetics in ionic solids,
a brief summary of the defect chemistry of ionic solids, and an English translation of
the main part of the pioneering classic paper by C. Wagner (Beitrag zur Theorie des
Anlaufvorgangs, Z. Phys. Chem. B21, 25 (1933)) which was inherited to the author
also from his mentors and would, otherwise, remain beyond reach due to the
language barrier depending on the reader. In addition, listed below are the classics
as references for advanced learning:
For in-depth diffusion mathematics
(1) J. Crank, Mathematics of Diffusion, 2nd edn. (Oxford University Press, 1975).
(2) H.S. Carslaw, J.C. Jeager, Conduction of Heat in Solids, 2nd edn. (Oxford
University Press, 1959).
For diffusion phenomena in general
(3) W. Jost, Diffusion in Solids, Liquids and Gases (Academic Press,
New York, 1960).
For diffusion, particularly in metals
(4) P.G. Shewmon, Diffusion in Solids (McGraw-Hill Book Co., Inc.,
New York, 1963).
For all solid state kinetics, particularly in ionic solids
(5) H. Schmalzried, Chemical Kinetics of Solids (VCH, Weinheim, 1995).
I first learned materials kinetics (then Course 3.21) from Prof. D. R. Sadoway,
isothermal and non-isothermal diffusion in ionic solids from Prof. B. J. Wuensch,
and kinetics in real world, electroceramics in particular, from Prof. H. L. Tuller all at
Massachusetts Institute of Technology, USA, and later, as an admirer of legend Carl
Wagner, the Onsagerian interference between ionic and electronic flows in mixed
ionic electronic conductor solids from Prof. H. Schmalzried at the University of
Hannover, Germany. I am very much indebted to these mentors for their teachings,
particularly to Prof. Schmalzried for letting me come across, in his own archive,
C. Wagner’s legendary lecture note, “Kinetics in Metallurgy,” with his own hand-
writings and drawings, taught at MIT in 1955. (The reader will encounter a couple of
photocopies of those in this textbook.)
I would like to thank the “Grande Petites,” Profs. J. Mizusaki (formerly, Tohoku
University, Japan), S. Yamaguchi (formerly, University of Tokyo, Japan), M. Martin
(RWTH Aachen University, Germany), J. Maier (Max-Planck Institute, Stuttgart,
Germany), T. Norby (University of Oslo, Norway), and K. D. Becker (formerly,
Technical University of Braunschweig, Germany), who have helped deepen and
Preface ix

sharpen my understanding of kinetics through enlightening discussions at the bien-


nial “Petite Workshop on Defect Chemical Nature of Advanced Materials” since it
was first nucleated at Seoul National University in 1989.
Special thanks go to Profs. M. Martin and R. de Souza both at Aachen University,
Germany, Prof. H. L. Tuller at MIT, and Prof. S. M. Haile at Northwestern
University, USA, for their exhaustive and critical reviews of the manuscript. Their
criticisms, suggestions, and corrections are gratefully acknowledged.
This textbook would not have taken its shape were it not for the psychological
pressure or stimuli from my former students in the Solid State Ionics Laboratory at
Seoul National University who were all taught this kinetics course. They had long
been continually encouraging, if not enforcing, me to write down a kinetics textbook
as being lectured in class. Furthermore, all the figures and illustrations have been
prepared mostly by my last PhD student, Dr. W. Joo, and partly by former PhD
student, Dr. J. Chun, and MS student, Ms. S.-W. Kim, whose encouragement and
sweat put in this project are greatly appreciated and acknowledged.
Last, but not the least of course, credit should go to my endearing and enduring
lifelong mate, my love.

Seoul, South Korea Han-Ill Yoo


December 2019
Contents

1 Diffusion in Continuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Fick’s First Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2.1 Definition of Diffusion and Fick’s First Law . . . . . . . . . 1
1.2.2 Mathematical Interlude: Gradient Operator . . . . . . . . . . . 2
1.2.3 Similar Linear Rate Laws . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.4 Limitations to Fick’s First Law . . . . . . . . . . . . . . . . . . . 4
1.2.5 Unit and Magnitudes of D . . . . . . . . . . . . . . . . . . . . . . 6
1.2.6 Application of the First Law . . . . . . . . . . . . . . . . . . . . . 7
1.3 Fick’s Second Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.1 Mass Conservation and Continuity Equation . . . . . . . . . 8
1.3.2 Mathematical Interlude: Divergence Operator . . . . . . . . . 10
1.3.3 Fick’s Second Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3.4 Evolution of c(x,t) with Time . . . . . . . . . . . . . . . . . . . . 13
1.3.5 Mathematical Similarity to Fourier’s Law . . . . . . . . . . . 14
1.4 Solutions to Fick’s Second Law . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4.1 Steady-State Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.2 Transient-State Solutions . . . . . . . . . . . . . . . . . . . . . . . 16
1.5 Diffusion from an Instantaneous Planar Source
or Thin-Film Source Solution . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5.1 Solution of Fick’s Second Law . . . . . . . . . . . . . . . . . . . 17
1.5.2 Salient Features of the Solution . . . . . . . . . . . . . . . . . . . 19
1.5.3 Leak Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.5.4 Mathematical Interlude: Error Function . . . . . . . . . . . . . 23
1.5.5 Back to the Leak Test . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.5.6 Variations of the Thin-Film Source Solution . . . . . . . . . 27
1.5.7 Reflection and Superposition Method . . . . . . . . . . . . . . 28
1.5.8 Application of the Thin-Film Source Solution . . . . . . . . 29
1.6 Semi-infinite Source Solution . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.6.1 Solution of Fick’s Second Law . . . . . . . . . . . . . . . . . . . 30
1.6.2 Salient Features of the Solution . . . . . . . . . . . . . . . . . . . 32
xi
xii Contents

1.6.3 Variations of the Error-Functional Solution . . . . . . . . . . 33


1.6.4 Generalization of Error-Functional Solutions . . . . . . . . . 35
1.6.5 Mathematical Interlude: Properties of erf(z) . . . . . . . . . . 36
1.7 Trigonometric Infinite Series Solution . . . . . . . . . . . . . . . . . . . . 36
1.7.1 Statement of the Problem . . . . . . . . . . . . . . . . . . . . . . . 36
1.7.2 Solution of Fick’s Second Law by
Separation-of-Variables Method . . . . . . . . . . . . . . . . . . 37
1.7.3 Mathematical Interlude: Orthogonality Theorem . . . . . . . 40
1.7.4 Evaluation of the Last Constant by the
Orthogonality Theorem . . . . . . . . . . . . . . . . . . . . . . . . 41
1.7.5 Application to the Trivial Initial-Condition Case . . . . . . 42
1.7.6 Salient Features of the Series Solution . . . . . . . . . . . . . . 44
1.7.7 First-Term Approximation or Long-Time Solution . . . . . 45
1.7.8 Average Concentration and the First-Term
Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
1.7.9 Homogenization Time . . . . . . . . . . . . . . . . . . . . . . . . . 47
1.8 Method of Laplace Transform . . . . . . . . . . . . . . . . . . . . . . . . . . 47
1.8.1 Definition of Laplace Transform . . . . . . . . . . . . . . . . . . 48
1.8.2 Application to Diffusion Problems . . . . . . . . . . . . . . . . 49
1.9 Solutions When D Is Not Constant . . . . . . . . . . . . . . . . . . . . . . . 51
1.9.1 When D ¼ D(t) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
1.9.2 When D ¼ D(c): Boltzmann-Matano Analysis . . . . . . . . 52
1.10 Diffusion in Higher Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . 56
1.11 Moving Boundary Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2 Atomic Theory of Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.2 A Naïve View of Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
2.3 Random Walk Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.4 Diffusion Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2.5 A Few General Relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
2.6 Derivation of Self-Diffusion Coefficient . . . . . . . . . . . . . . . . . . . 83
2.7 Defect (Vacancy) Diffusion Coefficient . . . . . . . . . . . . . . . . . . . 86
2.8 Thermodynamic Variables Dependence of DA . . . . . . . . . . . . . . 87
2.9 Correlation Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.10 Quantitative Treatment of the Correlation Factor . . . . . . . . . . . . . 93
2.11 Various Diffusivities So Far . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3 Chemical Reaction Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
3.2 Chemical Reaction and Rate Law . . . . . . . . . . . . . . . . . . . . . . . . 102
3.2.1 Reaction System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Contents xiii

3.2.2 Chemical Reaction Rate . . . . . . . . . . . . . . . . . . . . . . . . 102


3.2.3 Concentration Change in General . . . . . . . . . . . . . . . . . 103
3.2.4 Reaction Rate Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3.3 Simple Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.3.1 Integration Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.3.2 Half-Life Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.3.3 Method of Initial Rates . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.4 Complex Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.4.1 Reversible 1st-Order Reaction . . . . . . . . . . . . . . . . . . . . 110
3.4.2 Consecutive 1st-Order Reaction . . . . . . . . . . . . . . . . . . . 111
3.4.3 Concurrent Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
3.5 Mixed-Controlled Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.6 Thermodynamic Restrictions on the Rate Laws . . . . . . . . . . . . . . 117
3.7 Temperature-Dependence of the Rate-Law Constant . . . . . . . . . . 120
3.7.1 Absolute Reaction Rate Theory . . . . . . . . . . . . . . . . . . . 121
3.8 Adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3.8.1 Langmuir Adsorption (Ideal Adsorption) . . . . . . . . . . . . 124
3.8.2 Dissociative Adsorption . . . . . . . . . . . . . . . . . . . . . . . . 126
3.8.3 Competitive Adsorption . . . . . . . . . . . . . . . . . . . . . . . . 127
3.8.4 Multilayer Adsorption: BET Isotherm . . . . . . . . . . . . . . 128
3.9 Langmuir Evaporation (Ideal Evaporation) . . . . . . . . . . . . . . . . . 131
3.10 Gibbs-Langmuir Isotherm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.10.1 Thermodynamics of Interface and Gibbs Isotherm . . . . . 133
3.10.2 Gibbs-Langmuir Isotherm . . . . . . . . . . . . . . . . . . . . . . . 135
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4 Diffusion in Concentration Gradients . . . . . . . . . . . . . . . . . . . . . . . . 145
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.2 Kirkendall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
4.3 Darken’s Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.3.1 Diffusion and Drift . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.3.2 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.4 Thermodynamic Treatment of Diffusion . . . . . . . . . . . . . . . . . . . 154
4.5 Thermodynamic Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.6 Evidence for ∇μ, Rather than ∇c, Being
the Diffusional Driving Force . . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.7 Atomistic Interpretation of the Kirkendall Effect . . . . . . . . . . . . . 162
4.8 Summary of the Various Diffusivities . . . . . . . . . . . . . . . . . . . . . 167
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
5 Kinetics of Phase Transformation: Initial Stage . . . . . . . . . . . . . . . . 173
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
5.2 Homogeneous Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
5.2.1 Driving Force for a Phase Change . . . . . . . . . . . . . . . . . 174
xiv Contents

5.2.2 Critical Size and Energy Barrier for Nucleation . . . . . . . 175


5.2.3 Nucleation Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
5.2.4 Nucleation Involving Strain Energy . . . . . . . . . . . . . . . . 181
5.3 Heterogeneous Nucleation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.4 Spinodal Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
5.4.1 Phase Stability Against Compositional Fluctuation . . . . . 190
5.4.2 Gibbs Free Energy vs. Composition Diagram . . . . . . . . . 192
5.4.3 Fate of Composition Fluctuations . . . . . . . . . . . . . . . . . 193
5.4.4 How Phase Transition Initiates? . . . . . . . . . . . . . . . . . . 195
5.4.5 Origin of Interfacial Energy: Gradient Energy . . . . . . . . 196
5.4.6 Cahn’s Theory on Spinodal Decomposition . . . . . . . . . . 199
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6 Kinetics of Phase Transformation: Later Stage . . . . . . . . . . . . . . . . . 215
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6.2 Type I: Growth Without Phase Change and Without
Composition Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6.2.1 Driving Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
6.2.2 Growth Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
6.3 Type II: Growth With Phase Change and Without
Composition Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6.4 Type III: Growth With Phase Change and With
Composition Change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6.4.1 Normal Solidification of a Binary Liquid . . . . . . . . . . . . 222
6.4.2 Growth of Spherical Precipitates from
Solid Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
6.5 Overall Rate of Transformation . . . . . . . . . . . . . . . . . . . . . . . . . 231
6.6 Last Stage of Phase Transformation: Precipitate
Coarsening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
6.6.1 Thomson-Freundlich Equation . . . . . . . . . . . . . . . . . . . 234
6.6.2 Coarsening Kinetics: Exact Formulation . . . . . . . . . . . . 235
6.6.3 Coarsening Kinetics: Highly Dispersed Case . . . . . . . . . 236
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
7 Diffusion in Ionic Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
7.1 Introduction: System and Mobile Species . . . . . . . . . . . . . . . . . . 247
7.2 Mobile Species and Diffusion Mechanisms . . . . . . . . . . . . . . . . . 248
7.3 Flux Equation and a Few Definitions . . . . . . . . . . . . . . . . . . . . . 250
7.4 Self-Diffusivities in Stoichiometric Compounds . . . . . . . . . . . . . 254
7.4.1 Defect Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
7.4.2 Self-Diffusivities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
7.5 Self-Diffusivities in Nonstoichiometric Compounds . . . . . . . . . . 259
7.5.1 Defect Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
Contents xv

7.5.2 Self-Diffusivities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263


7.5.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
7.6 Diffusion in Concentration Gradients . . . . . . . . . . . . . . . . . . . . . 270
7.6.1 Component Flux Equations . . . . . . . . . . . . . . . . . . . . . . 271
7.6.2 Nonstoichiometry Flux . . . . . . . . . . . . . . . . . . . . . . . . . 273
7.6.3 Chemical Diffusion Coefficient . . . . . . . . . . . . . . . . . . . 274
7.7 Nonstoichiometry Re-Equilibration Kinetics . . . . . . . . . . . . . . . . 277
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case . . . . . . . . 285
8.1 Introduction: Tarnishing Reaction . . . . . . . . . . . . . . . . . . . . . . . 285
8.2 Thermodynamics of Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . 286
8.3 Kinetic Steps and Rate Laws Observed . . . . . . . . . . . . . . . . . . . 288
8.4 Parabolic Rate Law and Wagner’s Theory . . . . . . . . . . . . . . . . . 289
8.5 Remark on the Quasi-steady State Assumption . . . . . . . . . . . . . . 294
8.6 Telltale Marker Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
8.7 An Example: Oxidation of Co to CoO . . . . . . . . . . . . . . . . . . . . 296
8.8 Internal Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311

Appendix I: Defect Chemistry of Solid State Ionic Compounds . . . . . . . 313

Appendix II: English Translation: Carl Wagner, “Beitrag zur


Theorie des Anlaufvorgangs,” Z. Phys. Chem.
B21 (1933) 25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
Chapter 1
Diffusion in Continuum

1.1 Introduction

All materials can be processed only because their constituent atoms (or ions) are
mobile under the action of thermodynamic driving forces. The atomic mobility is
a prime mover of the processing but simultaneously provides a kinetic barrier to the
processing. Furthermore, the atomic or ionic mobility itself generates a variety of
technological applications of the materials but also causes the devices thereof to
degrade in function and eventually to fail. Understanding and controlling the atomic
mobility, thus, is one of the fundamentals of the discipline of materials science and
engineering in general.
A typical manifestation of the atomic mobility is the phenomenon of diffusion,
and the diffusion rate is a measure of the atomic mobility. In this chapter, we will
learn an empirical rate law describing the phenomenon of diffusion and the math-
ematics to solve the diffusion equation for the concentration as a function of position
and time. We will regard our system as a continuum, totally neglecting its atomic
nature for the time being.
At the end of this chapter, we will have come to understand the nature of solutions
to the diffusion equation and how to solve analytically or graphically the diffusion
equation appropriate for our system in question.

1.2 Fick’s First Law

1.2.1 Definition of Diffusion and Fick’s First Law

Diffusion refers to “a process leading in an inhomogeneous multi-component


one-phase system finally to an equalization of concentration (homogenization) by
non-convective flows of the components under the action of their concentration

© Springer Nature Switzerland AG 2020 1


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_1
2 1 Diffusion in Continuum

gradients” [1]. As early as 1855 [2], it was known empirically that when there is a
difference in concentration of a species along, say, x-direction, in a continuum at
uniform temperature and uniform pressure, there arises a flux or flow of the species
from a higher concentration to a lower one, and the flux is linearly proportional to the
concentration difference. The flux is defined as the amount of the species passing
through a unit area per unit time or amount/(m2s). Letting J be the flux and c the
concentration or amount of the species per unit volume, we can represent this
empirical observation as

∂c
J ¼ D ð1:1aÞ
∂x

or in a vectorial form independent of the coordinates systems, as

! !
J ¼ D∇c: ð1:1bÞ
! !
It is noted that the flux J and the concentration gradient ∇c are vector
quantities; hence, D should be a tensor quantity as discussed later. For the sake of
simplicity, we will often omit the vector notation (!) and treat D as a scalar in the
absence of any possibility of misunderstanding. Here, the proportionality constant1
D is called the diffusion coefficient or diffusivity in short. This empirical law is
referred to as Fick’s first law in honor of the discoverer of the phenomenon, A. E.
Fick [2].

1.2.2 Mathematical Interlude: Gradient Operator


!
The gradient operator in Eq. (1.1b), denoted as ∇ or “Grad,” is defined from the
elementary vector calculus depending on the coordinates systems as:
(i) In the rectangular coordinate system (x, y, z) with the three orthogonal unit
vectors (^i, ^j, ^
k),

! ∂^ ∂^ ∂^
Grad ¼ ∇ ¼ iþ j þ k: ð1:2aÞ
∂x ∂y ∂z

(ii) In the cylindrical coordinate system (r, θ, z) with the orthogonal unit vectors
(^er , ^eθ , ^ez ),

1
It looks like a proportionality “constant”, but, in general, can be a function of c at a fixed
temperature.
1.2 Fick’s First Law 3

! ∂ ∂ ∂
Grad ¼ ∇ ¼ ^er þ ^eθ þ ^ez : ð1:2bÞ
∂r r∂θ ∂z

(iii) In the spherical coordinate system (r, θ, φ) with the orthogonal unit vectors
(^er , ^eθ , ^eφ ), where r is the radial distance from the origin, θ the polar angle
measured from the z-axis, and φ the azimuthal angle measured from the xz
plane,

! ∂ ∂ ∂
Grad ¼ ∇ ¼ ^er þ ^eθ þ ^eφ : ð1:2cÞ
∂r r∂θ r sin θ∂φ

In the latter two curvilinear coordinate systems, you can easily figure out
why the denominators should take the form rdθ or rsinθdφ considering the
distance change with dθ and dφ in the corresponding infinitesimal volume
element.
These are nothing but a decomposition of the change of something scalar in
the direction of the steepest change, into its three orthogonal components in
each coordinates system. For example, the difference in concentration Δc
between the position (x, y, z) and (x + Δx, y + Δy, z + Δz) may be written as

∂c ∂c ∂c ! !
Δc ¼ Δx þ Δy þ Δz  ∇c  Δ r :
∂x ∂y ∂z
!
We see that the rate of change of c in any direction of Δ r is the component
! !
of ∇c in that direction. It follows that the direction of ∇c is that in which it has
the largest possible component or the direction in which c changes the most
steeply.

1.2.3 Similar Linear Rate Laws

In nature, there are some other rate laws that are quite similar to Fick’s law,
Eq. (1.1). The first example may be Fourier’s law of heat conduction, proposed
by J. Fourier in 1822. When there is a temperature gradient ∇T, then there arises a
heat flux Jq (“amount of heat”/unit area/unit time) down the gradient in proportion to
the magnitude of the temperature gradient,
* *
Jq ¼ κ∇T, ð1:3Þ

where the proportionality constant κ is the thermal conductivity.


When there is an electric potential gradient ∇ϕ, then there arises a charge flux
(amount of charge/unit area/unit time) or current density, i, down the gradient in
proportion to the magnitude of the gradient,
4 1 Diffusion in Continuum

* *
i ¼ σ∇Φ: ð1:4Þ

It is Ohm’s law of electric conduction, proposed by G. Ohm in 1827, and the


proportionality constant σ is the electrical conductivity.
Because all these fluxes are linearly proportional to the relevant gradient, they are
called the linear rate laws.

1.2.4 Limitations to Fick’s First Law

Here, we have to note a few things concerning Fick’s law and other linear rate
laws:
(i) The law is valid only when the cause for the flux, say, ∇c or driving force,2 is
not extremely large. Otherwise, the law would break down. One typical example
may be a ceramic variable-resistor or varistor. In this case, when an applied
electrical potential gradient is larger than a critical value, the current, which
would otherwise obey Ohm’s law, increases much more rapidly than linear or

i / ð∇ϕÞα with α  1: ð1:5Þ

Similar behavior may be expected for the other rate laws, too, when the
relevant gradient or driving force is very large.
One can actually write a flux J as a Taylor series in terms of its driving force X
(¼ ∇c, ∇ϕ, ∇T, etc.) as
  2 
∂J 1 ∂ J
J ¼ J0 þ Xþ X2 þ    ð1:6Þ
∂X X¼0 2! ∂X2 X¼0

where J0  0 identically because any flux must disappear when X ¼ 0 or in


equilibrium. We can then immediately recognize that Fick’s first law appears
only when the second- and higher-order terms are negligibly small compared
with the first-order term or the gradient is small enough:
 
∂J
J X ¼ D∇c: ð1:7Þ
∂X X¼0

2
In strict sense, the gradients, ∇c, ∇T and ∇ϕ themselves may not be called the “driving forces” per
se because they do not have a force unit, e.g., N(¼Newton). The thermodynamically correct driving
forces are defined later in Chaps. 4 and 7.
1.2 Fick’s First Law 5

Do not forget that all the linear rate laws are empirical and, strictly speaking,
are true only when the driving force X is small enough. As far as diffusion is
concerned, however, Fick’s first law remains valid almost always.
(ii) Secondly, you should recognize in Eq. (1.1b) that the diffusivity D links a
! !
vector J and another vector ∇c. In the rectangular coordinate system, for
example, a vector consists of three components in the three orthogonal direc-
!
tions x1(¼x), x2(¼y), and x3(¼z), say, J ¼ ðJ1 , J2 , J3 Þ. Similarly, the gradient
does, too. Therefore, D in Eq. (1.1b) couples the three components on the left-
hand side and another three on the right-hand side, thus, comprising nine
components in principle:

0 1 0 10 1
J1 D11 D12 D13 ∂c=∂x1
B C B CB C
@ J2 A ¼ @ D21 D22 D23 A@ ∂c=∂x2 A: ð1:8Þ
J3 D31 D32 D33 ∂c=∂x3

Such a coupling coefficient is called a tensor of the second rank compris-


ing the nine components. As the symmetry of the system increases, the number
of independent components of D is reduced. For the case of the cubic crystal
class, all the second rank tensor properties, the diffusivity D, heat conduc-
tivity κ, and electrical conductivity σ, turn isotropic, that is, the property is
independent of crystallographic directions, and only one element remains out
of the nine Dij (i,j ¼ 1, 2, 3) or

Dij ¼ δij D ð1:9Þ

where δij is the Kronecker delta or δij ¼ 0 if i 6¼ j and δij ¼ 1 if i ¼ j. In reality


in materials science and engineering, we are normally playing with polycrys-
talline materials. If those polycrystalline materials are random in grain orien-
tations, we may safely regard our system, even though each grain is
anisotropic, as isotropic with respect to the second rank tensor properties.
If isotropic, then Fick’s law in Eq. (1.1b) tells that the flux J arises in the
same direction as ∇c or perpendicular to the iso-concentration surfaces.
Otherwise, the flux and the gradient would not be parallel with each other, or
the flux direction would not be normal to the iso-concentration surfaces (see
Fig. 1.1). We will not pursue any further this orientation-dependent

! !
Fig. 1.1 J and ∇c in J (b)
isotropic (a) and J (a)
non-isotropic medium (b)
∇C
6 1 Diffusion in Continuum

characteristics of tensor properties. If you are interested in this sort of tensor


properties, you are referred to a classic textbook by F. Nye [3], which is
exhaustively dealing with this problem in really an elegant way. In any case,
it is emphasized that our D is formally a second-rank tensor property.
(iii) The third point is that Fick’s law is only true in the absence of the other
thermodynamic driving forces than a concentration gradient. Actually, a
mass flux can be driven by any driving force of the same tensorial character,
not only the concentration gradient ∇c but also a temperature gradient ∇T,
an electric potential gradient ∇ϕ, etc. or

J ¼ D∇c þ LT ∇T þ Lϕ ∇ϕ þ    ð1:10Þ

where LT and Lϕ are the coupling coefficients analogous to the diffusivity


D. While the mass flow driven by a concentration gradient is called the
(ordinary) diffusion, that driven by a temperature gradient is termed
thermomigration, thermotransport, or thermophoresis and that driven by
an electric potential gradient electromigration or electrotransport. The latter
indirect or cross phenomena are important subjects in the discipline of irre-
versible thermodynamics. If you are interested in these indirect, cross phe-
nomena, you are referred to Denbigh [4], de Groot [5], and Howard and
Lidiard [6], among others.

1.2.5 Unit and Magnitudes of D

What kind of unit does the diffusivity have? A dimensional analysis of Fick’s first
law, Eq. (1.1) indicates that

½ J ½quantity m2 =s
½D ¼ ¼ : ð1:11Þ
½∇c ½quantity m3 =m

When the identical units are given to the [quantity] or amount in the denominator
and in the numerator of Eq. (1.11), the diffusivity has the units of m2/s. As regards
the quantity for the concentration, we will employ the number of moles or atoms in
this book. In any case, the unit, m2/s, looks not so friendly, but you will come to
understand shortly what this unit means in connection to the physical meaning of the
diffusion coefficient.
It may now be helpful to get a feeling about the orders of magnitude of the
diffusivities. For most elemental metals and oxides, the diffusivities are typically
on the order of 1012 m2/s at their melting points; for elemental semiconductors,
say, Si, the diffusivity is on the order of 10–15 ~ 16 m2/s around their melting
points; In water, the diffusivities of ions are typically on the order of 10–8 ~ 9 m2/s
(¼~ 1 cm2/day) at room temperature; in gases, the diffusivity is on the order of
1.2 Fick’s First Law 7

105 m2/s at room temperature. All these diffusivities refer to the self-diffusivities,
the meaning of which will become clear in the next chapter.

1.2.6 Application of the First Law

One can employ Fick’s first law to determine the diffusivity in the steady state. Let
us consider, for instance, an infinite slab of iron with a thickness L. Suppose that
one side of the slab, say, x ¼ 0, is exposed to a carburizing gas with a fixed carbon
activity and the other side, say, x ¼ L, to another gas with a lower carbon activity.
If we keep the slab in this situation for a sufficiently long time, the carbon
concentrations at the surfaces x ¼ 0 and L are fixed at c ¼ c0 and c ¼ cL, in
equilibrium with the higher activity gas and lower activity one, respectively, and
carbon diffuses from x ¼ 0 to x ¼ L at a constant rate down the gradient of carbon
concentration that is also invariant with time. This kind of time-invariant state is
called a steady state and will be revisited later on. It is easy to expect that in order
to keep the flux of carbon J constant, the product of D and the steady-state gradient
dc/dx should be constant, suggesting the concentration profiles as illustrated in
Fig. 1.2, depending on whether D is constant (a), increasing (b), or decreasing
(c) with the carbon concentration c. From the concentration distribution in the
steady state, thus, one can tell how D varies with the concentration.

c c c
c0 c0 c0

cL cL cL

X X X

(a) (b) (c)

Fig. 1.2 Correlation between D and dc/dx for J ¼ constant in the steady state: when D is constant
(a), increasing (b), and decreasing (c) with c

The spatial distribution of the solute, say, carbon, can be readily determined
experimentally. Given the surface area of the slab, by analyzing the gas swept over
the lower activity surface for known period of time, one can determine the flux also,
thus evaluating the diffusivity as a function of the position across the slab or as a
function of the solute concentration.
This kind of steady-state experiment was carried out, e.g., by R. P. Smith [7] in
1953 to see the composition dependence of the carbon diffusivity in γ-iron. The
only difference is that in his experiment, a hollow cylinder, instead of a slab, was
employed with a carburizing gas passed externally and a decarburizing gas
8 1 Diffusion in Continuum

flown internally. The analysis of the similar experimental results is left as an


exercise for you at the end of this chapter.
This steady-state method is often used to determine the gas permeability of a
solid also, e.g., a glass. In this measurement, a solid slab with a thickness Δx is
subjected to a gas pressure difference ΔP¼P2P1. If the equilibrium
concentration of the gas in the system solid is given as c ¼ SP where S is the
proportionality constant called the solubility, Fick’s first law may be written for thin
enough a slab as

∂c ΔP
J ¼ D  DS : ð1:12Þ
∂x Δx

By measuring the permeation flux J in the steady state for given pressure difference
ΔP, one can determine the product DS, called the gas permeability. The gas
permeability apparently corresponds to the flux when a unit pressure difference is
imposed across a unit thickness.

1.3 Fick’s Second Law

Fick’s first law only says that when there is a concentration gradient, a mass flux
arises down the gradient. As diffusion proceeds, how does the local concentration
change with time?

1.3.1 Mass Conservation and Continuity Equation

Let us consider a volume V enclosed by a surface A (see Fig. 1.3). The total amount
Q of the species within the closed volume V is obviously the summation or
integration of the amount of the species in an infinitesimal volume element, cdV,
over the entire volume V
Z
Q¼ cdV: ð1:13Þ
V

Note here that the local concentration c is, in general, a function of position and
time, and as the volume element is of three dimension, namely, dV ¼ dxdydz in the
rectangular coordinates, for example, the volume integral here should be under-
stood to be a triple integral with respect to dx, dy, and dz. The time change of the
total amount Q is then
1.3 Fick’s Second Law 9

Fig. 1.3 A volume V J


enclosed by a surface “A.” n̂
Any decrease in the amount A
of species inside is due to
dA
the flux J
V

Z Z
dQ d ∂c
¼ cdV ¼ dV: ð1:14Þ
dt dt ∂t
V V

You know why the total derivative (d/dt) is altered to the partial derivative (∂/
∂t) upon jumping across the integral symbol.
If there is no source or sink within the volume V, then the change of the total
amount of the species in the volume will be only due to the flux of the species across
the surface. In vector algebra, the flux is always defined as positive when it comes
out of a closed volume. Thus, the total amount of the species coming out of the
closed volume per unit time should be the summation or integration of the normal
component of the flux times the surface area element dA, over the entire surface
A. Letting ^n denote the unit normal vector to the surface element dA, the temporal
change of the total amount of the species is then
Z
dQ !
¼ J  ^ndA: ð1:15Þ
dt
A

As the surface element is geometrically of two dimensions, say, dA ¼ dxdy, in


the rectangular coordinates, one should understand that this surface integral be a
double integral with respect to dx and dy.
Regarding such a surface integral, there is a very amusing theorem, known as the
Gauss Theorem in vector calculus. This theorem simply allows you to transform the
surface integral over the closed volume to a volume integral for that closed
volume, leading to the definition of a vector operator, divergence abbreviated as
!
div or ∇ . It goes in this way: Consider an infinitesimal parallelepiped at the
coordinates (x,y,z) with a volume ΔV¼ΔxΔyΔz (Fig. 1.4), where an
x-component flux Jx is coming into the volume ΔV across the surface ΔA ¼ ΔyΔz
at position x and out of the volume across the surface ΔA ¼ ΔyΔz at x+Δx and so on
!
in y and z direction, respectively. Noting that J  ^n ¼ Jx ðxÞ upon the surface ΔyΔz
!
at x and J  ^
n ¼ þJx ðx þ ΔxÞ upon the parallel surface at x + Δx and so on for the
rest of the surfaces, let us carry out the surface integral in Eq. (1.15) for an
infinitesimal volume element in Fig. 1.4:
10 1 Diffusion in Continuum

Fig. 1.4 A volume element


ΔxΔyΔz with a flux
component Jx(x) coming in J x ( x) J x ( x + Δx)
at x and a flux Jx(x + Δx) Δz
coming out at x + Δx and so
on with other faces Δy
( x, y, z ) Δx

Z
!
J ^
ndA ¼ ½Jx ðx þ ΔxÞ  Jx ðxÞΔyΔz
A ð1:16Þ
 
þ Jy ðy þ ΔyÞ  Jy ðyÞ ΔzΔx þ ½Jz ðz þ ΔzÞ  Jz ðzÞΔxΔy:

As Jx(x + Δx)  Jx(x) ¼ (∂Jx/∂x)Δx for small enough Δx and so on, Eq. (1.16)
now takes the form
Z   Z
! ∂Jx ∂Jy ∂Jz ! !
J ^
ndA ¼ þ þ ΔxΔyΔz  ∇  J dV: ð1:17Þ
∂x ∂y ∂z
A V

! !
Here, the vector ∇ operating on a vector quantity J as such is called the
divergence and often denoted equivalently as div. The definition of the divergence
is self-evident from this equation, which will nevertheless be dealt with in more
detail for different coordinate systems later on.
By combining Eqs. (1.14), (1.15), and (1.17), one finally obtains
Z Z
∂c ! !
dV ¼  ∇  J dV ð1:18aÞ
∂t
V V

or, by getting rid of the identical integral operators from both sides,

∂c ! !
¼ ∇  J : ð1:18bÞ
∂t

This is a rigorous representation of the mass conservation when there is no local


source or sink and often called a continuity equation. In other words, what goes in
and does not go out, stays there [8]. Note that this is always true in the absence of the
local sources or sinks.

1.3.2 Mathematical Interlude: Divergence Operator

By executing the surface integral for an infinitesimal volume element in the three
different coordinate systems, one can easily derive the expression for the divergence
operator as:
1.3 Fick’s Second Law 11

(i) In the rectangular coordinates (x, y, z) with the volume element dV ¼ dxdydz

! ∂^ ∂^ ∂ ^
div ¼ ∇ ¼ iþ j þ k: ð1:19aÞ
∂x ∂y ∂z

(ii) In the cylindrical coordinates (r, θ, z) with dV ¼ dr(rdθ)dz

! ∂ ∂ ∂
div ¼ ∇ ¼ r^er þ ^eθ þ ^ez : ð1:19bÞ
r∂r r∂θ ∂z

(iii) In the spherical coordinates (r, θ, φ) with dV ¼ dr(rdθ)(rsinθdφ)

! ∂ 2 ∂ ∂
div ¼ ∇ ¼ r ^er þ sin θ^eθ þ ^eφ : ð1:19cÞ
r2 ∂r r sin θ∂θ r sin θ∂φ

In executing the surface integral over an infinitesimal volume element, you


should not forget that, while all the six lateral areas of the infinitesimal volume
element remain unchanged as x increases to x + Δx, etc., in the rectangular
coordinates (see Fig. 1.4), they are changing as r increases to r + Δr for the rest
of the two curvilinear coordinates and even as θ to θ + Δθ for the spherical
coordinates in particular. Then, you will get the right expressions for the divergence
in the latter cylindrical and spherical coordinate systems. This derivation will be left
for your amusement.

1.3.3 Fick’s Second Law

Let us now substitute Fick’s first law in Eq. (1.18) for J to get
 ! 
∂c !
¼ ∇  D∇c : ð1:20Þ
∂t

This is called Fick’s second law. Note that while the continuity equation,
Eq. (1.18), is always true, this second law is still empirical as much as the first
law is.
By replacing the full expression for the gradient in Eq. (1.2) and for the
divergence in Eq. (1.19) depending on the coordinate systems, one can get the
full-fledged equation for the second law. Particularly when the diffusivity is not a
function of concentration or position, then the second law can be simplified as
12 1 Diffusion in Continuum

∂c
¼ D∇2 c: ð1:21Þ
∂t

Here, the inner product of divergence and gradient operators are abbreviated as
! !
∇  ∇  ∇2 and called the “del squared” or the Laplacian operator.
It is easy to show by combining Eqs. (1.2) and (1.19) that the Laplacian operator
takes the forms depending on the coordinate systems as:
(i) In the rectangular coordinates (x, y, z)

2 2 2
∂ ∂ ∂
div  grad ¼ ∇2 ¼ þ þ : ð1:22aÞ
∂x2 ∂y2 ∂z2

(ii) In the cylindrical coordinates (r,θ,z)

  2 2
1 ∂ ∂ ∂ ∂
div  grad ¼ ∇ ¼
2
r þ 2 2þ 2: ð1:22bÞ
r ∂r ∂r r ∂θ ∂z

(iii) In the spherical coordinates (r,θ,φ)

    2
1 ∂ 2∂ ∂ ∂ ∂
div  grad ¼ ∇ ¼ 2
2
r þ 2 sin θ þ 2 : ð1:22cÞ
r ∂r ∂r r sin θ∂θ ∂θ r sin θ∂φ2
2

If the concentration gradient is one-dimensional along, say, x-direction only,


Fick’s second law with the constant D, Eq. (1.21) takes the form

2
∂c ∂ c
¼D 2 ðin one dimensionÞ: ð1:23Þ
∂t ∂x

If the concentration gradient has a cylindrical symmetry or is only along the


radial direction, then the second law takes the form
 2 
∂c ∂ c 1 ∂c
¼D þ ðin cylindrical symmetryÞ: ð1:24Þ
∂t ∂r2 r ∂r

If the concentration has a spherical symmetry or the gradient is only along the
radial direction, then
 2 
∂c ∂ c 2 ∂c
¼D þ ðin spherical symmetryÞ: ð1:25Þ
∂t ∂r2 r ∂r
1.3 Fick’s Second Law 13

1.3.4 Evolution of c(x,t) with Time

How does the local concentration c(x,t) evolve with time? In order to get the answer,
let us first examine Eq. (1.23) with a constant diffusivity in one dimension. The
equation indicates that the local concentration increases or ∂c/∂t > 0 where the
spatial distribution of c is convex downward or ∂2c/∂x2 > 0 because more species
come in (Jin) and less go out (Jout) due to the gradient distribution and vice versa (see
Fig. 1.5). This interpretation is nothing but the story of divergence.

Fig. 1.5 The local c c


concentration increases (") J out J in J out J in
where its distribution is
locally convex downward ∂ 2c
(a) and decreases (#) where <0
∂c ∂x 2 ∂c
convex upward (b) >0 <0
∂t ∂ 2c ∂t
>0
∂x 2
x x
(a) (b)

Rigorously, the concentration of diffusing species as a function of position and


time, c(x,y,z;t) can be calculated by solving the differential equation, Eq. (1.20).
How to solve the equation? Suppose that you have a function c ¼ c(x,y,z,t), which is
often called the primitive function that is what you want to get by solving the
differential equation. Then this equation should satisfy Fick’s second law. But you
see, by differentiating with respect to time, you have lost the information on the time-
independent part. Also by differentiating twice with respect to positions x, y, z,
respectively, you have also lost six pieces of information. In order to completely
retrieve the primitive, the differential equation should be given together with
1 condition with respect to time t and 2 with respect to x, 2 with respect to y, and
finally 2 with respect to z. The condition with respect to time t ¼ 0 is called the initial
condition (IC), and the rest with respect to the position variables are called the
boundary conditions (BC). Then, in principle, you can retrieve the primitive
function by solving Fick’s second law. You will learn shortly how to solve the
equation, but solving Fick’s second law is not so trivial and more often makes
students in materials frustrated. But, one should not be frustrated. Solving differen-
tial equations is the job of mathematicians, not totally yours. They have worked very
hard to solve the equation for almost all the possible initial and boundary conditions
which you may be faced with in reality. An almost exhaustive compilation of the
solutions is the book by Crank [9]. Sometimes, you may have to work with the
initial and boundary conditions for which even Crank does not provide the solution.
The final resort will then be a classic by Carslaw and Jeager [10] which is actually
an exhaustive compilation of the solutions to the problem with heat conduction. The
heat conduction problem is mathematically exactly identical to the diffusion problem
as you will see now.
14 1 Diffusion in Continuum

1.3.5 Mathematical Similarity to Fourier’s Law

We have learned Fourier’s law of heat conduction as a linear rate law in Eq. (1.3).
As mass is conserved, energy is also conserved; therefore we also have a continuity
equation for energy simply by replacing the mass concentration “c” with the energy
density “u” and the mass flux J with the thermal energy or heat flux Jq in
Eq. (1.18b) as

∂u
¼ ∇  Jq : ð1:26Þ
∂t

Letting the specific heat (or molar heat capacity) at constant volume be CV and
the mass density (or molar density) be ρ, the energy density variation is simply

du ¼ ρCV dT: ð1:27Þ

Substituting Fourier’s law into the continuity equation, Eq. (1.26) and
rearranging, one has

∂T κ
¼ α∇2 T with α  ð1:28Þ
∂t ρCV

where we have assumed the thermal conductivity κ to be constant. You may call
this equation Fourier’s second law if you like. You can immediately recognize that
this equation is exactly of the same form as Fick’s second law for D being constant,
(Eq. 1.21) and then α takes the same unit as that of D. It is, thus, called the thermal
diffusivity.
You know that if the forms of differential equations are the same, their solutions
are also the same. The solution to the Fourier’s second law can be the solution to
Fick’s second law, of course, for the same initial and boundary conditions. All the
solutions in Carslaw and Jaeger can be transformed to those to the diffusion
problems simply by replacing α with D and T with c.

1.4 Solutions to Fick’s Second Law

You were told earlier that you do not have to worry about solving Fick’s second
law, a second-order partial differential equation, for yourself if you cannot.
Nevertheless, you should be prepared to know how to pick up the right solution to
your problem and how to read it once a solution is provided. In this section, you
will learn how to solve the differential equation and what types of solution
there are.
1.4 Solutions to Fick’s Second Law 15

If the boundary conditions in terms of the concentrations at the boundaries are


imposed upon our system, the concentration distribution starts to change from the
initial distribution and eventually reaches a time-independent distribution. The
changing period is called the transient state, and the final time-independent state
is called the steady state. Here we will start with the steady-state solutions and then
try to get the transient solutions.

1.4.1 Steady-State Solutions

The steady state is defined as the state in which a local property, the concentration in
the present case, no longer varies with time. Thus, the differential equation, Fick’s
second law, Eq. (1.20), which we will have to solve, will take the form at the steady
state as

∂c
¼ 0 ¼ ∇D∇c: ð1:29Þ
∂t

This is the situation in which, if in the rectangular coordinate system, the flux
J ¼ D∇c itself is constant as we considered earlier.3 If the diffusivity D is
independent of the concentration or not a function of the spatial coordinates, the
differential equation, Eq. (1.29), becomes

∇2 c ¼ 0: ð1:30Þ

This equation is called the Laplace equation which you have probably encoun-
tered in electrostatics concerning the electrostatic potential distribution, φ, in the
absence of space charge

∇2 φ ¼ 0: ð1:31Þ

You have to again recognize that because the forms of the equations are the same,
the solutions to the latter can be immediately adopted as the solutions to the steady-
state diffusion problem.
In general, Eq. (1.29) cannot be solved unless the concentration dependence of D
is known a priori. When D is a constant, however, we can get the general solution to
the steady-state distribution by integrating Eq. (1.30) as:
(i) In one-dimensional case

! ! !
3
Be aware that, when J ¼ constant, ∇  J 6¼ 0 in general in the curvilinear coordinates systems.
! !
For ∇  J ¼ 0 or, in the steady state, rJr ¼ constant and r2Jr ¼ 0 in cylindrically and spherically
symmetric diffusion, respectively, see Eq. (1.19) and also Eqs. (1.33) and (1.34).
16 1 Diffusion in Continuum

2
∂ c
∇2 c ¼ ¼ 0; cðxÞ ¼ A þ Bx: ð1:32Þ
∂x2

This corresponds to the case of Fig. 1.2a.


(ii) In cylindrical symmetry

  2
1 ∂ ∂c ∂ c 1 ∂c
∇ c¼
2
r ¼ 2þ ¼ 0; cðrÞ ¼ A þ Blnr: ð1:33Þ
r ∂r ∂r ∂r r ∂r

(iii) In spherical symmetry

  2
1 ∂ 2 ∂c ∂ c 2 ∂c B
∇ c¼ 2
2
r ¼ 2þ ¼ 0; cðrÞ ¼ A þ : ð1:34Þ
r ∂r ∂r ∂r r ∂r r

Here, the integration constants A and B are to be determined by the two


boundary conditions with respect to x or r.

1.4.2 Transient-State Solutions

The transient-state solutions, on the other hand, should be obtained by solving Fick’s
second law, Eq. (1.20), with the help of the appropriate initial condition (IC) and
boundary conditions (BC). The two situations may be distinguished: (i) when D is
a constant (i.e., D 6¼ D(c)) and (ii) when D depends on the position or concentration
(i.e., D ¼ D(c)). For the former case, the differential equation to solve is Eq. (1.21),
and for the latter, the general form of Fick’s second law, Eq. (1.20) must be solved.
We will solve for D ¼ constant first and then deal with the second case in one
dimension.
When D ¼ constant, what one should do is to solve Eq. (1.23)

2
∂c ∂ c
¼D 2
∂t ∂x

with the aid of one IC and two BCs to get back the primitive, c ¼ c(x,t).
The solutions may be either of the two types: the error-functional solution and
the trigonometric infinite series solution. The former is often called the short-time
solution because it describes the temporal variation of the concentration at short
times and the latter the long-time solution because it describes the temporal
variation at long times with appropriate approximation. You will see shortly why
they are so.
In order to understand the nature of the solutions, we will first study the error-
functional, short-time solutions: (i) the solution for an instantaneous planar
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 17

source or thin-film source solution and (ii) the solution for the semi-infinite source
solution. Subsequently follows (iii) the infinite trigonometric series solution.
Through this course of exposition, you will learn two different methods of solving
the diffusion equation, the method of superposition and the method of separation-
of-variables. We will begin with the thin-film solution under a separate heading.

1.5 Diffusion from an Instantaneous Planar Source or


Thin-Film Source Solution

1.5.1 Solution of Fick’s Second Law

Let us consider two semi-infinite bars with an instantaneous planar source at the
center x¼0 (Fig. 1.6a). As an example, you may imagine two (infinitely!) long
pieces of radish with their heads truncated flat. Apply a thin layer of ink upon the flat
end of a piece, and then attach the other piece to it, and you will have a
one-dimensional diffusion media with an instantaneous planar source of ink. A
plane is supposed not to have thickness by definition, but in reality any planar
source cannot be made to have zero thickness, always having a finite thickness
however thin it is made. It is, thus, called a thin-film source.

Fig. 1.6 (a) Instantaneous


planar source between two
semi-infinite bars. (b)
Concentration profiles
expected with time
(t ¼ 0 < t1 < t2 < t3) 0 x

(a)

c
t=0

t1

t2

t3

0 x

(b)
18 1 Diffusion in Continuum

What would happen then to this thin-film source (Fig. 1.6a)? This source will
diffuse away symmetrically from where it is placed initially, x ¼ 0. It will not be hard
to imagine that the concentration distribution with time may look like what is shown
in Fig. 1.6b.
The initial distribution of the solute or initial condition may be written as

ðiÞ c ¼ 1 at x ¼ 0 at t ¼ 0: ð1:35aÞ
ðiiÞ c ¼ 0 at x > 0 at t ¼ 0: ð1:35bÞ
ðiiiÞ c ¼ 0 at x < 0 at t ¼ 0: ð1:35cÞ

The boundary conditions are:

ðivÞ c ¼ 0 as x ! 1 at all t ð1:35dÞ


ðvÞ c ¼ 0 as x ! 1 at all t ð1:35eÞ

because we assume here that our diffusion medium is semi-infinite in both


directions.
We said earlier that to solve the diffusion equation we need only one IC and two
BCs for a one-dimensional problem, but here we have even five conditions instead of
three. Do not worry that the solution may be overdetermined and let us simply accept
it now.
If you solve Fick’s second law with the use of above “five” conditions, one can
get the solution, concentration distribution c(x,t), as
 
M x2
cðx, tÞ ¼ pffiffiffiffiffiffiffiffiffiffi exp  ð1:36Þ
4πDt 4Dt

where M is the amount of source per unit area of the planar source at x ¼ 0 (see
Fig. 1.6a).
How do we know this is the solution to Fick’s second law, Eq. (1.23)? For a
function to be the solution to a differential equation, it should satisfy the equation
itself as well as all the initial and boundary conditions. First try yourself to differ-
entiate this solution, Eq. (1.36), with respect to t and x, respectively, to see whether it
indeed satisfies the differential equation, Fick’s second law, and then try to confirm
that this solution indeed satisfies those five conditions in Eq. (1.35). In doing this,
you may recognize that all these conditions eventually get pyou ffi only three distin-
guishable conditions in terms of the single variable η ¼ x= t , namely, c ¼ 1 at
η ¼ 0 and c ¼ 0 as η ! 1. Do you have any difficulty with pffi the very first condition,
Eq. (1.35a)? Do not forget that x approaches 0 faster than t does, even if they both
approach 0.
We have now confirmed that Eq. (1.36) is indeed the solution to the thin-film
source problem describing the concentration distribution with time as in Fig. 1.6b.
This bell-shaped distribution is called a Gaussian distribution, normal distribu-
tion, or natural distribution as it was first invented by German mathematician
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 19

Fig. 1.7 Banknote of 10 German Mark before Euro, bearing the picture of Gauss and the Gaussian
curve with its equation

C. Gauss (1777–1855), called the prince of mathematics, and describes well any
natural distribution. This natural distribution equation,
pffiffiffiffiffiffiffi
ffi Eq. (1.36), with a bit modi-
fied by substituting f(x) for c(x,t)/M, σ for 2Dt, and (x  μ)2 for x2, was printed on
the front face of the old German 10-Mark (Zehn Deutsche Mark) note which was
(sadly!) removed from circulation with the advent of Euro. See Fig. 1.7 for your
amusement.

1.5.2 Salient Features of the Solution

Let us examine the thin-film source solution in more detail.


(i) Concentration profiles: You can see in Fig. 1.6b that the concentration distri-
bution c(x,t) gets more and more spread as time passes, but the area below it
should remain the same all the time. It is because the mass must always be
conserved or

Z þ1
cðx, tÞdx ¼ M: ð1:37Þ
1

(ii) Maximum concentration: The maximum concentration of the profile at x ¼ 0


decreases with time as

M
cð0, tÞ ¼ pffiffiffiffiffiffiffiffi / t1=2 : ð1:38Þ
2 πDt
20 1 Diffusion in Continuum

The half-power dependence on time, be it +1/2 or 1/2, is typical of


diffusion-controlled kinetics.
(iii) Flux: For a given concentration distribution at an instant (Fig. 1.8a), the local
flux J is calculated as

∂c
J ¼ D ,
∂x

that is as shown in Fig. 1.8b. The flux is antisymmetric, indicating that the mass
diffuses away symmetrically from the source at x ¼ 0. This is why the profile in
Fig. 1.8a appears symmetric and the flux at x ¼ 0 is 0. This is simply because
the diffusion medium is isotropic.
(iv) Local concentration change: The change of local concentration with time is
calculated as

2
∂c ∂ c
¼D 2,
∂t ∂x

that is as shown in Fig. 1.8c. You see, the concentration in the region with a
positive curvature or convex downward increases with time and vice versa,
compare with Fig. 1.8a.

Fig. 1.8 Concentration (a), ∂ 2c


flux (b), and local <0
∂ 2c ∂ x2 ∂ 2c
>0 >0
concentration change ∂ x2 ∂ x2
(c) vs. distance at a moment
(a) c(x, t)

− 2Dt 0 + 2Dt x

(b) J(x, t)

∂ c(x, t)
(c)
∂t

x
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 21

(v) Inflection points: The inflection points demarcating the boundary between the
convex-downward and convex-upward regions are defined as such that

 2 
∂ c
¼ 0: ð1:39Þ
∂x2 t

You may wish to calculate from the solution to find that those points are
pffiffiffiffiffiffiffiffi
x ¼  2Dt: ð1:40Þ

The width of the spread ofpdistribution


ffiffiffiffiffiffiffiffi at time t may be taken to p
beffi the
distance to an inflection point 2Dt, that is expanding in proportion to t. By
measuring the width of a Gaussian distribution, one can thus evaluate the
diffusion coefficient.
(vi) Mean displacement and mean square displacement: What is the mean
displacement of the solutes in time t? As the profile is symmetric with respect
to x ¼ 0, one can immediately guess that the average displacement will be zero.
It is shown mathematically in this way: Back to the Gaussian distribution, c
(x,t)dx/M corresponds to the probability of finding the diffusing species
between the position x and x+dx; thus the average value <x> for the displace-
ment is calculated as
Z þ1
cdx
< x >¼ x ¼0 ð1:41Þ
1 M

simply because the integrand is an odd function with respect to x ¼ 0. In


statistics, <x> is called the first moment of x.
Nevertheless, a concentration distribution has a spread, and what can be a
measure of this spread? In order to avoid being an odd function, let us instead
calculate the average of x2:
Z þ1
cdx
< x >¼
2
x2 ¼ 2Dt ð1:42Þ
1 M

due to the Gaussian distribution, Eq. (1.36). How to execute integration? This is
a well-known integral with an integrand of the form x2 eax , the answer of which
2

can be found from any table of Boltzmann integrals. You will also learn shortly
how to integrate it. One thing amusing is that the mean square displacement
corresponds to the square of the distance from the origin to an inflection point in
a Gaussian distribution (Fig. 1.8a)! The square root of the mean square displace-
ment, called the root-mean-square displacement, is nothing but the distance to
an inflection point
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
< x2 > ¼ 2Dt: ð1:43Þ
22 1 Diffusion in Continuum

What is the overall extent of a diffusion profile from x ¼ 0 at time t?


Probably two or three times the root meanpffiffiffiffiffisquare displacement? Well, it
appears so, and let us tentatively take 4 Dt as the diffusion distance or
extension. D E
In statistics, σ2x ¼ ðx < x >Þ2 is called the second moment with
respect to the average <x> or the variance and σx the standard deviation of
the measured value x. When you measure a physical quantity, say, the length of
your pencil, then your measurement results xi (data) should show a Gaussian or
normal distribution with respect to the algebraic mean <x> only if you measure
them with no bias at all or at random

1 <x>Þ2 =2σ2x
f ðxÞ ¼ pffiffiffiffiffiffiffiffiffiffi eðx ð1:44Þ
2πσ2x

where f(x) is the probability density of measuring a specific value x. This is


actually what is printed on the 10-German Mark note with <x> ¼ μ (see
Fig. 1.7). No surprise that all the basic error analysis starts from this distribution
function. You may now understand why your professor awards you an A-grade
if your score is, for example, above <x> +2σx.

1.5.3 Leak Test

Finally, we know that there exists nothing like a semi-infinite bar in this world:
anything is finite. If the length of our diffusion medium
pffiffiffiffiffi x ¼ L in Fig. 1.6 is shorter
than the overall extent of a diffusion profile, say, 4 Dt, the solution, Eq. (1.36), will
obviously start to break down. As long as the overall extent of diffusion remains
within the dimension L of our medium, we may practically regard our medium as
semi-infinite, and our thin-film solution remains valid. Is there then any quantitative
criterion to tell whether our medium dimension L may or may not be regarded as
semi-infinite? Yes, there is one, of course, which isp called
ffiffiffiffiffi the leak test:
Let us compare the so-called diffusion distance 4 Dt and thepphysical
ffiffiffiffiffi dimension
of the diffusion medium L (see Fig. 1.9). Intuitively, if L  4 Dt, the medium is
essentially semi-infinite; otherwise, the thin-film source solution (the solid curve in
Fig. 1.9) will break down (e.g., the dotted curve in Fig. 1.9)! In order to be more
quantitative, one may compare the leaked amount of the profile outside L (the
shaded areas in Fig. 1.9) with the total amount of the species M. If the leak is
negligibly small, then we may regard L as semi-infinite and continue to use the thin-
film source solution. The problem is how much tolerant we will have to be to the
leak. Let us say if the leak is 0.1% or less, then we will accept the solution or
R1
cdx
R ¼ RL1 0:001: ð1:45Þ
0 cdx
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 23

Fig. 1.9 Principle of the c(x, t)


leak test

leak

x
-L 0 L

This is called the leak test to judge whether the thin-film source solution is valid
for the given medium dimension.
In order to carry out the integrations, we have to learn the error function now.
We will here a bit digress from the main stream for a mathematical interlude.

1.5.4 Mathematical Interlude: Error Function

The main ingredient of the Gaussian distribution, Eq. (1.36) or Eq. (1.44), is

eη
2

in whatsoever form it takes. This is an even function and has a bell shape against the
variable η as shown again in Fig. 1.10a. Let us try to calculate the area I below the
curve over the entire range 1 < η < 1 as
Z þ1
eη dη:
2
I¼ ð1:46Þ
‐1

It looks beautiful, but how to integrate it? Do not worry! This is a very famous
integral haunting statistical mechanics in particular so that it has long been solved in
a truly amusing way.
I am sure you agree that
Z þ1  Z þ1  Z þ1 Z þ1
x2 y2
eðx þy2 Þ
2
I ¼
2
e dx  e dy ¼ dxdy: ð1:47Þ
1 1 1 1

That is, you have ended up with a surface integral over the entire surface. Then we
can transform the surface integral in terms of the polar coordinates with no
difficulty at all by using the relations x2 + y2 ¼ r2 and dxdy ¼ rdrdθ to get the
answer immediately:
Z 1 Z 2π
er rdrdθ ¼ π:
2
I ¼
2
ð1:48Þ
0 0

Isn’t it amusing enough? You can then immediately recognize that


24 1 Diffusion in Continuum

Fig. 1.10 Definition of erf 2

(z) (a) and erf(z) vs. z (b) e-η


erf(z)

erfc(z)

η
0 z
(a)

1.0

0.5

erf(z)
0.0

-0.5

-1.0
-2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5
z
(b)

Z 1 pffiffiffi
π
eη dη ¼
2
: ð1:49Þ
0 2

Could we then calculate the area below the curve only up to an arbitrary point
η ¼ z from 0? No, it is not possible to integrate algebraically as we did with the
definite integral, Eq. (1.49). The fraction of the area to this point is thus defined as the
error function abbreviated as erf(z),
Z z
2
eη dη,
2
erf ðzÞ  pffiffiffi ð1:50Þ
π 0

and its value depending on z can only be obtained numerically. The fractional area or
erf(z) looks like Fig. 1.10b, and its values are tabulated (see Table 1.1). Why is it
called the error function? Remember that the Gaussian distribution, Eq. (1.44),
indeed deals with the error distribution around the average value <x>. Naming the
error function makes sense.
Before going back to the main stream, let us define an auxiliary function to the
former, that is, a complimentary error function, erfc(z). It is obvious from
Fig. 1.10a that
Z z Z 1
2 2
eη dη þ pffiffiffi eη dη ¼ 1:
2 2
pffiffiffi ð1:51Þ
π 0 π z
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 25

Table 1.1 Error function table


Rz
erf ðzÞ ¼ p2ffiffiπ 0e
x2
dx
Hundredths digit of z
z 0 1 2 3 4 5 6 7 8 9
0.0 0.00000 0.01128 0.02256 0.03384 0.04511 0.05637 0.06762 0.07886 0.09008 0.10128
0.1 0.11246 0.12362 0.13476 0.14587 0.15695 0.16800 0.17901 0.18999 0.20094 0.21184
0.2 0.22270 0.23352 0.24430 0.25502 0.26570 0.27633 0.28690 0.29742 0.30788 0.31828
0.3 0.32863 0.33891 0.34913 0.35928 0.36936 0.37938 0.38933 0.39921 0.40901 0.41874
0.4 0.42839 0.43797 0.44747 0.45689 0.46623 0.47548 0.48466 0.49375 0.50275 0.51167
0.5 0.52050 0.52924 0.53790 0.54646 0.55494 0.56332 0.57162 0.57982 0.58792 0.59594
0.6 0.60386 0.61168 0.61941 0.62705 0.63459 0.64203 0.64938 0.65663 0.66378 0.67084
0.7 0.67780 0.68467 0.69143 0.69810 0.70468 0.71116 0.71754 0.72382 0.73001 0.73610
0.8 0.74210 0.74800 0.75381 0.75952 0.76514 0.77067 0.77610 0.78144 0.78669 0.79184
0.9 0.79691 0.80188 0.80677 0.81156 0.81627 0.82089 0.82542 0.82987 0.83423 0.83851
1.0 0.84270 0.84681 0.85084 0.85478 0.85865 0.86244 0.86614 0.86977 0.87333 0.87680
1.1 0.88021 0.88353 0.88679 0.88997 0.89308 0.89612 0.89910 0.90200 0.90484 0.90761
1.2 0.91031 0.91296 0.91553 0.91805 0.92051 0.92290 0.92524 0.92751 0.92973 0.93190
1.3 0.93401 0.93606 0.93807 0.94002 0.94191 0.94376 0.94556 0.94731 0.94902 0.95067
1.4 0.95229 0.95385 0.95538 0.95686 0.95830 0.95970 0.96105 0.96237 0.96365 0.96490
1.5 0.96611 0.96728 0.96841 0.96952 0.97059 0.97162 0.97263 0.97360 0.97455 0.97546
1.6 0.97635 0.97721 0.97804 0.97884 0.97962 0.98038 0.98110 0.98181 0.98249 0.98315
1.7 0.98379 0.98441 0.98500 0.98558 0.98613 0.98667 0.98719 0.98769 0.98817 0.98864
1.8 0.98909 0.98952 0.98994 0.99035 0.99074 0.99111 0.99147 0.99182 0.99216 0.99248
1.9 0.99279 0.99309 0.99338 0.99366 0.99392 0.99418 0.99443 0.99466 0.99489 0.99511
2.0 0.99532 0.99552 0.99572 0.99591 0.99609 0.99626 0.99642 0.99658 0.99673 0.99688
2.1 0.99702 0.99715 0.99728 0.99741 0.99753 0.99764 0.99775 0.99785 0.99795 0.99805
2.2 0.99814 0.99822 0.99831 0.99839 0.99846 0.99854 0.99861 0.99867 0.99874 0.99880
2.3 0.99886 0.99891 0.99897 0.99902 0.99906 0.99911 0.99915 0.99920 0.99924 0.99928
2.4 0.99931 0.99935 0.99938 0.99941 0.99944 0.99947 0.99950 0.99952 0.99955 0.99957
2.5 0.99959 0.99961 0.99963 0.99965 0.99967 0.99969 0.99971 0.99972 0.99974 0.99975
2.6 0.99976 0.99978 0.99979 0.99980 0.99981 0.99982 0.99983 0.99984 0.99985 0.99986
2.7 0.99987 0.99987 0.99988 0.99989 0.99989 0.99990 0.99991 0.99991 0.99992 0.99992
2.8 0.99992 0.99993 0.99993 0.99994 0.99994 0.99994 0.99995 0.99995 0.99995 0.99996
2.9 0.99996 0.99996 0.99996 0.99997 0.99997 0.99997 0.99997 0.99997 0.99997 0.99998
3.0 0.99998 0.99998 0.99998 0.99998 0.99998 0.99998 0.99998 0.99999 0.99999 0.99999
3.1 0.99999 0.99999 0.99999 0.99999 0.99999 0.99999 0.99999 0.99999 0.99999 0.99999
3.2 0.99999 0.99999 0.99999 1.00000 1.00000 1.00000 1.00000 1.00000 1.00000 1.00000

When something + something else makes 1, we say they are complementary to


each other. Thus
Z 1
2
eη dη:
2
erfcðzÞ ¼ 1  erf ðzÞ ¼ pffiffiffi ð1:52Þ
π z
26 1 Diffusion in Continuum

1.5.5 Back to the Leak Test

We have now come to be able to play with the integrals in Eq. (1.45) at our disposal.
The fraction of the leak R may now be rewritten as
R1  
ex =4Dt dx
2
L
R ¼ RL1 ¼ erfc p ffiffiffiffiffi : ð1:53Þ
0 ex2 =4Dt dx 2 Dt

Let us say that if R 0.001, for example, we will regard our diffusion media as
semi-infinite so that the thin-film solution is accurate within 0.1% error. It is matter
of our tolerance or generosity to error. For the case of the error bound R 0.001, for
pffiffiffiffiffi
example, L=2 Dt should be such that
 
L
erfc pffiffiffiffiffi 0:001: ð1:54Þ
2 Dt

You can get the numerical value for the argument of erfc from the error function
table, Table 1.1, as

L pffiffiffiffiffi
pffiffiffiffiffi
2:3 or L
4:6 Dt: ð1:55Þ
2 Dt

You remember thepestimate


ffiffiffiffiffi for the overall diffusion distance in a given time t that
was approximately 4 Dt. Now you recognize that how long your diffusion system
must be for the thin-film solution to remain valid p within
ffiffiffiffiffi 0.1% error. If youptolerated
ffiffiffiffiffi
up to 1% error (¼R) instead, you would have L=2 Dt
1:82 or L
3:64 Dt from
the error function table (so a shorter material would be permissible).
The condition Eq. (1.55) may be recast in terms of diffusion time t for given
dimension L as

L2 L2
t 2
 0:047 for 0:1 % error: ð1:56Þ
4:6 D D

For 1% error, t 0.075L2/D (and a longer time would be permissible). This


means that for a given length L, one should keep the diffusion time below a certain
upper limit for the thin-film solution to be valid. In this sense, the thin-film solution is
called the short-time solution. The shorter the diffusion time, the more accurate.
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 27

1.5.6 Variations of the Thin-Film Source Solution

Once you understand the nature of the thin-film solution a bit in detail, you can now
recognize that the same thin-film solution will make the solution for the cases of
(i) the instantaneous line source in an infinite plane (like a line of ink in an infinite
strip of paper) and (ii) the instantaneous point source upon an infinite line (like a
dot of ink on an infinite string of cotton). The common aspect is that all are
one-dimensional diffusion: from a planar source into a volume, from a line source
into a plane, and from a point source along a line. The difference is that M in
Eq. (1.36) should take a different dimension in accord with the geometric dimension
of the source: [M] ¼ mol/m2, mol/m, and mol for the planar, line, and point source,
respectively. The situation is summarized in Fig. 1.11.

source example
(a) Planar source

Semi- Semi- M / unit area An ink plane inside a stick of radish


infinite infinite

(b)
Line source
M / unit length An ink line on a strip of paper
in a plane

(c)
M / point An ink point on a piece of cotton string
Point source on a line

Fig. 1.11 A planar source (M/unit area) within an infinite bar (a), a line source (M/unit line) upon
an infinite strip (b), and a point source (M/point) on an infinite line (c)

What does the solution look like depending on the dimension of diffusion from an
appropriate instantaneous source? The thin-film solution is for the one-dimensional
diffusion from the two-dimensional plane source. One may systematically consider
the two-dimensional diffusion from an instantaneous line source (diffusion in
cylindrical symmetry) and the three-dimensional diffusion from an instantaneous
point source (diffusion in spherical symmetry). The solutions describing the con-
centration distribution in these cases are summarized in Table 1.2. You are asked to
examine carefully how the pre-exponential factor and the unit of M change with the
change of dimension.
28 1 Diffusion in Continuum

Table 1.2 Solutions for an instantaneous plane, line, and point source

1-dimensional diffusion from a planar source


∂c ∂ 2c
=D 2;
∂t ∂x
M
e − x /4Dt
2
c=
4πDt

2-dimensional diffusion from a line source


in cylindrical symmetry
∂c 1 ∂ ⎛ ∂c ⎞
=D ⎜ r ⎟;
∂t r ∂r ⎝ ∂r ⎠
M − r 2 /4Dt
c= e
4πDt

3-dimensional diffusion from a point source


in spherical symmetry

∂c 1 ∂ ⎛ ∂c ⎞
= D 2 ⎜ r2 ⎟;
∂t r ∂r ⎝ ∂r ⎠
M
e − r /4Dt
2
c=
(4πDt)3/2

1.5.7 Reflection and Superposition Method

Lastly we will learn an important method of getting the solution to Fick’s second law
that is called a reflection and superposition method. The thin-film solution,
Eq. (1.36), was for the diffusion in positive (+) as well as in negative () direction
from an instantaneous planar source. What would the solution look like if the planar
source is applied onto one semi-finite bar only, i.e., x
0 in Fig. 1.12. As there is no
medium to the left, all the source should remain in the right-hand side bar. You may
then immediately guess that the solution will be the reflection and superposition of
the left-hand side profile onto the right-hand side profile in Fig. 1.8a or two times the
solution in Eq. (1.36),

M
cðx, tÞ ¼ pffiffiffiffiffiffiffiffi ex =4Dt :
2
ð1:57Þ
πDt

Indeed this solution can be obtained formally by solving Fick’s second law with
the use of the initial and boundary conditions:

C ¼ 0 for x > 0 at t ¼ 0: ð1:58aÞ


C ¼ 1 at x ¼ 0 at t ¼ 0: ð1:58bÞ
1.5 Diffusion from an Instantaneous Planar Source or Thin-Film Source Solution 29

Fig. 1.12 Diffusion from a M − x 2 /4Dt


planar source into a single c(x, t) = e
πDt
semi-infinite medium

0 x→

C ¼ 0 at x ! 1 at any t: ð1:58cÞ

∂c
J ¼ 0 or ¼ 0 at x ¼ 0 for all t: ð1:58dÞ
∂x x¼0þ

The last boundary condition means that the boundary x ¼ 0 is impermeable, and it
is because there is no medium to diffuse outward. We have already observed this fact
in the thin-film solution where the plane x ¼ 0 behaves as if it were impermeable or
J ¼ 0 (see Fig. 1.8b).

1.5.8 Application of the Thin-Film Source Solution

The thin-film solution has often been employed to determine the tracer diffusion
coefficient of the host atoms or ions in a system. Here, we mean by the tracers the
same atoms as the host, but distinguishable from them, thus, traceable. We often use
radioactive isotopes or even stable isotopes. We normally put a thin-film of the
tracers between the two semi-infinite bars of specimen, Fig. 1.6a, or at the surface of
a semi-infinite bar, Fig.1.12, and let them diffuse. In this case, of course, one should
observe the short-time condition, Eq. (1.56) so that the thin-film-source solution
remains valid. In time t, one stops diffusion by quenching or so and then cut thin
slices normal to the diffusion direction (this is called sectioning, and it is a really
laborious job! Modern spectroscopic techniques, e.g., secondary-ion mass spectros-
copy (SIMS) (See, e.g., Yoo et al. [11]), of course, save much time and effort, but the
principle of sectioning remains the same!) and measure the concentration of the
tracers in each slice. If we let the thickness of the ith slice be Δxi, what we eventually
obtain is a histogram of the tracer amount of each slab along the diffusion direction.
By assigning the average concentration at the distance from the origin to the center
of each slice xi, one can get a data set of c(xi). When we had to do all analyses
afterward by using a slide rule or later a small hand-calculator, we normally plotted
ln c, if the tracer concentration c can be directly measured, or ln (radioactivity), as
the radioactivity is proportional to the concentration, vs. x2i , in accord with the thin-
film solution Eq. (1.36) or Eq. (1.57) and determine the tracer diffusivity from the
slope that is 1/4Dt.
This is all for the thin-film solution. We will next turn to a bit more complicated
situation, diffusion from a semi-infinite source into a semi-infinite medium.
30 1 Diffusion in Continuum

1.6 Semi-infinite Source Solution

1.6.1 Solution of Fick’s Second Law

Let us suppose that a semi-infinite source, with a concentration of c0, is put onto a
semi-infinite medium with no solute or c ¼ 0 (Fig. 1.13). You can intuitively
imagine that from the step-like distribution at t ¼ 0, the diffusion profile evolves
with time (t ¼ 0 < t1 < t2 < t3) as indicated in Fig. 1.13. The exact solution c(x,t) can
only be obtained by solving Fick’s second law

2
∂c ∂ c
¼D 2 ð1:59aÞ
∂t ∂x

with the use of the initial and boundary conditions:

cðx > 0, 0Þ ¼ c0 ,
cðx < 0, 0Þ ¼ 0,
ð1:59bÞ
cðþ1, tÞ ¼ c0 ,
cð1, tÞ ¼ 0:

Fig. 1.13 Diffusion from a C


semi-infinite source (x > 0) t=0
into a semi-infinite medium
(x < 0). The dotted lines are C′
t1 t2
expected diffusion profiles t3
c(x,t) at t ¼ 0 < t1 < t2 < t3,
respectively

0 x

Again you may


pffi be embarrassed by too many conditions, but be aware that in
terms of η ¼ x= t, they reduce to only two independent conditions:

c ¼ c0 as η ! þ1; c ¼ 0 as η ! 1: ð1:60Þ

Equipped with the knowledge of the thin-film solution, you may now wish to
solve this equation rigorously. Let us do it by using the superposition method. You
agree that the semi-infinite source may be regarded as a sum of infinite number of
thin-films. Let us slice the semi-infinite source into an infinite number of thin-films
i ¼ 1,2,3. . . . each with the same thickness Δξi ¼ Δξ see Fig. 1.14a. Each thin-film
source, with an amount per unit area Mi ¼ c0Δξ, will produce its own Gaussian
distribution with the center at ξi from x ¼ 0,
1.6 Semi-infinite Source Solution 31

c0 Δξ 2
ci ðx, tÞ ¼ pffiffiffiffiffiffiffiffi eðxξi Þ =4Dt , ð1:61Þ
2 πDt

mathematically irrespective of the other thin-films (Fig. 1.14a). When we add up all
the profiles produced from each thin-film source, you will end up with the overall
profile (Fig. 1.14b),

X
n Z 1
c0 2
cðx, tÞ ¼ lim ci ðx, tÞ ¼ pffiffiffiffiffiffiffiffi eðxξÞ =4Dt dξ: ð1:62Þ
n!1 0 2 πDt
ðΔξ ! 0Þ i¼1

c′
ith slab

ci (x, t)

0 ξ i ξ i + Δξ

(a)

Δξ
c′

c′Δξ − ( x −ξ2 )2 / 4Dt


c(x, t) e
2 πDt

0 ξ2

(b)

Fig. 1.14 (a) Gaussian distribution from a thin-slab source and (b) superposition of those from all
the slabs

How do we integrate it? We have already learned when studying erf(z). Change
the integration variable as

xξ dξ
η ¼ pffiffiffiffiffi ; dη ¼ pffiffiffiffiffi , ð1:63Þ
2 Dt 2 Dt
32 1 Diffusion in Continuum

and you will then readily integrate to obtain



 
c0 x
c¼ 1 þ erf pffiffiffiffiffi : ð1:64Þ
2 2 Dt

What a neat way it is!


To confirm this to be the solution to the present problem, you are recommended to
check yourself to see if the initial and boundary conditions Eq. (1.59) are all satisfied.
Of course, they should be satisfied because the solution has been obtained in such
a way.

1.6.2 Salient Features of the Solution

It is worthwhile to note some salient features of this solution:


(i) First thing is that the profile c(x,t) is exactly antisymmetric with respect to x ¼ 0
where the concentration is always fixed at c ¼ c0 /2 (Fig. 1.15a). It is just because
D is constant and the mass is conserved.
(ii) Try to calculate the flux J(x,t) by differentiating c(x,t) with respect to x due to
Fick’s first law. You must be able to get

pffiffiffiffi
c0 D
 pffiffiffiffi ex =4Dt
2
J¼ ð1:65Þ
2 πt

which is plotted in Fig. 1.15b. It is noted that the functional shape is Gaussian. It
is quite natural because c(x,t) is an integration of the Gaussian distribution.
Also note that the flux becomes maximum at x ¼ 0. You know why.
(iii) The temporal variation of the local concentration is also obtained from the
curvature of c(x,t) in Fig. 1.15a due to Fick’s second law, which looks like
what is shown in Fig. 1.15c. Note that ∂c/∂t ¼ 0 at x ¼ 0 because this position
corresponds to the inflection point of c(x,t) or ∂2c/∂x2 ¼ 0.

(iv) As we did with the thin-film solution, we ask how long should the bar be for the
solution to be valid? You can now tell without hesitation that again

pffiffiffiffiffi
L
4:6 Dt for 0:1 % error ð1:66Þ

as is obvious from the leaks in Fig. 1.16. What if you tolerate up to 1% error?
You are asked to find the inequality. This lower limit of the dimension L for
given diffusion time t can be transformed to the upper limit to t for given L, that
is, Eq. (1.56). In this sense, the present solution is also a short-time solution.
Now you will understand why the short-time solutions are of error-function
including the Gaussian distribution.
1.6 Semi-infinite Source Solution 33

Fig. 1.15 Concentration c


profile from a semi-infinite
source (a), flux (b), and c′
local concentration
change (c) c′
c′ / 2 • c(0, t) =
2

0 x
(a)

0
x

c′ D
J(x=0)= −
2 πt

(b)
∂c
∂t

0 x

(c)

Fig. 1.16 Comparison of c


the diffusion distance and
the length of the diffusion c′
medium L c(x, t)

c′/2

-L 0 L x

1.6.3 Variations of the Error-Functional Solution

In order to get familiar with the error function solutions, here you will practice how
to find the solution for given initial and boundary conditions only by intuition or
symmetry argument. The first example is diffusion out of a semi-infinite source (c
(x > 0, t ¼ 0) ¼ cs) with its surface concentration kept constant at c(0,t) ¼ 0 (see Case
(i) in Table 1.3). It is reminded that for the semi-infinite source solution, Eq. (1.64),
the concentration at x ¼ 0 is always fixed at c ¼ c0 /2. In the present example, the
Table 1.3 Error functional solutions to various initial and boundary conditions
(i) c Surface concentration
  fixed at c(0,t) ¼ 0 (1.67)
t=0 c ¼ cs erf pxffiffiffiffi
cs 2 Dt

t↑
c(x,t)
0
0 x
(ii) c Surface concentration
  fixed at c(0,t) ¼ 0 (1.68)
t=0 c ¼ cs erf pxffiffiffiffi
cs 2 Dt

t↑
c(x,t)
0
x 0
(iii) c Surfaceh concentration
 ifixed at c(0,t) ¼ cs (1.69)
xffiffiffiffi
c ¼ cs 1 þ erf 2 Dt
p
cs
c(x,t)

t↑

0
x t=0 0

(iv) c Surface concentration


  fixed at c(0,t) ¼ cs (1.70)
c ¼ cs erfc pxffiffiffiffi
cs 2 Dt
c(x,t)

t↑

0
0 t=0 x
h  i
(v) c cco
c0 co ¼ 12 1 þ erf 2pxffiffiffi
Dt
ffi (1.71)

c(x,t)

t↑
c0
t=0
0 x
 
(vi) c cco
¼ 12 erfc pxffiffiffiffi (1.72)
c0 co 2 Dt

c(x,t)

t↑
c0
t=0
0 x
1.6 Semi-infinite Source Solution 35

surface concentration is fixed at 0 all the way. One can easily get the solution for the
present case simply by translating Eq. (1.64) in the y direction by c0 /2 and then by
setting c0 /2 ¼ cs as
 
x
c ¼ cs erf pffiffiffiffiffi : ð1:67Þ
2 Dt

For another example, let us try to construct the solution for Case (ii) in Table 1.3,
where the solute diffuses out of the semi-infinite source (x < 0) at a concentration
c ¼ cs. The situation is simply nothing but a mirror reflection of Case (i) with
respect to the y-axis. The solution will then have to be the reflection of Eq. (1.67)
with respect to the y-axis. By replacing x with –x in Eq. (1.67), thus, you can obtain
the solution as
 
x
c ¼ cs erf pffiffiffiffiffi : ð1:68Þ
2 Dt

The solutions to various initial and boundary conditions including the above-
treated two cases, which can be all guessed from the main solution, Eq. (1.64), are
summarized in Table 1.3. You are strongly urged to practice to get the solution for
each case only on the basis of translational and/or reflection symmetry.

1.6.4 Generalization of Error-Functional Solutions

So, did you all succeed in getting the solutions for all the cases in Table 1.3? Of
course, you can follow the recipe described above, but here is a clever trick. First,
have a look at all those solutions in Table 1.3. You may immediately recognize that
all those solutions are taking the general form,

x
c ¼ A þ Berf p ffiffiffiffiffi , ð1:73Þ
2 Dt

involving two parameters A and B at most. What you have to do is then to evaluate
these two parameters by using the given initial and boundary conditions. p But,
ffi wait to
see the trial solution, Eq. (1.73), is a function of a single variable η  x= t. As long
as the initial and boundary conditions, no matter how many there are, can be
represented in terms of this single variable η, then you can determine A and B
uniquely. Do not worry that those two may be overdetermined. In this sort of short-
time problem, there are supposed to be neither more nor less conditions: there are
exactly two independent conditions always.
Let us apply this trick to find the solution for the case of Fig. 1.13. The initial and
boundary conditions, Eq. (1.59b), are reduced to only two independent conditions,
Eq. (1.60):
36 1 Diffusion in Continuum

cð η ! 1 Þ ¼ c0 and cðη ! 1Þ ¼ 0:

Apply these conditions to the two parameters trial solution, Eq. (1.73), and you
will get the two equations with respect to the two unknowns A and B. These two are
solved simultaneously to evaluate A and B, which in turn gets you Eq. (1.64) as the
exact solution. You may wish to try this trick to other initial and boundary conditions
in Table 1.3.

1.6.5 Mathematical Interlude: Properties of erf(z)

By now, you will have come to feel pretty much at home with the error function. It
is now time to summarize some useful properties of erf(z), which are simply listed
below if obvious, but otherwise, given a short comment.
(i) erf(z) ¼ erf(z) (by definition). !
2 X ð1Þ X ðxÞn
n
x
(ii) erf ðzÞ ¼ pffiffiffi z 2nþ1
as e ¼ :
π n¼0 n!ð2n þ 1Þ n!
  n¼0
2 z3 z5
(iii) erf ðzÞ  pffiffiffi z  þ þ    (for small z).
π 13 215
2 2 dz
(iv) dx d
erf ðzÞ ¼ pffiffiffi ez :
Z 1π dx
Z 1
(v) in erfcðzÞ  in1 erfcðηÞ dη; ierfcðzÞ  erfcðηÞ dη:
z z
1
(vi) ierfcð0Þ ¼ pffiffiffi (by integrating by parts).
π
2  
ez 1 1 13
(vii) erfcðzÞ ¼ p ffiffiffi  þ     (for large z).
π z 2z3 22 z5
With the very last one (vii), you may have to suffer a little headache,4 but these
are normally sufficient. For further detailed properties, however, you are referred to
the appendices in Crank or Carslaw and Jeager.

1.7 Trigonometric Infinite Series Solution

1.7.1 Statement of the Problem

We have learned that the error-functional solutions are the short-time solution by
nature, that is, the diffusion zone is limited to the near vicinity of the source, and

Z 1 Z
R1 ez 1 1 x2 1
2
1
ex dx ¼ xex
2 2
4
Noting that z dx ¼  xe dx by integrating by parts, one
z x 2z 2 z x3
can easily eliminate the last headache by continuing this procedure.
1.7 Trigonometric Infinite Series Solution 37

hence, a complete homogenization is never achieved because the diffusion medium


is practically infinite. Our question now is what would happen to the concentration
distribution with time if the condition for the short-time solution to be valid,
Eq. (1.56), breaks down or, in other words, the diffusion distance eventually exceeds
the boundary of the system?
Let us consider an infinite slab, of finite thickness L, with a uniform distribution
of solutes at the concentration of c ¼ co inside (0 x L). Now suppose that the
solute concentration at both surfaces (x ¼ 0, L) of this slab is suddenly lowered to
c ¼ 0 (see Fig. 1.17). Then the solutes will start to diffuse out of the slab exhibiting
pffiffiffiffiffi
concentration gradients. When elapsed time t is still such that L=2 > 4:6 Dt
(remember no greater than 0.1% error!), the concentration distribution may be
described by the short-time solutions as
 
x
cðx, tÞ ¼ co erf pffiffiffiffiffi for 0 x < L=2; ð1:74aÞ
2 Dt
 
Lx
cðx, tÞ ¼ co erf pffiffiffiffiffi for L=2 x < L: ð1:74bÞ
2 Dt
pffiffiffiffiffi
As time elapses so that L=2 < 4:6 Dt, L/2 is rendered to be no longer infinite,
and the short-time solutions will begin to fail. At a sufficiently long time, we might
expect the concentration distribution to look like half a wavelength sine function
with an amplitude that decays exponentially with time as is the case with any natural
decay,
 
πx t
cðx, tÞ / sin  exp  , ð1:75Þ
L τ

where τ is a constant characteristic time.


This is just a reasonable guess, but how could we get the solution which can
actually describe the concentration distribution c(x,t) from the beginning (t ¼ 0) to
the end, steady state (t ! 1)? In order to get the legitimate solution, we have to
solve Fick’s second law anyway. Let us do it now.

1.7.2 Solution of Fick’s Second Law by


Separation-of-Variables Method

Let us again consider the same, infinite slab with the surface concentration fixed at
0 but with an arbitrary, non-trivial initial distribution of the solute ϕ(x) (see
Fig. 1.18). Then, the solution is the primitive function c(x,t) such that

2
∂c ∂ c
¼D 2 ð1:76Þ
∂t ∂x
38 1 Diffusion in Continuum

Fig. 1.17 Concentration c


profiles c(x,t) at different
c0 t=0
times at t ¼ 0, t1(>0), t1
t2(>t1),. . . upon diffusion
out of an infinite slab with a t2
finite thickness L t3

0 x
0 L

Fig. 1.18 The slab with an c


arbitrary initial distribution
c0
of solute ϕ(x)
φ(x)

c(x, t)

0 x
0 L

with

IC: cð0 < x < L, 0Þ ¼ ϕðxÞ, ð1:76aÞ


BC: cð0, tÞ ¼ 0, ð1:76bÞ
BC: cðL, tÞ ¼ 0: ð1:76cÞ

In order to solve this differential equation, let us assume that the primitive
function c(x,t) can be expressed as a product of X(x) a function of x only and T
(t) another of t only:

cðx, tÞ ¼ XðxÞ  TðtÞ: ð1:77Þ

By substituting this product for c in Eq. (1.76) and dividing through by this
product, you will have an equation

1 dT 1 d2 X
 ¼  : ð1:78Þ
DT dt X dx2

Look! The left-hand-side is a function of t only and the right-hand-side a function


of x only! For them to equal each other, the only possibility is that they have to be a
constant, which we set –k2 to make sure it to be a negative constant because the
solution would otherwise diverge with t. Then, we have two differential equations,
one for T and the other for X:
1.7 Trigonometric Infinite Series Solution 39

1 d ln T
 ¼ k2 ð1:79Þ
D dt
d2 X
þ k2 X ¼ 0: ð1:80Þ
dx2

The first one is really easy to solve: the solution is

T ¼ To ek Dt
2
ð1:81Þ

where To is a constant.
The second one is a typical differential equation describing the wave motion.
Waves may be represented by sine or cosine functions. Thus, sin(kx), cos(kx), and
even their linear combination can be the solution, and furthermore, those combina-
tions for all k-values, too, noting the differential equation, Eq. (1.80), is linear. The
general solution may then be written as
X
X¼ A0n sin kn x þ B0n cos kn x ð1:82Þ
n

with A0n and B0n as the amplitudes of each component waves.


By combining Eqs. (1.81) and (1.82) according to Eq. (1.77), you have the
generic solution c(x,t) such that
X
ðAn sin kn x þ Bn cos kn xÞ ekn Dt
2
cðx, tÞ ¼ ð1:83Þ
n

where An ¼ To A0n and so on. To complete the solution, we have to determine the
three constants An, Bn, and kn for each n(¼1,2,3,. . .). There are actually an infinite
number of these, and how could we determine them? You remember we have still
the three conditions left, that is, one initial condition and the two boundary condi-
tions. Now, you will see how they work out.
The boundary condition c(0,t) ¼ 0, Eq. (1.76b), immediately dictates

Bn ¼ 0: ð1:84Þ

The other boundary condition c(L,t) ¼ 0, Eq. (1.76c), stipulates


X
ðAn sin kn LÞ ekn Dt ¼ 0:
2
cðL, tÞ ¼ ð1:85Þ
n

If An ¼ 0, then the condition will be satisfied, but this is only trivial. The
nontrivial solution is with kn such that sinknL ¼ 0 or


kn ¼ ðn ¼ 1, 2, 3, . . .Þ: ð1:86Þ
L
40 1 Diffusion in Continuum

Thus, the generic solution, Eq. (1.85), now takes the form

X  2 2 
nπx n π Dt
cðx, tÞ ¼ An sin exp  : ð1:87Þ
n L L2

In order to complete the solution, we still have to evaluate An. How do we do


that? Remember we still reserve the initial condition c(x,0) ¼ ϕ(x), Eq. (1.76a).
Thus,
X nπx
cðx, 0Þ ¼ An sin ¼ ϕðxÞ, ð1:88Þ
n L

but so what? How do we use this initial condition to solve for {An}? Again the
orthogonality theorem, a beauty of mathematics, saves us. We will pause here to
learn the theorem.

1.7.3 Mathematical Interlude: Orthogonality Theorem

It is first noted that the differential equation, Eq. (1.80) has a solution of sin knx for
each of the kn values in Eq. (1.86) that is due to the boundary condition,
Eq. (1.76c). This kind of problem is called a boundary-value problem, and kn
and the corresponding solution are called the characteristic (or eigen-)value and
characteristic (or eigen-)function, respectively, for the given boundary-value
problem.
Let us suppose that the differential equation of the same type of Eq. (1.80)

y00 þ k2 y ¼ 0 ð1:89Þ

has a set of characteristic values and functions {kn, yn}. Then any function, say, y1
and y2 corresponding to the eigenvalue k1 and k2, respectively, satisfy the differen-
tial equation as

y001 þ k21 y1 ¼ 0; y002 þ k22 y2 ¼ 0: ð1:90Þ

By multiplying the first through by y2 and the second by y1, subtracting each
other, and integrating from x ¼ 0 to L, you will obtain
Z L Z
L
0¼ y001 y2  y002 y1 dx þ k21  k22 y1 y2 dx: ð1:91Þ
0 0

You can integrate by parts the first term on the right-hand side to obtain
1.7 Trigonometric Infinite Series Solution 41

Z L Z
 L L
y001 y2  y002 y1 dx ¼ y01 y2  y02 y1 0  y01 y02  y02 y01 dx ¼ 0 ð1:92Þ
0 0

due to the boundary conditions which are normally yn or y0n ¼ 0 at x ¼ 0 and L for a
boundary-value problem. Finally, you are left with the identity
Z
2 L
k1  k22 y1 y2 dx ¼ 0: ð1:93Þ
0

This implies that if k1 ¼ k2, the equation is automatically satisfied, but


Z L
y1 y2 dx ¼ 0 if k1 6¼ k2 : ð1:94Þ
0

This relation is reminiscent of the inner product of two orthogonal vectors, thus
called the orthogonality theorem.

1.7.4 Evaluation of the Last Constant by the Orthogonality


Theorem

Let us now multiply Eq. (1.88) by an arbitrary characteristic function, sin(mπx/L),


and integrate from x ¼ 0 to L to obtain

X Z L
nπ x mπ x
Z L
mπ x
An sin sin dx ¼ ϕðxÞ sin dx: ð1:95Þ
n 0 L L 0 L

Could you recognize that if m 6¼ n, the integral on the left-hand side makes 0 due
to the orthogonality theorem, Eq. (1.94)? The integral is nontrivial only if m ¼ n as
Z L Z L
nπ x nπ x
An sin 2 dx ¼ ϕðxÞ sin dx: ð1:96Þ
0 L 0 L

So, we have come to evaluate An in Eq. (1.88) as


Z L
2 nπ x
An ¼ ϕðxÞ sin dx: ð1:97Þ
L 0 L

Isn’t it neat?
Actually, Eq. (1.88) is nothing but the Fourier-sine-series representation of a
function ϕ(x), and the procedure up to now was a mathematical trick to determine the
amplitude of each component sine function.
42 1 Diffusion in Continuum

Let us now apply this method to get the solution to the problem in Fig. 1.17 with a
specific initial condition (see Fig. 1.18):

ϕð x Þ ¼ c o : ð1:98Þ

In this case, Eq. (1.97) can be easily integrated to yield


Z L
2c nπ x 2c
An ¼ o sin dx ¼ o ½1  cos nπ ð1:99Þ
L 0 L nπ

which has a nontrivial value only when n is odd, or otherwise 0:

4co
An ¼ ðj ¼ 0, 1, 2, . . .Þ: ð1:100Þ
ð2j þ 1Þπ

By substituting this equation for An in Eq. (1.87), we finally get the solution for
the initial and boundary conditions as shown in Fig. 1.17:
 
4co X 1 ð2j þ 1Þπ x ð2j þ 1Þ2 π2 D t
cðx, tÞ ¼ sin exp  : ð1:101Þ
π j¼0 ð2j þ 1Þ L L2

Before examining the nature or characteristics of this solution, we will practice


this method with one more heuristic example problem.

1.7.5 Application to the Trivial Initial-Condition Case

Suppose that the initial and boundary conditions are as illustrated in Fig. 1.19 or

cð0 < x < L, 0Þ ¼ 0


cð0, tÞ ¼ c1 ð1:102Þ
cðL, tÞ ¼ c2 :

You may wish to follow the procedure we have taken earlier. Then, you would be
in trouble with the step to evaluate An by Eq. (1.97) simply because the initial
concentration in the medium is zero, i.e., ϕ(x) ¼ 0. How can we get out of this
dilemma? Here is a trick.
Any concentration distribution c(x,t) may be regarded as the sum of the time-
independent, steady-state solution V(x) and the time-dependent, transient solution
U(x,t), or

cðx, tÞ ¼ VðxÞ þ Uðx, tÞ: ð1:103Þ


1.7 Trigonometric Infinite Series Solution 43

Fig. 1.19 The case of trivial c1


initial condition c(x,0) ¼ 0
c(x, ∞)

c2

0 x→ L

As the diffusivity is constant, the steady-state solution is obvious, which is

c2  c1
VðxÞ ¼ c1 þ x ð1:104Þ
L

for the present case. Then the original initial and boundary conditions are recast in
terms of U(x,t), due to Eqs. (1.103) and (1.104):

cð0, tÞ ¼ c1 ¼ c1 þ Uð0, tÞ, thus Uð0, tÞ ¼ 0: ð1:105aÞ


cðL, tÞ ¼ c2 ¼ c2 þ UðL, tÞ, thus UðL, tÞ ¼ 0: ð1:105bÞ
cð0 < x < L, 0Þ ¼ VðxÞ þ Uðx, 0Þ ¼ 0, thus Uðx, 0Þ ¼ VðxÞ: ð1:105cÞ

You will have by now recognized the essence of the trick to avoid the trivial
initial condition. By transforming we have now nontrivial initial condition in terms
of U, Eq. (1.105c).
The diffusion equation, Fick’s second law, is also accordingly transformed, by
substituting Eq. (1.103) to

2
∂U ∂ U
¼D 2: ð1:106Þ
∂t ∂x

We solve this equation following the standard procedure to obtain


X nπ x n2 π2 Dt=L2
U¼ An sin e , ð1:107Þ
L

and the amplitudes An are evaluated due to Eq. (1.97) as


Z L
2 nπ x
An ¼ ½VðxÞ sin dx: ð1:108Þ
L 0 L

By using V(x) in Eq. (1.104), you can now calculate An. The calculation is left for
your fun.
44 1 Diffusion in Continuum

Instead, we will get the solution for a special case that c1 ¼ c2 ¼ co, and hence, V
(x) ¼ co. Equation (1.108) gets you the value for An, which is the same as
Eq. (1.100). You will finally obtain the solution
 
4co X 1 ð2j þ 1Þπx ð2j þ 1Þ2 π2 Dt
c ¼ V þ U ¼ co  sin exp 
π j¼0 ð2j þ 1Þ L L2
ð1:109Þ

for the initial and boundary conditions,

cð0 x L, 0Þ ¼ 0; cð0, tÞ ¼ cðL, tÞ ¼ co , ð1:110Þ

(see Fig. 1.19). Actually, you can get this solution directly from the solution in
Eq. (1.101) simply by comparing the initial distributions, Figs. 1.17 and 1.19. Here
are the operations to take: First reflect the solution Eq. (1.101) with respect to the
x-axis, and then translate along the y-axis by co.

1.7.6 Salient Features of the Series Solution

Let us here examine some salient features of the trigonometric infinite series
solution, Eq. (1.101).
(i) First note that the solution is symmetric with respect to x ¼ L/2. This means
that the plane x ¼ L/2 behaves as if it were impermeable as you have already
seen from the thin-film source condition (see Fig. 1.12 and Eq. 1.57). You
would therefore have obtained the same solution, Eq. (1.101), even when you
were instead given as the initial and boundary conditions:

IC: cð0 < x L=2, 0Þ ¼ co : ð1:111aÞ


BC: cð0, tÞ ¼ 0: ð1:111bÞ
BC: Jðx ¼ L=2, tÞ ¼ 0 or ð∂c=∂xÞ ¼ 0 at x ¼ L=2 at all t: ð1:111cÞ

You can imagine what shape of concentration profile develops when the source
diffuses out through only one boundary, say x ¼ 0, and the other surface x ¼ L
remains impermeable.
(ii) The concentration profile at an instant is nothing but a superposition of sine
functions with an odd number of half wavelengths fitting the diffusion
dimension L, a characteristic of the boundary value problem. You see, the
component waves meet the standing wave condition, like violin strings. Also
note that as the wavelength gets shorter or the number of half wavelengths j
1.7 Trigonometric Infinite Series Solution 45

Fig. 1.20 First two (j ¼ 0, c/c0


1) sine wave components of
1
the series solution,
Eq. (1.101) j=0

j=1
0 L

increases, its amplitude or contribution is smaller (see Eq. 1.101 and Fig. 1.20).
Do not forget that the solution is nothing but a Fourier sine series represen-
tation of an instantaneous concentration distribution.

(iii) Each wave component decays exponentially with time, and as j or the number
of crests increases, the faster it decays with time. How much faster? Let us
compare the maximum of the (j+1)th term with that of the jth term, which we
may call the ratio test:



j ðj þ 1Þth term j 2j þ 1 8ðj þ 1Þπ2 Dt
R¼ ¼ exp  : ð1:112Þ
j jth term j 2j þ 3 L2

One can see that the ratio, R gets smaller than 1 exponentially with time as well
as with j+1. It indicates that in some time later, the infinite series solution,
Eq. (1.101), may be represented only by the first few terms with sufficient
accuracy. This will lead to the first-term approximation or the so-called
long-time solution.

1.7.7 First-Term Approximation or Long-Time Solution

If the second term (j¼1) is negligibly small compared to the first term (j¼0) in
Eq. (1.101), then the infinite series may be safely approximated by the first term
again as a matter of accuracy:
 2 
4co πx π Dt
cðx, tÞ ¼ sin exp  2 : ð1:113Þ
π L L

Let us suppose that we allow an error no greater than say, 0.1%. Then


j 2nd term j 1 8π2 Dt
R ¼ st ¼ exp  2 0:001: ð1:114Þ
j 1 term j 3 L
46 1 Diffusion in Continuum

Thus,

pffiffiffiffiffi L2
L 3:7 Dt or t
0:074 : ð1:115Þ
D

ffiffiffiffiffi generous to the error, say, up to 10%, then R 0.1, which will lead
If we arepmore
to L 8:1 Dt or t
0.015L2/D.
Note that the validity region of the first-term approximation is lower-bounded
with respect to time. It is, thus, often called the long-time solution. This is because
when the time is long enough, the physical dimension of the diffusion medium L,
which would otherwise
pffiffiffiffiffi be a semi-infinite medium, becomes smaller than the diffu-
sion distance (4 Dt).
At this point, you may wish to compare the validity time ranges of the short-time
solution and the long-time solution for the same diffusion medium L in Fig. 1.17. If
we still insist a 0.1% relative error:
L2
The short-time or error-function solution is valid for t ;
4:62 D
2
L
The long-time or first-term approximation is valid for t
.
3:72 D
What do we have to do in order to describe accurately the evolution of the
concentration profile in between? Over the entire time range, it may be accurately
described by the infinite series solution, Eq. (1.101), but for the short time, the
simple error-function solution will be sufficient, and for the longer times, the first-
term approximation will be enough. If we are getting less and less tolerant to the
error, the gap between the validity time zones of the two types of solution gets wider
and wider.

1.7.8 Average Concentration and the First-Term


Approximation

The transient, trigonometric series solution, Eq. (1.101) describes the local con-
centration with time, c(x,t), but experimental determination of the local concentra-
tion per se is challenging because of the limited spatial resolution of any
measurement probe. You remember that in the sectioning technique to measure
the tracer diffusivity, what you experimentally measure is the amount of tracers in a
slice with a finite thickness, not the local concentration itself. In reality, what can be
easily determined experimentally is the overall spatial average of concentration at a
time t or hc(t)i, and hence, we often want to know how the mean concentration varies
with time.
For the local concentration variation c(x,t), Eq. (1.101), the (spatial) mean
concentration of the entire system at a moment t is calculated as
1.8 Method of Laplace Transform 47

ZL

1 8co X 1 ð2j þ 1Þ2 π2 Dt
hcðtÞi ¼ cdx ¼ exp  : ð1:116Þ
L π2 j¼0 ð2j þ 1Þ2 L2
0

Note that the mean concentration hc(t)i, which is now a function of t only, is still
given in terms of an infinite series with respect to time, and we may also take the
first-term approximation as
 
8co t
hci ¼ exp  : ð1:117Þ
π 2 τ

Here, τ, called the relaxation time of the system, is defined as

L2
τ¼ : ð1:118Þ
π2 D

Equation (1.117) indicates that hci ¼ 8co/π2, not equal to co, as t ! 0. You know
it is simply because Eq. (1.117) is the first-term approximation, and you can
estimate the validity time range of this approximation by using the ratio test,
depending on the accuracy you want.

1.7.9 Homogenization Time

You may now want to ask how long one has to let diffusion continue or to anneal the
system until hc(t)i ¼ 0 in the present case or complete homogenization. The first-
term approximation, Eq. (1.117), allows you to estimate the time for complete
homogenization. Neglecting the numerical pre-factor 8/π2 because it is quite close
to 1, you may recognize that when t ¼ 1τ, hci/co  0.37, or your system is
approximately 63% homogenized, and for t ¼ 3τ, hci/co  0.05 or 95% homoge-
nized and so on. If you want to homogenize by more than 99%, you should anneal
your system for t
5τ. You may thus estimate how long you have to wait for your
system to completely homogenize or equilibrate in a new thermodynamic condition.
Depending on the magnitude of the diffusivity and the size of the medium, the
complete homogenization may become impossible even in the lifetime of this
universe (~1018 s).

1.8 Method of Laplace Transform

Up until now, we have obtained the solution to Fick’s second law by the superpo-
sition method for an error function solution and by the separation-of-variables
method for an infinite trigonometric series solution. As regards the thin-film
48 1 Diffusion in Continuum

solution, you have been required to simply accept it. This sort of learning sometimes
makes you feel uneasy. Now you will get a relief by learning a very powerful method
to solve the diffusion equation called the method of Laplace transform.
A total differential equation is normally easier to solve than a partial differ-
ential equation. You may thus wish to have an ingenious device which allows you
to transform Fick’s second law involving the partial derivatives to a total differen-
tial equation, that is, the Laplace transform. The principle is to transform the
differential equation by this device to an easily solvable form, to get the solution in
the Laplace-transformed form and then to transform back to the untransformed
function.

1.8.1 Definition of Laplace Transform

The Laplace transform of a function f(t), ℒ{f(t)}, is defined as

Z1
ℒff ðtÞg  f ðpÞ ¼ f ðtÞept dt ð1:119Þ
0

where the parameter p is such that the integral is made always converge. What can be
achieved by this transform? By multiplying f(t) by a function ept, you can make
sure that the area below the new function f(t)ept converges. This area, now a
function of p, corresponds to the transform f ðpÞ.
Some examples of the Laplace transforms are listed in Table 1.4. You are
strongly encouraged to transform all these functions for yourself following the
recipe, Eq. (1.119). The functions in this table are simple, but they are useful to
know because they appear quite often in diffusion problems.

Table 1.4 A few examples of Laplace-transforms


f(t) L{f(t)}
#1 1 1
p
#2 λ (any constant) λ
p
#3 eiωt pþiω
p2 þω2
#4 sinωt ω
p2 þω2
#5 cosωt p
p2 þω2
#6 f0(t) pf ðpÞ  f ðt ¼ 0Þ
00
#7 f (t) pf 0  f 0 ðt ¼ 0Þ ¼ p2 f  pf ðt ¼ 0Þ  f 0 ðt ¼ 0Þ
1.8 Method of Laplace Transform 49

1.8.2 Application to Diffusion Problems

Why do we transform in this way? It is just to solve the partial differential equations
more easily, and the end always justifies the means after all. You will see now how it
works.
First of all, we have to Laplace-transform the diffusion equation Fick’s second
law itself as
  !
2
∂cðx, tÞ ∂ cðx, tÞ
ℒ ¼ Dℒ : ð1:120Þ
∂t ∂x2

By noting that f0(t) ¼ ∂c/∂t, the left-hand side takes the form (#6 in Table 1.4)
 
∂c
ℒ ¼ pcðxÞ  cðx, t ¼ 0Þ ð1:121Þ
∂t

where c(x,t ¼ 0) is the initial condition given. The right-hand side transforms as
 2  Z 1 2
∂ c ∂ c d2 c
ℒ ¼ ept dt ¼ 2 : ð1:122Þ
∂x2 0 ∂x 2
dx

Fick’s second law, thus, takes the transformed form

d2 c p cðx, t ¼ 0Þ
2
 cþ ¼ 0: ð1:123Þ
dx D D

Here, you have to recognize that the primitive c(x,t) is a function of x and t, but
the Laplace-transformed cðx, pÞ now becomes a function of x only. The partial
differential equation with respect to x and t has, thus, been transformed to a total
differential equation with respect to x. By the way, do not forget that in this process,
the initial condition c(x,t ¼ 0) has already been used up during transforming,
Eq. (1.121).
Solving the transformed diffusion equation, Eq. (1.123) is now a piece of cake!
You can immediately write down the solution as
pffiffip pffiffip cðx, t ¼ 0Þ
cðx, pÞ ¼ Ae Dx þ Be Dx þ : ð1:124Þ
p

It is emphasized that this is the general solution to Fick’s second law which is
applicable to any case with a constant D but in the Laplace-transformed form. Now
what you have to do is to determine the two integration constants A and B, and you
know that you have still reserved the two boundary conditions.
As an example, let us apply this method to the case with the initial and boundary
conditions:
50 1 Diffusion in Continuum

IC: cðx
0, t ¼ 0Þ ¼ 0 ð1:125Þ
BC: cðx ¼ 0, tÞ ¼ cs ; cðx ! 1, tÞ ¼ 0 ð1:126Þ

which is for the case (iv) in Table 1.3. Due to the initial condition, Eq. (1.125), the
last term on the right-hand side of Eq. (1.124) disappears. In order to apply the
boundary conditions, Eq. (1.126), on the other hand, these should also be Laplace-
transformed because the solution is given in terms of the Laplace-transform of c or
c . Following the transformation recipe, Eq. (1.119), they are respectively
transformed as

cs
c ð x ¼ 0Þ ¼ ; cðx ¼ 1Þ ¼ 0: ð1:127Þ
p

The second boundary condition immediately requires that A ¼ 0 in Eq. (1.124)


because the solution would otherwise blow up as x ! 1. The first boundary
condition then stipulates that B ¼ cs/p. Finally, the solution takes the form

cs pffiffiDp x
cðx, pÞ ¼ e : ð1:128Þ
p

So simple, but alas, this is not what we want! What we want is in the form of c ¼ c
(x,t). How can we proceed from here? Yes, we should look for the function c(x,t)
whose Laplace transform is cðx, pÞ . The mathematical procedure to retrieve the
primitive c from c or to un-transform the Laplace-transformed function is called the
inverse transform. The basic idea of the inverse transform is so simple, that is:

If ℒfcðx, tÞg ¼ cðx, pÞ, then ℒ1 fcðx, pÞg ¼ cðx, tÞ: ð1:129Þ

The mathematical procedure, however, is lengthy, but you do not have to worry
about it: our kind mathematicians have already carried out the inverse transform of
almost all Laplace transforms haunting the diffusion problems to prepare a table,
which you can find from Crank [9] or Carslaw and Jaeger [10]. Some examples of
the inverse Laplace transforms which you may more likely encounter in solving
the ordinary or thermal diffusion problems are listed in Table 1.5 as an extension of
Table 1.4.
Looking up the table of inverse transforms, you can pick up the inverse
transform corresponding to our transformed solution Eq. (1.128) (#11 in Table 1.5):
( pffiffip )  
e Dx
1 x
ℒ ¼ erfc pffiffiffiffiffi : ð1:130Þ
p 2 Dt

So, we have come to the solution for the initial and boundary conditions,
Eqs. (1.125) and (1.126), as
1.9 Solutions When D Is Not Constant 51

pffiffiffiffiffiffiffiffiffi
Table 1.5 Some examples of the inverse Laplace transforms. Here, q  p=D, and h and L are
positive constants like the thickness of a source and of a diffusion specimen, respectively
 
f ðpÞ f ðtÞ ¼ ℒ1 f ðpÞ
#8 1 eαt
pþα
#9 eqx pffiffiffiffiffiffiffi
x
ex =4Dt
2

2 πDt3
#10 eqx D 1=2 x2 =4Dt
q πt e
eqx
 
#11 pxffiffiffiffi
p erfc
#12 eqx Dt 1=2
2 Dt
 
pq 2 π e  x  erfc pxffiffiffiffi
x2 =4Dt
2 Dt
#13 eqx     t 1=2 x2 =4Dt
erfc pxffiffiffiffi  x πD e
2
p2 t þ 2D
x
2 Dt
 1=2 2  
#14 eqx D x pffiffiffiffiffi
ex =4Dt  hDehxþDth erfc pffiffiffiffiffi þ h Dt
2
qþh
πt 2 Dt
eqx
 pffiffiffiffiffi
pxffiffiffiffi
#15
þ h Dt
2
qðqþhÞ DehxþDth erfc
2 Dt
sinh qx P ð2jþ1Þπx 2 2
#16
q cosh qL
2D
L ð1Þj sin 2L  eð2jþ1Þ π Dt=4L2
j¼0
cosh qx cosh ðLxÞ P ð2jþ1Þπx 2 2
#17
q cosh qL 4D
L cos L  eð2jþ1Þ π Dt=L2
j¼0
#8 to #15 from Crank [9]; #16 and #17 from Lee and Yoo [12]

 
x
c ¼ cs erfc pffiffiffiffiffi ,
2 Dt

which is what we already obtained by inspection, Eq. (1.70), in Table 1.3. The
method is indeed neat, isn’t it? One may wish to continue to apply this method to
solve more amusing diffusion problems for, e.g., a thin-film source, a semi-infinite
source, oscillating boundary conditions, and so on, but we have to stop here. If you
are interested, you are referred to a classic textbook on electrochemistry by Bockris
and Reddy [13].

1.9 Solutions When D Is Not Constant

Up to this point, we have played with Fick’s second law involving a constant D. It
is, however, a special case: In general, D may be dependent on position and/or time.
For such general cases, one should solve the general form of Fick’s second law,
 
∂C ∂ ∂C
¼ D ,
∂t ∂x ∂x
52 1 Diffusion in Continuum

of course with the initial and boundary conditions. The situation may be distin-
guished into two cases: (i) when the diffusivity is a function of time only or D ¼ D(t)
and (ii) when the diffusivity is a function of concentration D ¼ D(c), the most
general case.

1.9.1 When D = D(t)

This situation is often encountered in materials processing. For example, when a


material with a composition-independent diffusivity is heated up or cooled down
with a temperature change such that T ¼ T(t), the diffusivity D, which is normally
exponentially dependent on temperature, becomes a function of time via temperature
or D ¼ D(T(t)). The diffusion equation in this case may be rewritten as

2
∂c ∂ c
¼ : ð1:131Þ
DðtÞ∂t ∂x2

If we let dτ ¼ D(t)dt,
Z t
τðtÞ ¼ DðtÞdt þ τðt ¼ 0Þ, ð1:132Þ
0

and the diffusion equation becomes

∂c ∂c2
¼ : ð1:133Þ
∂τ ∂x2

This looks like Fick’s second law with D ¼ 1. In order to solve the equation, the
initial and boundary conditions, of course, should be given in terms of x and τ. The
solution is then the same as what we have obtained with constant D but with D and t
replaced with 1 and τ, respectively,

c ¼ cðx, t ¼ τ; D ¼ 1Þ: ð1:134Þ

By substituting Eq. (1.132) for τ, you can finally get the solution to the problem in
terms of C¼C(x,t).

1.9.2 When D = D(c): Boltzmann-Matano Analysis

If D is a function of concentration c, it is often said that Fick’s second law is very


difficult to solve analytically, if not impossible. The analytic solutions are available
only for a few special cases in which the concentration dependence of D is, for
example, exponential or linear, i.e.,
1.9 Solutions When D Is Not Constant 53

D ¼ eαc or D ¼ ac þ b:

For the analytic solutions, look up Crank’s book [9].


For the most general case in which the concentration dependence of D is arbitrary,
the analytic solution is impossible, but one can do now the opposite: determine
D ¼ D(c) graphically from a concentration profile c ¼ c(x,to) that is frozen-in in a
diffusion time t ¼ to. The graphical method is called the Boltzmann-Matano
analysis, which is quite widely used to determine D(c).
Total differential equations are always easier to solve than partial differential
equations. This analysis starts by transforming Fick’s partial differential equation
into a total differential equation. Let us set

x
η ¼ pffi : ð1:135Þ
t

Then, due to the chain rule of differentiation, you may transform the differential
operators in terms of the variable η as

∂ d ∂η 1 d
¼  ¼ pffi ð1:136aÞ
∂x dη ∂x t dη
∂ d ∂η 1 d
¼  ¼ η : ð1:136bÞ
∂t dη ∂t 2t dη

Our partial differential equation, Fick’s second law,


 
∂c ∂ ∂c
¼ D ,
∂t ∂x ∂x

is subsequently transformed to a total differential equation in terms of η as


 
1 dc d dc
 η ¼ D : ð1:137Þ
2 dη dη dη

This transform was first invented by Boltzmann [14] and is called the
Boltzmann transform.
Let us pause here to consider what would happen if D ¼ constant. If D is constant,
then Eq. (1.137) takes the form

d2 c η dc
þ ¼ 0: ð1:138Þ
dη2 2D dη

It is really easy to solve this equation by rewriting it in the form of y0/y ¼  η/2D
where y ¼ dc/dη and integrating twice. Do not be surprised that the solution is
54 1 Diffusion in Continuum

 
x
c ¼ A þ Berf pffiffiffiffiffi
2 Dt

which is the two parameters form of the error-function solutions, Eq. (1.73), that
we have already met. In order to determine the two integration constants A and B,
you should have two conditions, which must be able to be expressed in terms of the
single variable η. Otherwise, the present solution would not work. Here, you may
remember that all the initial and boundary conditions for all those short-time
solutions we have learned could be written in terms of η. That is how those solutions
are obtained. You now understand why the short-time solutions can be written
down in the two parameters form of Eq. (1.73).
Coming back to the main stream, let us consider the diffusion with an arbitrary
diffusivity from a semi-infinite source to a semi-infinite medium, Fig. 1.13. The
initial and boundary conditions are still the same as in Eq. (1.59):

cðx > 0, t ¼ 0Þ ¼ co
cðx ¼ 1, tÞ ¼ co
ð1:139aÞ
cðx 0, t ¼ 0Þ ¼ 0
cðx ¼ 1, tÞ ¼ 0:

These four conditions can be represented by just two in terms of the single
variable η as

cðη ¼ 1Þ ¼ co
ð1:139bÞ
cðη ¼ 1Þ ¼ 0:

You are thus allowed to use Eq. (1.137). The expected diffusion profile at t ¼ to
may look like Fig. 1.21, and the diffusivity can be determined in this way:
Equation (1.137) is a total differential equation with respect to η; thus this
equation may be modified by eliminating dη from the both sides to
 
1 dc
 η dc ¼ d D : ð1:140Þ
2 dη

As the
pffiffiffidiffusion
ffi time is fixed as t ¼ to, the single variable is replaced back with
η ¼ x= to to obtain
 
1 x dc
 dc ¼ d D : ð1:141Þ
2 to dx

You see, the right-hand side of this equation is a total differential of Ddc/dx! Let
us integrate the both sides from c ¼ 0 up to any arbitrary concentration c ¼ c0:
1.9 Solutions When D Is Not Constant 55

Fig. 1.21 The diffusion c


profile for D ¼ D(c) at t ¼ to,
c0
expected from a semi-
infinite source

c¢ ⎛ dc ⎞
⎜ ⎟
⎝ dx ⎠c =c′
0
x

Z c0 Z c0      
1 dc dc dc
 xdc ¼ d D ¼ D  D : ð1:142Þ
2to 0 0 dx dx c¼c0 dx c¼0

Can you recognize here that dc/dx ¼ 0 as c ! 0? See an expected profile in


Fig. 1.21. The diffusivity at c ¼ c0, D(c0) can, thus, be determined as
R c0
0 1 xdc
Dðc Þ ¼   dc0 : ð1:143Þ
2to dx 0 c¼c

That is, once you have an experimental diffusion profile c vs. x in a diffusion time
to, you can determine the diffusivity at a specific concentration c0 by evaluating the
instantaneous slope right there, (dc/dx)c¼c0 as well as the area under the curve x vs. c
from c ¼ 0 to there (see Fig. 1.21). It sounds easy enough, but there is a tricky point.
In order to execute this integration to evaluate the area, one should know where the
origin of the x-coordinate x ¼ 0 falls, but where is it?
When we first prepared a diffusion couple with a semi-infinite source and a semi-
infinite medium in contact, we can naturally put the origin x ¼ 0 at the interface
between the two semi-infinite bars where the concentration changes stepwise. As
diffusion proceeds, however, the step-like initial distribution gradually decays lead-
ing to a sigmoidal concentration profile as shown in Fig. 1.21. Where is x ¼ 0 now?
The origin x ¼ 0 can be located on the basis of the mass-conservation principle,
namely, the mass lost by the source in x > 0 should equal the mass gained by the
medium in x < 0 as

Zco
xdc ¼ 0: ð1:144Þ
0

This fact has already been suggested by Eq. (1.142) for (dc/dx) ! 0 as c0 ! co.
The plane x ¼ 0 as such is called the Matano-plane or Matano-interface after
Matano [15] who first applied this method to the diffusion in metals. Geometrically
this equation means that the area under x vs. c to the right of the Matano interface is
equal to that to the left of the Matano interface (see Fig. 1.22):
56 1 Diffusion in Continuum

Fig. 1.22 Location of the c


Matano-plane, x ¼ 0, and
c0
evaluation of D(c) at c ¼ c0 . t=t0
The line segment represents
dc/dx at c ¼ c0


0
0 x

Matano interface

Z0 Z1
cdx ¼ ðco  cÞdx: ð1:145Þ
1 0

Once the Matano interface is located on the basis of this equation, the numerator
integral in Eq. (1.143) can be evaluated, and thus D can be determined as a function
of c in the range of c ¼ 0 to co.
Finally, let us consider the sign of D in Eq. (1.143) referring to the profile in
Fig. 1.22. Obviously the denominator slope is +, and the numerator area remains –
up to c ¼ co, thus D > 0 always.
In passing, it is mentioned that the Matano interface, x ¼ 0, actually defines
various reference frames of coordinates in diffusion. For example, if c is given in
terms of number density (number of atoms/m3) or molar density (mol/m3), the
number of moles lost and gained is equal to each other with respect to the Matano
interface. This frame of reference is called the number-fixed frame of reference.
Similarly, if c is given in terms of mass density or kg/m3, the Matano interface
x ¼ 0 defines the mass-fixed frame of reference. Finally, if c is given in volume
density, the volume lost equals to the volume gained with respect to the Matano
interface, which defines the volume-fixed frame of reference. This reference frame
is a geometric reference frame in which the geometric scale of the x-axis remains
unchanged. Rigorously speaking, Fick’s law is valid only in this geometric frame of
reference, thus often called Fick’s frame of reference. In the present treatment,
however, we assume that there will be no volume change with diffusion.

1.10 Diffusion in Higher Dimensions

Up to now, we have learned the diffusion mathematics only in one dimension. How
do we proceed in two or three dimensions? We will get an idea with the aid of an
example again assuming a constant D. Consider an infinite bar, with a rectangular
cross section a b, which has initially no solute inside, but the surface concentration
is fixed at cs (see Fig. 1.23a).
1.10 Diffusion in Higher Dimensions 57

Fig. 1.23 (a) Diffusion in y


two dimensions where the
b
shade represents the
concentration c(x,y). (b) An
actual example, sidewalk
blocks upon drying after
rainfall

0 x
0 a
(a) (b)

You can immediately imagine that the diffusion proceeds two-dimensionally


along the x-axis and y-axis. The differential equation, Fick’s second law, is
written as
 2 2 
∂c ∂ c ∂ c
¼D þ ð1:146Þ
∂t ∂x2 ∂y2

with the initial and boundary conditions

IC: cð0 < x < a, 0 < y < b, t ¼ 0Þ ¼ 0 ð1:147aÞ


BC: cðx ¼ 0, y, tÞ ¼ cs ð1:147bÞ
cðx, y ¼ 0, tÞ ¼ cs ð1:147cÞ
cðx ¼ a, y, tÞ ¼ cs ð1:147dÞ
cðx, y ¼ b, tÞ ¼ cs : ð1:147eÞ

By applying the separation-of-variables method, you may assume the primi-


tive c(x,y,t) to be a product of a function of x and t only, X(x,t) and another of y and t
only, Y(y,t):

cðx, y, tÞ ¼ Xðx, tÞ  Yðy, tÞ: ð1:148Þ

Substituting into Fick’s second law and dividing through XY, you may obtain
 2 2 
1 ∂X 1 ∂Y 1 ∂ X 1 ∂ Y
þ ¼D þ : ð1:149Þ
X ∂t Y ∂t X ∂x2 Y ∂y2

Each side is a linear combination of a function of x only and another of y only.


The equation may, thus, be decomposed into two as

2
∂X ∂ X
¼D 2; ð1:150aÞ
∂t ∂x
58 1 Diffusion in Continuum

2
∂Y ∂ Y
¼D 2: ð1:150bÞ
∂t ∂y

For short times such that t 0.047(a/2)2/D or 0.047(b/2)2/D to the error of no


greater than 0.1%, for example, you can show that the solution is
   
x y
cðx, y, tÞ ¼ cs erfc pffiffiffiffiffi erfc pffiffiffiffiffi : ð1:151Þ
2 Dt 2 Dt

As exercise, you are asked to draw the concentration profiles c(x,y,t) at different
times paying attention to the lines x ¼ a/2, y ¼ b/2, and y ¼ x (see Fig. 1.23b).
Finally, it is emphasized that in this treatment, one essential condition is that the
initial condition should be able to be expressed as a product of the initial conditions
of the partial solutions,

cðx, y, z, 0Þ ¼ Xðx, 0Þ  Yðy, 0Þ  Zðz, 0Þ: ð1:152Þ

For more discussions, you are referred to Crank [9].

1.11 Moving Boundary Problem

In materials study, you may often encounter the situation in which the boundary
between two neighboring phases is moving with diffusion, and you may need to
know how fast the boundary moves. You are now given an example as the final topic
in this chapter.
Let us consider a binary subsolidus phase diagram, Fig. 1.24a, which is actually
like the system of Fe–C. What sort of diffusion profile do you expect if the solute,
T c
cS
γ
*
α α+γ γ cγ
T1

t1 t2 α+ γ

c*α
α
Fe c*α c*γ cS C 0 s x
(a) (b)

Fig. 1.24 (a) Partial phase diagram of Fe–C. (b) Carbon concentration profiles expected at time t1
and a later time t2 when a semi-infinite bar of α-Fe is exposed to a solute carbon (C) concentration cs
at temperature T1. Note that the phase boundary “s” between γ-Fe and α-Fe moves inward with time
1.11 Moving Boundary Problem 59

say, carbon(C), is allowed to diffuse into a semi-infinite bar of pure α-iron (with an
infinite cross sectional area), from the surface at a fixed concentration c ¼ cS? We
assume here that the local thermodynamic equilibrium or phase equilibrium is
always prevailing inside the system. If the system made a complete solid solution
and D were constant, then you would have a simple complementary-error-functional
distribution which we have earlier dealt with (see Table 1.3). The present system,
however, exhibits two different phases, say, α and γ, having different saturation
solubilities. Therefore, there will be a compositional discontinuity at the phase
boundary, now between α- and γ-phase, and the profile at time t1 and t2(>t1) may
look as illustrated in Fig. 1.24b. The compositions in equilibrium at the discontinuity
or phase boundary are c γ and c α according to the phase diagram, Fig. 1.24a. As
diffusion continues with time, this phase boundary between γ-Fe and α-Fe will move
inward. The question is how fast this phase boundary moves inward.
Suppose that the phase boundary has moved from x¼s to s+ds in dt. If we let Jin
be the solute(C) flux coming in when the boundary is at x¼s and Jout the flux coming
out when the boundary has moved to s+ds, we may write, due to the mass conser-
vation principle,
 
ds
ðJout  Jin Þ ¼ c γ  c α : ð1:153Þ
dt

This reminds you of the continuity equation, ∇J ¼ ∂C/∂t, doesn’t it? This
equation tells you that the moving speed of the boundary ds/dt is determined by the
difference between the incoming flux from the γ-phase and the outgoing flux to the
α-phase.
The fluxes coming to and going away from the boundary will be given by Fick’s
first law, but with the concentration gradient to the left (x ¼ s) and to the right
(x ¼ s+) of the boundary s, respectively. Assuming that the diffusion coefficients of
the solute, say, carbon, are constant Dα and Dγ in both phases, respectively,
   
∂cγ ∂cα
Jin ¼ Dγ ; Jout ¼ Dα : ð1:154Þ
∂x x¼s ∂x x¼sþ

In this formulation, we have also assumed no volume change with diffusion.


The problem has now been reduced to how to evaluate the both slopes ∂c/∂x at
the phase boundary, and for these, you should know the concentration distributions
cγ(x,t) and cα(x,t) in the γ-phase (0 < x s) and α-phase (s+ x), respectively. The
concentration distributions are formally found by solving Fick’s second law

2
∂cγ,α ∂ cγ,α
¼ Dγ,α ð1:155Þ
∂t ∂x2

with the appropriate initial and boundary conditions for cγ(x,t) and cα(x,t), respec-
tively. By noting, however, that for cγ(x,t) in the γ-phase
60 1 Diffusion in Continuum

cγ ðx ¼ 0, tÞ ¼ cs ; cγ ðx ¼ s , tÞ ¼ c γ ð1:156Þ

and for cα(x,t) in the α-phase

cα ðx ¼ sþ , tÞ ¼ c α ; cα ðx ¼ 1, tÞ ¼ 0, ð1:157Þ

let us try the error-function solution in the two-parameter fashion, Eq. (1.73). If we
first assume
!
x
cγ ¼ A þ B erf pffiffiffiffiffiffiffi , ð1:158Þ
2 Dγ t

the first boundary condition in Eq. (1.156) already gives you

A ¼ cs , ð1:159Þ

and the second boundary condition gets you


!
s
c γ ¼ cs þ B erf pffiffiffiffiffiffiffi : ð1:160Þ
2 Dγ t

Here, s means mathematically sε as the positive-definite ε!0, thus ultimately



s ! s. Letting

s
β ¼ pffiffiffiffiffiffiffi , ð1:161Þ
2 Dγ t

B is evaluated in terms of β, and finally the solution takes the form


!
c γ  cs x
cγ ¼ cs þ erf pffiffiffiffiffiffiffi : ð1:162Þ
erf ðβÞ 2 Dγ t

Therefore, the concentration gradient at x ¼ s and hence the flux is evaluated as


  rffiffiffiffiffiffi
∂cγ Dγ cs  c γ
Jin ¼ Dγ ¼  : ð1:163Þ
∂x x¼s
πt eβ2 erf ðβÞ

For cα, we also set a trial solution similar to Eq. (1.158), and by using the two
boundary conditions, Eq. (1.157), you can determine the constants A and B as

c
A ¼ B ¼ α : ð1:164Þ
erfc psffiffiffiffiffi
2 Dα t
Problems 61

Recollecting the definition of β in Eq. (1.161), the argument of erfc may be a bit
beautified by introducing the diffusivity ratio ψ as

s pffiffiffiffi Dγ
pffiffiffiffiffiffiffi ¼ β ψ; ψ , ð1:165Þ
2 Dα t Dα

and you get the solution


 
c α x
cα ¼
pffiffiffiffi erfc pffiffiffiffiffiffiffi : ð1:166Þ
erfc β ψ 2 Dα t

This equation, thus, allows you to have


  rffiffiffiffiffiffi
∂cα Dα c α
¼ Dα ¼  2 ffi : ð1:167Þ
πt eβ ψ erfc βpffiffiffi
Jout
∂x x¼sþ ψ

From Eq. (1.161), you also have


rffiffiffiffiffiffi
ds Dγ
¼β : ð1:168Þ
dt t

By substituting Eqs. (1.163), (1.167), and (1.168) into Eq. (1.153), you finally
obtain, after some algebra,

pffiffiffi  cs  c γ c
π cγ  c α ¼  pffiffiffiffi 2 α pffiffiffiffi , ð1:169Þ
βeβ erf ðβÞ β ψeβ Ψ erfc β ψ
2

an implicit equation for β. This allows you to evaluate β numerically in terms of all
known parameters, the diffusivity ratio ψ, the conjugate concentrations c γ and c α ,
and the surface concentration imposed cs.
How does the interface move after all? Equation (1.161) says that
pffiffiffiffiffiffiffi
s ¼ 2β Dγ t ð1:170Þ

where β is such as in Eq. (1.169).

Problems

1. By applying the mass conservation law to an infinitesimal volume element in


each of the following coordinates systems, respectively, derive the expressions
!
for divergence (∇) and Laplacian (∇2) for each coordinate system:
(a) Rectangular coordinates; (b) cylindrical coordinates; (c) spherical coordinates
62 1 Diffusion in Continuum

2. A sphere of inner radius ri and outer radius ro is immersed in a bath of temperature


To. If the inner temperature is kept at Ti and heat is lost at a rate q,_ derive an
equation for the thermal diffusivity “α” in terms of ri, ro, Ti, To, and q_ . The
continuity equation for energy in spherical symmetry takes the form

 2 
∂T ∂ T 2 ∂T
¼α þ :
∂t ∂r2 r ∂r

3. Calculate that the mean square displacement <r2> for an infinite line source and
for a point source, respectively, in an isotropic medium.
4. An experiment similar to Smith’s (Acta Met.,1 (1953) 578) is performed on a
hollow iron tube with a 1.11 cm outer diameter, a 0.86 cm inner diameter, and a
length of 10 cm. In 100 h, 3.60 g of carbon passes through the tube. The variation
of carbon with radius in the tube is given below. Plot c vs. r , and from this
calculate and plot the diffusion coefficient over this range of carbon contents.
(Density of iron is 7.87 g/cm3.)

r/cm Wt. % carbon r/cm Wt. % carbon


0.553 0.28 0.491 1.09
0.540 0.46 0.479 1.20
0.527 0.65 0.466 1.32
0.516 0.82 0.449 1.42
0.503 0.96

5. It is desired to store hydrogen at 400 C. A steel tank is available, and you are
asked to calculate the steady-state rate of pressure drop for a cubical tank of 1 L
capacity with a wall thickness of 1 mm when the internal pressure is P ¼ 10 atm.
The tank is to operate in a vacuum. Assume that D ¼ 104 cm2/s, independent of
the state of stress, that the hydrogen in the steel is in equilibrium with the
atmosphere at both the high- and low-pressure sides, that the solubility is pro-
portional to P1/2, and that at P ¼ 1 atm, the solubility is 3 ppm (by weight). State
clearly any additional assumptions made in your analysis. Density of steel,
ρ ¼ 7.87 g/cm3.
6. A thin film of radioactive copper was electroplated onto the end of a copper
cylinder. After a high temperature anneal of 20 h, the specimen was sectioned and
the activity (A) of each section counted. The data are:

A/counts/min/mg x/102 cm
5012 1
3981 2
2512 3
1413 4
524.8 5
Problems 63

(a) Plot the data and determine D by fitting to the thin-film source solution.
(b) Evaluate D from the slope of log A versus x2 using a least-squares procedure,
and compare with that in Part (a).
7. For a hard brittle material, it is difficult to determine D by grinding off layers and
collecting the material removed. An alternative method is to count the activity
remaining in the sample after removing material by grinding.
(a) For the case in which a thin layer of tracer is placed on the original surface at
x ¼ 0, give the solution to the diffusion equation, c(x,t), after a diffusion
anneal, but before any material is ground off the surface. Draw c(x,t) vs. x.
(b) Show that the total amount of material left in the sample after the diffusion
layer has been ground away to a depth of d is:


 
d
qðdÞ ¼ B 1  erf pffiffiffiffiffi :
2 Dt

For an isotope whose radiation is adsorbed little in leaving the sample,


B is the total activity before and grinding is done. Note the determination of
q(d) allows the determination of D.
8. Prove that

Z 1

α 0 ðx  x0 Þ2
Cðx, tÞ ¼ f ðx Þ exp dx0
t1=2 1 4Dt

is a solution to the one dimensional diffusion equation

2
∂C ∂ C
¼D 2
∂t ∂x

where f(x0) is the initial distribution of the diffusing species at t ¼ 0.


9. Silicon can be doped with boron in two steps:
(a) The boron is introduced by heat treatment in a gas of BCl3 or B2O3. The
vapor pressure is high enough so that the surface concentration reaches the
maximum solubility of boron in silicon, CS, at the heat treatment
temperature.
(a-1) Using the appropriate solution of Fick’s equation, give an analytical
expression for the quantity of boron, M, introduced during a treatment of
duration t1 at temperature T1.
(a-2) For T1 ¼ 1000 C, t1 ¼ 6 min, CS ¼ 3 1020 atoms cm3, and
D ¼ 4 1013 cm2 s1, plot the penetration profile of boron over five
decades of concentration.
64 1 Diffusion in Continuum

(a-3) Calculate the amount of boron introduced. Give the thickness of this
layer in terms of (Dt)1/2.
(b) A subsequent diffusion treatment is carried out to decrease the surface
concentration and increase the depth of the doped region.
(b-1) Represent the doped zone of the previous paragraph as a thin layer at
the surface, and assume that there is no loss of boron from the surface.
What is the solution of the diffusion equation using the analytical results
of the last paragraph? What conditions on the time t2 and the temperature
T2 are necessary so that this solution can be used?
(b-2) With T2 ¼ 1200 C, t2 ¼ 8 or 80 min, and D ¼ 3 1012 cm2 s1,
establish the corresponding diffusion profiles. Compare the penetration
depths so obtained to that obtained in the first heat treatment at 1100 C.
Discuss the result.
(c) What diffusion time at 1200 C is needed to obtain a concentration of 1018
atoms cm3 at a depth of 8 μm? What is the concentration of boron at the
surface after such a treatment?
10. Two large plates of steel, each 1 cm thick, are friction-bonded and then annealed
at 1000 C. The carbon concentrations are 1% in the left plate and 0.2% in the
right plate before annealing. The diffusivity of carbon at this temperature is
DC ¼ 2 107 cm2/s.
(a) How long will it take for the carbon concentration to reach 0.3 wt% in the
right plate at a distance of 0.1 cm from the interface?
(b) When is the short-time solution in part (a) no longer valid to within an error
of 1%?
(c) Derive an expression for the velocity of the iso-concentration plane,
ρ ¼ 0.3 wt% C.
(d) Write an expression for the velocity of the same iso-concentration plane,
ρ ¼ 0.3 wt% C, suitable for long times.
(e) Plot the position of the iso-concentration plane, ρ ¼ 0.3 wt% C, versus time
on the interval 0 < x/cm < 1.0. Approximate the region between the short-
and long-time solutions.
11. Consider an infinite slab of thickness L, where a solute is initially at a uniform
concentration co from a surface (x ¼ 0) to a thickness “g” such that g  L.
(a) You can tell how the concentration profile c(x,t) develops with position and
time by solving Fick’s second law with the appropriate initial and boundary
conditions. Write the initial and boundary conditions.
(b) Using the separation of variables method, solve Fick’s second law to obtain
c(x,t).
(c) Calculate the concentration at the other surface (x ¼ L) of the slab as a
function of time, c(L,t).
(d) What will be the maximum possible value (cM) of c(L,t)?
Problems 65

(e) How long will it take for the concentration c(L,t) to reach half the
maximum?
(f) Discuss how to determine the diffusivity of the solute.
(This is actually the principle of the laser flash method of determining the
thermal diffusivity of a solid. One can instantaneously achieve a temperature of
To in a shallow surface region (0 x < g  L) by flashing laser onto the surface
of a specimen. By monitoring the temperature at the other end of the slab
specimen (x ¼ L) against time, one can determine the thermal diffusivity as in
part (d).)
12. To harden fully a sheet of low-alloy steel during a quench from an initial
temperature, T0 ¼ 1000 C, the sheet must be cooled at 100 Ks1 in the vicinity
of 600 C. It may be assumed that the surfaces immediately fall to the temper-
ature of the quenching medium, 0 C.
The power series solution to this problem is given as

"  2 #
4T0 X 1
1
ð2j þ 1Þπx ð2j þ 1Þπ
T¼ sin exp  αt
π j¼0 2j þ 1 h h

where h is the thickness of the sheet and α is the thermal diffusivity.


(a) Sketch the temperature at the sheet’s center as a function of time as estimated
by the first-term approximation of T(x,t). Under what condition is this
approximation valid?
(b) Develop an expression for the cooling rate at the center of the steel sheet and
determine when a one-term approximation is valid.
(c) Calculate the cooling rate at the sheet’s center when its temperature is
600 C.
(d) Estimate the maximum sheet thickness that can be hardened fully in this
quenching medium. α ¼ 0.1 cm2 s1. Comment on how your result depends
upon the steel’s initial temperature.
13. Consider the outgassing of an infinite plate of thickness, 2 L, where the rate of
loss of diffusing substance is controlled by evaporation from the surface. If ρ is
the concentration in equilibrium with the surrounding atmosphere, the flux from
the surfaces of the plate may be represented as

Jsurface ¼ kðρ  ρ Þ at x ¼ L

where k is the evaporation rate constant. By the method of separation of vari-


ables, the solution is found to be

ρ  ρ X
1
2 sin ðλn LÞ
¼ exp λ2n Dt cos ðλn xÞ
ρ0  ρ λ L þ sin ðλn LÞ cos ðλn LÞ
n¼1 n
66 1 Diffusion in Continuum

where ρ0 is the initial concentration and λn’s are the roots of

k
λn tan ðλn LÞ ¼ :
D

(a) Consider the dimensionless group kL/D.


(i) Show that, when this quantity is large, the usual Fourier sinusoidal series
solution applies where ρ(L,t) ¼ ρ is the boundary condition.
(ii) Show that when this quantity is small, the concentration is uniform
across the plate.
(iii) Estimate values of kL/D such that within an error of 5% one can view
the outgassing as being either diffusion controlled or evaporation
controlled.
(b) Even when the ratio of kL/D puts the process in the regime of diffusion
control, at short times, the rate will be limited by evaporation from the
surface. Estimate the time interval over which the latter is the case.
14. A block of glass contains helium at a concentration of 1 ppm (part per million by
mass). The partial pressure of helium in equilibrium with this concentration is
0.1 atm. It may be assumed that Peq He is proportional to the ρHe in the glass over
the composition range of interest, the constant of proportionality denoted by
k. The glass block is placed in a furnace at 500 C in an atmosphere of flowing
neon which carries away the helium as it diffuses out of the glass.
In this problem, rather than assuming a constant zero surface concentration of
helium as the boundary condition, assume that there is a gas boundary layer next
to the surface of the glass. The rate of transfer of helium across the gas boundary
layer is given by

J ¼ αðPbulk  Psurface Þ

where Pbulk is the helium pressure in the bulk flow of the gas phase. In this case
Pbulk ¼ 0, and Psurface is the helium pressure at the surface of the glass. Assume
Psurface to be in equilibrium with ρHe at the glass surface at any time.
(a) Write the differential equation with initial and boundary conditions to
express the concentration of helium in the glass at short times.
(b) Solve by the Laplace transform method to obtain

 
ρðx, tÞ x x pffiffiffiffiffi
¼ erfc pffiffiffiffiffi  exp hx þ h Dt erfc pffiffiffiffiffi þ h Dt
2
ρðx, 0Þ 2 Dt 2 Dt

where h ¼ α/D.
(c) Plot the helium flux at the glass surface for 0 < t/s < 100.
Problems 67

(d) Compare the magnitude of resistance to flow in the gas phase boundary layer
to that in the glass block.
pffiffiffiffi
(e) How do your results change as the ratio α= D increases?
The density of the glass is 3.5 g cm3.
The mass transfer coefficient of helium, α, is 3 107 g cm2 s1 atm1.
The diffusion coefficient of helium in the glass block at 500 C is
2 108 cm2 s1.
15. A large steel pressure vessel 5 cm in thickness is used to contain a process that
generates hydrogen by a chemical reaction. To avoid problems with hydrogen
embrittlement of the steel, the pressure vessel is periodically purged and the
hydrogen baked out. After purging and bake-out, the hydrogen concentration in
the steel is negligibly low. Reactants are then introduced to the vessel, and the
kinetics of hydrogen production are such that the surface concentration on the
inside of the vessel increases linearly with time, i.e.,

ρH ð0, tÞ ¼ kt

where k ¼ 106 g cm3 s1.


(a) Derive a solution for ρH (x,t) that will be valid for a 1-day period after the
reaction starts. Please orient your solution with the origin (x ¼ 0) at the inner
surface of the wall. (In case, you may refer to Table 1.5 for inverse Laplace
transform.)
(b) A specification to prevent cracking of the vessel requires that the reaction be
stopped and the vessel purged of hydrogen if the average hydrogen concen-
tration exceeds 0.05 g/cm3 in the layer extending from the inner surface to a
depth of 1 mm. Considering this requirement, estimate if it is safe to run the
reaction for a 24-hour period before purging the vessel of hydrogen.
Data: Diffusion coefficient of hydrogen in steel, DH ¼ 106 cm2 s1.

16. An iron plate, 25 cm thick, contains H in the amount of 500 ppm (parts per
million by mass). During annealing, the surface concentration of hydrogen is
fixed at zero. The diffusion coefficient of H, DH ¼ 104 cm2/s.
(a) Calculate the degassing rate of hydrogen from the plate after 30 s. State your
assumptions.
(b) What does diffusion theory give for the degassing rate at t ¼ 0? Comment on
the physical relevance of this prediction.
(c) (i) Calculate the time for the plate to reach an average H concentration of
50 ppm. (ii) What is the degassing rate at this time?
17. As a result of the way that a block of austenite steel has been prepared,
“banding” is present in the structure. This banding consists of the concentration
of carbon having the form of a square wave in the x-direction:
68 1 Diffusion in Continuum

c
c1

c2

The width of the bands with the higher concentration is 50 microns each, and
that of those with the lower concentration is also 50 microns each. The concen-
trations are:

c1 ¼ 0:4%C, c2 ¼ 0:2%C:

(a) How long will it take to anneal the specimen at 1000 C


(DC ¼ 2 107 cm2/s) so that no point in the structure has a composition
that is more than 0.01% C away from the mean composition?
(b) If the original concentrations are c1 ¼ 0.35%C and c2 ¼ 0.25%C, what time
is required at 1000 C such that no point in the structure has a composition
that is more than 0.01%C away from the mean composition?
(c) How will the time of anneal necessary to obtain the same fractional com-
pletion of the process depend on the original values of c1 and c2 in this
structure? Assume that the diffusion coefficient is independent of
composition.
18. In a pure gold quenched from 700 C, it is thought that the supersaturation of
vacancies is relieved by adsorption of vacancies at dislocation lines.
(a) Considering the dislocation lines to be fixed cylindrical sinks of constant
radius ro, derive an equation giving the time dependence of the ratio of the
average vacancy concentration cðtÞ to the initial concentration co (for
0.8 < cðtÞ/co < 1), which could be used to check this hypothesis.
(b) Derive an equation for the case in which planar grain boundaries act as sinks
for the vacancies.
19. A rod of pure copper was jointed to a rod of a 29.4% Zn – 70.6% Cu alloy. After
annealing for 360 h, the % Zn vs. distance data were plotted on probability
paper, and the following values were picked off from the best line through the
data:

Atomic % Zn x/102 cm Atomic % Zn x/102 cm


0.3 50.05 23.5 36.55
1.5 48.15 25.0 34.05
4.4 46.45 26.5 30.75
8.8 44.95 27.9 25.15
14.7 43.15 28.8 18.95
20.6 39.65 29.1 14.95
Problems 69

Determine the position of the Matano interface, and calculate D(c) at 5% Zn


intervals across the couple.
20. The figure below shows a plot of composition (in atomic % A) vs. distance or
interdiffusion profile for an A–B diffusion couple which has been heated to a
temperature T for 100 hours. At this temperature the A–B system consists of the
terminal solid solutions, A in B and B in A, and two intermediate phases α(76

at % A
42) and β(25
at % A
22.5).
Construct a plot of the concentration dependence of the diffusion coefficient
over the concentration range 100> at % A >42 paying particular attention to the
behavior of the curve near phase boundaries.

21. A bar of steel 10 cm 10 cm 100 cm has an initial carbon content of


7.6 103 g/cm3. At time t ¼ 0, the surface of the bar is exposed to a
carburizing gas for which the equilibrium composition is 1.52 102 g/
cm3. Assume that the surface composition is the equilibrium value and that
the diffusion coefficient for the steel in this composition range is
2.5 107 cm2/s.
(a) Write the differential equations and initial and boundary conditions for the
diffusion of carbon into the bar. Look in sources such as Crank or Carslaw
and Jaeger for appropriate solutions to aid in answering the rest of this
problem.
70 1 Diffusion in Continuum

(b) Discuss the validity of the “semi-infinite” solution for the penetration of
carbon in this situation. At what point does this simplified treatment break
down?
(c) What is the rate of penetration of the iso-concentration line, ρ ¼ 102 g/cm3,
along the corner diagonal of the mid-plane placed through the bar perpen-
dicular to its long axis after a 10-minute exposure to the gas?
(d) Compare this rate with the rate along a line in the plane drawn perpendicular
to a mid-point on a side of the bar.
22. Referring to the Fe–C phase diagram below (Fig. 1.25), consider a semi-infinite
homogeneous bar of two phase mixture of α-Fe + Fe3C at a temperature below
723 C with an initial analytical carbon concentration co. From t ¼ 0 on, the
surface of this bar is exposed to decarburizing gas with a well-defined carbon
activity so that α-Fe with a definite carbon content cs is the only stable phase at
the surface. Then, a single-phase layer of α-Fe grows from the surface into the
two-phase region of the mean composition co. The interface, ξ, is now located
where the concentration of carbon in the surface α-phase reaches c α, the carbon
concentration in equilibrium with Fe3C at that temperature. It is assumed that
the carbon diffusivity in α-Fe, Dα, is independent of composition and that
interface reactions are fast enough so that only diffusion controls the rate of
movement of ξ.
(a) Draw the carbon concentration profile c(x,t) vs. x expected in time t,
and indicate the relevant concentrations, cs, c α , and co, and the moving
boundary ξ.
(b) Derive the equation which gives the position of the interface between single-
phase α and α + Fe3C two-phase region, ξ, as a function of t.

Fig. 1.25 Phase diagram of Fe–Fe3C hand-drawn and handwritten by C. Wagner (From the lecture
note of Wagner [16]. Courtesy of Prof. H. Schmalzried), where for the sake of clarity, the field of the
α-phase is shown somewhat larger than it is. The abscissa denotes weight percent of C

23. Again referring to Fig. 1.25, consider another semi-infinite bar between 723 and
910 C with a mean composition co within the (α + γ) field. This bar is
carburized by imposing a definite carbon concentration cs at the surface where
only γ-phase is stable. Carbon diffuses into the interior of the bar as far as the
References 71

γ-phase with a variable carbon content is present. Draw the carbon distribution
expected, and indicate the relevant carbon concentrations. The carbon contents
of α- and γ-phase in equilibrium are c α and c γ , respectively, at the present
temperature.

References

1. H. Schmalzried, Solid State Reactions, 2nd edn. (Verlag Chemie GmbH, Weinheim, 1981),
p. 59
2. A.E. Fick, Ann. Phys. Chem. 94, 59 (1855).; A.E. Fick, Philos. Mag. 10, 30 (1855)
3. J.F. Nye, Physical Properties of Crystals (Oxford University Press, Oxford, 1979)
4. K.G. Denbigh, The Thermodynamics of the Steady State (John Wiley & Sons, Inc., New York,
1965)
5. S.R. de Groot, Thermodynamics of Irreversible Processes (North Holland Publishing Co.,
Amsterdam, 1951)
6. R.E. Howard, A.B. Lidiard, Matter Transport in Solids. Rep. Prog. Phys. 27, 161–240 (1964)
7. R.P. Smith, The diffusivity of carbon in iron by the steady-state method. Acta Metall. 1,
578 (1953)
8. L.S. Darken, Formal basis of diffusion theory, in Atom Movements, (American Society for
Metals, Cleveland, 1951), pp. 1–25
9. J. Crank, The Mathematics of Diffusion, 2nd edn. (Oxford University Press, Oxford, 1975)
10. H.S. Carslaw, J.C. Jaeger, Conduction of Heat in Solids, 2nd edn. (Oxford University Press,
New York, 1959)
11. H.-I. Yoo, B.J. Wuensch, W.T. Petuskey, Oxygen self-diffusion in single-crystal MgO:
Secondary-ion mass spectrometric analysis with comparison of results from gas-solid and
solid-solid exchange. Solid State Ionics 150, 207 (2002)
12. K.-C. Lee, H.-I. Yoo, J. Phys. Chem. Solids 60, 911–927 (1999)
13. J.O’M. Bockris, A.K.N. Reddy, Modern Electrochemistry (Plenum Press, New York, 1973).
Chap. 4
14. L. Boltzmann, Wien. Ann. 53, 959 (1894)
15. C. Matano, On the relation between the diffusion-coefficients and concentrations of solid metals
(the nickel-copper system). Jap. J. Phys. 8, 109 (1933)
16. C. Wagner, Kinetics in Metallurgy (Lecture note, MIT, Cambridge, MA, 1955), p. 22
Chapter 2
Atomic Theory of Diffusion

2.1 Introduction

According to Fick’s first law, the (negative) ratio of the flux of matter to its
concentration gradient defines the diffusion coefficient, or

J
D¼ ð2:1Þ
∂C=∂x

and thus, D has the unit of (length)2(time)1 or m2/s. Why does diffusivity have such
rather unusual units? In Chap. 1, we learned that the mean square displacement or
spread is <x2 > ¼ 2Dt in one-dimensional diffusion from an instantaneous planar
source as shown in Fig. 2.1, <r2> ¼ 4Dt in two-dimensional diffusion from an
instantaneous line source, and <r2> ¼ 6Dt in three-dimensional diffusion from an
instantaneous point source. Thus,

< x2 > < r2 > < r2 >


D¼ ðin 1‐dimÞ; D ¼ ðin 2‐dimÞ; D ¼ ðin 3‐dimÞ ð2:2Þ
2t 4t 6t

This means that diffusivity is a measure of the mean square displacement for
given diffusion time t. This is why diffusivity has the unit of m2/s. It is the very
meaning of this diffusivity unit we have dug out from the mathematical analysis of
diffusion in a continuum.
In reality, however, solids comprise discrete atoms or ions, and thus, our diffu-
sion medium is not a continuum, but consists of discrete entities. What does
diffusivity with the unit of m2/s mean from this discrete, atomic point of view? In
this chapter we will explore the answer to this question. Consequently, you will
understand how diffusion occurs in an assembly of discrete atoms (or ions) and how
it varies with the thermodynamic variables of the system.

© Springer Nature Switzerland AG 2020 73


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_2
74 2 Atomic Theory of Diffusion

Fig. 2.1 D is a measure of c


the spread or mean square
displacement, e.g.,
<x2> ¼ 2Dt from an
instantaneous planar source
for a given time t. Note that
the mean square
displacement <r2> ¼ 4Dt
and ¼ 6Dt from an
instantaneous line and point
source, respectively x
– 2Dt 0 2Dt

2.2 A Naïve View of Diffusion

What is the atomistic meaning of the diffusivity unit, m2/s? Here is a very naïve but
inclusive and heuristic view. We know that a crystal comprises atomic planes with a
regular spacing. Consider two such neighboring planes, say, #1 and #2, separated by
a distance α; see Fig. 2.2. Suppose that there are n1 atoms per unit area at plane #1
and n2 at plane #2. Let us imagine that all these atoms are incessantly jumping back
and forth. Let us also assume that each atom on plane #1 successfully makes jumps
to plane #2 over the distance α by Γ12 times per second and each on plane #2 also
successfully jumps to plane #1 by Γ21 times per second. Then the number of atoms
jumping successfully from plane #1 to #2 through the unit area of the (imaginary)
reference plane located in-between per unit time, J12 will be

J12 ¼ n1 Γ12 : ð2:3aÞ

Fig. 2.2 Nearest- Atomic Reference Atomic


neighboring atomic planes plane #1 plane plane #2
#1 and #2 with a spacing α
in a crystal

Γ12

Γ21

n1 atoms n2 atoms
unit area unit area
2.2 A Naïve View of Diffusion 75

Similarly the number of atoms jumping successfully from plane #2 to #1 per unit
area per unit time, J21 is

J21 ¼ n2 Γ21 : ð2:3bÞ

The net flux of atoms in +x direction, J, is then given as the difference between
these two opposing fluxes or

J ¼ n1 Γ12  n2 Γ21 : ð2:4Þ

If the temperature and pressure are all the same at plane #1 and #2, there will have
to be no difference in jump frequency from #1 to #2 and from #2 to #1. If each atom
jumps Γ times per second in total, namely, in both +x and –x directions, then

1
Γ12 ¼ Γ21 ¼ Γ: ð2:5Þ
2

Note that if temperature and pressure are not uniform, you may expect that the
jump frequency differs depending on the jump direction.
At this point, we like to represent the number of atoms n1 and n2 in terms of the
concentration c1 and c2. A discrete quantity n to a continuous quantity c? It sounds a
bit strange or even illogical, but we may write it with a grain of salt as

n1 ¼ αc1 ; n2 ¼ αc2 ð2:6Þ

considering the volume, (unit area)  α, that surrounds each plane.


Combining all these things together, we will have

1
J ¼ Γαðc1  c2 Þ: ð2:7Þ
2

Now the “continuous” quantity c2 may be represented as

∂c
c2 ¼ c1 þ α ð2:8Þ
∂x

because the interplanar spacing α may be regarded as small enough. Thus, Eq. (2.7)
finally takes the form

1 ∂c
J ¼  Γα2 : ð2:9Þ
2 ∂x

This is exactly of the same form as Fick’s first law, isn’t it? You can, therefore,
identify the diffusion coefficient as
76 2 Atomic Theory of Diffusion

1
D ¼ Γα2 ð2:10aÞ
2

in one dimension. The numerical factor 1/2 is attributed to the fact that a half of the
total jumps back and forth Γ are directed to the diffusion direction in
one-dimensional diffusion. What about the higher dimensions, then? You bet they
will take different numerical factors according to

1
D ¼ Γα2 ðin two dimensionsÞ; ð2:10bÞ
4

1
D ¼ Γα2 ðin three dimensionsÞ: ð2:10cÞ
6

This derivation was all too simple and intuitive, but even in a more sophisticated
theory as you will see shortly, you would have the same results. You should,
however, note that these are true only for an isotropic medium. Otherwise, Γ and α
would have taken different values depending on jumping directions.
Here, we call Γ the successful (total) jump frequency, that is, the total number of
successful jumps made by an atom per unit time, and α the jump distance which is
the nearest distance between the two equivalent sites in a given crystal. The latter is
fixed once the crystal structure is given. The diffusion coefficient is, thus, a direct
measure of the successful (total) jump frequency Γ, given the crystallographic
structure of a solid. This is why D has the units of m2/s, where m2 is for α2 and s1
for Γ.
Let us take a numerical example to get a feeling about the magnitude of the
successful jump frequency Γ. Self-diffusion coefficients of many elemental metals
take values on the order of D  1012 m2/s close to their melting temperatures. If we
take the jump distance to be on the order of the lattice parameter, α  0.2~0.3 nm,
then from Eq. (2.10c), Γ  105~8 s1. Can you imagine how hectic an atom is by
jumping one hundred thousand to one hundred million times per second? This jump
frequency indicates that an atom resides at a lattice site only for 10–5~ 8 s! The
residence or lifetime of an atom at one site is extremely short, really ephemeral. But,
on the atomic scale, this successful jump is a very painstaking endeavor. The typical
value for the natural vibration frequency of an atom in a crystalline solid is on the
order of magnitude of the Debye or Einstein frequency, 1013~14 s1. The atom can,
thus, make only one jump successful out of 109 ~ 105 attempts, and the atom resides
at its site for most of its lifetime.
This naïve, simple picture of atomic diffusion gives you a feeling of how
diffusion proceeds in a discrete crystalline lattice. We will now develop a more
sophisticated theory of diffusion.
2.3 Random Walk Theory 77

2.3 Random Walk Theory

Suppose that a particle, be it a dust particle (executing Brownian motion) in air or an


atom in a crystal, jumps “n” times for a time duration t, each jump in a completely
random direction. As the jump direction is completely random, one can never know
its displacement, that is, where the particle will finally be found after n jumps. But
could we estimate how far the particle will be away from the starting point, on
average, after n-jumps? It sounds almost hopeless because all jumps are random.
But, there is an ingenious as well as amusing way to do it, which is called the
random walk theory invented by A. Einstein [1] in 1905.
Let us assume that a particle jumps n times absolutely at random from the origin
O with each jump distance ri (i ¼ 1,2,. . .n); see Fig. 2.3.
!
Then, the displacement vector Rn of the particle after n jumps is simply a vectorial
sum of all the jump vectors or

* X n
Rn ¼ *ri : ð2:11Þ
i¼1

As the direction of each jump is absolutely random, you never know the direction
!
of the displacement vector Rn , but its magnitude Rn may be calculated by making
an inner product with itself:
! !
* * X
n
!
X
n
!
R2n ¼ Rn  Rn ¼ ri  ri
i¼1 i¼1
! ! ! ! ! !
¼ r 1  r 1 þ r 1  r 2 þ        þ r 1  r n ð2:12aÞ
! ! ! ! ! !
þ r 2  r 1 þ r 2  r 2 þ        þ r 2  r n
þ                          
! ! ! ! ! !
þ r n  r 1 þ r n  r 2 þ        þ r n  r n

or

X
n n1 X
X ni
R2n ¼ r2i þ 2 ri *
* riþj : ð2:12bÞ
i i¼1 j¼1

Fig. 2.3 A particle executes r


random jumps rn
! ! !
r 1 , r 2 ,   , r n, resulting in a r
!
displacement Rn Rn
r r
O r1 r r3
r2
78 2 Atomic Theory of Diffusion

The first term on the right-hand side of Eq. (2.12b) is the sum of all the diagonal
terms and the second the sum of all the off-diagonal terms in Eq. (2.12a) by noting
! ! ! !
that r i  r j ¼ r j  r i . But, aren’t the indices on the double sum in the second term
not so transparent? If so, try to write down the terms, and you will immediately
recognize why j is running from 1 to ni and i from 1 to n1.
In general, the magnitudes of jump distances are not necessarily the same as each
other as, e.g., in a system of dust particles executing Brownian motion. Let us here
assume that the jump distances are all identical in magnitude as in an isotropic
crystalline solid, or
 
! 
 r i  ¼ ri ¼ α: ð2:13Þ

Then, Eq. (2.12b) becomes

X
n1 X
ni
R2n ¼ nα2 þ 2α2 cos θi,iþj ð2:14Þ
i¼1 j¼1

! !
where θi,i+j denotes the angle between the ith jump r i and (i+j)th jump r iþj . What
will then be the average of R2n or the mean square displacement < R2n >? We saw
above that the successful jump frequency is on the order of 108 per second which is
itself already a large number. We may, thus, take the total number of jumps made by
an atom even in a finite time to be infinite. Let us calculate the average as n
approaches practically to infinity as
" * +#
2 X
n1 X
ni
< R2n >¼ nα2 1 þ lim cos θi,iþj : ð2:15Þ
n n!1 i¼1 j¼1

Alas, the second term within the brackets looks hopelessly complicated to
evaluate. But the absolute randomness saves us! Can you agree that if all the atomic
jumps are absolutely random or independent of each other or in other words, a jump
does not remember any earlier jump or the ith jump and the (i+j)th jump are
uncorrelated, then a positive value of cosθi, i+j is equally probable as its negative
value? Therefore, to one’s surprise,
* + * + * +
X
n1 X
ni X
n1 X
ni X
1
lim cos θi,iþj ¼ lim cos θi,iþj ¼n cos θj
n!1 n!1
i¼1 j¼1 i¼1 j¼1 j¼1

X
1  
¼n cos θj ¼ 0: ð2:16Þ
j¼1

Here, θj is the angle between any preceding jump vector (e.g., the ith) and the jth
jump vector therefrom (e.g., (i+j)th). It is emphasized that it is solely because the
2.3 Random Walk Theory 79

atomic jumps are absolutely random or uncorrelated. Otherwise, it would not be zero
as we will see shortly.
For random walk, thus, the mean square displacement after n jumps is

< R2n >¼ nα2 ð2:17Þ

or the root mean square displacement is


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi
< R2n > ¼ nα: ð2:18Þ

In order to get a physical feeling of the root mean square displacement or


spread after n random jumps, let us here compare it with the total travel distance of
a specific atom. Suppose that an atom has jumped n ¼ 108 times for 1 second with
the jump distance of, say, α ¼ 0.2 nm. Then, the root mean square displacement
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffi  
< R2 > ¼ 108  2  1010 m ¼ 2 μm:

The total distance travelled ℓ, on the other hand, has been


 
ℓ ¼ nα ¼ 108  2  1010 m ¼ 2 cm:

It is like a total travel distance of as far as 1 km only results in a net displacement


or spread of 10 cm!
An amusing story: Suppose that you are totally drunken so that each step you
take is absolutely random (otherwise, you would not be totally drunken!). Now you
can calculate on average how far you would be away, after making n random walks
each with, say, 50 cm step distance, from the lamp post where you started the
drunkards’ walk. The following famous cartoon was drawn by G. Gamow [2]
himself, and the random walk is often called the “drunkards’ walk” after him
(Fig. 2.4).
So what? We learned in Chap. 1 that the mean square displacement is related to
the diffusion coefficient as

< R2 >¼ 6Dt ð2:19Þ

in a three-dimensional isotropic space. By combining with Eq. (2.17), we get

1n 2
D¼ α : ð2:20Þ
6 t

If we let the successful jump frequency Γ be


80 2 Atomic Theory of Diffusion

Fig. 2.4 Drunkards’ walk (left) and their statistical distribution from the starting point, the lamp
post (right). (From G. Gamow [2])

n
Γ¼ , ð2:21Þ
t

then,

1
D ¼ Γα2 : ð2:22Þ
6

Diffusivity is now measuring the successful jump frequency Γ/s1 for a given
structure with the fixed jump distance squared α2/m2. This is indeed the atomistic
meaning of diffusivity, a measure of the successful jump frequency (s1), and why
it has the unit of m2/s.
We now understand the unit of diffusivity, but how does diffusion takes place?
Let us consider diffusion mechanisms or vehicles by which atoms move around.

2.4 Diffusion Mechanisms

Let us first consider an ideal crystal structure of an elemental metal A. The ideal
crystal is an infinite lattice with each lattice point occupied by what there should be,
an atom, and may look like Fig. 2.5. You may then immediately envisage that an
atom may move around in the lattice by directly exchanging its positions with a
nearest-neighboring atom; see Fig. 2.5a. Yes, it should be possible in principle, but it
appears to be not so likely because, for the two nearest-neighboring atoms to
exchange their positions, they should jump simultaneously, but along different
paths to avoid a head-on collision. How could they synchronize their jumping
with each other? It would be extremely improbable as well as energetically
2.4 Diffusion Mechanisms 81

Fig. 2.5 Jump mechanisms


in an ideal crystal lattice
without point defects in two
dimensions. (a) Direct
exchange mechanism and
(b)
(b) ring mechanism for
diffusion
(a)

expensive because two atoms have to simultaneously escape their own minimum
binding energy traps. Before the presence of point defects in a crystalline lattice was
realized, people first conjectured this would be the possible mechanism of diffusion:
it is called the direct exchange mechanism.
People then invented a more sophisticated mechanism, a ring mechanism, where
instead of exchanging the sites directly with a nearest-neighboring atom, a ring of
atoms at equal jump distance apart may collectively jump around, say, clockwise;
see Fig. 2.5b. But again, synchronizing the jumps of a gang of atoms is even more
unlikely and more expensive energetically.
Today we know that point defects such as vacancies and interstitials are
thermodynamically stable.1 Now suppose that there is a vacancy (□) next to an
atom (•); see Fig. 2.6a. The jump of an atom into the vacant site will be obviously a
lot easier and much cheaper energetically than direct exchange. Diffusion by
exchanging sites with neighboring vacancies is referred to as a vacancy mechanism.
Likewise, an interstitial atom may easily diffuse by jumping into a neighboring
empty interstitial site. The latter is called an interstitial mechanism (Fig. 2.6b). An
interstitial atom may also jump to a nearest lattice site by kicking off the atom there

Fig. 2.6 Principal


elementary diffusion
mechanisms in real crystals:
(a) vacancy mechanism, (c)
(b) interstitial mechanism, (a)
(c) collinear interstitialcy
mechanism, (d) curvilinear
interstitialcy mechanism (d)

(b)

1
For equilibrium defect structures, see Appendix I at the end of this book.
82 2 Atomic Theory of Diffusion

into a neighboring vacant interstitial site. This mechanism is called an interstitialcy


mechanism. The direction of an interstitial jump to the lattice site and that of the
lattice atom into the interstitial site may be collinear (Fig. 2.6c) or curvilinear
(Fig. 2.6d), thus called a collinear interstitialcy mechanism and a curvilinear
interstitialcy mechanism, respectively. Even more complicated mechanisms, e.g.,
diffusion by a crowd of atoms or a crowdion mechanism, may be possible, but not
so likely.

2.5 A Few General Relations

All the likely diffusion mechanisms are basically based on vacancies or interstitials
as vehicles for atoms. For the simple vacancy mechanism in a pure metal A
(Fig. 2.6a), it is straightforward to deduce that the flux of A-atoms, JA, should be
counterbalanced by the flux of vacancies (V), JV, in counter directions or

JA þ JV ¼ 0: ð2:23Þ

This constraint indicates that the density of the lattice sites (A-sites and V-sites)
remains invariant while A self-diffusion proceeds via the vacancy mechanism.
Referring to Fig. 2.6a, Eq. (2.23) means (as JA ¼ |JV|)

nA Γ A ¼ nV Γ V ð2:24aÞ

or due to Eq. (2.6),

cA Γ A ¼ cV Γ V ð2:24bÞ

where cA and cV are the concentrations of A-atoms and A-vacancies in metal A,


respectively. The successful jump frequencies ΓA and ΓV are measured by the
corresponding diffusivities, self-diffusivity DA and defect (vacancy) diffusivity
DV as Eq. (2.10). Equation (2.24b) may, thus, be rewritten equivalently as

cA DA ¼ cV DV : ð2:24cÞ

Even though there is no concentration gradient of A, A jumps around by means of


vacancies.
For the interstitial mechanism, on the other hand, any flux of A-lattice atoms as
observed from outside has actually been due to the same amount of A-interstitials
(I) or

JA ¼ JI ð2:25Þ

which is nothing but a mass conservation constraint. It thus follows, as in the


vacancy mechanism, that
2.6 Derivation of Self-Diffusion Coefficient 83

cA Γ A ¼ cI Γ I ð2:26aÞ

or equivalently,

cA D A ¼ cI D I ð2:26bÞ

where DI is referred to as the defect(interstitial) diffusivity.


It is interesting to observe from Eqs. (2.24b) and (2.26a) that

cA
Γd ¼ Γ ðd ¼ V, IÞ ð2:27Þ
cd A

The equilibrium defect fraction, cd/cA, is normally very much smaller than
1, typically cd/cA  104 for most elemental metals. It means that in order to
make an A-atom jump once successfully, its vehicle, defect d, has to jump at least
104 times more! This is simply because the number of vehicles (d) is too small
compared to the number of passengers (A); thus the vehicles are that much busy.
Finally, what would self-diffusivity be like when both types of defects, vacancies
and interstitials, are operating simultaneously? As the defect concentrations are
normally very small compared to the lattice atom concentration, it is easy to imagine
that both contributions may be simply superposed or
X
cA D A ¼ cd Dd ¼ cV DV þ cI DI ð2:28Þ
d

where “d” denotes all kinds of defects serving as the vehicles for A-atoms and Dd is
generically called the defect (d) diffusivity. This relation serves as the very link
between the self-diffusivity and the equilibrium defect structure2 of a given crys-
talline solid, which allows you to predict the atomic diffusivity or mobility from the
latter.

2.6 Derivation of Self-Diffusion Coefficient

Let us consider a pure elemental metal A at uniform temperature T and pressure


P. Even when there is no concentration gradient of A, A-atoms incessantly jump
around in their vehicles, defects, be they vacancies or interstitials, as described by
Eqs. (2.24) and (2.26) or Eq. (2.28). The self-diffusion coefficient, DA, is a measure
of such a successful jump frequency ΓA in the absence of its own concentration
gradient.

2
For a concise treatment of equilibrium defect structures, see Appendix I.
84 2 Atomic Theory of Diffusion

How could we measure ΓA or DA? The simplest way will be to measure the mean
square displacement <R2> for a given time duration t according to Eq. (2.19). In
this case, one should be able to trace a specific atom to measure its square displace-
ment R2 in time t and repeat the procedures with sufficiently many such atoms to get
the mean value <R2> for the square displacement; see Fig. 2.4(right). But, A atoms
in A are all alike so that one can never distinguish one from another, and hence, one
can never trace the jumping trajectory of a specific atom either. Therefore, one needs,
for example, to label or mark each such atom to be able to trace their trajectories. A
typical technique is to put radioactive or even stable isotopes of A-atoms as their
tracers and then measure their mean square displacement with the aid of artificial
eyes, e.g., radioactivity counters for the former or mass spectrometers for the
latter. The diffusivity as measured from the mean square displacement of such
tracers is especially referred to as the tracer diffusivity of A to distinguish from
the self-diffusivity.
Assuming that these tracers jump around absolutely at random like the drunkards
in Fig. 2.4, let us derive the expression for the self-tracer-diffusivity by the vacancy
mechanism by using Eq. (2.22). We may take a cubic-close-packed or face-
centered cubic (fcc) crystal, e.g., Au; see Fig. 2.7.

Fig. 2.7 An fcc crystal


lattice with a tracer at the
center. ao, lattice parameter;
α, tracer jump distance

a0
α

tracer

First, the jump distance α: This should be the distance between the two nearest-
neighboring equivalent sites in the given structure. Letting ao be the lattice constant
of the present fcc lattice,

a
α ¼ poffiffiffi : ð2:29Þ
2

Next, the successful jump frequency Γ: For an atom to jump over a distance α to
a nearest equivalent site, the atoms in the neighborhood of the jumping path in
between should first open their own window wide enough so that the jumping atom
may pass through, which requires work, thus making a Gibbs energy barrier for
migration, Δgm, between the two equivalent sites, or Gibbs free energy of migra-
tion, at given T and P, see Fig. 2.8.
2.6 Derivation of Self-Diffusion Coefficient 85

Fig. 2.8 Atomic window


and energy landscape along
the jump from an
equilibrium site (a) through

Energy
the atomic window (b) to a (a) (b) (c)
nearest equivalent site (c).
(From Shewmon [3])

∆g m

(a) (b) (c)

An atom sitting at a site is incessantly vibrating at a total frequency νo with a


spectrum of energy values (ε), and only a fraction of νo with an energy no smaller
than Δgm, ν(Δgm), can go over the energy barrier (b) in Fig. 2.8. Such jump
frequency may be called the attempt frequency because it is energetically qualified
to attempt the nearest-neighboring equivalent site (c) by satisfying the necessary
condition, ε  Δgm. Assuming the Boltzmann distribution of vibration frequen-
cies against ε and letting kB denote the Boltzmann constant, we may write the
attempt frequency fraction as
R 1 ε=k T
νð Δgm Þ Δg e dε
B

¼ R 1m ε=k T ð2:30aÞ
νo 0 e
B dε

or

Δgm
ν ¼ νo exp  : ð2:30bÞ
kB T

Here, the total vibration frequency νo may be taken as the Einstein (νE) or
Debye frequency (νD) or the thermal frequency (kBT/h):

kB T
νo  νE  νD  : ð2:31Þ
h

Now, the sufficient condition for an atom to successfully jump to a next site is that
the very next site should be vacant; otherwise any attempt to this site would turn
futile. Letting Pv be the probability of a site being vacant, the successful jump
frequency to one of the nearest-neighboring sites is νPv. There are 12 nearest-
86 2 Atomic Theory of Diffusion

neighboring sites (coordination number) for an fcc lattice; the total successful
jump frequency of the tracer will then be

Γ ¼ 12νPv : ð2:32Þ

By combining Eqs. (2.29) and (2.32) according to Eq. (2.22), we obtain the self-
diffusivity as

DA ¼ a2o νPv : ð2:33Þ

If the vacancy distribution is random, then Pv should be equal to the vacancy


fraction Nv. If local defect equilibrium prevails or the defect concentration is
uniquely determined by the local thermodynamic variables, T and P, in the present
single-component case,

ΔgV
PV ¼ NV ¼ exp  ð2:34Þ
kB T

where ΔgV denotes the vacancy formation free energy that is itself a function of T
and P.
By substituting Eqs. (2.34) and (2.30b) into Eq. (2.33), we finally obtain
 
Δg Δg
DA ¼ a2o νNV ¼ a2o νo exp  m exp  v : ð2:35Þ
kB T kB T

This result is specifically for an fcc lattice. In general, the self-diffusivity may be
written as

DA ¼ γa2o νNV ð2:36Þ

in which “ao” may be taken as a representative lattice parameter and “γ” the
geometric factor specific to the given structure, which nevertheless takes a value
on the order of 1.

2.7 Defect (Vacancy) Diffusion Coefficient

Let us now trace a vacancy instead of the tracer. Its jump distance is obviously the
same as the jumping tracer, Eq. (2.29), and the attempt frequency, the necessity for
a successful jump, the same as that of the jumping atom, Eq. (2.30), because a
vacancy jump is equal to an atomic jump in counter directions due to Eq. (2.23).
What about the sufficiency for the jump? For the tracer, the sufficiency was for the
next equivalent site to be vacant. Likewise, for the vacancy, the sufficiency is for the
2.8 Thermodynamic Variables Dependence of DA 87

next site to be vacant from the vacancy point of view or to be occupied by an atom.
Thus, the successful jump frequency for a vacancy ΓV should be

ΓV ¼ 12νð1  PV Þ ¼ 12νð1  NV Þ  12ν ð2:37Þ

as NV << 1. Then, due to Eq. (2.22), the vacancy diffusivity DV takes the form

DV ¼ a2o ν: ð2:38Þ

By comparing with the self-diffusivity, Eq. (2.33) or Eq. (2.35), we can imme-
diately recognize that

DA ¼ DV NV : ð2:39Þ

This relationship is nothing but the local lattice conservation constraint,


Eq. (2.24c) as PV ¼ NV ¼ cV/(cA + cV)  cV/cA because cv << cA.

2.8 Thermodynamic Variables Dependence of DA

The self-diffusion coefficient D is often represented in the Arrhenian form as



Qa
D ¼ Do exp  ð2:40Þ
kB T

where Do is called the pre-exponential factor and Qa the activation energy. When
lnD is plotted against 1/T, we normally obtain a straight line with a negative slope,
which corresponds to –Qa/kB, as exampled in Fig. 2.9.
Let us examine Do and Qa by comparing with Eq. (2.35). Noting that
Δg ¼ Δh  TΔs, we may rewrite Eq. (2.35) as
 
Δsm þ ΔsV Δhm þ ΔhV
DA ¼ a2o νo exp exp  : ð2:41Þ
kB kB T

We can then immediately find that



Δsm þ ΔsV
Do ¼ a2o νo exp ; Qa ¼ Δhm þ ΔhV : ð2:42Þ
kB

The defect concentration may often be fixed or frozen-in by doping, e.g.,


aliovalent impurities for ionic compounds,3 or by quenching the crystal to low

3
For the defect structure of ionic solids, see Appendix I at the end of this book.
88 2 Atomic Theory of Diffusion

Fig. 2.9 An example of


Arrhenian behavior of the
self-diffusivity, DAu vs. 1/T
for Au, as measured with the
tracer (Au198) method.
(From Okkerse [4])

temperatures where the internal defect equilibrium is frozen-in. In such a case, the
defect concentration is fixed at PV, and therefore, Eq. (2.33) should be written as
 
Δsm Δhm
DA ¼ a2o PV νo exp exp  ð2:43Þ
kB kB T

and hence,

Δsm
Do ¼ a2o νo PV exp ; Qa ¼ Δhm : ð2:44Þ
kB

You should be aware that depending on whether the defect concentration is fixed
or varied with the thermodynamic variables, the information contents of Do and Qa
become quite different.
The self-diffusion coefficient is, in general, a function of the thermodynamic
variables of the system given, T and P, in the present case via Eq. (2.35) or (2.36).
Let us see how D varies with these variables.
First, the temperature dependence: By differentiating Eq. (2.35) with respect to
1/T, we obtain, due to the Gibbs-Helmholtz equation,
2.8 Thermodynamic Variables Dependence of DA 89


∂ ln D 1 ∂½ðΔgm þ ΔgV Þ=T Δh þ ΔhV Q
¼  ¼ m ¼ a: ð2:45Þ
∂ð1=TÞ P
kB ∂ð1=TÞ kB kB

That is, the activation energy Qa comes out. Depending on whether the defect
concentration is variable or fixed, of course, the defect formation enthalpy ΔhV
makes a contribution to Qa or not. In any case, Qa > 0 as Δhm > 0 and ΔhV > 0; thus,
D increases exponentially as T increases: it is thus called a thermally activated
process. Table 2.1 lists the experimental values for these enthalpies for some
selected elemental metals, from which you can get a feeling for their magnitudes.
Second, the pressure dependence: Differentiation of Eq. (2.35) with respect to P
leads, due to the thermodynamic identity (∂g/∂P)T ¼ v, to

∂ ln D 1 ∂ðΔgm þ ΔgV Þ Δv þ ΔvV V
¼  ¼ m ¼ a : ð2:46Þ
∂P T k B T ∂P k B T k BT

Here, Δvm corresponds to the volume change upon a jumping atom passing
through the intermediate activated state (b) in Fig. 2.8, thus called the volume
change of migration or migration volume, and Δvv the volume change upon
generating a vacancy or the volume of vacancy formation. The sum of these two
volume changes is called the volume of activation or the activation volume Va to
keep the parallelism with the energy of activation or activation energy Qa.

Table 2.1 Experimental Point defect energies/eV


values for the enthalpies of
Metal Δhva Δhmb
migration (Δhm) and vacancy
formation (Δhv) for some bcc
selected elemental metals Mo 3.0 1.50
Nb 2.65 1.00
Ta 2.8 1.45
V 2.1 1.30
W 4.0 1.5
fcc
Al 0.68 0.68
Ag 1.12 0.60
Au 0.89 0.84
Cu 1.29 0.78
Fe 1.4 1.26
Ni 1.78 1.32
Pb 0.57 0.59
Pd 1.85 0.91
Pt 1.32 1.37
From Schaefer [5]
a
Δhv from positron annihilation
b
Δhm ¼ Qa – Δhv
90 2 Atomic Theory of Diffusion

Table 2.2 The as-measured Lattice Metal Va/Ωa


values for activation volume
bcc K 0.55
(relative to the atomic volume)
for selected elemental metals Na 0.41
Li 0.26
hcp Cd 0.63
Zn 0.65
Mg 0.77
fcc Pb 0.80
Au 0.85
Pt 1.09
Ni 0.80
From Brown and Ashby [6]
a
Ω: molar (or atomic) volume of the metal

Imagining things happening when an atom is passing through the atomic window
along its jumping path (see Fig. 2.8), we may roughly guess that these volume
changes will be no larger than the atomic volume, Ω. If the gas constant R is used
instead of the Boltzmann constant kB, then the comparison is to the molar volume
of the metal A, Vm(¼ NAΩ). In reality, indeed

Δvv , Δvm < Ω: ð2:47Þ

The rule of thumb is that Δvv  0.5 Ω and Δvm  0.15 Ω for most elemental
metals. It is because once a vacancy is formed in an ideal crystal, the created volume
should be equal to its atomic volume, but the lattice relaxation that happens after a
vacancy is generated makes it normally smaller than the atomic volume. Table 2.2
lists the as-measured values for some selected elemental metals.
Normally, Δvm > 0 and Δvv > 0; thus Va > 0. The self-diffusivity, therefore,
should decrease exponentially with increasing P. An example is shown in Fig. 2.10.

2.9 Correlation Effect

In the random walk theory, we assumed Eq. (2.16). This means that the jth-jump
following any preceding jump happens in all possible directions with equal proba-
bility 1/z (Fig. 2.11), where z being the coordination number, because the jumps
are absolutely random or uncorrelated, thus,

< cos θj >¼ 0: ð2:48Þ

Can jumps really be absolutely random, or uncorrelated, for the tracer diffusion?
Suppose that a tracer has just exchanged its position with the neighboring vacancy;
see Fig. 2.12a. Remember that vacancies are really scarce, on the order of only one
2.9 Correlation Effect 91

Fig. 2.10 Pressure


dependence of the self-
diffusivity of Ge at various
temperatures, as measured
with the tracer (Ge71)
method. (From Werner
et al. [7])

Prob.

1/z

0 θj π

Fig. 2.11 Random walk theory assumes that the next jump of an atom takes place in any direction
(viz., to any nearest-neighboring site) with the equal probability, 1/z, where z is the coordination
number

Fig. 2.12 (a) The tracer (●)


Prob.
has just exchanged its
position with the vacancy 1/z
(□). (b) The most probable
jump of this tracer is to go
back to its original position
(θ ¼ π) with the probability
1/z because vacancy jumps 0 θj π
are absolutely random
(a) (b)
92 2 Atomic Theory of Diffusion

out of 10,000 at maximum in, e.g., metals. It is then obvious that the atom that has
just jumped to the present site is much more likely to jump back to the previous
position because the vacancy is now located where the atom was before the jump.
Namely, the probability of jumping back (θ ¼ π) or cancelling the previous jump
will be the highest of all possible jumps. If the coordination number is z, then this
maximum probability will be 1/z. If we now consider other directions, the probabil-
ity of the next jump should vary with θ schematically as shown in Fig. 2.12b, thus,

< cos θj >< 0: ð2:49Þ

Why does this kind of thing happen to the tracer? It is simply because the
surrounding of the tracer has been changed with a jump: the vacancy used to be to
the right of the tracer before jumping, but now to the left! That is why the tracer
jumps cannot be random, but are instead correlated upon the ride of vacancies.
Let us now see this situation from the vacancy’s point of view. Because it is so
scarce, its surrounding remains essentially the same before and after any jump:
When a vacancy jumps from right to left, its surrounding used to be occupied fully
by atoms still remains fully occupied by them; see Fig. 2.12a. The vacancy will thus
jump to any of the nearest sites with the equal probability, 1/z. Vacancies are
jumping around really at random or uncorrelatedly. Likewise, it is also expected
that for the interstitial mechanism, there should be no correlation either because the
surrounding before and after a jump remains the same, viz., the nearest-neighboring
interstices around a jumping interstitial remain essentially always vacant.
So what? Such correlated jumps yield an interesting consequence. If jumps are
correlated in this way, the spread or root mean square displacement of the tracers
turns out to be smaller than otherwise because the correlation effects the cancellation
of successive jumps by tending to make a tracer jump back to its immediately
previous site; see Fig. 2.13.

c
D*(actually observed)

0 * x
x2 x2

Fig. 2.13 Comparison of the spreads when uncorrelated, <x2> ¼ 2Dt, and when correlated,
<x2> ¼ 2D t
2.10 Quantitative Treatment of the Correlation Factor 93

If uncorrelated, the mean square displacement would have to be


 
R2n ¼ nα2 ¼ 6Dt: ð2:50Þ

In the tracer diffusion, however, the mean square displacement turns out to be
smaller due to the jump cancelling effect or correlation effect (n < n):
 2 
Rn ¼ n α2 ¼ 6D t: ð2:51Þ

This effective diffusion coefficient D as measured by a tracer experiment is, thus,


smaller than the self-diffusion coefficient by as much as the jump cancelling effect
or
 2 
D R n α2 Γ
¼  n2  ¼ ¼ ¼ f < 1: ð2:52Þ
D Rn nα2 Γ

This as-measured “self-diffusivity” D is, thus, specifically called the “tracer


diffusivity” in order to distinguish from the “self-diffusivity” D, and the ratio f(<1)
is called the correlation factor.

2.10 Quantitative Treatment of the Correlation Factor

Let us try to evaluate this correlation factor f crudely first and then more elegantly.
Consider a two-dimensional hexagonal lattice, Fig. 2.14, for example.
The probability for the tracer atom to jump back to its original position, now the
vacant site 1, is 1/z(¼ 1/6) because it is equal to the probability of the vacancy
jumping back to the tracer site now and the vacancy jump is random. Once the tracer
jumps back, it effectively cancels the two consecutive previous jumps and hence,

Fig. 2.14 A
two-dimensional hexagonal
lattice where the tracer at the 2 3
site 7 has just exchanged its
positions with the vacancy
1 7 4
now sitting at the site 1

6 5
94 2 Atomic Theory of Diffusion

n 2
f¼ 1 ð2:53Þ
n z

which gives the numerical values f ¼ 2/3 for the present case of the coordination
number z ¼ 6.
The correlation factor is more elegantly, but laboriously, calculated in the
following way. The definition of the correlation factor, Eq. (2.52), in association
with Eq. (2.15) and Eq. (2.16), yields
 2 
R X1
f
 n2  ¼ 1 þ 2 < cos θj > : ð2:54Þ
Rn j¼1

The problem, thus, converges to calculate the mean value of cos θj. Can you
imagine that as j increases, the ith and (i+j)th jumps are getting less and less
correlated, and hence, <cos θj> will have to converge to zero as j!1? The series
in Eq. (2.54) will then be dominated by the first few terms. Let us calculate this
leading term <cos θ1> for the hexagonal lattice in Fig. 2.14.
Look at Fig. 2.14! The tracer sitting now at the site 7, which has just landed from
the site 1, can make its next jump (j ¼ 1) to each of the six nearest-neighboring sites
numbered k ¼ 1(θ1 ¼ π), 2(θ2 ¼ 2π/3), 3(θ3 ¼ π/3), 4(θ4 ¼ 0), 5(θ5 ¼ π/3), and 6
(θ6 ¼ 2π/3), of course, with different probabilities Pk because the degree of vehicle
or vacancy availability is different depending on the site k. The mean value may then
be written as

X
6
< cos θ1 >¼ Pk cos θk
k¼1



2π π ð2:55aÞ
¼ P1 cos ðπÞ þ P2 cos þ P3 cos þ P4 cos ð0Þ
3 3



π 2π
þ P5 cos þ P6 cos :
3 3

Due to the symmetry of the given lattice, P2 ¼ P6 and P3 ¼ P5. Thus, Eq. (2.55a)
is rewritten as



1 1
< cos θ1 >¼ P1 ð1Þ þ 2P2  þ 2P3 þ P4 ð1Þ: ð2:55bÞ
2 2

We know that vacancy jumps are uncorrelated, and hence, its jumping probability
to any nearest-neighboring site is equally 1/z or 1/6 for the present. The vacancy at
the site 1 may make the tracer at the site 7 jump to the site k by jumping itself once
(n ¼ 1), twice (n ¼ 2), thrice (n ¼ 3), etc. Then,
2.10 Quantitative Treatment of the Correlation Factor 95

X
n
1
Pk ¼ νkn ð2:56Þ
n¼1
6

where νkn is the number of possible paths which allow the vacancy, now at the site
1, to reach the site k at the (n1)th jump. Carried out to the 4th jump (n ¼ 4) of the
vacancy, we obtain



2
3
4
1 1 1 1
P1 ¼ 1 þ0 þ5 þ8 ¼ 0:1960
6 6 6 6


2
3
4
1 1 1 1
P2 ¼ P6 ¼ 0 þ1 þ1 þ 11 ¼ 0:0409
6 6 6 6


2
3
4
1 1 1 1
P3 ¼ P5 ¼ 0 þ0 þ1 þ2 ¼ 0:0062
6 6 6 6


2
3
4
1 1 1 1
P4 ¼ 0 þ0 þ0 þ2 ¼ 0:0015:
6 6 6 6

You see, P1, the probability of jumping back, is nearly 200 times larger than P4,
the probability of jumping forth! Putting back these numbers into Eq. (2.55b), we
finally have

< cos θ1 >¼ 0:2262: ð2:57Þ

What about the higher-order terms, say, <cos θ2>, <cos θ3>, . . . .. in Eq. (2.54)?
One may evaluate them similarly, but it must be, no doubt, extremely laborious. To
our surprise, Compaan and Haven [8] have shown that

< cos θj >¼< cos θ1 >j ð2:58Þ

for a vacancy mechanism. Therefore Eq. (2.54) may be rewritten as

< cos θ1 > 1þ < cos θ1 >


f ¼1þ2 ¼ : ð2:59Þ
1 < cos θ1 > 1 < cos θ1 >

By using the value in Eq. (2.57), we can, thus, get the value f ¼ 0.6311 for the
hexagonal lattice in Fig. 2.14. The correlation factors for some selected lattices, as
calculated by Compaan and Haven, are listed in Table 2.3 in comparison with their
crude estimates due to Eq. (2.53).
96 2 Atomic Theory of Diffusion

Table 2.3 Correlation factors Lattice type f z 1–2/z


for diffusion via free
2-dimensional
vacancies from Compaan and
Haven(f), in comparison with Square 0.46705 4 0.5000
the rough estimates (1–2/z), Hexagonal 0.56006 6 0.6667
where z is the coordination 3-dimensional
number of a given lattice type Diamond 1/2 4 0.5000
Simple cubic 0.65549 6 0.6667
Body-centered cubic 0.72149 8 0.7500
Face-centered cubic 0.78146 12 0.8333
Hexagonal close packing 0.78146 12 0.8333

2.11 Various Diffusivities So Far

On the basis of the random walk theory, we have learned so far the three kinds of
diffusivities, the self-diffusivity D, the tracer diffusivity D , and the defect diffu-
sivities DV and DI. Particularly for the vacancy mechanism,

D
D ¼ NV DV ¼ : ð2:60Þ
f

Diffusivities are often classified by the locality where diffusion occurs as:
Locality Diffusivity
Surface: Surface diffusion coefficient
Grain boundary: Grain boundary diffusion coefficient
Dislocation pipe: Dislocation pipe diffusion coefficient
Volume or lattice: Volume or lattice diffusion coefficient
What we learned in this chapter, D, Dv, DI, and D are obviously the volume or
lattice diffusivities.
Finally, wherever diffusion occurs, its atomic picture remains the same as in the
random walk theory (Fig. 2.4).

Problems

1. Derive an equation for the D of an interstitial solute in a binary alloy with the
cubic lattice parameter ao for the situation in which

ao ∂Γ ∂c
Γ12  Γ21 ¼   6¼ 0:
2 ∂c ∂x
Problems 97

2. One hundred jumping beans are placed along the center line of a gymnasium floor
at 6 cm intervals. Twelve hours later the distance of each from the line is
measured, and the sum of the squares of the distance divided by 100 is 36 cm2.
(a) Calculate the diffusion coefficient of the jumping beans.
(b) If the mean jump (or roll) distance of a bean is equal to 0.1 cm, estimate the
mean jump (or roll) frequency of a bean.
3. A jumping particle makes a series of n jumps each of length L.
(a) From your knowledge of the random walk problem, write a general form of
the relation between n, L, and the mean distance moved R2.
(b) In three totally different experiments, it is found that: in one case R2 ¼ nL2, in
a second R2 ¼ 0 though n >> 0 and L > 0, and in a third nL2 < R2 < n2L2.
Explain the different relationships that must exist between the successive
jump directions for each of the three cases.
4. If at t ¼ 0, a quantity of solute is located along an infinite line r ¼ 0 in a three-
dimensional medium, the concentration of solute at any point r from the origin,
after time t, is


γ r2
cðr, tÞ ¼ exp  :
t 4Dt

(a) Give the probability (normalized to 1) of finding an atom in a cylindrical shell


dr thick at r from the origin.
(b) What is the mean square value of r, that is, <r2>, for the solute after time t?
(c) Using the results of part (b) and the random walk equation < r2n >¼ nα2 ,

show that

D ¼ Γα2 =4

where Γ ¼ n/t.
5. (a) Calculate the geometric factor γ for a tracer in a pure bcc metal where γ is
defined by the equation

D ¼ γa2o νNv :

(b) Calculate γ for an interstitial solute in a dilute bcc binary alloy.


6. (a) Empirically it is found that D ¼ Doe-Q/RT. Express Do and Q in terms of
fundamental quantities for tracer diffusion by a vacancy mechanism and for
diffusion of an interstitial solute in dilute concentrations.
98 2 Atomic Theory of Diffusion

(b) The D for C in Fe is given by


2 1 20, 000 cal=mol
D=cm s ¼ 0:008 exp  in α  Fe ðbccÞ
RT

and

33, 800 cal=mol
D=cm2 s1 ¼ 0:21 exp  in γ  Fe ðfccÞ:
RT

Calculate D at 800 C and explain the differences in the two expressions


and the magnitudes in terms of the quantities discussed in (a).
7. (a) In H2 at 1 atm and 25 C, the average molecular velocity is 1.3  104 cm/s
and the mean free path is 1.9  105 cm. Calculate the diffusion coefficient
of the gas. (Take the average velocity to be the same as the root-mean-square
velocity).
(b) Calculate the ratio of the equilibrium atom fractions of interstitial atoms and
vacancies in copper at 1000 C. Take the formation enthalpy ΔH to form
copper interstitial atoms to be 210 kcal/mol and ΔH to form vacancies as
30 kcal/mol. Assume that the entropy of formation is the same for both
defects.
(c) In the temperature range 70 to 400 C, the diffusion coefficient for carbon
in α-Fe is D ¼ 0.020 exp(20,100/RT) cm2/s, where the numerator of the
exponential is in cal/mol. The average vibrational frequency of a carbon
atom in the lattice is 1012 Hz. Calculate the entropy of migration for carbon
transport by diffusion.
8. In the pure metal M, the dominant diffusion mechanism is thought to be an
interstitialcy mechanism. A self-diffusion experiment shows that at 1000 K a
pressure of 104 bar increases D by a factor of 8.
(a) Is the experimental result qualitatively consistent with an interstitialcy
mechanism?
(b) Calculate ΔVa for M.
9. Calculate the correlation factor f for a vacancy mechanism in a two-dimensional
square lattice by counting up to the 5th jump. The true value is f ¼ 0.46705.
10. Calculate and plot f for an impurity-vacancy pair in an fcc lattice for ω1/ω2 ¼ 10,
5, 1, 1/5, and 1/10, where ω1 and ω2 are the attempt frequencies of the vacancy to
exchange with a host atom neighboring the impurity and with the impurity itself,
respectively.
11. (a) Using the data given below, make a plot of log D versus 1/T, and estimate, by
eye, the best straight line through the points.
References 99

(b) Calculate ΔH and Do for this line.


(c) Calculate ΔH and Do using a least-squares procedure, assuming all error to
be in the values of D. Plot the least-squares line on the graph of part (a).

D/cm2s1 108 109 1010 1011


T/K 1350 1100 950 800

12. You wish to shorten the time for a low-temperature diffusional process (e.g.,
precipitation) which depends upon quenched-in vacancies in a metallic crystal.
The non-equilibrium vacancy concentration resulting from the quench is CV.
Assume that the equilibrium, or intrinsic, concentration of vacancies at the aging
temperature is negligible and that the quenched-in vacancies last only until they
have diffused to grain boundaries. The grain size is λ. Species A is known to
diffuse by a vacancy mechanism.
(a) What is the diffusion coefficient, DA, at t ¼ 0, i.e., immediately following
the quench from high temperature?
(b) Estimate the mean lifetime of the vacancies.
(c) Sketch the profile of DA as it varies across a grain at some later time, t > 0,
where the metal has been held at a constant aging temperature since the
quench.
(d) Sketch the profile of DA as it varies across a grain at long times, t >> 0,
where the metal has been held at a constant aging temperature since the
quench.
(e) How is the diffusion of A affected by:
(i) The high temperature Tq from which the specimen is originally
quenched?
(ii) The aging temperature Ta?
In each case, state how the choice of temperature influences the magnitude of
DA and its variation with time. You may use figures if necessary and try to be as
quantitative as possible to justify your answers.

References

1. A. Einstein, Ann. Phys. 17, 549 (1905)


2. G. Gamow, One Two Three. . .Infinity (Bantam Books, Inc., New York, 1961), p. 201
3. P.G. Shewmon, Diffusion in Solids (McGraw-Hill Book Co. Inc., New York, 1963), p. 58
4. B. Okkerse, Phys. Rev. 103, 1246 (1956)
5. H.E. Schaefer, in Positron Annihilation, ed. by P. G. Coleman, S. C. Sharma, (North Holland,
1982), p. 369
6. A.M. Brown, M.F. Ashby, Acta Metall. 28 (1980)
7. M. Werner et al., Phys. Rev. B 32, 3930 (1985)
8. K. Compaan, Y. Haven, Trans. Faraday Soc. 52, 786 (1956)
Chapter 3
Chemical Reaction Kinetics

3.1 Introduction

So far, we have learned the mathematics of diffusion, together with the atomistic
meaning of the diffusivity, which allows us to evaluate the spatial and temporal
variation of the concentration, c(x,y,z;t) within a system material. There is another
kind of kinetics, also leading to a temporal variation of concentration or composi-
tion, which is chemical reaction. In materials, this kinetics is quite often combined
with diffusion kinetics in parallel or in tandem.
Examples are carburization (or decarburization) of an iron sheet in an atmo-
sphere of CH4/H2 gas mixtures by virtue of the chemical reaction,

CH4 ¼ C þ 2H2

and oxygenation (or deoxygenation) of a nonstoichiometric oxide ceramic in a


H2O/H2 buffer gas atmosphere by virtue of the chemical reaction,

H2 O ¼ O þ H2 ,

where C and O denote C and O within the system of interest, the iron sheet and the
nonstoichiometic oxide, respectively. Putting aside the gas-phase transport kinetics
of CH4 or H2O as the C or O donor, respectively, to the surface of the system sheet or
ceramic, typically the following consecutive kinetic steps are involved: (a) Chemical
reaction to produce C or O on the surface of the system according to the respective
reaction above; (b) Incorporation of C or O across the surface of the system;
(c) Diffusion of C or O from the surface into the interior of the system.
Here, we will briefly review the basic concepts of chemical reaction kinetics
which we often encounter in materials engineering as above. Therefore, it cannot be
exhaustive by any means. If you are interested in more diversified, in-depth

© Springer Nature Switzerland AG 2020 101


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_3
102 3 Chemical Reaction Kinetics

treatments, you may refer to the relevant references (for example House [1] and
Amdur and Hammes [2]). Let us begin by defining our reaction system first.

3.2 Chemical Reaction and Rate Law

3.2.1 Reaction System

In order to make our discussion simpler, let us assume for the time being that our
system is closed with respect to matter exchange and has a fixed volume. Otherwise,
there would be concentration changes not only due to chemical reactions, but also
due to mass exchange or diffusion across the boundary and volume change, respec-
tively. Obviously, it is much easier to treat such a closed and constant volume
system.
Secondly, we assume that our reaction system is held under isothermal condi-
tions. This means that heat transfer is very much faster than the chemical reactions
taking place inside, and thus our entire reaction system is kept at a constant
temperature. As the other extreme, if the chemical reactions are much faster no
matter whether exothermic or endothermic, then we may treat the reaction system as
adiabatic.
Thirdly, we assume that our reactions are homogeneous with respect to the
reaction sites. In reality, chemical reactions occur much more often at heteroge-
neous reaction sites, e.g., catalysts. For simplicity’s sake here, we treat only
homogeneous reactions taking place within the interior of the system.

3.2.2 Chemical Reaction Rate

Let us consider the chemical reaction,

aA þ bB þ    ¼ cC þ dD þ    ð3:1aÞ

where A, B,. . . stand for the reacting species and a, b, . . . the stoichiometric
coefficients. The reacting species on the left hand side are normally called the
reactants and those on the right-hand side the products.
It is often convenient to write this chemical reaction as
X
0 ¼ ðaÞA þ ðbÞB þ    þ cC þ dD þ    ¼ νs S ð3:1bÞ
S

where S denotes the reacting species A, B, . . ., and νS the stoichiometric coefficient


of S. Then, it turns out that a stoichiometric coefficient does have an algebraic sign:
νS < 0 for the reactants and νS > 0 for the products.
3.2 Chemical Reaction and Rate Law 103

The reaction rate of a chemical reaction Eq. (3.1), R, is usually defined as the
decreasing rate of a reactant or the increasing rate of a product to make it formally
(or officially) positive or, denoting the concentration of a reacting species S(¼A,
B,. . .C,D,..) as cS,

dcA dc dc dc
R¼ ;  B ;   ; C ; D ;    ð3:2Þ
dt dt dt dt

Then, the reaction rate of a given reaction should take different values depending
on the choice of S, and hence, looks somewhat awkward. We rather like to have a
common value for the given reaction, irrespective of the choice of S. You may notice
that once divided by a reacting species’ own stoichiometric coefficient, the rate R
turns universal or

1 dcS dξ
R¼   , ð3:3Þ
νS dt dt

where ξ as such is called the extent or advancement of reaction. The rate of a


chemical reaction is normally defined in this way.

3.2.3 Concentration Change in General

As the reaction rate under a closed, constant-volume condition has now been
appropriately defined, let us think about the concentration change in general,
namely, when the boundary is left open with respect to material exchange as well
as the volume is allowed to change. The concentration change of a species dc is then
the sum of the changes due to chemical reaction (rxn), diffusion (diff), and volume
change (vol) or
     
∂c ∂c ∂c
dc ¼ dt þ dt þ dV, ð3:4aÞ
∂t rxn ∂t diff ∂V vol

or dividing through by dt,


     
dc ∂c ∂c ∂c dV
¼ þ þ : ð3:4bÞ
dt ∂t rxn ∂t diff ∂V vol dt

We have just learned that the concentration change by a chemical reaction is, due
to Eq. (3.3),
 
∂c
¼ νR: ð3:5Þ
∂t rxn
104 3 Chemical Reaction Kinetics

The change by diffusion, which we have learned in Chap. 1, is written as


 
∂c
¼ ∇D∇c ð3:6Þ
∂t diff

where D is the appropriate diffusion coefficient.


As c ¼ n/V with n being the number of moles of the species in the system, its
concentration change due to volume change is
 
∂c dV c dV
¼ : ð3:7Þ
∂V vol
dt V dt

Thus, Eq. (3.4) finally takes the form in general,

dc c dV
¼ νR þ ∇D∇c  : ð3:8Þ
dt V dt

We will now be dealing with a constant-volume system with no diffusion.

3.2.4 Reaction Rate Law

In general, the reaction rate R as defined in Eq. (3.3) is a function of the concentra-
tions of all the species present in a reaction system or
 
R ¼ f c1 , c2 , . . . , cj , . . . : ð3:9Þ

This sort of relationship is called the rate law for a given reaction. The functional
form of the rate law is still to be determined by experiment and cannot be a priori
predicted from the stoichiometric coefficients of the reacting species of a given
reaction, Eq. (3.1). Generally, it has nothing to do with the stoichiometric coeffi-
cients and everything to do with the reaction mechanisms. An extreme example [3] is
that for the reaction,

H2ðgÞ þ Br2ðgÞ ¼ 2HBr ðgÞ , ð3:10Þ

for which the rate law is experimentally found to be

1=2
1 dcHBr kcH2 cBr2
R¼ ¼ ð3:11Þ
2 dt 1 þ k0 cHBr =cBr2

where k and k0 are constants. You see, it has, indeed, nothing to do with the
stoichiometric coefficients of the reacting species.
3.3 Simple Reactions 105

In many cases, though, it has been found that the rate law may be represented
empirically as
m
R ¼ kcm
1 c2   cj   
1 m2 j
ð3:12Þ

where the exponent mj is called the partial order of reaction or partial reaction
order with respect to the component j, and m(¼∑mj), the overall order of reaction
or overall reaction order. The partial reaction orders mj are not necessarily integers
and are normally determined empirically. Here, the proportionality constant, k is
called the rate law constant or simply, rate constant. Because this corresponds to
the rate R when c1 ¼ c2 ¼ . . . ¼ cj ¼ . . . ¼ 1 (for whatsoever unit of cj), it is often
called the specific reaction rate. It should be noted that the specific rate or rate
constant, k has a unit which depends on the overall reaction order m of the given
reaction.

3.3 Simple Reactions

One of the major tasks of chemical reaction kinetics is to determine the reaction
order because it is important in practice and sometimes gives some hints as to the
reaction mechanisms as well. We will learn the typical methods for doing this by
looking at some simple reactions.

3.3.1 Integration Method

The rate law, Eq. (3.12) in association with Eq. (3.3), is basically a differential
equation. As already implied by the name of the method, this method is to integrate
the rate law proposed or conjectured and to read off the reaction order when the
integrated rate law fits the observation.
1st-order reaction
Let us consider a simple, one-way or irreversible reaction,

A ! B: ð3:13Þ

We assume that this reaction will be of the 1st order with respect to the reactant
A. The rate law is, then, written as

dcA
R¼ ¼ kcA : ð3:14Þ
dt

Upon integration with the initial condition cA ¼ cA,o at t ¼ 0, we obtain


106 3 Chemical Reaction Kinetics

Fig. 3.1 lncA vs. t of the ln cA


one-way reaction A ! B if
the first order -k

0 t

cA
ln ¼ kt or cA ¼ cA,o ekt : ð3:15Þ
cA,o

It is often more convenient to express the rate law in terms of the amount reacted,

x ¼ cA,o  cA , ð3:16Þ

which actually corresponds to the amount of B generated in the present case. Then,
the rate law, Eq. (3.14) may be rewritten as

dx
¼ kðcA,o  xÞ, ð3:17Þ
dt

followed by the solution,

cA,o  x  
ln ¼ kt or x ¼ cA,o 1  ekt : ð3:18Þ
cA,o

If this is the case, the plot of ln cA or ln (cA,o – x) vs. t should be linear with the
slope equal to –k as shown schematically in Fig. 3.1.
If the observed data set, {cA,t} does not fit this “first-order-reaction solution,” one
may then wish to try a further possibility, second order and others.
2nd-order reaction
If the reaction, Eq. (3.13) is of the 2nd order, the rate law will be

dcA
R¼ ¼ kc2A ð3:19aÞ
dt

or, in terms of the amount reacted x,

dx
¼ kðcA,o  xÞ2 : ð3:19bÞ
dt

The integration of this equation leads to the solution,

1 1 1 1
 ¼ kt or  ¼ kt ð3:20Þ
cA cA,o cA,o  x cA,o
3.3 Simple Reactions 107

Fig. 3.2 1/cA vs. t for the 1/cA


second order reaction
A!B k

1/cA,0

0 t

By plotting 1/cA or 1/(cA,o – x) against t, one should get a straight line with the
slope k as in Fig. 3.2.
Here is another simple reaction of the second-order:

dcA
A þ B ! C; R ¼  ¼ kcA cB ð3:21Þ
dt

where cA ¼ cA,o, cB ¼ cB,o and cC ¼ 0 at t ¼ 0, for example. In this case, it is much


more convenient to rewrite the rate law in terms of the amount reacted x or

dx
R¼ ¼ kðcA,o  xÞðcB,o  xÞ: ð3:22Þ
dt

Separating the variables and integrating, one obtains the solution as

1 c ð c  xÞ
ln B,o A,o ¼ kt ð3:23Þ
ðcA,o  cB,o Þ cA,o ðcB,o  xÞ

If the observation fits this integrated rate law, one may say that the reaction is of
the overall 2nd order.
mth-order reaction
If the reaction, Eq. (3.13) is of the mth-order in general, the rate law takes the form in
terms of the amount reacted x as

dx
¼ kðcA,o  xÞm : ð3:24Þ
dt

Integration leads to the solution,


" #
1 1 1
 ¼ kt: ð3:25Þ
m  1 ðcA,o  xÞm1 cm1
A,o

One may recognize that this general law works for all reaction orders except for
m ¼ 1.
108 3 Chemical Reaction Kinetics

3.3.2 Half-Life Method

Another method to determine the overall reaction order of the simple reaction,
Eq. (3.13), is to examine the dependence of the half-life, t1/2, of a reactant, the time
elapse for the amount of the reactant, say, A to reach one-half of its initial amount.
1st-order reaction
At t ¼ t1/2, cA ¼ cA,o/2 or x ¼ cA,o/2. The integrated rate law, Eq. (3.18), may, thus,
be rewritten as

ln 2
ln 2 ¼ kt1=2 or t1=2 ¼ ð3:26Þ
k

The half-life turns out to be independent of the initial amount cA,o. In other
words, if one measures t1/2 as a function of the initial amount cA,o to find it
independent of cA,o or constant, one may then say that the reaction in question is
of the 1st order.
2nd-order reaction
Again as cA ¼ cA,o/2 ¼ x at t ¼ t1/2, the integrated, 2nd-order rate law, Eq. (3.20) may
be rewritten as

2 1 1
 ¼ ¼ kt1=2 ð3:27Þ
cA,o cA,o cA,o

The half-life t1/2 is inversely proportional to the initial concentration cA,o. Upon
plotting t1/2 against 1/cA,o, one should get a straight line with the slope 1/k.
mth-order reaction
In general, the mth-order rate law, Eq. (3.25) may be rewritten in terms of t1/2 as

1 2m1  1
t1=2 ¼  : ð3:28Þ
kð m  1Þ cm1
A,o

The dependence of t1/2 on cA,o allows one to evaluate the reaction order m. Note
again that this equation does not work for m ¼ 1.

3.3.3 Method of Initial Rates

This is a simple method to determine the partial reaction orders of the rate law.
Let us consider a simple reaction,
3.3 Simple Reactions 109

A þ B þ C þ  ! D þ E þ F þ  ð3:29Þ

for which the rate law is given as

dx
R¼ ¼ kcm
A cB cC   
A mB mC
ð3:30Þ
dt

We wish to determine the partial reaction order, e.g., mA. How do we do that?
It is noted that
 
∂ ln R
¼ mA ð3:31Þ
∂ ln cA cB ,cC ,...

This means that if one measures the reaction rate R at different cA’s while keeping
the other concentrations, cB, cC, etc. fixed, then the partial reaction order, mA may
be evaluated via this equation. The method goes like this: You may prepare two
reacting systems {A, B, C,. . . .} with different A-concentrations, cA,1 and cA,2 while
others are all kept the same, and measure the amount reacted x against time t for
these two systems, see Fig. 3.3.
Then, the initial slope, dx/dt as t ! 0 meets Eq. (3.31) viz., corresponds to the rate
R with other concentrations, cB, cC, etc. fixed or

dx
R1 ¼ lim A,1 cB cC   
¼ kcm A mB mC
ð3:32aÞ
t!0 dt
dx
R2 ¼ lim ¼ kcm
A,2 cB cC   
A mB mC
ð3:32bÞ
t!0 dt

These two equations are solved simultaneously for the partial reaction order,
mA, as

ln R1  ln R2
mA ¼ ð3:33Þ
ln cA,1  ln cA,2

Here, you should understand why only the initial slopes as t ! 0, Eq. (3.32), must
be taken. Otherwise, the condition for Eq. (3.31), the other concentrations being kept
fixed, would be violated. It is why this method is called the initial rate method.

Fig. 3.3 The initial slopes,


x
R ¼ dx/dt as t ! 0, at two
R2
different initial
concentrations of a reactant,
R1 cA,2
say, cA,1 and cA,2
cA,1
0 t
110 3 Chemical Reaction Kinetics

3.4 Complex Reactions

So far, so simple! We will now examine some not-so-simple reactions as above,


which may be encountered rather more often in our playground, materials
engineering.

3.4.1 Reversible 1st-Order Reaction

Let us consider a reversible reaction between the two reacting species A and B,

kf
A) *B ð3:34Þ
kr

where kf and kr denote the specific rate of the “forward” reaction and “reverse”
reaction, respectively. If the reaction is of the 1st-order, the reaction rate R may be
written as

dcB
R¼ ¼ kf c A  kr c B ð3:35aÞ
dt

or in terms of the amount reacted, x(¼cA,o  cA ¼ cB  cB,0), as

dx
R¼ ¼ kf ðcA,o  xÞ  kr ðcB,o þ xÞ ð3:35bÞ
dt

where cS,o refers to the initial concentration of the reacting species S(¼A,B).
Upon integration, one obtains the integrated rate law as

kf ðcA,o  xÞ  kr ðcB,o þ xÞ k c  kr c B
¼ f A ¼ eðkf þkr Þt : ð3:36Þ
kf cA,o  kr cB,o kf cA,o  kr cB,o

For the case of cA,o > cB,o, for example, cA and cB may vary with time t as shown
schematically in Fig. 3.4.
It is noted that as t ! 1, the reaction, Eq. (3.34) should approach to the chemical
reaction equilibrium and Eq. (3.36) leads to

c B kf
kf c A  kr c B ¼ 0 or ¼ : ð3:37Þ
c A kr

Actually, at the chemical reaction equilibrium of Eq. (3.34),


3.4 Complex Reactions 111

Fig. 3.4 Temporal cB


variation of cA and cB for the cA,0
reversible 1st-order reaction
AÐB
cA
cB,0

0 t

dx
R¼ ¼ 0: ð3:38Þ
dt

It follows from Eq. (3.35) that


 
kf cB,o þ xeq cB,eq ΔGo
¼ ¼ ¼ Keq ¼ exp  ð3:39Þ
kr cA,o  xeq cA,eq RT

where the subscript “eq” means its host values (X and C) being for the equilibrium
state and ΔGo is the standard reaction free energy of the reaction, A ¼ B. The
reaction equilibrium constant Keq turns out to be corresponding to the dynamic
equilibrium of the reaction, Eq. (3.37), that is, the forward reaction rate is exactly
balanced by its reverse reaction rate.

3.4.2 Consecutive 1st-Order Reaction

Here is a reaction consecutively changing to another and to other with a specific


rate, k1 and k2, respectively,

k1 k2
A ! B ! C: ð3:40Þ

If the first order, the rate law may be written as:

dcA
¼ k1 cA ; ð3:41aÞ
dt
dcB
¼ k1 c A  k2 c B ; ð3:41bÞ
dt
dcC
¼ k2 c B : ð3:41cÞ
dt

Let us suppose that at t ¼ 0,


112 3 Chemical Reaction Kinetics

cS ¼ cS,o ðS ¼ A, B, CÞ: ð3:42Þ

The solution to these rate laws may go formally and strictly as follows:
Eq. (3.41a) is first integrated as

cA ¼ cA,o ek1 t : ð3:43Þ

Substitution of this solution into Eq. (3.41b) yields

dcB
¼ k1 cA,o ek1 t  k2 cB : ð3:44Þ
dt

It does not appear so straightforward to solve for cB, but we may guess that the
solution will take the form

cB ¼ aek1 t þ bek2 t ð3:45Þ

with “a” and “b” being the constants to be determined. By substituting this so-called
trial solution into Eq. (3.44) and comparing the both sides with the aid of the initial
condition, Eq. (3.42), you may determine “a” and “b” to have the solution as
   
k1 k1
cB ¼ cA,o ek1 t þ cB,o  cA,o ek2 t : ð3:46Þ
k2  k1 k2  k 1

For the present consecutive reaction, mass conservation stipulates that


X dc
S
¼ 0: ð3:47Þ
S¼A, B, C
dt

Thus, cC is finally obtained as

cC ¼ ðcA,o þ cB,o þ cC,o Þ  ðcA þ cB Þ


   
k2 k1 :
¼ ðcA,o þ cB,o þ cC,o Þ  cA,o ek1 t  cB,o  cA,o ek2 t
k2  k1 k2  k1
ð3:48Þ

Neglecting initial transients, it is often observed that


     
dcA  dcC  dc 
 ,     B  ð3:49Þ
dt dt dt

We may then approximate as

dcB
 0: ð3:50Þ
dt
3.4 Complex Reactions 113

15 15 15
CA CA CA
10 CB 10 CB 10 CB
CC CC CC
C C C
5 5 5

0 0 0

0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
t t t

(a) (b) (c)

Fig. 3.5 cA(t), cB(t) and cC(t) due to the formal solutions, Eqs. (3.43), (3.46) and (3.48) as
calculated with the arbitrary numerical values: (a) k2 ¼ 100k1; (b) k2 ¼ 10k1; (c) k2 ¼ 2k1 with
k1 ¼ 1 and the initial conditions: cA,o ¼ 10, cB,o ¼ 5 and cC,o ¼ 0. Note the time when the stationary-
state approximation, Eq. (3.50) starts to work depending on the rate-constant ratio k2/k1

This is called the steady-state approximation. When this is the case, the solution
to Eq. (3.41) may be immediately obtained even by inspection as:

cA ¼ cA,o ek1 t
k
cB  1 cA,o ek1 t ð3:51Þ
k2
cC  ðcA,o þ cB,o þ cC,o Þ  cA,o ek1 t

Temporal variations of cA, cB and cC depending on the ratio of k2/k1 are as shown
in Fig. 3.5 when cA,o > cB,o. You may wonder how and when the steady-state
approximation, Eq. (3.50) can be met depending on the ratio k2/k1.
This sort of steady-state approximation sometimes makes the problem
extremely easy to solve as we have just seen by comparing with the formal solution
above. It is noted, however, that the approximation may be justified only when

k1  k2 : ð3:52Þ

This means that the reaction A ! B is extremely slow or a very rare event
compared to the reaction B ! C, thus, once formed, B immediately transforms to
C. You may check yourself that if this is the case, the formal solutions, Eqs. (3.43),
(3.46) and (3.48), will indeed converge to Eq. (3.51).
Let us see how the steady-state approximation makes an otherwise complicated
solution much simpler with a consecutive reversible reaction,

k1 k2
A )* B )* C: ð3:53Þ
k1 k2

Assuming the first-order, the rate law is written as


114 3 Chemical Reaction Kinetics

dcA
¼ k1 cA þ k1 cB ð3:54aÞ
dt
dcB
¼ k1 cA þ k2 cC  k1 cB  k2 cB ð3:54bÞ
dt
dcC
¼ k2 cB  k2 cC ð3:54cÞ
dt

Apparently, direct integration of the law is quite difficult. Now assume the steady
state with respect to B or

dcB
0 ð3:55Þ
dt

Then, one may eliminate cB with the aid of Eq. (3.54b) to obtain

dcA dc
 k01 cA þ k02 cC   C ð3:56aÞ
dt dt

where

k1 k2 k1 k2
k01 ¼ ; k02 ¼ ð3:56bÞ
k1 þ k2 k1 þ k2

The approximation has rendered the reaction to look like a first-order reversible
reaction between A and C,

k01
A )* C ð3:57Þ
k02

3.4.3 Concurrent Reaction

Suppose that A transforms to B, C and D concurrently with the rate constant kB, kC,
and kD, respectively, or

kB
A ! B,
kC
A ! C, ð3:58Þ
kD
A ! D:

Assuming each reaction to be of the 1st order, the rate laws may be written with
respect to A as
3.5 Mixed-Controlled Kinetics 115

dcA
 ¼ ðkB þ kC þ kD ÞcA  kcA ð3:59aÞ
dt

and with respect to the rest, B, C and D, as

dcS
¼ kS c A ðS ¼ B, C, DÞ: ð3:59bÞ
dt

Upon integration with the initial conditions cS(t ¼ 0) ¼ cS,o for all the reacting
species, the solutions take the shapes

cA ¼ cA,o ekt ; ð3:60aÞ


kS  kt

cS ¼ cS,o þ c 1e ðS ¼ B, C, DÞ: ð3:60bÞ
k A,o

Notice how the reacted amount of the reactant, cA,o  cA is divided into each
product.

3.5 Mixed-Controlled Kinetics

Consider a reaction in a condensed phase, say, solid,

AþB!C ð3:61Þ

for which the overall reaction rate law may be written as

dcC
¼ kcA cB : ð3:62Þ
dt

For this reaction to occur, the reactants A and B should first be transported
through the medium via diffusion to meet to react. If the reaction medium were
gaseous, transport of the reactants would provide little difficulty to the overall
kinetics. The overall reaction may now be governed by diffusion or transport itself
rather than by reaction. We will often encounter this sort of situation from the kinetic
processes in materials.
We may conjecture that the reaction Eq. (3.61) consists of the following elemen-
tary steps:
(i) A and B diffuse to bring themselves in contact, thus, to make an intermediate
species AB:

kd
A þ B ! AB ð3:61aÞ
116 3 Chemical Reaction Kinetics

(ii) AB may, then, either depart away from each other again via diffusion or react to
produce C:

k0d
AB ! A þ B, ð3:61bÞ
kr
AB ! C: ð3:61cÞ

We may write the rate law with respect to the intermediates AB as

dcAB  
¼ kd cA cB  k0d þ kr cAB : ð3:63Þ
dt

Assuming the steady state or dcAB/dt ¼ 0,

kd
cAB ¼ cA c B ð3:64Þ
k0d þ kr

which leads to the production rate of C as

dcC kk
¼ kr cAB ¼ 0 r d cA cB : ð3:65Þ
dt kd þ kr

Upon comparison with the given rate law, Eq. (3.62), one obtains the overall
or effective rate constant keff as

kr kd
keff ¼ ¼ k: ð3:66Þ
k0dþ kr

The overall reaction rate turns out to be determined by two different kinetic
steps, diffusion and chemical reaction in this case, which is called the mixed-
controlled kinetics.
The two limiting cases may be distinguished:
(i) If k0d  kr or diffusion rate is much faster than the reaction,

kd
keff  kr Keq ; Keq ¼ ð3:67Þ
k0d

where Keq corresponds to the reaction equilibrium constant for the reversible
elementary reaction, Eq. (3.61a) and (3.61b) or A + B ¼ AB. Thus, the overall
reaction rate is controlled or governed by kr or the slowest reaction step,
Eq. (3.61c), thus called a reaction-controlled kinetics.
3.6 Thermodynamic Restrictions on the Rate Laws 117

(ii) If k0d  kr or reaction is much faster than diffusion, Eq. (3.66) reads

keff  kd ð3:68Þ

The overall reaction is controlled or governed by the slowest step, diffusion,


thus, called the diffusion-controlled reaction. We will see many diffusion-
controlled examples later on.

3.6 Thermodynamic Restrictions on the Rate Laws

For the rate law of a chemical reaction to be valid over the entire range of
concentration of the product, it should be consistent with the conditions that are
stipulated by the thermodynamics of chemical equilibrium. When a reaction
achieves the chemical equilibrium, the concentration cS of each reacting species
S is rendered to be fixed at the equilibrium value cS ¼ cS,eq, thus

dcS
¼0 ð3:69Þ
dt

In this sense, the rate law, e.g., Eq. (3.14) for the reaction A ! B (Eq. (3.13)), is
valid only when cA  cB far away from the equilibrium. At the equilibrium between
A and B, thermodynamics requires that their chemical potentials should be rendered
equal to each other or

μA ¼ μB : ð3:70Þ

Assuming the ideal solution behavior, it follows that


 
cB,eq ΔGo
¼ Keq ¼ exp  : ð3:71Þ
cA,eq RT
 
where ΔGo ¼ μoB  μoA is the standard reaction free energy of the reaction,
A ¼ B.
This means that for the rate law, Eq. (3.14), in association with the thermody-
namic requirement, Eq. (3.69), to be consistent with the thermodynamic equilibrium
condition, Eq. (3.71), the rate law should take the form of a difference between the
forward reaction rate involving the reactant concentrations and the reverse reaction
rate involving the product concentrations or, e.g., Eq. (3.35) leading to Eq. (3.39) or
(3.71). So, at the chemical equilibrium, the forward reaction should be balanced by
its reverse reaction.
118 3 Chemical Reaction Kinetics

With this balance in mind, let us consider a triangular tautomeric reaction,

N$& N%$
N&$ N$% ð3:72Þ

N%&
& %
N&%

where a certain substance may coexist in a homogeneous medium in three different


forms A, B, and C that are mutually transforming to each other. Assuming the first
order reactions, one may write the rate laws for this reversible triangular reaction as

dcA
¼ ðkBA þ kCA ÞcA þ kAB cB þ kAC cC ;
dt
dcB
¼ kBA cA  ðkAB þ kCB ÞcB þ kBC cC ; ð3:73Þ
dt
dcC
¼ kCA cA þ kCB cB  ðkAC þ kBC ÞcC :
dt

At the chemical reaction equilibrium, the requirement Eq. (3.69) yields the
nontrivial solutions with respect to the equilibrium concentrations cS,eq:

cA,eq kAC kAB þ kAC kCB þ kBC kAB


¼ , ð3:74aÞ
cC,eq kCB kBA þ kCA kAB þ kCA kCB
cB,eq kBC kBA þ kBC kCA þ kBA kAC
¼ : ð3:74bÞ
cC,eq kCB kBA þ kCA kAB þ kCA kCB

These are obviously consistent with the equilibrium conditions,

μA ¼ μB ¼ μC ð3:75aÞ

or, assuming the ideal behavior,

cA,eq cB,eq cB,eq


¼ Keq,AC ; ¼ Keq,BC ; ¼ Keq,BA : ð3:75bÞ
cC,eq cC,eq cA,eq

where Keq,AC etc. stand for the reaction equilibrium constant for the reaction C Ð
A etc., respectively.
Mother Nature now requires one more thing to be observed at the equilibrium,
that is, each elementary reaction should be balanced by itself or the forward rate of
each elementary reaction should be balanced exactly by its reverse reaction rate. This
is a consequence of the time-reversal invariance or microscopic reversibility in
mechanics, which is called in chemistry the principle of detailed balance. This
principle requires for the reaction, Eq. (3.72) that
3.6 Thermodynamic Restrictions on the Rate Laws 119

kBA cA,eq kAB


A ! B : kBA cA,eq ¼ kAB cB,eq or ¼ ;
kAB cB,eq kBA
kBC cB,eq kBC
B ! C : kCB cB,eq ¼ kBC cC,eq or ¼ ; ð3:76Þ
kBC cC,eq kCB
kAC cC,eq kCA
C ! A : kAC cC,eq ¼ kCA cA,eq or ¼ :
kCA cA,eq kAC

It immediately follows that

kAB kBC kCA ¼ kAC kCB kBA ð3:77Þ

By substituting this relationship into Eq. (3.74), you may obtain

kBC kAB
cA,eq kAC kAB þ kCB þ kAC
¼  ; ð3:78aÞ
cC,eq kCA k k
kAB þ kCB þ CB BA
kCA
kBA kAC
cB,eq kBC kBA þ kCA þ kBC
¼  , ð3:78bÞ
cC,eq kCB k k
kBA þ kCA þ AB CA
kCB

which, as the second factors herein are equal to 1, have turned out to be exactly equal
to Eq. (3.76).
Summing up, a chemical reaction equilibrium is distinguished not only by that
the concentrations of reacting species should not vary with time, Eq. (3.69), but also
by that the principle of detailed balance should be met, Eq. (3.76). A unidirectional
circular reaction, for example,

N$& N%$
ð3:79Þ

& %
N&%

may satisfy the first condition of concentration time-invariance if

kBA cA ¼ kCB cB ¼ kAC cC , ð3:80Þ

but does not satisfy the principle of detailed balance. If the detailed balance is not
met, the concentrations of reacting species oscillate to approach the time-
independent values, violating the free energy minimum principle of thermody-
namics [4], as demonstrated in Fig. 3.6 by numerically solving Eq. (3.73) with
120 3 Chemical Reaction Kinetics

Fig. 3.6 Temporal variations of concentrations cS(¼[S]; S ¼ A,B,C) for the reversible circular
reaction, Eq. (3.72) (a) and for the unidirectional circular reaction, Eq. (3.79) (b). For numerical
calculations, kAB ¼ kBA ¼ kBC ¼ kCB ¼ kCA ¼ kAC ¼ 1 (a) and kBA ¼ kCB ¼ kAC ¼ 1 (b) are
assumed with the common initial condition, cA,0 ¼ 1; cB,0 ¼ cC,0 ¼ 0. The lower figures are for a
larger-scale y-axis. Pay attention to the variations of [B](dotted curves). (Reproduced from
Alberty [5])

kAB ¼ kBA ¼ kBC ¼ kCB ¼ kCA ¼ kAC ¼ 1 (a) and kBA ¼ kCB ¼ kAC ¼ 1;
kAB ¼ kBC ¼ kCA ¼ 0 (b) both for cA,0 ¼ 1; cB,0 ¼ cC,0 ¼ 0.
Historically, L. Onsager [6] first conceived the idea of the reciprocity between
two different irreversible processes of the same tensorial character, e.g., heat con-
duction and diffusion from the principle of detailed balance for the chemical
reaction, Eq. (3.72), which awarded him a Nobel Prize in chemistry more than
30 years later.

3.7 Temperature-Dependence of the Rate-Law Constant

We have so far learned the concentration-dependence of the reaction rate R at


constant temperature. What about the temperature dependence of the rate? It is
empirically known that the rate law constant k, whether forward rate or reverse
rate, increases exponentially with temperature to take usually the form,

E
k ¼ A exp  a ð3:81Þ
RT

where A represents the pre-exponential factor, R the universal gas constant, Ea the
activation energy. Thus, lnk vs. 1/T is linear with the slope of –Ea/R as shown in
Fig. 3.7.
3.7 Temperature-Dependence of the Rate-Law Constant 121

Fig. 3.7 An Arrhenian plot,


ln k
ln k vs. 1/T Ea
-
R

1/T

Fig. 3.8 Free energy


landscape along the reaction
G*
coordinate
(AB)∗

0
A,B
Grxn
C
reaction coordinate

This equation is called the Arrhenius equation after S. Arrhenius [7] who first
proposed this equation in 1889. The equation is sometimes modified by taking into
account the temperature dependence of the pre-exponential factor itself as

E
k ¼ ATn exp  a ð3:81bÞ
RT

where the exponent “n” takes a value typically in the range of 1 < n < 1. How come
such a form? Let us explore its inner-working.

3.7.1 Absolute Reaction Rate Theory

Let us consider the reaction, Eq. (3.61) with the rate law, Eq. (3.62) again,

dcC
A þ B ! C; ¼ kcA cB : ð3:82Þ
dt

You can imagine that for the product C to form, the reactants A and B should first
encounter each other, whether via diffusion, flight or whatsoever, at the nearest
possible distance or collide, and only when they meet with high enough an energy or
are sufficiently activated, they may turn to C. This intermediate entity at the
transition state with high enough an energy is called an activated complex,
(AB). Letting ΔG denote the activation free energy and ΔGrxn the reaction
free energy, the free energy landscape along the reaction coordinate may look like
Fig. 3.8.
122 3 Chemical Reaction Kinetics

The theory considers the reaction as a consecutive one, with the transition state
or activated complex (AB) in between, or

k1þ
AþB ! ðABÞ !
k2
C ð3:83Þ
k1

with the first step, activation reaction being reversible while the second step being
one-way traffic or irreversible. Subsequently, the rate laws may be written as

dcðABÞ
¼ k1þ cA cB  k1 cðABÞ  k2 cðABÞ ; ð3:84aÞ
dt
dcC
¼ k2 cðABÞ : ð3:84bÞ
dt

Assuming the steady state with respect to the activated complex,

dcðABÞ
¼0 ð3:85Þ
dt

which, via Eq. (3.84a), leads to


 
cðABÞ k1þ k1þ  ΔG
¼  ¼ Keq ¼ exp  : ð3:86Þ
cA cB k1 þ k2 k1 RT

Here, it has been assumed that

k1  k2 : ð3:87Þ

This means that the overall reaction is rate-controlled by k2 or the second step
(AB) ! C is a very rare event. This assumption eventually renders the first
activation step to appear to be in reaction equilibrium or A + B ¼ (AB) with the
equilibrium constant Keq.
Substitution of Eq. (3.86) into Eq. (3.84b) yields the rate law Eq. (3.82) as

dcC
¼ k2 cðABÞ ¼ k2 Keq cA cB  keff cA cB : ð3:88Þ
dt

The theory endows the specific rate k2 with an absolute value,

kB T
k2 ¼ ð3:89Þ
h

where kB is the Boltzmann constant and h the Planck constant. It dimensionally


corresponds to frequency (s1), which may be called the thermal frequency.
3.7 Temperature-Dependence of the Rate-Law Constant 123

Originally it has been derived by means of statistical mechanics, but it may be


derived in rather a schematic way, without losing the fundamental physics behind, as
follows.
Suppose that once activated or formed, the activated complexes (AB) stay upon
the col extending over a distance ℓ along the energy landscape in Fig. 3.8 for a time
span or life time τ such that


τ¼ ð3:90Þ
v

where v may be taken as the (1-dimensional) root-mean-square velocity. The


specific rate or frequency k2 to form the product C may then be written as

1
k2 ¼ : ð3:91Þ

The numerical factor 1/2 counts for the fact that only half of the total frequency (i.e.,
rightward vibration) leads to C. As the complex is confined over ℓ, its momentum p
(¼mv) will then be, due to the standing wave condition (ℓ ¼ λ/2 with λ being the
matter wave length),

h h
2ℓ ¼ λ ¼ ¼ ð3:92Þ
p mv

Noting the equipartition principle for a one-dimensional translational motion,

1 2 1
mv ¼ kB T, ð3:93Þ
2 2

the combination of Eqs. (3.90), (3.91), (3.92), and (3.93) leads to Eq. (3.89).
Finally, one has
 
kB T ΔG
k ¼ k2 Keq ¼ exp  ð3:94aÞ
h RT

or, due to the thermodynamic identity ΔG ¼ ΔH  TΔS,


   
kB T ΔS =R ΔH
k¼ e exp  : ð3:94bÞ
h RT

This is how the specific rate k usually takes the Arrhenian form, Eq. (3.81), and
the theory is, thus, called the absolute reaction rate theory, activated complex
theory, or transition state theory [8].
124 3 Chemical Reaction Kinetics

3.8 Adsorption

If you monitor as a function of time the mass of a piece of single crystal MgO, for
example, which has previously been completely dried in an oven, on a balance in a
normal ambient condition, you may observe that its mass continually increases with
ever decreasing rate towards saturation. This is because water vapor in air adsorbs
onto the crystal surface and it often makes the casual measurement of, e.g., mass or
density of a hygroscopic system, in particular, not trivial.
The phenomenon of adsorption is normally classified into two types depending
on the energetic origin of adsorption, that is, physisorption and chemisorption. In
the former, adsorbed atoms or molecules (called adsorbates) are bound to the
substrate (called adsorbent) due to the mutual attraction of physical origin, that
is, the van der Waals force, which is weak in nature. The heat of adsorption
normally takes a value in the range of –(20 ~ 40) kJ/mol. In the latter, on the other
hand, there is formed a sort of chemical bond between an adsorbate and the
adsorbent, which is of chemical origin. The adsorbate is rather strongly bonded
with the heat of adsorption on the order of the bond enthalpy, typically in the range
of –(80 ~ 240) kJ/mol, and hence, monolayer adsorption is normally preferred.
If temperature is low enough, the surface will have to be fully covered because the
enthalpy effect is overwhelming, but as temperature increases, the entropic
disordering effect increases. As temperature goes up, therefore, there arises a kind
of trade-off between the ordering enthalpy of adsorption and the disordering entropy
to minimize the free energy of our system. Kinetics of adsorption may, thus, be
treated from the chemical reaction point of view and hence, the equilibrium surface
coverage as a kind of chemical equilibrium.

3.8.1 Langmuir Adsorption (Ideal Adsorption)

Let us suppose that a substrate or adsorbent is exposed to an ideal gas G at a pressure P


at a fixed temperature T. The gas atoms or molecules subsequently adsorb onto the
surface of the adsorbent as illustrated in Fig. 3.9. We assume that the gas G adsorbs
only by a single layer and that there is no interaction among the adsorbates themselves.
Could we then predict the equilibrium surface coverage θ by the gas G?
We may envisage that there are dangling bonds or adsorbing sites S on the
adsorbent and they are always hungry for the adsorbates G. The process of adsorp-
tion may then be written like a reversible reaction as

G
| |
kf
G + S S ð3:95Þ
kr

(P) (1-θ) (θ)


3.8 Adsorption 125

Fig. 3.9 Monolayer


adsorption of gas G onto the G
surface of an adsorbent or
substrate

absorbent

where the quantity corresponding to the thermodynamic activity of each reacting


species is given within the parentheses just beneath it, e.g., P the pressure of the
adsorbate G. Letting θ denote the fraction of the dangling bonds or surface sites
which are occupied or adsorbed by G, the rate law may be written, by analogy to a
bi-molecular reaction A + B ¼ C, as:

ðAdsorption rateÞ ¼ kf Pð1  θÞ, ð3:96aÞ


ðDesorption rateÞ ¼ kr θ: ð3:96bÞ

Then, at the dynamic equilibrium,

ðAdsorption rateÞ ¼ ðdesorption rateÞ

or

θ k
¼ f ¼ Kad : ð3:97Þ
Pð1  θÞ kr

Here, Kad corresponds to the equilibrium constant for the adsorption


 reaction,
Eq. (3.95) with the standard adsorption free energy ΔGoad ¼ ΔHoad  TΔSoad
such that
 
ΔGoad
Kad ¼ exp  : ð3:98Þ
RT

Solving Eq. (3.97) for the equilibrium coverage θ, we finally obtain

Kad P
θ¼ : ð3:99Þ
1 þ Kad P

This is called the Langmuir isotherm after its inventor, I. Langmuir, in 1916.
If you plot θ vs. P, it may look as in Fig. 3.10. It should be noted that

lim θ ¼ Kad P; lim θ ¼ 1 ð3:100Þ


P!0 P!1
126 3 Chemical Reaction Kinetics

Fig. 3.10 The Langmuir


isotherm 1

Kad
0
P

Noting that the adsorption is innately an exothermic process or ΔHoad < 0, you
may tell, on the basis of the Le Chatlier principle, whether the equilibrium
coverage increases or decreases with temperature at the given pressure.

3.8.2 Dissociative Adsorption

Now, suppose that our gas G, e.g., H2, first dissociates and then adsorbs. The overall
kinetics may then be regarded as a consecutive reversible reaction, viz., dissocia-
tion followed by atomic gas adsorption or

H2 ⇄ 2H; ð3:101aÞ
H
| kf |
ð3:101bÞ
H+ S S.
kr

Due to the Langmuir isotherm for the dynamic equilibrium, the equilibrium
coverage by H is given as

Kad PH
θ¼ ð3:102Þ
1 þ Kad PH

where PH denotes the pressure of H-gas and Kad(¼kf/kr) the equilibrium constant for
the adsorption reaction, Eq. (3.101b). The dissociation reaction equilibrium,
Eq. (3.101a), stipulates

P2H
Kd ¼ : ð3:103Þ
PH2

Thus, Eq. (3.102) is rewritten, in terms of the pressure of H2-gas, PH2 , as

Kad ðKd PH2 Þ1=2


θ¼ ð3:104Þ
1 þ Kad ðKd PH2 Þ1=2
3.8 Adsorption 127

1=2 1=2
In this case, θ vs. PH2 should look like Fig. 3.10 with the slope of Kad Kd as
PH2 ! 0.

3.8.3 Competitive Adsorption

Next, consider the situation in which there are two different gases A and B above an
adsorbent with the partial pressures PA and PB, respectively, and hence, A and B
are competing for the same sites on the adsorbent as illustrated in Fig. 3.11.

Fig. 3.11 Competitive


adsorption of two different
gases A(○) and B( )

absorbent

The overall adsorption kinetics may be described as concurrent reactions as

A
A+ S S; ð3:105aÞ
(PA) (1-θA-θB) (θA)

B
B+ S S ð3:105bÞ
(PB ) (1-θA-θB) (θB )

where the quantity within the parentheses beneath each reacting species corresponds
to its activity.
Letting θA and θB denote the surface coverage by A and B, respectively, the
empty site fraction is then (1  θA  θB). At dynamic equilibria, the respective
equilibrium constants KA and KB may be written as

θA θB
KA ¼ ; KB ¼ ð3:106Þ
PA ð1  θA  θB Þ PB ð1  θA  θB Þ

which lead to the equilibrium coverages of A and B as

KA PA KB PB
θA ¼ ; θB ¼ : ð3:107Þ
1 þ KA PA þ KB PB 1 þ KA PA þ KB PB
128 3 Chemical Reaction Kinetics

3.8.4 Multilayer Adsorption: BET Isotherm

So far, we have considered the mono-layer adsorption no matter whether ideal,


dissociative or competitive. In reality, however, the mono-layer adsorption is not so
common because it is “ideal.” It is more natural to adsorb in multitude upon already-
adsorbed layers, viz., the 2nd layer upon the 1st layer, the 3rd upon the 2nd, and so
on. Obviously, the adsorption nature of the 1st layer should differ from that of the
other upper layers, because while the 1st-layer sits directly upon the absorbent of a
different physico-chemical nature, the 2nd and higher layers upon the immediately
lower layer of the same nature. The 1st layer may, thus, be expected to be more
chemisorbed and the others more physisorbed. How can we deal with this more
natural, multi-layer adsorption? Here is an ingenious theory invented by
S. Brunauer, P. H. Emmett, and E. Teller, thus, known as the BET isotherm
[9]. Let us explore its underlying ideas, which are very amusing.
Suppose that a gas G is naturally adsorbed in multilayers upon the adsorbent S
under its own pressure p at a given temperature. Once the adsorption equilibrium is
achieved, the situation may be illustrated “ideally” as in Fig. 3.12, where all the
molecules (or atoms) randomly adsorbed in each layer are all swept into a continuous
layer to the left end of the substrate S.
The adsorption of gas G upon S to make the 1st-layer G1S, upon this G1S-layer to
make the 2nd-layer G2S, and on and on, may be each represented as a reversible
reaction with the corresponding adsorption equilibrium constant
Kn(n ¼ 1,2,3,. . . .), like the Langmuir adsorption equilibrium constant, Kad in
Eq. (3.97), as

θ1
G þ S Ð G1 S : K1 ¼ ; ð3:108aÞ
pθv
θ2
G þ G1 S Ð G2 S : K2 ¼ ; ð3:108bÞ
pθ1
θ3
G þ G2 S Ð G3 S : K3 ¼ ; ð3:108cÞ
pθ2
                          
θn
G þ Gn1 S Ð Gn S : Kn ¼ : ð3:108dÞ
pθn1

Here, θv denotes the vacant surface-site fraction on the adsorbent S, and θn


(n ¼ 1,2,3,. . .) the site-fraction occupied by the n-fold adsorption layer GnS, see
Fig. 3.12. Considering the same nature of the interlayers for n > 1, namely, G
upon G, you may accept, with no difficulty, that all the corresponding equilibrium
constants should be the same or

K2 ¼ K3 ¼ K4 ¼    ¼ Kn  K: ð3:109Þ
3.8 Adsorption 129

Fig. 3.12 Multilayer θn


adsorption by n-fold n θn-1
(n ¼ 1,2,3,. . . .n) upon a n-1
substrate S: θv, the vacant-
site fraction upon S; θn, the
site fraction occupied by the
n-fold layers (nP ¼ 1,2,. . . .). 3 θ2
Note that θv þ 1 1 θn ¼ 1 2 θ1
1 θv
S

Furthermore, as the adsorption of gas G upon an already adsorbed layer G may be


regarded as the condensation reaction equilibrium of gas G(g) to liquid G(liq) or

GðgÞ Ð GðliqÞ : ð3:110aÞ

Thus, this all-the-same equilibrium constant K should correspond to the equilibrium


constant for this condensation reaction or

1
K¼ ð3:110bÞ
po

where po stands for the saturation vapor pressure of pure liquid G(liq) at the given
temperature. The equilibrium coverage of the n-fold layer, θn (n > 1) and the vacant-
site fraction, θv may then be all solved, in terms of θ1, from Eq. (3.108) as

1
θv ¼ θ; θn ¼ ðKpÞn1 θ1 ð3:111Þ
K1 p 1

Now, note from Fig. 3.12 that

X
1
θv þ θn ¼ 1 ð3:112Þ
n¼1

The infinite series in Eq. (3.112) converges as

X
1 X
1
1
θn ¼ ðKpÞn1 θ1 ¼ θ ð3:113Þ
n¼1 n¼1
1  Kp 1

as Kp ¼ p/po < 1 due to Eq. (3.110b). Equation (3.112) may then be rewritten all in
terms of θ1 as

1 1
θ þ θ ¼1 ð3:114aÞ
K1 p 1 1  Kp 1
130 3 Chemical Reaction Kinetics

or

K1
Kpð1  KpÞ
θ1 ¼ K  ð3:114bÞ
K
1  Kp 1  1
K

Now, letting N denote the total number of G-molecules adsorbed and Nm the total
number of adsorption sites both per unit mass of the adsorbent S, it is obvious that

N ¼ Nm ðθ1 þ 2θ2 þ 3θ3 þ   Þ ð3:115aÞ

or, due to Eqs. (3.111) and (3.113),


h i
N ¼ Nm θ1 1 þ 2ðKpÞ þ 3ðKpÞ2 þ   
d X
1
¼ N m θ1 ðKpÞn ð3:115bÞ
dðKpÞ n¼1
1
¼ N m θ1 :
ð1  KpÞ2

The total amount of the adsorbate N is usually measured gravimetrically or


volumetrically. Eq. (3.115b) associated with Eq. (3.114b) then takes the form, e.g.,
in terms of volume V as measured, say, at standard temperature and pressure (STP)
as

K1
Kp
V ¼ Vm h K i ð3:116aÞ
K1
ð1  KpÞ 1 þ  1 Kp
K

or rearranging a bit,
 h  i
Kp 1 K K1
¼ 1þ  1 Kp , ð3:116bÞ
Vð1  KpÞ Vm K1 K

where Vm is the volume of Nm molecules or the volume of the adsorbate gas when
fully covering the adsorbent by a mono-molecular layer, i.e., θ1 ¼ 1. Isn’t it neat?
This is the BET isotherm.
This isotherm allows you to determine K1 and Vm by measuring the adsorbed
gas volume, V against Kp(¼p/po) on the basis of Eq. (3.116b) as shown in Fig. 3.13.
Once you determine Vm in this way or so, then you may estimate the specific
surface area, or the surface area per unit mass, A of the adsorbent S, by using the
relationship Nm ¼ NAVm/(0.02241 m3 mol1) with NA being the Avogadro num-
ber, as A ¼ Nma with “a” being the area covered by a molecule or adsorption cross
section. This serves as the measurement principle of the specific surface area of fine
powders.
3.9 Langmuir Evaporation (Ideal Evaporation) 131

Fig. 3.13 Linear plot of the


Kp
BET isotherm, Eq. (3.116).
V (1 − Kp )
K1 and Vm can be
determined from the slope
and intercept values 1 ⎛ K ⎞ ⎛ K1 ⎞
⎜ ⎟⎜ − 1⎟
combined Vm ⎝ K1 ⎠ ⎝ K ⎠

1 ⎛K⎞
⎜ ⎟
Vm ⎝ K1 ⎠

Kp

3.9 Langmuir Evaporation (Ideal Evaporation)

We know from experience that if we leave a bowl of water in a dry atmosphere, the
water evaporates. A pure material in vacuum at a given temperature evaporates
eventually to saturation by its own vapor, This sort of evaporation occurs to any
condensed material. We wish to know how fast it will happen.
Let us suppose that a bulk material M is placed in its own vapor atmosphere of
pressure pM at a given temperature T as illustrated in Fig. 3.14.

Fig. 3.14 Evaporation (J) pM


of a bulk material and
condensation (J+) of its J+ J-
vapor (pressure PM) onto the
bulk M

bulk M

As the material boundary is left open, the substrate M and its surrounding vapor
continually exchange M-particles, Letting hvz+i be the mean velocity of the vapor
M-particles approaching the substrate M from above, and cM the concentration of
M-particles in the vapor, the condensation rate or flux of M, J+ may be written as

Jþ ¼ cM < vzþ> ð3:117Þ

By using the Maxwell distribution of one-dimensional velocity vz, the mean


velocity <vz+> may be calculated as

Z1 rffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffi
m mv2z =2kB T kB T
hvzþ i ¼ vz e dvz ¼ , ð3:118Þ
2πkB T 2πm
0

where m is the atomic mass of M. Due to the ideal gas law, the concentration is
given as
pM
cM ¼ : ð3:119Þ
kB T
132 3 Chemical Reaction Kinetics

The condensation rate may, thus, take the form,

pM
Jþ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð3:120Þ
2πm kB T

Letting Ns denote the number of M-atoms per unit area of the M-surface, on
the other hand, the evaporation rate J may be written, assuming the first order
reaction, as

J ¼ kNs ð3:121Þ

where k is the rate constant.


At the dynamic equilibrium,

Jþ ¼ J ð3:122Þ

and as this equilibrium is nothing but the condensed phase/vapor, two-phase


equilibrium,

ðsatÞ
pM ¼ pM ð3:123Þ

ðsatÞ
where pM is the saturation vapor pressure of M at given temperature. Assuming
Ns to be constant, Eqs. (3.120), (3.121), (3.122), and (3.123) lead to

ðsatÞ
pM
kNs ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð3:124Þ
2πm kB T

and hence, the net-vaporization rate, Jvap, finally takes the form,

ðsatÞ
p  pM
Jvap ¼ J  Jþ ¼ pMffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð3:125Þ
2πm kB T

One thing should be noted here: In this formulation, we have assumed that every
atom or molecule coming from above will stick to the substrate M irrespective of the
geometric configurations of the molecules as well as the surface sites. In other words,
the accommodation rate is 100% or accommodation factor α ¼ 1, but in reality,
some molecules may fail to settle onto the surface sites mainly due to their geometric
mismatch. This accommodation factor is, thus, equally called the steric factor as
well and normally takes a value much less than 1 or α  1. Eq. (3.125) should, thus,
be corrected against the possible failure of accommodation as

ðsatÞ
α pM  pM
Jvap ¼ J  Jþ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð3:126Þ
2πm kB T
3.10 Gibbs-Langmuir Isotherm 133

This is called the Langmuir evaporation equation.


You can immediately see that the rate becomes maximum when evaporating into
vacuum (pM ¼ 0) or

ðsatÞ
αpM
Jvap ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð3:127Þ
2πm kB T

If water contained in an insulated jar is let evaporate into vacuum (pH2 O ¼ 0), for
example, the evaporating water carries its own latent heat of vaporization (ΔHvap)
with it at the rate of ΔHvapJvap, thus, cooling down the jar. By using the energy
conservation equation, you may readily estimate how fast the interior of your jar
cools down. This is indeed a method of getting a low temperature.

3.10 Gibbs-Langmuir Isotherm

Here, we have a solution where the mechanical strain effect may be neglected, say,
an iron melt with sulfur dissolved as impurities. Could we tell whether such solute
impurities segregate to or desegregate from the surface, and if segregated, then how
much? It is also quite often observed in polycrystalline ceramics that certain impu-
rities segregate to and others desegregate from the grain boundaries. To answer this
question, we will start by briefly reviewing the thermodynamics of surface or
interface.

3.10.1 Thermodynamics of Interface and Gibbs Isotherm

Consider the two bulk phases α and β in contact with a planar boundary σ, each with
C-components at common temperature T and pressure P, as illustrated in Fig. 3.15.
Letting U, S, V, ni, and A denote the internal energy, entropy, volume, and mole
number of component i(¼1 to C), and interfacial area, respectively, the energetic
fundamental equations for the total, composite system (α + β + σ) and the two
subsystems α and β may be written, respectively, as

X
C
Total system ðα þ β þ σÞ : dU ¼ TdS  PdV þ μi dni þ γdA ð3:128aÞ
i¼1

X
C
ðαÞ
Subsystem α : dUðαÞ ¼ TdSðαÞ  PdVðαÞ þ μi dni ð3:128bÞ
i¼1

X
C
ðβÞ
Subsystem β : dUðβÞ ¼ TdSðβÞ  PdVðβÞ þ μi dni ð3:128cÞ
i¼1
134 3 Chemical Reaction Kinetics

where μi and γ are respectively the chemical potential of component i and the
interfacial or surface energy. Noting that V(σ) ¼ 0, the fundamental equation for the
interface σ, dU(σ) is subsequently obtained as

Fig. 3.15 Two bulk phases


α and β in contact with a
planar boundary σ α

σ (planar interface)

Interface σ: dU(σ) ¼ dU  dU(α)  dU(β)


X
C
ðσÞ
¼ TdSðσÞ þ μi dni þ γdA: ð3:129Þ
i¼1

As U is a first order homogeneous function with respect to other extensive


variables,

X
C
ð σÞ
UðσÞ ¼ TSðσÞ þ μ i ni þ γA ð3:130Þ
i¼1

due to Euler [10]. Differentiating Eq. (3.130) and comparing with the fundamental
equation Eq. (3.129), we obtain the Gibbs-Duhem equation for the boundary phase
or interface, σ, as

X
C
ðσÞ
SðσÞ dT þ ni dμi þ Adγ ¼ 0, ð3:131Þ
i¼1

which is nothing but a relationship among the intensive variables for the phase of
interface (σ). Particularly in an isothermal condition (dT ¼ 0), this equation reduces
to

XC ðσÞ
ni XC
dγ ¼  dμi   Γi dμi ð3:132Þ
i¼1
A i¼1

ðσÞ
where Γi(¼ni =A) is called the surface excess of component i. This equation is
referred to as the Gibbs equation.
Macroscopically, an interface or phase boundary, σ is regarded as a singularity
ðαÞ ðβÞ
(thus, V(σ) ¼ 0) where, e.g., ni changes discontinuously from, say, ni to ni , but
microscopically, it should vary continuously across a transition zone of atomic
planes. This is indeed the origin of the surface tension or energy that is, thus,
3.10 Gibbs-Langmuir Isotherm 135

often referred to as the gradient energy.1 Obviously, the extent of variation will
differ depending on the kind of component i and the variation itself is not necessarily
antisymmetric as well. The interface is, thus, normally defined as the plane where the
surface excess of the solvent (i ¼ 1) becomes nil or

ð σÞ
n1
Γ1 ¼ ¼0 ð3:133Þ
A

which is the position x across the interface such that

ð βÞ
Zn1
xdn1 ¼ 0: ð3:134Þ
ðαÞ
n1

You see, this is nothing but the Matano-interface (Eq. (1.144) in Chap. 1) with
respect to the solvent 1.
For the case of binary solution (C ¼ 2), the solute 2 may then have to redistribute
with respect to the interface, due to Eq. (3.132), as

dγ ¼ Γ2 dμ2 ¼ Γ2 RTd ln a2 ð3:135aÞ

or
 
1 ∂γ
Γ2 ¼  ð3:135bÞ
RT ∂ ln a2 T,P

where a2 is the activity of the solute. This equation is called the Gibbs isotherm.
The Gibbs isotherm tells that if ∂γ/∂lna2 < 0 or the surface energy γ decreases as
the solute activity a2 or concentration increases, then Γ2 > 0, i.e., the solute
segregates to the surface or grain boundaries, and vice versa. If you measure the
surface tension γ against the activity of the solute a2, you can actually determine the
surface excess Γ2 due to Eq. (3.135b).

3.10.2 Gibbs-Langmuir Isotherm

Now, suppose that the solute C in the system segregates to the surface as illustrated
in Fig. 3.16.
The situation of segregation is somewhat similar to the adsorption of gas G onto
the surface sites S of an adsorbent as illustrated in Fig. 3.9. We may thus write the

1
The inner working of the gradient energy is given in Sect. 5.3.5.
136 3 Chemical Reaction Kinetics

Fig. 3.16 Segregation of VXUIDFH


the solute C to the surface
looks like a gas adsorption
onto a substrate

& VROXWH

process of solute(C) segregation to the surface as a chemical reaction like a gas


adsorption reaction, Eq. (3.95) or

CþS Ð S
j j : ð3:136Þ
C

The equilibrium coverage θ is then given as the Langmuir isotherm or

Ka2
θ¼ ð3:137Þ
1 þ Ka2

where K is the equilibrium constant for the adsorption or segregation reaction,


Eq. (3.136). Letting Γo be the surface excess at full coverage or saturation, the
equilibrium coverage is given as

Γ2
θ¼ ð3:138Þ
Γo

By substituting Eq. (3.138) into Eq. (3.137) and combining with the Gibbs
isotherm, Eq. (3.135b), one obtains

Ka2 1 ∂γ
Γ2 ¼ Γ ¼ ð3:139Þ
1 þ Ka2 o RT ∂ ln a2

Upon integration of the second equality, one has

Zγ Za2
Ka2
dγ ¼ RTΓo d ln a2 ð3:140aÞ
1 þ Ka2
γo a2 ¼0

or

γ ¼ γo  RTΓo ln ð1 þ Ka2 Þ ð3:140bÞ

where γo is the surface tension of the pure solvent (a2 ¼ 0). This equation is called
the Gibbs-Langmuir isotherm. By fitting to this isotherm the raw data set {a2,γ} as
measured at a given temperature, you may determine the surface tension of the pure
solvent γo and its surface excess at the full coverage Γo as well.
Problems 137

Problems

1. The kinetics of reduction of FeO in the slag by carbon dissolved in molten pig
iron are as follows:

ρFeO/wt % 20.0 11.5 9.4 7.1 4.4


Time/s 0 60 90 120 180

(a) Calculate the order of reaction with respect to FeO concentration in the slag.
(b) How long will it take for ρFeO to reach 2.2 wt% in the slag? State your
assumption.
2. Consider the decomposition of gas A2 at 438 C

A2 ! 2A:

The rate is to be measured by observing the pressure (P) change in a constant


volume system; assume that the gas mixture is ideal.
(a) Express the rate of reaction d(ξ/V)/dt, in terms of dP/dt, where ξ is the extent
of reaction.
(b) Let P1 be the pressure in the system after the A2 is completely decomposed
(at t ¼ 1). If the reaction is first order in the concentrations of A2, derive the
relation between the pressure and time. What function of pressure should be
plotted against time to determine the rate constant?
(c) If the rate constant is 2.48
104 s1, calculate the half-life, and the time
required for 98% of the A2 to decompose.
(d) What will the value of P/P1 be after 2.0 hours?
3. The rate of the reaction, 2NO + 2H2 ! N2 + 2H2O, has been studied at 826 C.
Some of the data are:

Initial pressure H2 Initial pressure NO Initial rate


Run (PH2)o/kPa (PNO)o/kPa (dP/dt)(kPa/s)
1 53.3 40.0 0.137
2 53.3 20.3 0.033
3 38.5 53.3 0.213
4 19.6 53.3 0.105

(a) What are the orders of the reaction with respect to NO and with respect to H2?
(b) Assume that the gas mixture is ideal and find the relation between the rate of
reaction per unit volume and dP/dt, where P is the total pressure. The volume
is constant.
138 3 Chemical Reaction Kinetics

(c) Combine the results of (a) and (b) to find the relation between dP/dt and the
pressure. Initially, the total pressure is Po, the mole fraction of NO is xo, that
of H2 is 1  xo.
4. The rate of the reaction
H2ðgÞ þ I2ðgÞ ¼ 2HIðgÞ

is given by

d½HI
¼ k1 ½H2 ½I2  k2 ½HI 2
dt

where at 400 K, k1 ¼ 6.3


108 and k2 ¼ 8.0
1011 cm 3 mol1 s1.
(a) Calculate the apparent equilibrium constant for this reaction and compare it
with the value determined from a knowledge of ΔGo.
(b) An equimolar mixture of H2(g) and I2(g) at a total pressure of 1 atm is allowed
to react at 400 K. Determine the concentration of each species at equilibrium.
(c) Calculate the rate of reaction of the mixture in Part (b) at t ¼ 0, and when half
of the H2 and I2 have reacted.
(d) The activation energy of the forward reaction is 39 kcal/mol. Calculate the
initial rate of reaction at 500 K for the mixture of Part (b).

H2(gas) I2(gas) HI(gas) I2(solid)


ΔHo298,form =kcal  mol1 0 14.9 6.3 0
So298 =cal  mol1  K1 31.21 62.25 49.4 17.9

5. The homogeneous decomposition of ozone, O3, has a number of complications,


one of which is featured here. The reaction is catalyzed by various gases, one of
which is CO2. Table 3.1 gives the variation of the total pressure of a mixture
initially consisting of ozone and CO2.
(a) (i) Show what the apparent order of reaction is for experiment 1.
(ii) Calculate the apparent chemical rate constant, k.
(b) (i) Assume the complete rate law to be of the form

Table 3.1 The variation of Experiment 1 Experiment 2


total pressure with time for the
[CO2] ¼ 0.01 mol/L [CO2] ¼ 0.005 mol/L
reaction 2O3 ¼ 3O2 at
T ¼ 323 K Time (min) Ptotal (mm Hg) Time (min) Ptotal (mm Hg)
0 400 0 300
30 450 30 330
60 475 60 350
1 500 120 375
1 400
Problems 139

d[O3]/dt ¼  k[O3]a[CO2]b.

What must be the value of the exponent, b?


(ii) Calculate k.
(c) A rise in temperature of 10 K quadruples k. What is the activation energy of
the process?
(d) A proposed mechanism for the decomposition of ozone in the absence of a
catalyst is
k1
2O3 ! O3 þ O2 þ O;
k2
k3
O3 þ O ! 2O2 :

Making the necessary assumptions derive the rate law which corresponds to this
mechanism.
6. Gas, L, adsorbs strongly on the surface of a metal and appears to satisfy a simple
Langmuir adsorption behavior for a monatomic gas. At 300 K the Langmuir
adsorption coefficient is K ¼ 10 Pa1. A sample of metal is cleaved to expose
clean surface in a vacuum in which the pressure is 108 atm of L atoms.
(a) What is the final fraction of the surface sites which are covered at equilibrium
adsorption, θeq?
(b) What is the initial rate of adsorption, i.e., at t ¼ 0 (in #/m2-s)? Assume the
sticking coefficient α ¼ 0.8.
(c) What time is required to give coverage of 80% of the equilibrium saturation
coverage or θ ¼ 0.8θeq? There are 10+19 sites/m2 on the metal surface.
Assume diffusion of L into the metal is negligibly slow. First derive θ as a
function of time θ ¼ θ(t), and then calculate the required time for θ/θeq ¼ 0.8.

Molar weight of L ¼ 30 g=mol; Avogadro number ¼ 6:023


1023 molecules=mol

7. The absorption of N2 on TiO2 has been studied at a temperature of 75 K, at which


the saturation vapor pressure po of liquid nitrogen is 0.76 bar. The following data
have been reported for the volume of the adsorbed gas (V), measured at 25 C and
1 bar, per gram of TiO2.

p/po 0.002 0.024 0.080 0.153 0.224 0.288 0.359


V/mg1 600 720 822 935 1050 1150 1250

(a) Using the BET isotherm, deduce the volume of nitrogen to form a mono-
molecular layer Vm on 1 g of this TiO2 and the adsorption equilibrium
constant K1.
(b) Estimate the specific area of this TiO2.
140 3 Chemical Reaction Kinetics

8. The decomposition of N2O gas is catalyzed by solid In2O3. The rate law is

dcN2 O
¼ kθN2 O
dt

where cN2 O is molar concentration and θN2 O is the fraction of surface covered by
adsorbed N2O. Assume that the Langmuir isotherm applies. Show that the
decomposition of N2O (i) at low pressure is first order in PN2 O and (ii) at high
pressures is independent of pressure.
9. (a) The equilibrium vapor pressure of molten iron at 1550 C is 2.75
105
atm. The vapor species is monatomic. The accommodation coefficient for
condensation has been found to be α ¼ 0.97. Calculate the rate of evapora-
tion of Fe for a sample of pure metal placed in a vacuum chamber at
1550 C. Assume that the surface is uncontaminated. State the assumptions
necessary for your calculation.
(b) Oxygen is known to have a pronounced effect on the surface tension of
molten iron. The Gibbs-Langmuir isotherm for oxygen in liquid iron at
1550 C is found to be [11].

γ=Jm2 ¼ 1:788  0:240 ln ð1 þ 220aO Þ

where Henry’s law is assumed and 1 weight percent (w/o) is taken as the
standard state, i.e., aO ¼ w/o. What effect on the evaporation rate of Fe will
an oxygen impurity level of 10 ppm have? What about 500 ppm? Justify
your conclusion.

Molar weight of Fe, MFe ¼ 55:85 g=mol; density of Fe, ρFe ¼ 7:87 g=cm3

10. The desulphurization of molten iron at 1450 C into a flowing inert gas is
thought to occur by the reaction

1
Sin Fe ! S2 ðgasÞ ð1Þ
2

The kinetics of this reaction have been measured experimentally to be

d½%S
 ¼ k½%S 0:70 ð2Þ
dt

The non-integral value of the exponent reveals little about possible mechanisms.
However, eq. (2) deals with bulk concentrations of S, a species well known to be
surface active in molten iron. Indeed, the concentration dependence of the
surface tension of Fe-S melts at 1450 C has been measured to be
Problems 141

γ ¼ γo  β½%S 0:35 ð3Þ

where γo is the surface energy of pure iron and β is a positive constant.


(a) (i) Show that the rate of desulphurization of Fe-S melts at 1450 C can be
expressed as 2nd order in surface concentration of S. Is this reasonable?
(ii) How do you expect k in eq. (2) to be related to the Langmuir vaporization
rate?
(b) Other investigators have argued that desulphurization is controlled not by
chemical reaction kinetics, but by mass transport across a boundary layer in
the melt just beneath the surface. In this case, the flux of S to the gas phase
from the well stirred bulk melt is given by

JS ¼ αΔ½%S

where α is the mass transfer coefficient. Write an expression for the bath
composition as a function of time. State your assumptions.
11. Metal Q is heated in a furnace to 800 K. In order to deposit a layer of Q onto the
surface of a thick piece of metal M, the latter is exposed to Q vapor at a small
port in the furnace. During this exposure the temperature of M is constant at
500 K. Q vapor is known to be monatomic, condensing with an accommodation
coefficient, α, of 0.95. At 500 K, Q dissolves up to a limit of 1 atomic per cent
in M, obeying Henry’s law. M is effectively insoluble in Q at this temperature.
(a) (i) After an exposure time of 10 minutes will most (or all) of the deposited
metal Q be in the form of an identifiable surface layer or will it have
dissolved in the base metal M?
(ii) If there is a surface layer, what is its approximate thickness?
(b) Approximately how deep into the base metal will the concentration of Q be
raised to 0.5 atomic per cent?
MW ¼ 60 g/mol for M; 46 g/mol for Q
D ¼ 109 cm2/sec for Q in M
Molar volumes VM ¼ VQ ¼ 10 cm3/mole
Vapor pressure of Q ¼ 10 Pa at 800 K
Vapor pressure of Q ¼ 102 Pa at 500 K
Vapor pressure of M is negligible.
12. Gwankium(Gw) hydride decomposes at 1000 C to produce Gw-metal and
hydrogen gas according to the reaction

GwH2 ðsolidÞ ! Gw ðsolidÞ þ H2 ðgasÞ ðiÞ

The equilibrium hydrogen pressure over GwH2 and Gw at 1000 C is 2 atm.


142 3 Chemical Reaction Kinetics

To measure the diffusivity of hydrogen in copper the cell shown in illustration


below, has been constructed. The cell consists of two chambers separated by
copper foil of thickness, δ. The hydrogen pressure in the left chamber is generated
by reaction (i) and in the right chamber is fixed at 104 atm by a vacuum pump.
The rate of production of hydrogen by reaction (i) has been measured and found
to be zeroth order with a specific chemical rate constant for the left chamber of
3.5
1010 mol H2/s.
(a) What is the minimum value of δ that will allow a steady state hydrogen
concentration profile to be established across the copper foil? Neglect
hydrogen losses through the external walls of the chambers.
copper foil

H2 at P1 H2 at P2
to vacuum pump
P=10–4 atm

GwH2/Gw

(b) The overall rate of hydrogen transport from the left chamber to the right
chamber can be controlled by a variety of mechanisms. A partial list follows:
1. Diffusion of H through the copper foil
2. Chemical reaction at the surface of the foil to dissolve atomic hydrogen in
the copper
3. Dissociation of GwH2 to form H2
4. Mixed control by 1 and 2
5. Mixed control by 1 and 3
For each of items 1 through 5 above make a separate graph of the
hydrogen concentration profile in the copper foil. On each of graph draw
two curves: one valid at short times and one valid at long times. Pay strict
attention to relative magnitudes of concentration, relative slopes, intercepts at
surfaces, and other such graphical details.
Data:
Volume of left chamber, V ¼ 1 cm3
Area of the copper foil, A ¼ 1 cm2
Diffusivity of hydrogen in copper, DH ¼ 106 cm2 s1
The relationship between atomic hydrogen dissolved in copper and molecular
hydrogen gas in the atmosphere of the chamber is given by the reaction
References 143

H2 =2ðgasÞ ¼ Hðin copperÞ:

At 1000 C, the equilibrium constant for this reaction has the value,
K ¼ 1.4 ppm atm1/2.
Density of Cu: 8.93 gcm3
Density of Gw: 13.3 gcm3
Atomic mass of Cu: 63.55 g/mol
Atomic mass of Gw: 178.49 g/mol

References

1. J.E. House, Principles of Chemical Kinetics, 2nd edn. (Elsevier Inc., Amsterdam, 2007)
2. I. Amdur, G.G. Hammes, Chemical Kinetics (McGraw-Hill Book Co., New York, 1966)
3. C. Wagner, Lecture note, Kinetics in Metallurgy (MIT, Cambridge, 1955)
4. K.G. Denbigh, The Thermodynamics of the Steady State (Wiley, New York, 1965), p. 31
5. R.A. Alberty, J. Chem. Edu. 81, 1206 (2004)
6. L. Onsager, Phys. Rev. 37, 405 (1931)
7. S.A. Arrhenius, Z. Phys. Chem. 4, 96 (1889).; 4, 226 (1889).
8. H. Eyring, J. Chem. Phys. 3, 107 (1935)
9. S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers. J. Am.
Chem. Soc. 60, 309–319 (1938)
10. H.B. Callen, Thermodynamics and an Introduction to Thermostatics, 2nd edn. (Wiley,
New York, 1985), p. 28
11. G.R. Belton, Metall. Trans. B 78, 35–42 (1976)
Chapter 4
Diffusion in Concentration Gradients

4.1 Introduction

Suppose that we have a diffusion couple of a homogeneous 50:50 A/B alloy with its
left part involving, say, 2 atom % of A-tracers (A) together with 48 atom % normal
A and the right counterpart involving 2 atom % B-tracers (B) together with 48 atom
% normal B, as illustrated in Fig. 4.1a. When this couple is diffusion-annealed for a
time duration t, then we will have two antisymmetric, error-functional diffusion-
profiles for these two types of tracers, NA(x,t) and NB(x,t), in terms of mole
fractions, as shown schematically in Fig. 4.1b or
" !# " !#
x x
N B ¼ 0:01 1 þ erf pffiffiffiffiffiffiffi
ffi ; NA ¼ 0:01 1  erf pffiffiffiffiffiffiffi
ffi ð4:1Þ
2 DB t 2 DA t

We can subsequently determine thereby the tracer diffusivities, DA and DB ,
respectively. In general, DA 6¼ DB, of course; hence the spreads of the profiles differ
from each other.
Now, we have an interdiffusion couple A|B comprising pure A as the left part and
pure B as the right counterpart (see Fig. 4.2a). If we anneal it for a time duration t at
the same temperature, we will have a concentration profile NB(x,t) or NA(x,t)
(¼1NB), as shown in Fig. 4.2b. By using the Boltzmann-Matano analysis, we
can, then, determine one and only one interdiffusion or chemical diffusion coef-
ficient De at the same composition as the above tracer diffusion couple of homoge-
neous NB ¼ NA ¼ 0.5 alloy (Fig. 4.1a) as
R NB ¼0:5
1 xdNB
e ¼   NB ¼0
D ð4:2Þ
2t dN B
dx N ¼0:5 B

© Springer Nature Switzerland AG 2020 145


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_4
146 4 Diffusion in Concentration Gradients

Fig. 4.1 (a) A tracer diffusion couple of homogeneous 50:50 A/B alloy with its left part involving
2 a/o A and right counterpart 2 a/o B. (b) The tracer diffusion profiles expected in time t, NA(x,t),
and NB(x,t) in terms of mole fraction

NB NB
1 1

A B 0.5 dN B
dx N B =0.5

0 0
0 x x
(a) t=0 (b) t>0

Fig. 4.2 (a) Interdiffusion couple comprising pure A and pure B at t ¼ 0; (b) interdiffusion profile
R1
NB(x,t) therefrom in time t > 0, where x ¼ 0 is such that 0 xdNB ¼ 0 due to Matano

You see, from the tracer experiments in Fig. 4.1, we obtain the two tracer
diffusivities DA and DB corresponding to the overall alloy composition NB ¼ 0.5
but from the interdiffusion experiment in Fig. 4.2, only one interdiffusion or
chemical diffusion coefficient D e corresponding to the same composition
NB ¼ 0.5! In the latter, components A and B are both diffusing obviously with
different diffusivities, as in the tracer diffusion in Fig. 4.1, but only one common
diffusivity has been resulted! Why is that, and, if any, what is the relationship among
these three diffusivities, DA , DB , and D e ? We will explore the answer to this
question in this chapter. The road to the answer, we will see, is full of intellectual
beauty. Indeed, this is the climax of the diffusion story in solids.

4.2 Kirkendall Effect

It was in 1947 that Smigelskas and Kirkendall [1] reported a landmark discovery in
the history of diffusion in solids. They took a bar of 7:3 brass (0.7Cu + 0.3Zn),
wound tightly thin Mo-wires around it, and then electroplated pure copper upon the
4.2 Kirkendall Effect 147

Fig. 4.3 The interdiffusion



couple between pure Cu and • • • • • • • • • •
7:3 brass (0.7Cu + 0.3Zn)
with the inert Mo-markers jœGG^W
(•) in between. (From ‹
•GGZW
Smigelskas and Kirkendall
[1]) • • • • • • • • • •

Mo wire

bar with the Mo-wires wound so that the cross section may look as illustrated in
Fig. 4.3.
They then diffusion-annealed the bar at 785  C for a time duration t, cut a slice out
of the bar to measure the separation distance “d” between the two rows of Mo-wires
across the cross section of the brass bar, and repeated the procedure at different time
durations to find that the separation distance “d” ever decreases with time t. It was
then known that by alloying Cu with Zn, the mean molar volume increases. This
means that if Cu diffuses into or Zn diffuses out of the brass in Fig. 4.3, the volume
of the brass should shrink, thus, making “d” appear to decrease. Even after correcting
against this volume contraction, however, Smigelskas and Kirkendall found that “d”
still decreases with time as
pffi
d ¼ do  β t ð4:3Þ

where do is the initial separation distance at t ¼ 0 and β the proportionality constant.


In their experiment, for example, the two Mo-wire rows shifted toward each other by
0.125 mm in 56 days or β ¼ 0.00167 cm/day1/2 at 785  C.
Why is that? It was then a mystery. When the inner-working of a new phenom-
enon is not immediately clear, it is often called “such and such effect.” This
phenomenon has since been called the Kirkendall effect.1
Anyway, they could determine the interdiffusion coefficient as a function of
composition by using the Boltzmann-Matano analysis. For example, the row of
Mo-wires happened to be fixed at the composition of 0.225Zn, where D e
9
¼ 4.45  10 cm /s. You are strongly urged to read their original paper [1]
2

together with further amusing stories about the discovery and acceptance of this
effect [2].
But, this mystery did not survive that long. It is just 1 year later that the mystery
was cracked by Darken [3]. Let us see how he deciphered the mystery.

1
Why not called the “Smigelskas” effect? A joke goes, “It is because her name (Alexandra
Smigelskas) was too long.”
148 4 Diffusion in Concentration Gradients

t1 t2 t3 t4
t1
t2
t3
t4

(a) (b)

Fig. 4.4 Dispersion of milk as observed on the boat drifting together with the water stream (a) and
as observed from the bank (b), at different times t1 < t2 < t3 < t4

4.3 Darken’s Analysis


4.3.1 Diffusion and Drift

He starts his analysis by asking, “What is diffusion?” His argument is that diffusion
and drift should be distinguished. Suppose that water in a river flows at a velocity
v relative to the bank. You are now sitting on a boat that is drifting together with the
water stream at the same velocity v. At a certain point, you pour a bucketful of milk
onto the stream. The milk then spreads in the water stream due to its own concen-
tration gradient. You, on the boat drifting together with the stream at the same
velocity, will see a circle of milk ever expanding concentrically with its center fixed
relative to yourself. If you are sitting on the bank, on the other hand, things appear
different to you: the circle of milk, while ever expanding, drifts together with the
stream (see Fig. 4.4).
Now imagine that you put into the stream a unit reference plane to measure the
flux of milk, J. If you do it sitting on the boat, i.e., in the moving frame of reference
(denoted by the subscript “L”), what you measure will be solely due to diffusion
(concentric dispersion) or

e
¼ D∇c, ð4:4Þ
LJ

where c denotes the milk concentration. Back to our crystalline solids, the diffusion
medium which is drifting is the atomic lattice, and you, sitting on the alloy-lattice,
are observing Cu and Zn atoms jumping back and forth, thus denoted as “L”
referring to as the lattice frame of reference.
Sitting on the bank or in the laboratory frame of reference (denoted by the
subscript “F”), on the other hand, what you measure will be not only due to diffusion
but also due to the stream v itself or
4.3 Darken’s Analysis 149

e
¼ D∇c þ cv: ð4:5Þ
FJ

Why “cv” for the flux due to stream? It is because all the amount of milk within
the distance |v| normal to the unit reference plane, cv, passes this unit area per unit
time. Equation (4.5) is quite general when there is a drift or stream for whatsoever
causes, e.g., an electrical potential gradient ∇ϕ and a temperature gradient ∇T.
How could we then measure the stream or drift velocity, v? You may put
something insoluble and floating, or inert markers, e.g., corks, leaves, etc., on the
stream and measure its drift velocity relative to the bank. If these markers of different
nature drift at the same velocity, then you can say, for sure, that it should correspond
to the drift velocity of the water stream. Such inert markers are often called
Darken’s chips.

4.3.2 Analysis

Back to the Kirkendall effect, the rows of Mo-markers are drifting as Darken’s
chips do, thus, indicating that the matrix in the interdiffusion zone is drifting
somehow at a certain velocity v, like the water stream. Denoting Zn and Cu as
1 and 2, respectively, their fluxes as observed in the moving frame of reference L
are solely due to diffusion or

e1 ∂c1 e2 ∂c2
L J1 ¼ D ; L J2 ¼ D ð4:6Þ
∂x ∂x

for the present case of one-dimensional diffusion. Here, the diffusivity, denoted as
e k (k ¼ 1,2), is something new because there is now a gradient of concentration or
D
composition: Let us call it the intrinsic diffusivity or component diffusivity of
component k.
To the observer sitting on the bank or in the laboratory frame of reference, F, on
the other hand, these fluxes should appear with the drift effect superposed or.

e 1 ∂c1 þ c1 v;
¼ D e 2 ∂c2 þ c2 v:
¼ D ð4:7Þ
F J1 F J2
∂x ∂x

Here, we are assuming that there is no volume change with composition at all or,
in other words, the partial molar volumes of the components 1 and 2 are respec-
tively equal to the mean molar volume of the alloy, Vm or

V1 ¼ V2 ¼ Vm : ð4:8Þ
150 4 Diffusion in Concentration Gradients

Actually, this may be regarded approximately as the case in reality. Otherwise, there
would be a third effect on Jk due to the volume change itself via the diffusion
coordinate x and concentration ck.
Then, it is always true to an observer sitting on the bank that no matter whatsoever
happens internally,

1
c1 þ c2 ¼ co ¼ , ð4:9Þ
Vm

in terms of molar concentrations. This is a constant, thus,

∂ðc1 þ c2 Þ ∂  
¼ F J1 þ F J2  0 ð4:10Þ
∂t ∂x

It means that the sum of the fluxes as observed in the laboratory reference
frame (F) should be a constant I anywhere. Now you see there happens nothing far
outside the diffusion zone around the inert-markers, say, at each end of the
diffusion couple. Therefore, this constant I has to be 0 or

F J1 þ F J2 ¼ I ¼ 0: ð4:11Þ

This is actually nothing but the condition of no volume change with diffusion,
i.e.,
 
V1 F J1 þ V2 F J2 ¼ Vm F J1 þ F J2 ¼ 0: ð4:12Þ

If there were a net volume change, then the dimension of our interdiffusion
couple would have changed. Strictly speaking, Fick’s law is valid only for the case
of Eq. (4.12). The reference frame as defined by Eq. (4.12) is accordingly called the
volume-fixed reference frame or Fick’s reference frame, thus, denoted as “F.” Do
you remember that this equation has defined the Matano-interface in the
Boltzmann-Matano analysis of an interdiffusion profile (Fig. 4.2b)?
Substituting Eq. (4.7) into Eq. (4.11), one obtains.
 
1 e ∂c1 e ∂c2
v¼ D1 þ D2 ð4:13Þ
c1 þ c2 ∂x ∂x

or, noting that c1/(c1 + c2)¼N1, the mole fraction of component 1, and ∇c1 + ∇c2 ¼ 0,
   
v¼ D e 2 ∂N1 ¼ D
e1  D e 1 ∂N2 :
e2  D ð4:14Þ
∂x ∂x
4.3 Darken’s Analysis 151

This means that if D e1 > De 2, v is in the direction of increasing N1, and vice versa.
That is, the diffusion matrix or local lattice drifts up the gradient of the faster
component or against the faster diffusing component. In the Kirkendall experi-
ment, the Mo-markers moved toward the brass end, thus indicating that Zn is faster
than Cu. You see, this is an unavoidable consequence of Eq. (4.11) or Eq. (4.12): If
you let the faster component diffuse faster, say, from right to left in Fig. 4.2, an
emptied volume or vacant lattice sites would have to be created to the right and a
deficit of volume for the excess material to the left. The only way to keep the overall
system contiguous is to let the matrix shift rightward as a whole by as much as the
emptied volume so that the surplus amount of the faster component may be accom-
modated exactly into the left. By doing that, the faster component moving leftward
appears to be retarded and the slower component moving rightward to be expedited
so that the two components diffuse against each other with the apparently identical
diffusivity, and this is why one has a single, common diffusivity in the interdiffusion
experiment. This kind of drift phenomenon is considered to be rather general than
special. The same thing should happen even in a liquid-state interdiffusion couple.
Finally, substitution of Eq. (4.14) into Eq. (4.7) leads to
 
e e ∂c1
J
F 1 ¼  N D
2 1 þ N D
1 2 ¼  F J2 ð4:15Þ
∂x

Comparing with Fick’s first law, you can recognize that there appears a single
apparent diffusivity for both components:

D e 1 þ N1 D
e ¼ N2 D e 2: ð4:16Þ

Upon comparison of Eq. (4.15) with Eq. (4.7), you may immediately recognize
that this apparent De involves the information on the diffusion of each component and
the drift effect of the matrix as well. This apparent diffusivity is called the interdif-
fusion coefficient because it is a measure of interdiffusion rate or chemical diffu-
sion coefficient because it is a measure of diffusion under chemical gradient. As we
have already expected, the interdiffusion coefficient is neither larger than the diffu-
sivity of the faster nor smaller than the slower. Combination of the two component
diffusivities in this way is said to be a Darken type. (There is another type of
combination called a Nernst-Planck type due to the flux coupling to keep the local
charge neutrality when diffusing components are electrically charged as in ionic
solids, as discussed later in Chap. 7.)
We can then evaluate the component or intrinsic diffusivities D e 1 and De 2 from
this interdiffusion measurement by measuring the drift velocity v and the interdif-
fusion coefficient D e via Eqs. (4.14) and (4.16). Let us do it from the Smigelskas-
Kirkendall experiment of Fig. 4.3. We should first evaluate the drift velocity
v. Their interdiffusion couple may be regarded as comprising a semi-infinite bar of
pure Cu and another of 7:3 brass as long as the diffusion zone is narrow enough
compared to the overall length of the couple, which is fixed at the laboratory frame
152 4 Diffusion in Concentration Gradients

of reference. Letting xm denote the position of, say, the upper row of Mo-markers in
Fig. 4.3 as measured from its initial position (xm ¼ 0) at t ¼ 0 in the laboratory
frame of reference, the as-observed result, Eq. (4.3) may be rewritten as

1 1 pffi
xm ¼ ð do  dÞ ¼ β t : ð4:17Þ
2 2

You remember? The Boltzmann transform of the variables x and t into a single
variable η(x/t1/2) has turned Fick’s second law,
 
∂c ∂ ∂c
¼ D , ð4:18aÞ
∂t ∂x ∂x

into an ordinary differential equation in terms of the single variable η as


 
1 dc d dc
 η ¼ D : ð4:18bÞ
2 dη dη dη

pffi ¼ c(η) and, hence, if η is fixed, c is fixed. Now, dividing


This implies that c(x,t)
Eq. (4.17) through by t, you find that

xm 1
pffi ¼ ηm ¼ β: ð4:19Þ
t 2

That is, the Mo-markers stay all the way at constant ηm ¼ β/2, thus, at the same
composition c(ηm) ¼ c(β/2). According to the original observation, N1(ηm) ¼ 0.225.
The velocity vm of the markers is then obtained by differentiating Eq. (4.17) as

dxm β x β2
vm ¼ ¼ pffi ¼ m ¼ : ð4:20Þ
dt 4 t 2t 8xm

You see, the markers, which were initially placed at the original interface or join
of the couple, slow down with time as t1/2 (i.e., vm/1/t1/2) or with shift as xm (i.e.,
vm/1/xm).
Smigelskas and Kirkendall found that xm ¼ 0.125 mm at t ¼ 56 days; thus, due
to Eq. (4.20)

xm
vm ¼ ¼ 1:29  109 cm=s ð4:21Þ
2t

at the composition of 0.225Zn (N1 ¼ 0.225) where the slope of the interdiffusion
profile was read to be dN1/dx ¼ 0.43 cm1. Via the Matano analysis, the interdif-
fusion coefficient at this composition was evaluated as.
4.3 Darken’s Analysis 153

e ¼ 4:45  109 cm2 =s:


D ð4:22Þ

Equations (4.14) and (4.16), thus, finally yield the numerical values for the
intrinsic diffusion coefficients at this specific composition as.

e Cu ¼ 2:2  109 cm2 =s;


D e Zn ¼ 5:1  109 cm2 =s:
D ð4:23Þ

As expected from the direction of marker movement or Eq. (4.14), the intrinsic
diffusion coefficient of Zn is indeed larger by 2.3 times than that of Cu at the
composition 0.225Zn.
It is noted from Eq. (4.16) that

e¼D
lim D e Zn ð4:24Þ
NZn !0

which, by definition, should correspond to the tracer diffusivity of Zn, DZn in


Cu. Likewise, DCu in Zn as NCu ! 0. The as-measured tracer diffusivity of Zn in Cu
is DZn ¼ 3  1010 cm2/s. You can see that D e Zn has increased by 17-fold as NZn
increases from 0 to 0.225. Can you tell why?
Let us here consider the relationship between the original interface of the diffu-
sion couple where the Mo-markers were implanted (xm ¼ 0) and the Matano-
interface which was located through Matano-analysis of an interdiffusion profile.
Recall that the Matano-interface is defined in the present case, Fig. 4.3, as
Z 0:3
xdNZn ¼ 0: ð4:25Þ
0

It means that across the Matano plane x ¼ 0, the number of Zn-atoms or, due to
Eq. (4.8), the volume transferred leftward is exactly balanced by the number of
Cu-atoms or volume transferred rightward, just describing Eq. (4.11) in words.
Therefore, the Matano-interface as located in this way is exactly the original
interface or join where the Mo-markers were implanted before diffusion anneal or
xm ¼ 0.
Final thing to note: Strictly speaking, the reference frame in which the net flux of
particles becomes nil or the particle fluxes are balanced, i.e., Σk NJk ¼ 0, is called the
number-fixed reference frame (N); one Pin which the net flux of volume disappears
or volume fluxes are balanced, i.e., Vk F Jk ¼ 0 , is called the volume-fixed
reference frame. In the present case, these are the same as we have assumed no
volume change with composition or Eq. (4.8). Otherwise, they would differ, of
course.
154 4 Diffusion in Concentration Gradients

4.4 Thermodynamic Treatment of Diffusion

We have learned that the self-diffusivity of a species is a measure of its total


successful jump frequency, Γ, in a homogeneous medium (i.e., in the absence of
any composition gradient), and it is normally measured in terms of the tracer
diffusivity, which is somewhat smaller than the self-diffusivity due to the possible
correlation effect for a given mechanism in a given crystallographic structure. And,
we have just seen that the component diffusivity of Zn in Cu increases by 17-fold as
NZn increases from 0 to 0.225. Wouldn’t it mean that there is a sort of bias which
makes Zn jump more frequently than at random down its own concentration
gradient? Let us now explore the interrelationship and inner-working.
Any diffusion flux of, say, component k, LJk in Eq. (4.6) may be written, in
general, as

L Jk ¼ c k vk : ð4:26Þ

Here, vk is the drift velocity of k-type particles, but is not to be confused with the
drift velocity “v” of a diffusion matrix in Eq. (4.7). According to mechanics, the
velocity of a particle in the stationary state, following a brief transient state of
acceleration under the action of an external force, is linearly proportional to the
force Fk acting on the particle,2 and the force is equal to the negative gradient of
potential energy of the particle, which, in a thermodynamic system under isothermal
and isobaric conditions, is the electrochemical potential of k, ηk, in general, or
chemical potential μk if k is not charged as in metals3 in particular, or

vk ¼ Bk Fk ¼ Bk ð∇μk Þ: ð4:27Þ

The proportionality constant Bk is called the mechanical mobility having dimen-


sions4 of (m/s)/(Newton/mol). We recollect that the chemical potential μk is the
partial molar Gibbs free energy of k. If we represent the force Fk per-particle
instead of per-mole by dividing the chemical potential by the Avogadro number NA,
Eq. (4.27) may be rewritten equivalently as
 
1
vk ¼ bk  ∇μk : ð4:28Þ
NA

2
Consider the stationary state (dvk/dt ¼ 0) solution to the equation of motion, Fk ¼ mk dvdtk þ τk vk ,
where mk is the mass of the particle k and τk(¼1/Bk) the friction coefficient of the matrix or
medium.
3
For the diffusion in charged particles systems, see Chap. 7.
4
Here, the SI unit of force is written in its full-fledged form, “Newton” instead of its abbreviation N
not to be confused with the mole fraction or others.
4.4 Thermodynamic Treatment of Diffusion 155

Here, bk is basically the same thing as the mechanical mobility but with a bit
different units (m/s)/(Newton/particle) and called a bit differently the fundamental
mobility. Do not forget, however, that these two mobilities differ by a factor of the
Avogadro number, NA ¼ 6.02  1023, i.e., bk ¼ NABk!
Due to Eq. (4.27), or equivalently Eq. (4.28), therefore, Eq. (4.26) is rewritten as.

L Jk ¼ ck Bk ∇μk ð4:29Þ

under the isothermal and isobaric conditions.5 This means that a component diffu-
sion (LJk) is induced by a gradient of its chemical potential (∇μk) or escaping
tendency, not necessarily by its concentration gradient (∇ck), in the direction of
lowering escaping tendency (∇μk), not necessarily down the concentration gradi-
ent (∇ck). This is actually the thermodynamically legitimate definition of diffusion.
In an isothermal and isobaric condition,

∇μk ¼ RT∇ ln ak ð4:30Þ

where ak is the thermodynamic activity of component k. By substituting Eq. (4.30)


into Eq. (4.29) and transforming the result into the form of Fick’s first law, we
obtain
 
∂ ln ak e k ∇ck :
J
L k ¼ c B
k k ∇μ ¼ Bk RT ∇ck  D ð4:31Þ
k
∂ ln ck T,P

for the present binary system (e.g., k ¼ Cu, Zn) with a single composition variable ck
(i.e., either Cu or Zn) at given temperature and pressure. The component diffusivity
is, thus, read as
 
e ∂ ln ak
Dk ¼ Bk RT  Bk RTf k ðθÞ ð4:32Þ
∂ ln ck

As we have suspected, the component diffusivity D e k indeed involves a unfamil-


iar factor, fk(θ), viz., the composition (ck) dependence of the thermodynamic
activity of that specific component, ak. We call it the thermodynamic factor as
both ak and ck are the thermodynamic quantities. Note that it has popped up in

5
If non-isothermal, the component flux would be written, according to irreversible thermodynamics,
as
 μ   
1
Jk ¼ Lkk T∇ k þ Lkq T∇ ,
T T

where Lkk and Lkq are the Onsager coupling coefficients. Interested readers are referred to Lee
et al. [4], and the references therein.
156 4 Diffusion in Concentration Gradients

Eq. (4.31) simply as a conversion factor of the thermodynamic force, ∇μk, to the
Fickian force, ∇ck, at constant temperature and pressure.
In a condensed solution, a component activity is generally given as

ak  γk Nk ¼ γk ck Vm ð4:33Þ

where γk and Nk are the activity coefficient and the mole fraction of k, respectively.
The thermodynamic factor, fk(θ), may, thus, be factored into
     
∂ ln ak ∂ ln γk ∂ ln Vm
f k ð θÞ  ¼1þ þ : ð4:34aÞ
∂ ln ck T,P ∂ ln ck T,P ∂ ln ck T,P

In diffusion systems, we usually assume the constant molar volume Vm or no


volume change; hence,
 
∂ ln γk
f k ð θÞ ¼ 1 þ : ð4:34bÞ
∂ ln ck T,P

Note that fk(θ) ¼ 1 exactly (under the assumption of Vm being constant) if our
alloy system is ideal (i.e., γk ¼ 1) or Henrian (i.e., γk¼γok , constant).
Particularly for a binary system, e.g., Cu-Zn, at given temperature and pressure,
the Gibbs-Duhem equation stipulates that

c1 dμ1 þ c2 dμ2 ¼ 0 ð4:35aÞ

or, due to Eq. (4.30),

c1 d ln a1 þ c2 d ln a2 ¼ 0: ð4:35bÞ

Dividing through by dc1 and noting that dc2 ¼ dc1 when the molar volume Vm is
constant, Eq. (4.35b) takes the form

∂ ln a1 ∂ ln a2 ∂ ln a1 ∂ ln a2
c1 þ c2 ¼  ¼ 0: ð4:36Þ
∂c1 ð∂c2 Þ ∂ ln c1 ∂ ln c2

Thus,

f 1 ð θÞ ¼ f 2 ð θÞ  f ð θÞ ð4:37Þ

for a binary system.


Now, let us describe the tracer diffusion experiment in Fig. 4.1 in the light of
Eq. (4.31). The diffusion flux of tracer k(¼A or B) and of normal k may be
respectively written, due to Eqs. (4.31) and (4.34b), as
4.4 Thermodynamic Treatment of Diffusion 157

" 
 #
∂ ln γk
Jk ¼ Bk RT
1þ ∇ck ; ð4:38aÞ
∂ ln ck ck þck
"   #
∂ ln γk
Jk ¼ Bk RT 1 þ ∇ck : ð4:38bÞ
∂ ln ck ck þck

We know that a mixture of k and k is just an ideal mixture of, say, white and red,
otherwise identical, balls,6 and ∇ck þ ∇ck ¼ 0 in the present homogeneous alloy.
The activity coefficient γk should depend on the overall chemical composition or
total concentration of k, ck þ ck , not on the isotope ratio ck =ck . That is, γk ¼ γk
¼ constant as long as ck þ ck remains constant or
   
∂ ln γk ∂ ln γk
¼0¼ : ð4:39Þ
∂ ln ck ck þck ∂ ln ck ck þck

Equation (4.38) is, thus, rewritten as

Jk ¼ Bk RT∇ck ; Jk ¼ Bk RT∇ck , ð4:40Þ

and in comparison with Fick’s first law, we can read off the tracer diffusivity and
self-diffusivity as

Dk ¼ Bk RT; Dk ¼ Bk RT ð4:41Þ

Noting that Jk + Jk ¼ 0 and ∇ck + ∇ck ¼ 0 in the present tracer experiment in
Fig. 4.1, it turns out that Dk ¼ Dk and Bk ¼ Bk . We know, however, that the two
should differ by the correlation factor f or Dk ¼ fDk , suggesting something is
missing in Eq. (4.31). It is indeed so in the strict sense. Everything would turn to fit
with each other, however, in the full-fledged irreversible thermodynamic treatment
[5, 6]. To make the story simpler now, we will neglect the correlation effect for the
time being and keep using the tracer diffusivity Dk and the self-diffusivity Dk
interchangeably.
Nevertheless, you may still wonder how the relations in Eq. (4.41) can be,
because, as is seen in Eq. (4.27), the mobility Bk is a measure of the directional
walk or drift in the direction of the force applied (Fk ¼ ∇μk), but the self-diffusivity
Dk is a measure of nondirectional, random walk (Dk ¼ Γkα2/6). Well, the higher the

6
These isotopes are chemically identical but have different nuclear masses and, thus, have different
mobilities as the heavier isotope moves more slowly. According to a simple harmonic oscillator
model, the natural vibrational frequency, and thus diffusion jump frequency, is inversely propor-
tional to the square root of isotope mass. This is called the isotope effect, but unless the atomic mass
is small and hence the difference in isotope masses is relatively large, say, H and D, the effect is
usually smaller than the usual experimental error. For a detailed treatment, see, e.g., H. Mehrer,
Diffusion in Solids, Springer-Verlag, 2007, Chap. 9.
158 4 Diffusion in Concentration Gradients

Fig. 4.5 General trend of ln γ k


variation of lnγk vs. mole
fraction Nk in a condensed (c)
ln γ ok
solution when ideal (a),
(a)
negatively deviated (b), and 0
positively deviated (c). Note (b)
that (∂lnγk/∂Nk) ¼ 0 if ln γ ok
ideal, >0 if negatively
deviated, and < 0 if
positively deviated 0 Nk 1

jump frequency Γk, the more mobile no matter whether homogeneous or inhomo-
geneous! This unexpected relationship between the directional walk (Bk) and the
random walk (Dk), Dk ¼ BkRT in Eq. (4.41) was first recognized independently by
A. Einstein in 1905 and M. von Smoluchowski in 1906 in their works [7] on
Brownian motion, thence is called the Einstein relation (or Einstein-
Smoluchowski relation).
So, we have now come to know the relation between the component diffusivity
and self-diffusivity, due to Eqs. (4.32) and (4.41), as

e k ¼ Dk f k ðθÞ,
D ð4:42Þ

and hence, the chemical diffusivity in a binary system is

e ¼ ðN1 D2 þ N2 D1 Þf ðθÞ:
D ð4:43Þ

It has become clear from Eq. (4.42) that the component diffusivity in a concen-
tration gradient is apparently enhanced by the thermodynamic factor. In this
respect, the thermodynamic factor is often referred to as the diffusion enhance-
ment factor. Whether enhanced or suppressed does, of course, depend on whether
f(θ) > 1 or < 1.

4.5 Thermodynamic Factor

We have seen the thermodynamic factor, Eq. (4.34b), being introduced as a


conversion factor of a chemical potential gradient ∇μk to the corresponding concen-
tration gradient ∇ck and appear to enhance (or suppress) the self-diffusivity as is in
Eq. (4.42). What does this factor mean after all?
You remember? As a solution gets dilute with respect to component k or as
ck ! 0, the activity coefficient approaches a constant value, Henry’s law constant
or γk ! γko, and as ck ! 1, γk ! 1 due to Raoult’s law. Noting that 0 < γk < 1 for the
negative deviation of the solution behavior and γk > 1 for the positive deviation,
lnγk varies against mole fraction Nk schematically as shown in Fig. 4.5.
4.5 Thermodynamic Factor 159

When the solution is negatively deviated or its components, solute and solvent
like, thus, attract each other, then fk(θ) > 1 (see Eq. (4.34b)); When positively
deviated, they dislike, thus, repel each other, fk(θ) < 1. In the case of negative
deviation, the diffusing species is attracted by the other component, and hence, it
should feel a bias toward further mixing, thus, making it jump more frequently down
its concentration gradient to enhance the interdiffusion and vice versa for the positive
deviation to suppress the interdiffusion. Indeed, the thermodynamic factor acts as a
thermodynamic bias to otherwise random jumps, stemming from the nature of
mutual interaction between the end member components.
An extreme situation of the positive deviation of a binary solution is clustering or
splitting into the two end member phases. For this to happen, each component
particles flock together instead of intermingling, thus, by diffusing up their concen-
tration gradients! How can it be?
According to solution thermodynamics, the relative integral molar free
energy of mixing, ΔGM of a binary solution is given in terms of the relative partial
M
molar free energy of mixing of each component, ΔGk (k ¼ 1,2) as

M M
ΔGM ¼ N1 ΔG1 þ N2 ΔG2 ¼ RTðN1 ln a1 þ N2 ln a2 Þ: ð4:44Þ

Differentiating both sides with respect to N1 at given temperature and pressure, we


have

1 ∂ΔGM
 ¼ ln a1  ln a2 ð4:45Þ
RT ∂N1
M M
as dN1 + dN2 ¼ 0 and N1 dΔG1 þ N2 dΔG2 ¼ 0 due to the Gibbs-Duhem equa-
tion. Differentiating once more Eq. (4.45) with respect to N1, we have

2
∂ ln a1 ∂ ln a2 1 ∂ ΔGM
þ ¼  : ð4:46aÞ
∂N1 ∂N2 RT ∂N21

A little modification, with Eq. (4.36) in mind, leads to


  2
∂ ln a1 ∂ ln a2 1 1 1 ∂ ΔGM
þ ¼ þ f ð θÞ ¼  ð4:46bÞ
N1 ∂ ln N1 N2 ∂ ln N2 N1 N2 RT ∂N21

or

2
N1 N2 ∂ ΔGM
f ð θÞ ¼  : ð4:47Þ
RT ∂N21
160 4 Diffusion in Concentration Gradients

You see, the thermodynamic factor is nothing but the curvature of the integral
mixing free energy ΔGM vs. composition N1 or N2(¼1-N1) curve: If convex
downward, f(θ) > 0; if convex upward, f(θ) < 0.
In order to better understand this aspect, let us take a regular solution for which
the analytic expression for the mixing free energy is analytically available as

ΔGM ¼ RTðN1 ln N1 þ N2 ln N2 Þ þ αN1 N2 , ð4:48Þ

where α7 is a constant. The thermodynamic factor is calculated due to Eq. (4.47) as

f ðθÞ ¼ 1  2α0 N1 N2 ð4:49Þ

where

α
α0  : ð4:50Þ
RT

If α ¼ 0, the solution is ideal; if α < 0, the solution is negatively deviated from the
ideal behavior. In these cases, the mixing curve is convex downward across the
entire composition range, thus, with a positive curvature or

2
∂ ΔGM
> 0, ð4:51Þ
∂N21

and hence,

f ðθÞ ¼ 1  2α0 N1 N2  1 ð4:52Þ

The sharper the curve, the larger the thermodynamic factor and, hence, the more
enhanced the diffusivity is.
The regular solution shows the mixing free energy vs. composition curve as in
Fig. 4.6 when α0> α0cr ¼ 2 or T is lower than the critical temperature Tcr(¼α/2R).
Actually, the curvature starts to become negative at NA ¼ NB ¼ 1/2 as soon as α’
exceeds the critical value, α0cr ¼ 2.
2
The boundary between the region of positive curvature (∂ ΔGM =∂N2A > 0) and
2 2
negative one (∂ ΔG =∂NA < 0), or the inflection point where ∂ ΔGM =∂N2A ¼ 0,
M 2

is called a spinodal. You can see that f(θ) < 0 between the two spinodals, thus,
e < 0,
D that is, the interdiffusion coefficient becomes negative! When
2
∂ ΔGM =∂N2A < 0, the system is thermodynamically unstable with respect to
minor fluctuation and, hence, decomposes into the two nearby equilibrium compo-
sitions via so-called the spinodal decomposition. For this decomposition to occur,
components A and B should diffuse up the gradients of their own, respectively,

7
Not to be confused with the jumping distance α in Chap. 2
4.6 Evidence for –∇μ, Rather than –∇c, Being the Diffusional Driving 161

Fig. 4.6 The mixing free


energy vs. composition of a
regular solution when α0 > 2 ∆G M

A B

∂ 2 ΔG M ∂ 2 ΔG M ∂ 2 ΔG M
>0 <0 >0
∂ N 2A ∂N 2A ∂N 2A

rendering the diffusivity to appear negative. This uphill-diffusion-involved spinodal


decomposition will be treated in detail in Chap. 5.

4.6 Evidence for 2—μ, Rather than 2—c, Being


the Diffusional Driving Force

Are you now convinced by Eq. (4.29), viz., that the diffusion flux Jk is, in general,
caused by a negative gradient of chemical potential, ∇μk, rather than by a negative
gradient of concentration, ∇ck, as Fick’s first law indicates? It was actually first
suggested by Einstein [8, 9].
If you see a component k really diffusing even when ∇ck ¼ 0 while ∇μk 6¼ 0, then
you will be, no doubt, convinced by Eq. (4.29). Wait! If ∇ck ¼ 0, then ∇μk ¼ 0? Yes,
but it is the case only for an isothermal, isobaric binary system because there is only
one composition variable ck. For a ternary or higher systems, however, there are two
or more composition variables. Therefore, even if ∇ck ¼ 0, ∇μk 6¼ 0 as long as
∇cj 6¼ k 6¼ 0.
Here is a beautiful experimental demonstration of this fact: If an iron-silicon-
carbon bar is welded to an iron-carbon bar of the same carbon content and heat-
treated, the initially uniform carbon content is disturbed to induce a nonuniform
carbon distribution as shown in Fig. 4.7. Note that this is a ternary system. Even
though there was no carbon concentration difference across the interdiffusion couple
(Fe, Si, C)|(Fe, C) initially, carbon has diffused from (Fe, Si, C) to (Fe, C), even
creating a concentration discontinuity across the join. This is because C prefers Fe to
Si or, thermodynamically speaking, Si has increased μC, thus, causing ∇μC 6¼ 0,
while ∇cC ¼ 0 across the join, and hence, C is escaping from (Fe, Si, C) to (Fe, C).
This elegant experiment has been made by Darken [9, 10].
162 4 Diffusion in Concentration Gradients

carbon concentration/ %
0.6
0.586% C
3.80% Si
0.5

0.478% C 0.441% C
0.4

0.3 0.315% C
-2 -1 0 1 2
distance from weld/ cm

Fig. 4.7 Nonuniform distribution of carbon produced from initially uniform distribution. Carbon
migrates from Fe-Si-C alloy to Fe-C alloy for 13 days at 1050  C. Redrawn from L. S. Darken
[9, 10]

4.7 Atomistic Interpretation of the Kirkendall Effect

We have learned that when the intrinsic diffusion coefficients of the counter-diffusing
species are different, the diffusion matrix drifts against the faster species to keep the
contiguity of the matrix or, in other words, to keep the overall volume constant.
Macroscopically, things happening would be as illustrated in Fig. 4.8: Suppose
that we have a diffusion couple comprising pure A and B and, say, D ~ A ¼ 2D~B
(Fig. 4.8a). In time t, “A” would then have diffused toward the B-side two times
more than “B” toward the A-sides, thus, say, 2 units of volume, 2 V moved by A to the
right and 1 unit of volume, 1 V by B to the left, leaving 1 V vacant in the left-hand side
and 1 V surplus in the right-hand side (see the schematic in Fig. 4.8b). In order to keep
the contiguity of the system as a whole, the interdiffusion zone would then have to shift
leftward as much as to fill up the vacant room 1 V to the left, thus accommodating the
surplus volume 1 V to the right. By doing that the interdiffusion couple has been
rendered contiguous again (Fig. 4.8c).
The drift velocity of the diffusion zone, v, would then be, assuming the constant
molar volume of the solution Vm of course,

A B (a)

(b)

(c)

Fig. 4.8 Phenomenological inner-working of the Kirkendall effect. (a) Interdiffusion couple A|B at
t ¼ 0; (b) assuming D ~ B , 2 units of A would have diffused to the right and 1 unit of B to the
~ A ¼ 2D
left in time t, thus, leaving 1 unit vacant to the left and 1 surplus unit to the right; (c) the entire
system is rendered to be contiguous by shifting the interdiffusion zone to the left by 1 unit, as
indicated by the arrow in (b)
4.7 Atomistic Interpretation of the Kirkendall Effect 163

Fig. 4.9 (a) Concentration


profile cB(x,t) at time t in an
interdiffusion couple A|B;
(b) corresponding fluxes
LJA, LJB, and LJV such that
LJA + LJB + LJV ¼ 0; (c)
temporal variation, ∂CV/∂t
(¼  r) along the diffusion
couple if vacancies were to
be conserved(dotted line)
and the generation (r > 0)
and annihilation rate(r < 0)
to keep the local defect
equilibrium (solid line)

 
v ¼  L JA þ L JB Vm : ð4:53Þ

You see, (LJA + LJB)Vm would correspond to the net volume transported per unit
time if everything were set free, but this volume transfer should be cancelled out by
the matrix sweeping back by a distance v per unit time.
How come would this kind of thing happen in a crystalline lattice? It is conve-
nient to think of a vacancy mechanism. When A and B are diffusing via vacancies,
then it should be always true that

L JA þ L JB þ L JV ¼ 0 ð4:54Þ

because the density of lattice points should remain constant in a given crystallo-
graphic structure in whatsoever way A and B are diffusing or

1
cA þ cB þ cV  co ¼ ð4:55Þ
Vm

in terms of molar concentrations. This is called the principle of local lattice


conservation. Note that normally cV < <cA + cB.
For the diffusion couple A|B as in Fig. 4.8a, the concentration profile CB(x,t)
may look like Fig. 4.9a in time t. The flux of A, LJA to the right and that of B, LJB to
the left will then be schematically as shown in Fig. 4.9b due to Fick’s first law: As
164 4 Diffusion in Concentration Gradients

~A > D
D ~ B, |LJA| > |LJB|, thus, resulting in a net flux of vacancies, LJV to the left due to
Eq. (4.54) or
 
L JV ¼  L JA þ L JB < 0: ð4:56Þ

This net vacancy flux, induced only because D ~ A 6¼ D


~ B , would alter the local
vacancy concentration cV(x,t), if conserved, at a rate

∂cV ∂ JV
¼ L ð4:57Þ
∂t ∂x

which is schematically as shown in Fig. 4.9c (dotted line): The net vacancy flux
tends to increase the vacancy concentration to the left and reduce to the right. The
vacancy concentration, however, should remain fixed at the local equilibrium value
cV ðx, tÞ ¼ ceq
V ðxÞ as determined by the local composition at given temperature and
pressure if and only if the local equilibrium prevails. Any excess vacancy should
therefore be annihilated and any deficit compensated by generation to keep the local
defect equilibrium. The annihilation of vacancies annihilates the corresponding,
already-existing lattice points, and the generation of vacancies generates new lattice
points, thus, rendering the lattice in the diffusion zone to drift, as a whole against the
faster, or leftward in the present case, to fill up the space otherwise being left vacant.
For the vacancies which may be generated or annihilated, i.e., not conserved, the
continuity equation may be written as.
 
∂cV ∂cV
¼ ∇L JV þ r; r ð4:58Þ
∂t ∂t generation

where r denotes the generation rate defined as such. As ∂cV =∂t ¼ ∂ceqV =∂t ¼ 0 due
to the local equilibrium, thus, the vacancy generation rate is rendered to

r ¼ ∇L JV : ð4:59Þ

One can see in Fig. 4.9c where vacancies should be generated (r > 0) and where
annihilated (r < 0), thus, pushing the matrix in between toward the latter at a certain
velocity v. The vacancy flux may then be equated to v as

v ¼ V m  L JV ð4:60Þ

or, due to Eq. (4.56),

   
v ¼ Vm L JA þ L JB ¼ D ~ B ∂NA
~A  D ð4:61Þ
∂x
4.7 Atomistic Interpretation of the Kirkendall Effect 165

by noting VmcA ¼ NA. You should now understand that the vacancy mechanism is
not the necessity for the matrix drift, thus for the Kirkendall effect, which actually
may happen in liquids and gases. If interested you may refer to Hartley and Crank
[11], who independently derived the Darken-type diffusion coefficient and coined
the “intrinsic diffusion coefficients.”

Fig. 4.10 Illustration of the


Kirkendall effect via edge
dislocation climb. When
the atom at the end of the
edge dislocation (⊥) in the
A-side diffuses to the end of
^
the dislocation ( ) in the
B-side, the former shrinks
(as indicated by a small
arrow), and the latter
elongates (as indicated by
another arrow), thus,
making the lattice in
between shift to the left or
against the faster moving
species

Let us consider at this point how vacancies keep being annihilated or generated to
keep the local defect equilibrium (cV ðx, tÞ ¼ ceq V ðxÞ ) and how effectively they
do. For vacancies to be generated or annihilated, they need their sources or sinks
such as dislocations, grain boundaries, and surfaces whether internal or external,
which are called the repeatable growth sites. An idealized heuristic picture may be
via edge dislocation climb. Suppose that there is an edge dislocation or an extra
^
atomic plane (⊥, ) in each side of an interdiffusion couple A|B, in parallel with its
initial join, where D~A > D ~ B . If you like, you may imagine a single atomic plane
^
diffusion couple, instead, with an extra atomic line (⊥, ) at each side, as illustrated
in Fig. 4.10a. As D~A > D~ B, there should be a net flow of atoms from the A-side to the
B-side. Thus, suppose that the atoms sitting at ⊥ (marked in red) in the A-side
^
diffuse, one by one from the tip of ⊥ to the B-side to land at the tip of one by one
^
(marked in red), thus, shortening ⊥ and elongating (see Fig. 4.10b). This process
^
eventually makes ⊥ disappear and grow to a full atomic line or plane. That is, a
line of atoms disappears in the A-side, and an extra line of atoms appears in the
B-side, thus, shifting the matrix in the diffusion zone shift to the A-side or against the
faster moving component A. This is like that the vacancies are generated (or A-atoms
^
settle down) at the tip of the dislocation , thus, expanding the extra plane in the
B-side and annihilated by landing (or A-atoms leaving) at the tip of the dislocation
166 4 Diffusion in Concentration Gradients

⊥, thus, shrinking the extra plane in the A-side. For a more realistic and amusing
discussion, you may be referred to Bardeen and Herring [12].
In any case, for the excess vacancies to annihilate, they have to first diffuse to the
repeatable growth sites. If our diffusion media is endowed with many enough grain
boundaries and dislocations as in fine-grained polycrystalline alloys, then local
defect equilibrium may be achieved with relative ease particularly at elevated
temperatures, thus, letting the inert markers drift as they should do according to
Eq. (4.61). For the experimental demonstration, you may refer to da Silva and
Mehl [13].

Fig. 4.11 (a) Contour


alteration (black line to red L JV
one) and inert marker shift
(black dots to red ones) of an (a)
interdiffusion couple A|B
when the only source or sink
were the external surface.
Note that the surface bulges
in the B-side and sinks in the
A-side because D ~A > D ~ B.
(b) If vacancy diffusion to
and from the surface were J (b)
L V
completely blocked, then
Kirkendall pores (indicated
by ○) would form instead of
the contour alteration as in
(a) with no marker-shift. (c)
In reality, however, excess
vacancies annihilate partly
at the surface altering the J
L V
contour and partly by (c)
forming pores inside

What would otherwise happen? Let us consider a perfect single crystal which has
no repeatable growth site internally. Then, the only available sink or source of
vacancies should be the external surface of the crystal: Vacancies should be gener-
ated from the surface of the B-side, resulting in a bulge (as diffusing-in atoms
accumulate there), and diffuse to the surface of the A-side, resulting in a sink
(as atoms there diffuse away), thus, changing the crystal contour as illustrated in
Fig. 4.11a. As a consequence, only the inert markers located near the surface would
have shifted, thus, curving the line of markers as indicated in the figure.
It is also painstaking to diffuse to the surface if the diffusion distance is too long.
Then, the excess vacancies exceeding their equilibrium concentration, or supersat-
urated vacancies, may likely condense to form pores, instead of changing the
external contour as in Fig. 4.11b. In this case, the markers would not move at all.
In reality, excess vacancies may annihilate themselves partly at the surface and partly
by forming pores as illustrated in Fig. 4.11c.
4.8 Summary of the Various Diffusivities 167

This kind of pores formed as such is often called the Kirkendall pores. It should
be noted that once the Kirkendall pores are formed, Eq. (4.61) is no longer valid so
that one may not determine the intrinsic diffusivities appropriately due to Eqs. (4.14)
and (4.16).

4.8 Summary of the Various Diffusivities

We have so far encountered a variety of lattice diffusivities: the self-diffusivity of


component k, Dk; tracer diffusivity of k, Dk ; defect diffusivity of defect d
(¼interstitials, vacancies, etc), Dd; the intrinsic (or component) diffusivity of k,
e k; and the interdiffusion or chemical diffusivity (for a binary, i.e., k ¼ 1,2), D.
D e Let
us close this chapter by summarizing the interrelationships among these with the
definition of each.
The self-diffusivity Dk is a measure of the total, successful jump frequency Γk
of k in a homogeneous medium, i.e., in the absence of chemical gradient or

1
Dk ¼ Γk α2 : ð4:62Þ
6

This is normally measured in terms of the tracer diffusivity Dk by measuring the
mean square displacement < r2 > of the k-tracers (k) in time t as

1 < r 2 >
Dk ¼  : ð4:63Þ
6 t

This is correlated to the self-diffusivity via the correlation factor f depending on


diffusion mechanisms and crystallographic structures of the medium as

Dk ¼ fDk : ð4:64Þ

The defect diffusivity Dd is the self-diffusivity of defects of type d themselves,


which is related to the self-diffusivity as
X
Ck Dk ¼ Cd Dd : ð4:65Þ
d

The intrinsic or component diffusivity of k is a measure of thermodynamically


biased jump under chemical gradient. For, e.g., a binary interdiffusion system

e k ¼ Dk f ðθÞ:
D ð4:66Þ

The interdiffusion or chemical diffusion coefficient is a measure of interdiffu-


sion rate in chemical gradient or, e.g., an A|B binary,
168 4 Diffusion in Concentration Gradients

e ¼ NA D
D e B þ NB D
eA ð4:67Þ

representing the diffusion rate with the matrix drift effect superposed.

Problems

1. Two markers are placed in a diffusion couple formed of two semi-infinite regions
of A and B – one at the initial weld interface and one a short distance away. Show
qualitatively how the position of each varies with time if DA > DB. Derive these
curves by plotting N(x) and ∂N/∂x vs. x and using the equation

∂NA
v ¼ ðDA  DB Þ :
∂x

2. It is frequently stated that if a reaction or mixing is diffusion controlled, any given


composition, e.g., an α-β interface, will move as t1/2 even when D varies with
composition. Show under what conditions this is true for the case of D ¼ f(c).
(It may be helpful to first assume D ¼ const. and show that the above is true for
one general set of boundary conditions but is not true for another general set.)
3. A gold-nickel diffusion couple of limiting compositions (in mole fractions)
NNi ¼ 0.0974 and NNi ¼ 0.4978 is heated at 925  C for 2.07  106 s. Layers
0.003 cm thick and parallel to the original interface are machined off and
analyzed.
(a) Using the data tabulated below, calculate the diffusion coefficient at 20, 30,
and 40 at % nickel.
(b) Suppose that markers are inserted at the original interface and move along
during the diffusion process at a composition of 30 atom fraction nickel.
From this, determine the intrinsic coefficients of gold and nickel at 30 atom
fraction nickel.

Slice No. a/o Ni Slice No. a/o Ni Slice No. a/o Ni Slice No. a/o Ni
11 49.78 22 33.17 29 21.38 39 12.55
12 49.59 23 31.40 30 20.51 41 11.41
14 47.45 24 29.74 31 19.12 43 10.48
16 44.49 26 25.87 32 17.92 45 9.99
18 40.58 27 24.11 33 16.86 47 9.74
19 38.01 28 22.49 35 15.49
20 37.01 37 13.90
21 35.10 38 13.26
Problems 169

4. A diffusion couple is prepared by welding a long bar of A-B alloy (ρB ¼ 0.072
moles/cm3) end-to-end with a long bar of pure A (ρA ¼ 0.144 moles/cm3). The
welded piece is then annealed at 1000  C.
(a) Set up the differential equation and initial and boundary conditions for the
transport process. Use the term D e in your equations. Be certain to identify all
terms used. (Assume that D e ¼ 1010 cm2 =s, if necessary.)
(b) Indicate the nature of the solution to the differential equation to be obtained
for part (a). Sketch carefully the composition profiles for this solution along
the axial centerline in the vicinity of the weld for several annealing times.
(c) Isotopic measurements show that DA =DB ¼4 in the A-B system in the range
of 0.072 > ρB > 0, and it is known that the systems is ideal (Raoultian). How
do you expect the actual composition profiles to differ from the calculated
ones in part (b)? How is the “Matano” interface shifted relative to the weld?
Be specific and use sketches where appropriate.
(d) Two sets of markers are placed in the couple, one at the interface and one a
short distance away. Show qualitatively how each moves with time (1) if
DA =DB ¼1 and (2) if DA =DB ¼4.
(e) Will the markers move with positions of constant composition?
5. A thin-walled brass pipe is exposed to zinc vapor both inside and outside at
1100 K: Pinside ¼ P1 ¼ 0.035 atm and Poutside ¼ P2 ¼ 0.060 atm. The pipe has an
outside diameter, d, of 10 cm and a wall thickness,ℓ, of 1 mm.
(a) Calculate the rate at which zinc permeates (g/cm2s) the pipe wall at steady
state.
(b) At “steady-state” permeation, the diameter of the pipe has been observed to
increase at a rate of 1.5  10–7 cm/s. Compute from the data given the change
in diameter that might arise from diffusion effects. Can the observed dimen-
sional change be explained this way? Discuss.
(c) Sets of inert markers are attached to both the inner and outer surfaces of the
pipe before it is exposed to the zinc. What will happen to each set as a result of
the exposure of the pipe to the zinc vapor?
Data at 1100 K (0.2 x 0.4)
For α-Brass (assume aZn / XZn in this range)

XZn aZn
0.30 0.06
0.38 0.13

Saturation vapor pressure of pure liquid Zn, PZn ¼ 0.48 atm.


Self-diffusion coefficients: DCu ¼ 2.2  109 cm2/s; DZn ¼ 5.1  109 cm2/s
Density of α-brass, ρBrass ¼ 7 g/cm3
Molar weight of Zn ¼ 65.4; molar weight of Cu ¼ 63.5
170 4 Diffusion in Concentration Gradients

6. A diffusion couple consisting of copper and a copper-zinc alloy is shown before


and after annealing. The fact that the inert markers did not move with respect to
the zinc-rich end has been explained by suggesting that pores have formed.

Inertmarkers Inertmarkers

1cm
1cm
1cm 1cm
5cm 5cm 5.05cm 5cm

t=0 after diffusion anneal

Fig. 4.12 Diffusion couple, Cu-Zn, before and after annealing

(a) Sketch a concentration profile for this diffusion couple after a “short time”
anneal, i.e., concentrations at the ends of the couple unchanged from their
initial values.
(b) On your sketch, identify the region of maximum porosity.
(c) Write, but do not attempt to solve, the equation which describes the rate of
change of total pore volume.
Data: DZn > DCu
7. Consider interdiffusion in a two-component system, 1–2, which behaves as a
regular solution, i.e., parabolic enthalpy and ideal entropy.
(a) (i) Show how D e 1, the intrinsic diffusivity of component 1, is related to its self-
diffusivity D1 and the regular solution parameter Ω such that the relative
integral molar enthalpy of mixing, ΔHM ¼ ΩN1N2, with Ni being the mole
fraction of the component i(¼1,2). (ii) How does D e 1 vary with temperature?
(iii) What is the apparent activation energy for interdiffusion of species 1?
(b) (i) Using the regular solution model, show that if dNi/dx ¼ 0, then the
chemical potential gradient dμi/dx ¼ 0. (ii) Does it necessarily follow that if
dNi/dx > 0, then dμi/dx > 0?
8. A Kirkendall-type experiment was performed in the A-B alloy system, which is
known to make regular solutions with relative integral molar free energy of
mixing

ΔGM ¼ RTðNA ln NA þ NB ln NB Þ  0:5RTNA NB :

Markers placed at the original interface were found to move with the
iso-concentration front, NA ¼ 0.50, in mole fraction. After a 60-hour anneal,
the following data were obtained:
References 171

Concentration gradient at NA ¼ 0.50, ∂NA/∂x ¼ 4 cm1


Self-diffusion coefficient of A at NA ¼ 0.50, DA ¼2.5  108 cm2/s
Self-diffusion coefficient of B at NA ¼ 0.50, DB ¼1.5  108 cm2/s
(a) Set up the differential equation and initial and boundary conditions for the
transport process. Be certain to identify the diffusion coefficient used.
(b) If DA ¼ DB, what kind of diffusion profile would you get? Write the solution,
and sketch carefully the composition profiles for this solution along the axial
centerline in the vicinity of the weld for several annealing times.
(c) How do you expect the actual composition profiles to differ from the calcu-
lated ones in part (b)? You may sketch the two profiles to compare.
(d) Discuss the difference in activation energy for self-diffusion and intrinsic
(component) diffusion of A.
(e) How is the Matano interface shifted relative to an end of the diffusion couple?
(f) How will the marker move with time? Explain why. Calculate the marker
displacement from the initial weld. State your assumptions if any.
(g) Suppose now that the marker displacement of part (a) were zero due to
porosity. (i) Calculate the pore volume per unit cross sectional area at
t ¼ 60 h. (ii) In which part of the specimen do you expect the pores to form?
9. A Kirkendall-type experiment was performed in the A-B alloy system. Markers
placed at the original interface were found to move toward the A-rich end with the
iso-concentration front, NA ¼ 0.35. After a 100-hour anneal, the following data
were obtained:
Concentration gradient at NA ¼ 0.35, ∂NA/∂x ¼ 2.0 cm1.
Marker velocity, vm ¼ 3.2  1010 cm/s.
e
Interdiffusion coefficient D¼1.1  1010 cm2 s1.
Molar volume, Vm ¼ 12 cm /mol.
3

Calculate the vacancy flux across the plane of the markers at t ¼ 100 hr.

References

1. A. Smigelskas, E. Kirkendall, Zinc diffusion in alpha brass. Trans AIME 171, 130 (1947)
2. H. Nakajima, The discovery and acceptance of the Kirkendall effect: the result of a short
research career. JOM 49, 15 (1997)
3. L. Darken, Diffusion, mobility and their interrelation through free energy in binary metallic
systems. Trans. AIME 174, 184 (1948)
4. T. Lee, H.-S. Kim, H.-I. Yoo, From Onsager to mixed ionic electronic conductors. Solid State
Ionics 262, 2 (2014)
5. J.R. Bardeen, C. Herring, Diffusion in alloys and the Kirkendall effect, in Atom Movements,
(American Society for Metals, Cleveland, 1951), pp. 87–111
6. A.R. Allnatt, A.B. Lidiard, Atomic Transport in Solids (Cambridge University Press, Cam-
bridge, 1993). Chap. 5
7. A. Einstein, Annal. Phys. 322(8), 549 (1905); M. von Smoluchowski, ibid., 326(14) 756 (1906)
172 4 Diffusion in Concentration Gradients

8. A. Einstein, Ann. Phys. 17, 549 (1905), quoted from L. S. Darken [9]
9. L.S. Darken, Formal basis of diffusion theory, in Atom Movements (American Society for
Metals, Cleveland, 1951), pp. 1–25
10. L.S. Darken, Diffusion of carbon in austenite with a discontinuity in composition. Trans. AIME
180, 430 (1949)
11. G.S. Hartley, J. Crank, Some fundamental definitions and concepts in diffusion processes.
Trans. Faraday. Soc. 45, 801–818 (1949)
12. J. Bardeen, C. Herring, Diffusion in alloys and the Kirkendall effect, in Imperfections in Nearly
Perfect Crystals, (Wiley, New York, 1952), pp. 261–288
13. L.C.C. da Silva, R.F. Mehl, Interface and marker movements in diffusion in solid solutions of
metals. Trans. AIME 191, 155–173 (1951)
Chapter 5
Kinetics of Phase Transformation: Initial
Stage

5.1 Introduction

We know from experience that water vapor turns to liquid water and the latter to
solid ice as temperature is lowered at a fixed pressure. We wonder how liquid water
comes into existence out of its vapor and solid ice out of liquid water and how fast
they grow larger in extent.
Let us see rather a common phase diagram of limited solubility between A and
B as shown in Fig. 5.1. The phase diagram indicates that when a homogeneous
solution α of composition “N” is cooled down from temperature T2 to T1, this used-
to-be homogeneous solution (phase) decomposes into two conjugate solutions
(phases) α (of composition Nα) and β (of composition Nβ) with the equilibrium
(amount) fractions Xα and Xβ (¼1Xα), respectively. For this kind of phase change
to occur, the new phase β has first to be conceived out of α somehow, thus coming
into existence, and then to grow up to the equilibrium fraction Xβ, followed by
coarsening or homogenization. These consecutive processes each involve diffusion
of, say, component B through phase α under chemical potential gradients and
accommodation into β via the surface reaction. How fast will each process and,
hence, the overall process of phase transition proceed?
We will learn in this chapter the classic, basic ideas on the kinetics of the
daughter-phase (β) conception, nucleation and spinodal decomposition, and in
the following chapter the kinetics of growth and coarsening or homogenization,
again from the old wisdom [1–4] as we have done so far.

© Springer Nature Switzerland AG 2020 173


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_5
174 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.1 Partial phase


diagram of system A–B with
limited mutual solubility N
T2

temperature
a a+b b

T1
Xb Xa

A Na Nb B
NB
composition

5.2 Homogeneous Nucleation


5.2.1 Driving Force for a Phase Change

Consider a phase transition from phase α to β with no composition change like the
condensation of water vapor(α) to liquid water(β) below the boiling point, the
equilibrium transition temperature Ttr under a fixed pressure, say, P ¼ 1 atm. If
we plot the molar Gibbs free energy or chemical potential of α and β, μα and μβ,
respectively, against temperature under the fixed pressure, it may look as shown in
Fig. 5.2.

Fig. 5.2 Schematic of the m


chemical potential of phase a
α (vapor), μα, and of β
(liquid), μβ, vs. temperature ∆gβ/α
b
at a fixed pressure, say,
P ¼ 1 atm, where Ttr is the
equilibrium transition
temperature, the normal
boiling point in the
present case
T Ttr

You know, each μϕ vs. T curve should be convex upward with negative slopes
because (∂μϕ/∂T)P ¼ sϕ < 0 and (∂2μϕ/∂T2)P ¼ (∂sϕ/∂T)P ¼ cP,ϕ/T < 0,
where sϕ is the molar entropy and cP,ϕ the molar heat capacity at constant
pressure of phase ϕ (¼α,β). As is seen, at T < Ttr, μβ < μα or the escaping tendency
of water molecules is higher in the vapor(α) than in the liquid(β); thus, the free
energy difference per mole of the system water
5.2 Homogeneous Nucleation 175

Δgβ=α ¼ μβ  μα < 0 ð5:1Þ

drives the water molecules to escape from the vapor phase(α) to the more stable,
liquid phase(β), viz., phase transformation from vapor(α) to liquid(β), or the vapor
condenses to form liquid. The situation turns vice versa at T > Ttr where μβ > μα, or
phase transforms from the liquid(β) to the more stable vapor(α), i.e., water boils. At
the equilibrium transition temperature, T ¼ Ttr, obviously Δgβ/α ¼ 0 or μβ ¼ μα, and
thus, the two phases are in thermodynamic equilibrium.
You see, the driving force for phase transition |Δgβ/α| is getting larger, the farther
away from the equilibrium transition temperature, Ttr. How does it change with
temperature? Noting that μϕ ¼ hϕTsϕ with hϕ and sϕ being the molar enthalpy and
entropy of the phase ϕ (¼α,β), respectively,
 
    T
Δgβ=α ¼ hβ  hα  T sβ  sα  Δhβ=α  TΔsβ=α ffi Δhtr 1 
o
ð5:2Þ
Ttr

if Δcp ¼ cp,βcp,α ffi 0 or if the degree of supercooling or superheating is not large


enough, i.e., |TTtr|/Ttr << 1 so that μϕ ¼ μϕ(T) may be regarded as linear around
T ¼ Ttr. Otherwise, due to Kirchhoff’s law [5],
Z Z
T
Δhotr T
ΔcP
Δhβ=α ¼ Δhotr þ ΔcP dT; Δsβ=α ¼ þ dT: ð5:3Þ
Ttr Ttr Ttr T

Here, Δhotr is the molar enthalpy of transition or latent heat for α ! β at the
equilibrium transition temperature Ttr and the corresponding molar entropy of
transition, Δsotr ¼ Δhotr =Ttr .

5.2.2 Critical Size and Energy Barrier for Nucleation

It is empirically known that condensation or boiling never occurs right at the


equilibrium boiling temperature Ttr, say, 100  C for water: It always requires some
degrees of supercooling for water droplets to appear within the vapor phase and
some degrees of superheating for water vapor bubbles to appear within the liquid.
Why is that?
Imagine that a water(β) droplet of radius r forms somehow within the homoge-
neous vapor(α) phase. Here, we assume that the liquid(β) droplet is large enough to
define the surface or interfacial energy, γβ/α. The formation of a droplet is always
accompanied by the formation of a new surface Aβ ¼ 4πr2, thus tending to increase
the overall free energy of the system, the liquid droplet and vapor, by the surface
energy, γβ/αAβ. If the matrix where a new phase nucleates is rigid like crystalline
solids, then there may arise even a strain energy due to the volume difference or
lattice mismatch between the matrix α-phase and the droplet β-phase, which also
176 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.3 Schematic of


ΔGr vs. r upon forming a ∆ Gr r2
spherical β-droplet of radius
r within the α-phase. The
volume free energy
increases negatively as r3 ∆ Gr*
and the surface energy r
increases positively as r2. r*
The sum of the two makes ∆ Gr
ΔGr in Eq. (5.4)
r3

tends to increase the free energy of the total system. For the droplet formation or
nucleation process to proceed spontaneously, the overall free energy change for the
process, ΔGr, should be negative, and hence, the volume free energy gain due to
Eq. (5.1) should be negatively large enough to compensate these increases due to the
new interface in any case and the volume difference as well if the medium is rigid.
This is why the phase transformation cannot happen just at the thermodynamic
equilibrium temperature Ttr and always requires some degrees of supercooling or
superheating. Let us quantify this energetic scenario of the nucleation.
Suppose that a spherical droplet of liquid water(β) forms to a radius r at a given
temperature T(<Ttr) and pressure. Letting Vm,β denote the molar volume of the liquid
(β), the overall Gibbs free energy change, ΔGr, accompanying this process should
depend on r as

4 Δgβ=α
ΔGr ¼ πr3  þ 4πr2  γβ=α ð5:4Þ
3 Vm,β

neglecting the strain energy in the present case. Remember that Δgβ/α < 0 when
T < Ttr, but γβ/α > 0 always. As is schematically shown in Fig. 5.3, the first term on
the right-hand side, the volume free energy, increases negatively as r3 and the
second term, the surface or interfacial energy, increases positively as r2. The
overall free energy change ΔGr, thus, first increases positively as r increases up to
a certain critical value r and then decreases leaving a maximum ΔGr at r ¼ r. This
means that a droplet, once formed, tends to disappear if r < r and to grow if r > r, to
reduce the overall free energy of the system. A droplet is called a subcritical
embryo if r < r, a critical embryo or critical nucleus if r ¼ r, and a supercritical
nucleus if r > r. The process of passing over the critical size to form nuclei is called
the nucleation. When the nucleation occurs within a homogeneous mother phase (α)
as now, it is called the homogeneous nucleation.
How large can the critical size, r, be after all? You may immediately guess that
this should become larger as γβ/α increases because, in order to compensate now the
larger surface energy, one needs more negative volume free energy (due to Δgβ/α),
which is proportional to the volume of the droplet. As ∂ΔGr/∂r ¼ 0 at the maximum
in Fig. 5.3, you may obtain from Eq. (5.4) and Eq. (5.2):
5.2 Homogeneous Nucleation 177

2γβ=α 2γβ=α Vm,β


r ¼  ffi ð5:5aÞ
Δgβ=α =Vm,β Δh tr ðΔT=Ttr Þ
o

16πγ3β=α 16πγ3β=α V 2m,β


ΔGr ¼  2 ffi  o 2 : ð5:5bÞ
3 Δgβ=α =Vm,β 3 Δhtr ΔT=Ttr

Here, the equality-or-approximations (ffi) are due to Eq. (5.2) for Δcp ffi 0 or for the
supercooling or superheating, ΔT(TTtr), being such that |ΔT|/Ttr ¼ |TTtr|/
Ttr << 1.
You may wish to know how many water(β) molecules comprise a critical-size
embryo. A droplet of radius r bears “i” molecules such that

4 3
πr ¼ iΩβ ð5:6Þ
3

where Ωβ (¼Vm,β/NA with NA being the Avogadro number) is the atomic or


molecular volume of the daughter β-phase. This leads to the critical number of
molecules i comprising a critical nucleus of size r as

32πγ3β=α
i ¼  3 : ð5:7Þ
3 Δgβ=α =Vm,β Ωβ

It is estimated that the number of molecules in a critical nucleus for “sensible”


homogeneous condensation of liquids, say, 1 critical nucleus per 1 cm3, is generally
in the range of 100 molecules, see the footnote 1 with Eq. (5.9b). For example, the
nucleus for sensible condensation of water at 0  C should contain about i  70
molecules [1].

5.2.3 Nucleation Rate

We now know that in the supercooled state, water droplets with a size r < r or i < i,
i.e., subcritical embryos, tend to disappear because the system energy has increased
by ΔGr, Eq. (5.4), with their formation, but those with the size r ¼ r or critical
nuclei may grow to supercritical nuclei, thus reducing the system free energy. How
do these embryos form to various sizes then despite the increase in the system free
energy? The answer to this question was provided by Volmer and Weber [6].
It is well-known that in a homogeneous phase, say, water vapor at a temperature
higher than the equilibrium boiling point, its local concentration always fluctuates.
Likewise, in a supercooled vapor below the equilibrium transition temperature,
similar concentration fluctuation may occur, but once concentrated in a region, the
region may be regarded as the daughter phase, say, liquid water. What is happening
178 5 Kinetics of Phase Transformation: Initial Stage

is like ceaseless formation and disintegration of variable-size liquid clusters at


random. While the former is called the homogeneous fluctuation, the latter in a
supercooled vapor may, thus, be called the hetero-phase fluctuation. These
embryos may then be stabilized to have a quasi-equilibrium or stationary concen-
tration due to the entropic effect.
Let nr denote the number/unit volume of the embryos of size r (or corresponding
number of molecules i); ΔGr the energy increase with the size r, Eq. (5.4); and nα the
number/unit volume of single molecules in the mother phase α. The total free
energy change of this hetero-system of unit volume or mixing free energy, ΔGM,
may, then,
 be written, by taking into account the configurational entropy
P   
¼ kB ln nα þ nr != nα !Πnr ! , as
r r

X X n
ΔGM ¼ nr ΔGr þ kB T nr ln r ð5:8Þ
r r n α

assuming the ideal mixing of the variable-size(r) embryos and single molecules α,
and nα >> nr. At equilibrium, ΔGM should be minimum with respect to nr or ∂ΔGM/
∂nr ¼ 0. Thus, the quasi-equilibrium concentration of size-r embryos, nr, is given
as
 
ΔGr
nr ¼ nα exp  : ð5:9aÞ
kB T

This means that the concentration of critical embryos (r ¼ r), nr, should be1
 
ΔGr
nr ¼ nα exp  ð5:9bÞ
kB T

where ΔGr is the energy barrier as given in Eq. (5.5b); see Fig. 5.3.
How many out of these nr critical embryos will turn to nuclei per unit time?
This question sounds rather embarrassing, but Volmer and Weber [6] conjecture
that all these clusters or embryos of variable sizes form by succession of bimolec-
ular collision processes. You see, other possible mechanisms such as simultaneous

1
For “sensible” nucleation rate, one should have, say, at least 1 critical nucleus/cm3 or nr 1/cm3.
Noting from Eq. (5.5) that ΔGr ¼ 43 πγβ=α r2 , Eq. (5.9b) implies that

 1=2
3kB T ln nα
r

4πγβ=α

where the single molecule concentration nα should be in number/cm3. This gets you the critical
number of molecules i(¼4πr3/3Ωβ) in the range of 100 molecules. For more extensive discussion,
see D. A. Porter and K. E. Eastering [4].
5.2 Homogeneous Nucleation 179

collision of i molecules to form a size-r cluster will be extremely improbable. In


other words, a single molecule α1 collides with another α1 to form a bimolecular
cluster α2; this collides with other single molecule α1 to form a trimolecular cluster
α3, and on and on or

α1 þ α1 ⇄ α2
α2 þ α1 ⇄ α3
α3 þ α1 ⇄ α4
   ð5:10Þ
  
  
αi 1 þ α1 ⇄ αi

Each step is considered to be a reversible reaction so that sort of equilibrium


concentration of each cluster may be defined as in Eq. (5.9).
Now, we may conjecture that once a single molecule is stuck to a critical-size
cluster or embryo, αi, via collision, this immediately turns irreversibly to a nucleus
αi+1 as

αi þ α1 ! αi þ1 : ð5:11Þ

This means that, as we have learned earlier in Chap. 3, the collision process in
Eq. (5.11) has to be very infrequent one compared to the other reversible reactions in
Eq. (5.10). This is the very nucleation reaction, and hence, the nucleation rate
should correspond to the reaction rate of this irreversible reaction.
How do we estimate the nucleation rate “I” in terms of the number of nuclei/unit
volume/unit time? Let us assume that once a single molecule α1 touches at least the
surface of a critical embryo αi of the radius r, it immediately sticks to the latter with
100% efficiency, i.e., the sticking or accommodation factor f¼1, to turn the
critical embryo αi to a nucleus αi+1. The point is then how many critical embryos
are collided by single molecules per unit time in unit volume. Let r1 be the radius of a
single molecule α1. Once an α1 comes within a radius of r + r1 with an αi at the
center, it is said, by definition, that the critical embryo is collided by a single
molecule. The area

σ ¼ πðr þ r1 Þ2  πr
2
ð5:12Þ

is called the collision cross section. A critical embryo is so bulky compared to a


single molecule or r> > r1, and hence, it may be regarded as standing still, but those
single molecules are zigzagging ceaselessly with a mean speed <v>. Given this
chaotic situation, how could we count how many critical embryos make a collision in
unit time? Imagine, in the opposite way, that all these single molecules are standing
still in space and a critical embryo is moving, instead, at the mean speed <v>. You
may then recognize that each αi collides with all the single molecules within the
180 5 Kinetics of Phase Transformation: Initial Stage

volume swept by itself, σ<v > per unit time. The collision frequency, Z, of each αi
may then be written as

Z ¼ σ hvinα : ð5:13Þ

This means that each critical embryo, αi turns to a nucleus, αi+1 in 1/Z sec, and
hence, the theoretical nucleation rate, Ith, should be

nr
Ith ¼ ¼ σ hvinα nr ð5:14Þ
1=Z

as there are nr critical embryos per unit volume; see Eq. (5.9b). Assuming the ideal
gas behavior of our vapor(α) and the Maxwell distribution of v,
rffiffiffiffiffiffiffiffiffiffiffi
p 8kB T
nα ¼ ; h vi ¼ ð5:15Þ
kB T πm

where p is the pressure of water(α) vapor at T < Ttr and m the mass of a water(α)
molecule.
Substituting Eqs. (5.9b) and (5.15) into Eq. (5.14), we finally obtain
pffiffiffi  
2 2p ΔGr
Ith ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi σ  nα exp  ð5:16Þ
πm kB T kB T

where another nα is kept as is only for the cosmetic reason, but, of course, it may also
be replaced by the ideal gas law in Eq. (5.15), if you want.
The equation looks very elegant, but there is no way to prove it experimentally.
Nevertheless, it is easy to recognize that this theoretical nucleation rate “Ith” has
been excessively overestimated in two aspects: (i) In this formulation, we assumed
that any single molecule, once collided, is stuck or accommodated 100% to the
critical embryo or the sticking, steric, or accommodation factor, f ¼ 1. However,
it will not be the case in reality mainly for the steric reason; thus, f << 1. (ii) We also
assumed the one-way or irreversible traffic with regard to the nucleation reaction,
Eq. (5.11), but there should occur the reverse reaction, too, thus rendering the
nucleation rate to be smaller by a factor of g(<1). You may then correct the
theoretical nucleation rate Ith in Eq. (5.16) to Icorr as

Icorr ¼ g  f  Ith : ð5:17Þ

That is, the two factors g and f are to blame for any overestimation of Ith, no matter
however large. This correction was first considered by Becker and Doering [7] and
expounded later by Reiss [8].
5.2 Homogeneous Nucleation 181

5.2.4 Nucleation Involving Strain Energy

Free Energy Barrier


So far, we have considered the nucleation kinetics in a compressible fluid, say,
vapor; thus, any possible strain energy may be released away. How is then the
nucleation kinetics affected if the strain energy due to the volume change or lattice
mismatch cannot be released? Let us consider the nucleation of β out of an
incompressible, condensed phase α. This process involves, as before, (i) the volume
free energy decrease due to Δgβ/α(<0), Eq. (5.2), at T < Ttr or ΔGvol ¼ Δgβ/αVβ/Vm,β
with Vβ being the volume of a β-embryo and Vm,β the molar volume of the β-phase
and (ii) the surface energy increase due to the interfacial energy γβ/α(>0) or
ΔGsurf ¼ γβ/αAβ with Aβ being the newly formed surface area of the β-embryo. In
addition, we now have (iii) the strain energy due to the volume change or lattice
mismatch between the incompressible phases α and β, ΔGstrain ¼ (σε/2)Vβ > 0 with
σ and ε being the stress and strain induced, respectively. If a β-embryo is spherical
with a radius r, then the overall Gibbs free energy change ΔGr is written in terms of
size r or number of molecules i, due to Eq. (5.6), as follows:

ΔGr ¼ ΔGvol þ ΔGsurf þ ΔGstrain


Δgβ=α 4 3 1 4
¼  πr þ γβ=α  4πr2 þ σε  πr3 ð5:18aÞ
Vm,β 3 2 3
 
i  2=3 1  
¼ Δgβ=α  þ γβ=α  ð4πÞ1=3  3Ωβ i þ σε  Ωβ i : ð5:18bÞ
NA 2

Each of these three kinds of energy contributions and their sum, ΔGr, is shown
against the number of molecules i this time in Fig. 5.4.

Fig. 5.4 ΔGvol, ΔGsurf, ∆Gr


ΔGstrain, and their sum ∆ Gsurf ∝ i2/3
ΔGr vs. number of
molecules i in a spherical
∆ Gstrain ∝ i
β-embryo in α-phase
∆Gr*
i
i*
∆ Gr

∆ Gvol ∝ i
182 5 Kinetics of Phase Transformation: Initial Stage

Then, it follows that

2γ 16πγ3
r ¼  ; ΔGr ¼  2 : ð5:19Þ
Δgβ=α 1 Δgβ=α 1
 þ σε 3  þ σε
Vm,β 2 Vm,β 2

In comparison with Eq. (5.5) for the case of ΔGstrain ¼ 0, you may immediately
recognize that you need a larger degree of supercooling or superheating for the same
critical size r (or equivalently i) or a larger r (or i) at the same supercooling or
superheating, in order to compensate the additional free energy increase due to the
strain energy (σε/2 per unit volume), if not completely released.
Compromise Between Surface and Strain Energy in Reality
When the α-phase of a crystallographic structure transforms to the β-phase of
another structure, of course, with no composition change, there normally appear
two extreme types of α/β interfaces: One type is that the crystallographic lattice of
the daughter β-phase is more or less continuous with the surrounding mother α-phase
and the other the β-phase forms detachedly with no lattice continuity with the mother
α-phase. The former is called the “coherent interface” and the latter the “incoher-
ent interface.” As the different crystal structures have different lattice parameters,
“a,” the coherent interface should be stressed to the maximum or the strain that is
proportional to the lattice parameter difference |Δa| is maximum because the two
different lattices are forced to be continuous. The incoherent interface, on the other
hand, is obviously free from such a stress because the two different lattices are set
free from each other, and hence, the strain ε  0. These two extremes or distinct
situations are illustrated in Fig. 5.5.

Fig. 5.5 Coherent (a) and incoherent (b) precipitate of β (•) in α (о), surrounded by the coherent
and incoherent interfaces, respectively. Due to the difference in lattice parameter or lattice
mismatch, the unit cells are strained in (a), but strain-free in (b)
5.2 Homogeneous Nucleation 183

What about the concentration change Δc across each of these two interfaces? For
the coherent interface, the lattice constants are enforced to have the same value;
thus, the concentration difference across this interface should be negligible or |∇
c|  0, but across the incoherent interfaces, |∇c| will be the maximum.
You see, ε/|Δa| being maximum, but |∇c|  0 for the coherent nucleation
(Fig. 5.5a); ε  0, but |∇c| being maximum for the incoherent nucleation
(Fig. 5.5b)! It is known2 that the interfacial energy γβ/α is proportional to (∇C)2,
thus called the gradient energy, and we know that the strain energy/unit volume σε/
2/ε2 and ε/|Δa| or

σε
γβ=α / ð∇CÞ2 ; / ðΔaÞ2 : ð5:20Þ
2

This means that for the case of coherent nucleation, you have negligible surface
energy, but the maximum possible strain energy, and vice versa for the incoherent
nucleation. Mother Nature always tends to choose the lower possible Gibbs free
energy option: If ΔGsurf (for the incoherent nucleation) > ΔGstrain (for the coherent
nucleation), She let the embryo choose to form coherently to take the lower strain
energy option and if ΔGsurf < ΔGstrain incoherently to take the lower surface energy
option. It is how the β-precipitates form coherently or incoherently in the α-matrix.
Nucleation Rate
We may again assume the quasi-equilibrium state with respect to the concentration
of critical embryos, nr, Eq. (5.9b), but the concentration of single molecules nα
should now be understood to be that of atoms or molecules in the condensed
α-phase. The critical energy barrier ΔGr is as given in Eq. (5.19), which may,
however, depend on whether coherent or incoherent.
The nucleation rate may also be formulated in similar way to the earlier case,
Eq. (5.14): Once an α-phase atom or molecule, sitting on the surface or in the
periphery of a critical β-embryo, jumps across the α/β boundary and sticks to this
critical β-embryo, then the latter turns to a β-nucleus. How often will it happen? Let
ns denote the number of α-phase atoms or molecules in the periphery of a β-critical
embryo and νgb the successful jump frequency (s1) of an α-molecule crossing the
α/β boundary or grain boundary. Then, a critical embryo turns to a nucleus in
1/nsνgb second assuming 100% sticking efficiency or accommodation factor f ¼ 1.
As there are nr critical embryos, the theoretical nucleation rate Ith may be written
as

Ith ¼ νgb ns nr : ð5:21Þ

Upon comparison with Eq. (5.14), you may recognize the similarity in idea behind.
We have learned from Chap. 2 that when an atom or molecule crosses a boundary,
which is a grain boundary now, it should pass over a migrational free energy

2
The physico-chemical origin of the gradient energy will be explained in Sect. 5.4.5
184 5 Kinetics of Phase Transformation: Initial Stage

barrier, Δgm,gb, and this successful jump frequency νgb is a measure of the grain
boundary diffusivity, Dgb (across the grain boundary), or
 
Δgm,gb Dgb
νgb ¼ νogb exp   2 ð5:22Þ
kB T δ

where νogb is the vibrational frequency to attempt the energy barrier and δ denotes
the thickness of the grain boundary which now corresponds to the jump distance
to cross the grain boundary. By substituting Eq. (5.22) and Eq. (5.9b) into
Eq. (5.21), we obtain
   
Δgm,gb ΔGr
Ith ¼ νogb exp   ns  nα exp  ð5:23aÞ
kB T kB T
 
Dgb ΔGr
 2  ns  nα exp  : ð5:23bÞ
δ kB T

The approximations, Eqs. (5.22) and (5.23b), are due to that we have neglected
the geometric factor γ, of order of 1, from the relation Dgb ¼ γνgbδ2. In any case,
there is no way to prove experimentally this pre-exponential factor as in Eq. (5.16),
and obviously this factor is overestimated. You may, thus, want to correct it against
non-perfect accommodation (f << 1) and possible jumping back (g < 1), as we did in
Eq. (5.17), or

Icorr ¼ g  f  Ith ¼ g  f  ns νgb nr : ð5:24Þ

You may derive the successful jump frequency across the grain boundary, νgb, a
bit more elaborately or sophisticatedly by using the absolute reaction rate theory
which you learned in Chap. 3. The free energy landscape that is to be experienced by
a jumping α-atom or molecule along the reaction coordinate, say, from the α-matrix
to the β-embryo across the boundary of width δ, may look like Fig. 5.6.
!
The jump rate from α to β, ν , is then written as
 
! Δg
ν ¼ νogb exp  m ð5:25Þ
kB T

and that in the reverse reaction


 
Δgm  Δgβ=α
ν ¼ νogb exp  ð5:26Þ
kB T

because the barrier height is different depending on the direction of jump; see
Fig. 5.6. The net jump rate νgb is then
5.2 Homogeneous Nucleation 185

Fig. 5.6 Free energy


variation across the
boundary between the
matrix α and the embryo β.
The free energy of β is lower
than α by |Δgβ/α| ¼ Δgβ/α,
thus driving the atomic jump
from α to β. The jumping
atom, however, should cross
the migrational energy
barrier Δgm from α to β, but
the higher barrier Δgm-Δgβ/α
from β to α. Note that
Δgβ/α < 0 at T < Ttr

    
! Δg Δgβ=α
νgb ¼ ν  ν ¼ νogb exp  m  1  exp , ð5:27Þ
kB T kB T

which may be approximated in two extreme cases as:


 
Δgβ=α Δgm
ðiÞ If >> 1, then νgb ¼ νgb exp 
o
; ð5:28aÞ
kB T kB T
   
Δgβ=α Δgm Δgβ=α
ðiiÞ If << 1, then νgb ¼ νgb exp 
o
 : ð5:28bÞ
kB T kB T kB T

The first case is for a larger supercooling (T << |Δgβ/α|/kB; thus eΔgβ=α =kB T  0)
and the second for small enough a supercooling (T >> |Δgβ/α|/kB; thus eΔgβ=α =kB T 
1 þ Δgβ=α =kB T ); see Eq. (5.2). Here, you remember, the vibrational frequency
may have the absolute value, νogb ¼ kB T=h, according to the absolute reaction rate
theory.
Temperature Dependence
Equation (5.21) indicates that the nucleation rate, I, in the condensed phase is
governed by two factors, (i) the successful jump frequency νgb or diffusivity across
the grain boundary Dgb and (ii) the concentration of the critical embryos nr. The
former decreases (Eq. (5.27)) and the latter increases (Eq. (5.9b)) exponentially with
decreasing temperature; thence the nucleation rate, I, should exhibit a maximum at a
specific temperature depending on the system; see Fig. 5.7.
186 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.7 Schematic


variation against reciprocal
temperature of the jump ngb I ମn gbnr* nr*
frequency νgb, the number of
critical embryos nr, and the
nucleation rate I

1/T

5.3 Heterogeneous Nucleation

Do you have ever watched water boiling closely? Upon heating a water-full beaker,
for example, water vapor bubbles normally always start from the surface of the
beaker in contact with water, much earlier than bubbles start to form homogeneously
within the water. Upon cooling a glass chamber containing water vapor, liquid
droplets also start to form preferentially on the chamber surface in contact with the
vapor before liquid water nucleates internally. When diamonds are artificially grown
in a carbon-containing gas atmosphere, e.g., Si-wafers with many scratches on are
usually provided because diamond starts to form preferentially upon them. That is,
the heterogeneous surfaces work as a catalyzer for nucleation. This sort of
initiation of a second phase upon heterogeneous sites, e.g., the vessel surface and
Si-wafers, is called the heterogeneous nucleation in contradistinction with the
homogeneous nucleation. It is also known that such a heterogeneous nucleation
may occur at much smaller degree of supercooling or superheating compared to the
homogeneous nucleation case, of course, depending on substrates. Why is that?
This fact suggests that the nucleation energy barrier should be lower for the
heterogeneous nucleation. Knowing that the barrier is due to the surface energy
increase with the formation of a new surface, you may immediately guess that the
surface energy increase should be smaller for the heterogeneous nucleation.
Let us examine how interfacial energy of the system varies when a liquid-water
(β) droplet forms on a heterogeneous site. We assume, for simplicity sake, that
water-vapor(α) condenses to a liquid(β) drop with a radius of curvature R and
contact angle θ on the substrate surface S, as illustrated in Fig. 5.8. We can see
that there form three different interphase areas or interfaces Aβ/α, Aβ/S, and Aα/S
between any two out of three different phases α, β, and S.
Now suppose that the contact area between β and S, Aβ/S, virtually expands
infinitesimally by δAβ/S, as delineated by the dotted line in Fig. 5.8. Then, it is easy
to recognize from the figure that the other two interfaces accordingly vary as

δAβ=α ¼ δAβ=S  cos θ; δAα=S ¼ δAβ=S ð5:29Þ

The free energy variation δG accompanying these interfacial area variations is


5.3 Heterogeneous Nucleation 187

a
gb/a Ab/a
Aa/S
q b
ga/S g b/S Ab/S
S dAb/S

R
q

Fig. 5.8 Formation of a β-phase droplet, with a radius of curvature R and contact angle θ, upon a
heterogeneous surface S within the α-phase, and the force balance among three interfacial tensions
γβ/α, γα/S, and γβ/S. Dotted line upon the β-droplet delineates the virtual expansion of its surface Aβ/α
with an infinitesimal increase δAβ/S of the interfacial area between β and S

δG ¼ γβ=S δAβ=S þ γβ=α δAβ=α þ γα=S δAα=S


 
¼ γβ=S þ γβ=α cos θ  γα=S δAβ=S ð5:30Þ

at given temperature and pressure, where γβ/S denotes the interfacial energy
between β and S and so on. At equilibrium, G should be minimum with respect to
the virtual variation of the interfacial area or δG/δAβ/S ¼ 0; thence Eq. (5.30)
leads to

γα=S ¼ γβ=S þ γβ=α cos θ: ð5:31Þ

This looks like the equation of vectorial force equilibrium, namely, the sum of
the forces in +x direction is equal to that in –x direction. In this light, a surface or
interface energy is equally called the surface or interfacial tension.
We are now ready to calculate the Gibbs free energy change, ΔGR, with the
formation of a β-embryo, with the radius of curvature R, upon an area Aβ/S (with the
interface energy γβ/S) replacing the used-to-be interface Aα/S (with γα/S) and creating
a new interface Aβ/α (with γβ/α) (see Fig. 5.8):
 
ΔGR ¼ Δgv Vβ þ γβ=α Aβ=α þ γβ=S  γα=S Aβ=S ð5:32Þ

where Δgv (¼Δgβ/α/Vm,β) represents the transition free energy per unit volume or
volume free energy and Vβ the volume of the β-embryo.
Noting that the β-embryo in question is a part of the sphere with radius R, we
can easily calculate the interfacial area Aβ/S as that of a simple circle of radius Rsinθ,
Aβ/α, by adding up the infinitesimal annular area 2πRsinθRdθ, and volume Vβ by
adding up the infinitesimal disc volume π(Rsinθ)2(Rdθsinθ) in Fig. 5.8 as
188 5 Kinetics of Phase Transformation: Initial Stage

Aβ=S ¼ πðR sin θÞ2 ; ð5:33aÞ


Z θ
Aβ=α ¼ 2πðR sin θÞðRdθÞ ¼ 2πR2 ð1  cos θÞ; ð5:33bÞ
0
Z θ
π  
Vβ ¼ πðR sin θÞ2 ðR sin θdθÞ ¼ R3 2  3 cos θ þ cos 3 θ : ð5:33cÞ
0 3

By substituting these into Eq. (5.32), one obtains


 
Δgv R3
ΔGR ¼ 4π þ γβ=α R2  f ð θÞ ð5:34Þ
3

where

ð2 þ cos θÞð1  cos θÞ2


f ð θÞ ¼ : ð5:35Þ
4

We can, thus, calculate the critical size R and nucleation energy barrier ΔGR as
in the homogeneous nucleation as

2 γβ=α 16 πγ3β=α
R ¼ ¼ r ; ΔGR ¼  f ðθÞ ¼ ΔGr  f ðθÞ ð5:36Þ
ðΔgv Þ 3ðΔgv Þ2

It is interesting to see that the critical size R is the same as that for the
homogeneous nucleation r, but the energy barrier height is equal to that of the
homogeneous nucleation ΔGr modulated by a factor, f(θ), which is a function of the
contact angle θ between β and the substrate S only. Compare with Eq. (5.19) for the
homogeneous nucleation.
When a phase β comes in contact with the substrate S, there may be two extreme
situations: β spreads out to wet the S-surface completely (complete wetting), thus
making the contact angle θ ¼ 0, and β never spreads even a bit, thus keeping itself as
spherical as possible to make the contact area zero (non-wetting) or θ ¼ π. In reality,
the contact angle takes a value in between depending on the nature of the two phases
to be in contact; thus 0
θ
π.
The contact angle function f(θ) in Eq. (5.35) varies against θ in its entire range as
shown in Fig. 5.9.
It is clearly seen that as θ ! 0 or our β completely wets the substrate S, then
f(θ) ! 0; thus, ΔGR ! 0! You would have no energy barrier at all; thus, α ! β
transition may take place even at the equilibrium transition temperature. As
θ ! π or our β completely non-wets the substrate, on the other hand, f(θ) ! 1 or
the situation is exactly the same as the homogeneous nucleation, requiring a certain
degree of supercooling or superheating for the transformation. Now, you can
understand why the heterogeneous nucleation is energetically cheaper than the
5.4 Spinodal Decomposition 189

Fig. 5.9 Contact angle


function f(θ) vs. contact
angle θ. Note that f(θ) ! 0
as θ ! 0 (complete wetting)
and f(θ) ! 1 as θ ! π
(complete unwetting)

homogeneous nucleation and why you need Si substrates to nucleate diamonds


artificially. If there are enough heterogeneous sites, a phase transition may take place
quite near to the equilibrium transition temperature depending on the nature of the
substrate.

5.4 Spinodal Decomposition

Do you have ever heard about the history of duralumin, an aluminum alloy
typically containing 3.5~5.5% Cu plus less than 1% Mg and Mn, which is still
widely used as the structural material for the fuselage of modern aircrafts? Metal Al
is light in weight (only 1/3 as heavy as steel) and in its unalloyed form, one of the
weakest of all metals. In 1906, A. Wilm [10] in Prussian Berlin discovered that this
aluminum alloy could be age-hardened even at room temperature upon being
quenched from an elevated temperature. He got this aluminum alloy patented, and
this alloy was started to be produced commercially with the brand name Duralumin
(¼company name Durener + Aluminum).
The inner working of this age-hardening, however, remained a mystery. It was
then known that a soft metal could be hardened by dispersing precipitates (thus
referred to as precipitation hardening) and its strength is inversely proportional to
the inter-precipitates distance. In the case of Duralumin, however, full strengthening
was achieved even before the precipitates appeared in sight, and the theory indicated
that the inter-precipitates distance should be 20~40 nm. It was in 1938 that
B. Guinier [11] in France and G. D. Preston [12] in England found “independently”
by means of small angle x-ray diffractometry that there were indeed dispersed
copper-rich coherent clusters or aggregates, which have since been called the GP
zones. People, thus, tried to understand how these clusters could form at an
interspacing of 20~40 nm. You may immediately consider the nucleation to be
responsible for those clusters and the native dislocations as energetically the most
favorable heterogeneous nucleation sites. Normally, the native dislocation density
190 5 Kinetics of Phase Transformation: Initial Stage

of a metal amounts to 107/cm2. If these were the heterogeneous nucleation sites,


then the spacing between nuclei would be expected to be on the order of
(107/cm2)-1/2 or 3.2 μm, two orders of magnitude larger than the value observed,
20~40 nm!! Furthermore, for the GP zones to form, Cu should diffuse uphill its own
concentration gradient, quite contrary to the then well-established Fick’s law!! What
will then be the mechanism for the formation of these GP zones after all? It should be
something different from the conventional nucleation mechanism. It is now known
as the spinodal decomposition due to M. Hillert [13] and J. Cahn. [14] We will
explore it now on the basis of the old wisdom [14, 15].

5.4.1 Phase Stability Against Compositional Fluctuation

You may remember from thermodynamics that for a system with fixed composition
to be stable, its internal energy U should be minimum for given entropy S and
volume V due to the energy minimum principle. This principle may be translated
into the mathematical jargon in this way: The variation of the internal energy δU(S,
V) as a function of S and V may be expanded into the Taylor series in terms of the
variations of entropy δS and volume δV as
 
∂U ∂U 1
δUðS, VÞ ¼ δS þ δV þ
∂S ∂V 2
 2 2 2 
∂ U 2 ∂ U ∂ U 2
δS þ 2 δSδV þ δV þ    ð5:37Þ
∂S2 ∂S∂V ∂V2

with the 3rd- and higher-order terms neglected. For U to be minimum, then its first
differential dU should disappear, as the necessity, for the virtual variations of S and
V or

∂U ∂U
dU ¼ δS þ δV ¼ 0 ð5:38Þ
∂S ∂V

and as the sufficiency, the second differential d2U should be positive definite for the
virtual variations of S and V or
 2 2 2 
1 ∂ U 2 ∂ U ∂ U 2
d U¼
2
δS þ 2 δS δV þ δV > 0 definite: ð5:39Þ
2 ∂S2 ∂S∂V ∂V2

You see, d2U takes a quadratic form in terms of the variable δS/δV by dividing
by δV2(>0) or, equivalently, in terms of the variable δV/δS by dividing by δS2(>0)
instead. For d2U to be positive definite, these two quadratic functions should be each
convex downward and stay above the axis of the variable, thus not to have the real
5.4 Spinodal Decomposition 191

roots. In other words, the coefficient of the second-order term should be positive and
the discriminant of the quadratic equation be negative:

2 2  2 2  2  2 
∂ U ∂ U ∂ U ∂ U ∂ U
> 0, > 0;  < 0: ð5:40Þ
∂S2 ∂V2 ∂S∂V ∂S2 ∂V2

These are called the intrinsic stability criteria. Due to the thermodynamic
identities, the first two conditions are rewritten, respectively, as
 2     2  

∂ U ∂T T ∂ U ∂P 1
¼ ¼ > 0; ¼ ¼ > 0: ð5:41Þ
∂S V
2 ∂S V c V ∂V2 S ∂V S κ SV

That is, for a system to be intrinsically stable, its heat capacity at constant
volume cV and adiabatic compressibility κS should be positive definite. Otherwise,
the energy minimum principle (or equivalently, entropy maximum principle)
would be violated, viz., heating and pressurizing would cause your system to cool
down and expand, respectively. At this point, you may wish to see for yourself what
kind of stability criterion pops out from the third condition in Eq. (5.40).
What about the system stability against composition fluctuation at given tem-
perature and pressure? At constant temperature and pressure, the energy minimum
principle stipulates that the Gibbs free energy should be minimum. As we played
with the U-minimum principle above, the variation of Gibbs free energy δG)T,P
induced by the concentration fluctuations δci of component i at given T and P may be
written as

X ∂G 1 XX ∂ G
2
δGÞT,P ¼ δci þ δc δc þ   : ð5:42Þ
i
∂ci 2 j i ∂cj ∂ci i j

For G to be minimum,

X ∂G
dG ¼ δci ¼ 0 ð5:43Þ
i
∂ci

as the necessity, and

XX ∂ 2 G
d2 G ¼ δci δcj > 0 definite ð5:44Þ
j i
∂cj ∂ci

as the sufficiency. It follows for, e.g., a binary system (i ¼ 1,2), as in Eq. (5.40), that
against any composition fluctuation, the system is:
192 5 Kinetics of Phase Transformation: Initial Stage

 2 
∂ G
Stable if > 0; ð5:45aÞ
∂c2i T,P
 2 
∂ G
Unstable if < 0; ð5:45bÞ
∂c2i T,P
 2 
∂ G
Indeterminate if ¼ 0: ð5:45cÞ
∂c2i T,P

Can you recognize that the stability criterion, Eq. (5.45a), is nothing but the Le
Chatelier principle? That is, its chemical potential should increase as the amount
of a component increases, just as temperature increases when heat flows in, if the
system is stable. If volume V is fixed instead of pressure P, the Helmholtz free
energy F, instead of G, should be minimum.

5.4.2 Gibbs Free Energy vs. Composition Diagram

Now, let us look at a hypothetical, immiscibility phase diagram of system A–B


with the upper consolute temperature, TC, Fig. 5.10a. At a temperature above TC,
say, T ¼ T3, our system exists in a single-phase solution “s” across its entire
composition range 0
NB(¼cBVm)
1. Thus its relative integral molar free
energy of mixing, ΔGM, should be convex downward across the entire composition
range as shown at T ¼ T3 in Fig. 5.10b. At a temperature lower than TC, say, T ¼ T2,
our system exists in two phases s1 (with composition “α”) and s2 (with the compo-
sition “β”) across the immiscibility gap between the composition α and β in
Fig. 5.10a. Recollecting the thermodynamic identity, ∂ΔGM/∂T ¼ -ΔSM, as tem-
perature decreases from T3, the convex-downward ΔGM-curve starts to increase
fastest at the composition where the relative integral molar entropy of mixing
ΔSM is largest, normally the most mixed-up composition (NB ¼ 1/2 in the present
symmetric case). The ΔGM vs. composition (NB) curve, ΔGM(NB), may, thus, take
the shape at, e.g., T ¼ T2 and T1 as illustrated in Fig. 5.10b. There should then be a
temperature demarcating the convex-downward and convex-upward ΔGM(NB), that
is, T ¼ TC, the consolute temperature, where ΔGM(NB) exhibits rather flat a
minimum. The common tangent to ΔGM(NB) at, e.g., T1 results in the two points
of contact “α” and “β” in Fig. 5.10b, which correspond to the conjugate composi-
tions “α” and “β” in Fig. 5.10a. The trace of such common-tangent contact points
“α” and “β” at different temperatures generates the immiscibility dome as delineated
by the solid line in Fig. 5.10a. You may recognize that ΔGM(NB) at a temperature
below TC, say, T ¼ T1, consists of the convex-downward regions (A < NB < α’ and
β’ < NB < B) where ∂2ΔGM/∂NB2 > 0 and the convex-upward region (α’ < NB < β’)
where ∂2ΔGM/∂NB2 < 0. According to the stability criteria with respect to small
composition fluctuations, Eq. (5.45), our system is stable in the former and unstable
5.4 Spinodal Decomposition 193

Fig. 5.10 (a) Hypothetical T s


phase diagram with an T3
immiscibility gap below the Tc
consolute temperature TC. s1 s2
(b) Corresponding “relative
molar Gibbs free energy of T2
mixing” ΔGM vs. composition α α′ β′ β
s1+s2
NB diagram at selected T1
temperatures T1, T2, TC, α α′ β′ β
and T3
A NB B
(a)

(T=T1)

B
0 A
α′ β′
α β
α′ β′ β
α
(T=T2)

(T=Tc)

(T=T3)
ȟG M
RT
(b)

in the latter. The borders between these two regions, i.e., points α’ and β’ where
∂2ΔGM/∂NB2 ¼ 0, are termed the spinodes. The loci of the spinodes (α’ and β’)
against T in Fig. 5.10b are denoted by the dashed curve in Fig. 5.10a, which is
termed the spinodal curve.

5.4.3 Fate of Composition Fluctuations

Our system, whether as a whole or in part, is continually interacting with its


surrounding, and hence, minor disturbance or fluctuation of the local state variables
is unavoidable. Let us now examine what would happen to a composition
fluctuation depending on whether the initial composition is in the convex-upward
(∂2ΔGM/∂NB2 < 0) or convex–downward (∂2ΔGM/∂NB2 > 0) region; see Fig. 5.11.
Suppose that we have a homogeneous solution 1 (composition c1) in the convex-
2
downward (∂ ΔGM =∂N2B > 0) region. Obviously, this solution is not in the abso-
lute thermodynamic equilibrium: This should decompose eventually into the two
solutions of composition “α” and “β,” thus to decrease the free energy from 1 to 1000 .
Now, imagine that this composition 1 is rendered to fluctuate by a small degree, say,
between 10 and 100 . Then, you can see that this fluctuation increases the mean free
194 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.11 Free energy ∂2 G M ∂2 G M ∂2 G M


GM >0 <0 >0
ΔGM vs. composition CB ∂N B2 ∂N B2 ∂N B2
(¼NB/Vm) diagram at a 0
temperature, e.g., 0''
T ¼ T1 < TC in Fig. 5.10 0'

1''

1' 1
1''' 0'''

C C1 C0 C CB (=NB/Vm)

energy of the system, rather than decreases, as indicated by an arrow upward at 1.


The natural direction of change at given temperature and pressure (or volume) is to
decrease the free energy of the system. Any composition fluctuation of this sort
should, thus, die away spontaneously to reduce the overall free energy back to 1.
You may recognize, however, that if the fluctuation is large enough to the right
(of the intersection of the tangent at 1 with the ΔGM-curve), the mean free energy
gets lower instead. In this light, our system 1 is said to be stable with respect to
minor fluctuations of composition. It is, nevertheless, not absolutely stable ther-
modynamically, thus referred to as metastable because it has still to go further down
to 1000 by decomposing into the two conjugate solutions α and β.
What about the solution 0 (composition co) within the convex-upward
2
(∂ ΔGM =∂N2B < 0) region (in Fig. 5.11) then? Again suppose that its composition
fluctuates to separate into compositions 00 and 000 . Unlike the previous case, the free
energy always decreases as indicated by an arrow downward at the composition 0 no
matter whether small or large in fluctuation. Change in the decreasing direction of
the Gibbs free energy is natural, spontaneous, and irreversible. This fluctuation
should, thus, continue to increase, decreasing the system free energy further. In
this light, our system 0 (composition co) is said to be unstable with respect to
minor composition fluctuation. This kind of composition segregation spontane-
ously continues eventually to α and β, thus achieving the thermodynamic equilib-
rium (0000 ). In this light, the system 0 is intrinsically unstable, thence even cannot be
made into existence at all, unlike a metastable state, 1. The spinodal (∂2ΔGM/
∂NB2 ¼ 0), thus, corresponds to the limit of the metastability (∂2ΔGM/∂NB2 < 0) as
defined earlier by Gibbs.3

3
The Scientific Papers of J. Willard Gibbs, p.105, Dover, 1961
5.4 Spinodal Decomposition 195

Fig. 5.12 Schematic c classical nucleation and growth


temporal evolution of
concentration profiles: (a)

Upon nucleation followed (a)
by growth via downhill
(arrows) diffusion; (b) upon c1
spinodal decomposition

Early Later Final
with uphill diffusion
(arrows) from early through distance
later to final. (Redrawn from c spinodal decomposition
Cahn [14])

(b)
co

Early Later Final

5.4.4 How Phase Transition Initiates?

Any homogeneous solution within the composition range cα < c < cβ, if any, would
not be in thermodynamic equilibrium: it should separate into two conjugate
equilibrium phases α and β so that the system free energy is lowered eventually
down to the common tangent (e.g., 0000 and 1000 ). In the composition range where
∂2ΔGM/∂NB2 > 0, however, our system is not in the absolute thermodynamic
equilibrium, but stable with respect to minor composition fluctuations. It is then
obvious that an unavoidable minor fluctuation of composition can never lead to
phase separation into α and β because our system is resilient to any minor
composition fluctuation. There should then be large enough a fluctuation to initiate
the phase separation, that is, the nucleation of phase “β” under the action of the
driving force, the free energy decrease from 1 to 1000 in Fig. 5.11 as we learned
earlier.
When the β-phase nucleates in the matrix of otherwise homogeneous composition
c1, the composition profile along the spatial coordinates may be as shown in
Fig. 5.12a (Early). A nucleus is small in spatial extent, but the composition differ-
ence between the nucleus and its immediate surrounding, cβ–cα, is extreme. This
process of nucleation is, thus, referred to as “small in extent (size); large in degree
(composition difference).” Once a nucleus forms, the composition right at its
external surface should be cα, the composition of its conjugate phase α (due to
the local equilibrium). There, thus, forms a concentration gradient in the immediate
surrounding of the nucleus, and hence, diffusion occurs down the concentration
gradient (downhill diffusion) as indicated by arrows in Fig. 5.12a (Early), which
causes the nucleus to grow (Fig. 5.12a (Later)) till the completion of phase separa-
tion as shown in Fig. 5.12a (Final). This is the scenario of the nucleation and
growth mechanism of phase separation from the beginning to the end.
In the composition range where ∂2ΔGM/∂NB2 < 0, i.e., within the spinodal, on
the other hand, our system of, e.g., composition co in Fig. 5.11 is already unstable
196 5 Kinetics of Phase Transformation: Initial Stage

with respect even to minor fluctuations in composition. Once initiated, thus, a minor
fluctuation may continue to grow with time by diffusion up the concentration
gradient or uphill diffusion, as indicated by the arrows in Fig. 5.12b (Early and
Later), which eventually let the initially homogeneous phase 0 decompose into α and
β as shown in Fig. 5.12b (Final). This is called the mechanism of spinodal decom-
position. In the very initial stage of decomposition, composition fluctuation occurs
globally everywhere or over large spatial extent, but with a small difference in
composition. The initial stage of spinodal decomposition is, thus, referred to as
“large in extent; small in degree,” which is exactly opposite to the case of the
nucleation (Fig. 5.12a (Early)). How diffusion proceeds uphill the concentration
gradient? You are reminded that the thermodynamic factor within the spinodal is
negative, and hence, the intrinsic component diffusivities are rendered to be
negative.
We have learned that the nucleation always requires an energy cost, viz., the
interfacial energy and even the strain energy, if not totally relieved, which should
be paid by the volume free energy gain, thus requiring a certain degree of
supercooling or superheating. No energy cost in the spinodal decomposition?
Once a composition gradient is introduced into an otherwise homogeneous matrix,
there should then be the similar energy cost to pay, viz., the interfacial energy and
strain energy. From this point on does J. Cahn’s classic seminal theory start. We
will begin by trying to understand the origin of the interfacial energy before jumping
into his theory on spinodal decomposition.

5.4.5 Origin of Interfacial Energy: Gradient Energy

We know that when two different phases, say, α and β, are in contact, there is an
interface with an excess energy, called the interfacial energy, over the total energy
which, if separated, the two homogeneous bulk phases would have. Macroscopi-
cally, the interface is regarded as a compositional singularity or the plane where the
composition changes discontinuously. Microscopically, however, the composition
should change gradually across the interphase as schematically illustrated in
Fig. 5.13.
For a system comprising components A and B, let εAA, εBB, and εAB denote the
binding energy of A–A, B–B, and A–B bond, respectively, where εAA, εBB < 0 by
nature. It is then easy to imagine, recollecting the pairwise potential model of
solutions [16], that, if ω  εAB–(εAA + εBB)/2 > 0 or each atom likes the atoms of the
same kind more than those of different kind, the solution tends to cluster into the
A-rich and B-rich phases, thus forming an interface in between; if ω < 0 or each atom
likes more the atoms of different kind, the solution tends to mix together, thus
tending to eliminate the interface. Referring to Fig. 5.13, let us now calculate the
binding energy Ep,p+1, per unit area, of the two neighboring atomic planes p and p+1
with composition (in atom fraction of B) Np and Np+1, respectively, by counting the
5.4 Spinodal Decomposition 197

Fig. 5.13 Compositional


variation across a diffuse
interface. (Reproduced from
Kingery et al. [9])

number of A–A, B–B, and A–B bonds, PAA, PBB, and PAB, respectively, between
these two planes as

Εp,pþ1 ¼ PAB εAB þ PAA εAA þ PBB εBB : ð5:46Þ

Assuming that there are “m” atoms per unit area of each plane, an atom in a plane
is coordinated by z atoms in the two nearest-neighboring planes (i.e., z/2 per each
plane), and the atoms are distributed at random in each plane,

  1  
PAA ¼ m 1  Np  z 1  Npþ1 ;
2
1
PBB ¼ mNp  zNpþ1 ; ð5:47Þ
2
1     1
PAB ¼ mNp  z 1  Npþ1 þ m 1  Np  zNpþ1 ,
2 2

as for PAA, for example, z(1Np + 1)/2 A-atoms in the plane p+1 are bonded to each
of m(1Np) A-atoms in the plane p.
198 5 Kinetics of Phase Transformation: Initial Stage

Similarly, you can calculate Ep,p, the energy, per unit area, of two contiguous
planes with the same composition Np (by replacing Np for Np+1 in Eq. (5.47)) and
Ep+1,p+1 with the same composition Np+1 (by replacing Np+1 for Np in Eq. (5.47)).
The excess energy ΔE per unit interphase area due to the composition difference is
then obtained as
 2
1  1 2 ∂N
ΔΕ ¼ Εp,pþ1  Εp,p þ Εpþ1,pþ1 ¼ m zωa ð5:48Þ
2 2 ∂x

where letting “a” denote the interplanar spacing (see Fig. 5.13), we have
approximated

∂N
Npþ1  Np ¼ a: ð5:48aÞ
∂x

Noting that this excess energy ΔE per unit interphase area is nothing but the
interfacial energy, this equation may be rewritten in the form independent of the
coordinates as

ΔE ¼ κð∇cÞ2 ð5:49Þ

where the proportionality constant, κ, is such that

κ / ω > 0: ð5:50Þ

This surface excess energy or interfacial energy is attributed to the composition


gradient ∇c, thus often termed the gradient energy. It is noted that the condition,
Eq. (5.50), may be regarded as another kind of intrinsic stability criterion: If κ < 0,
the two phases α and β in contact would not be stable because A and B tend to get
uniformly mixed up. This neat picture of the interfacial energy was first invented by
R. Becker [17].
We learned that within the spinodal, our system is unstable with respect to minor
fluctuations of composition (Fig. 5.11), and hence, the fluctuations may continue to
grow eventually leading to separation into the two equilibrium phases α and β
(Fig. 5.12b). If there were no energy cost at all with a compositional fluctuation,
any fluctuation component, once started, would then continue to grow with no
constraint at all, like the nucleation with no energy barrier. Now, there comes an
energy increase or barrier due to the compositional fluctuation itself, which is the
gradient energy, Eq. (5.49), and this should act to subdue the composition fluctu-
ations. In addition, there may also arise a strain energy due to the fluctuation if the
molar volume of the solution is dependent on composition and the solution is rigid.
We will neglect this coherency strain energy for the time being by assuming that
the molar volume of our system is independent of composition.
5.4 Spinodal Decomposition 199

Now suppose that there sets in a composition fluctuation of a wave length λ,


which decreases the volume free energy (see Fig. 5.11), but simultaneously intro-
duces the excess gradient energy (see Eq. (5.49)). For a fluctuation across a given
concentration difference Δc, the concentration gradient induced may be approxi-
mated roughly as
Δc
∇c  : ð5:51Þ
λ

The smaller λ, the larger ∇c, thence the larger the gradient energy. If the latter is
larger than the volume free energy gain due to the compositional instability,
Eq. (5.45b), such a wave length fluctuation will spontaneously die out and otherwise
will continue to grow. There will then be a critical wave length λcr for which the
gradient energy is exactly balanced by the volume free energy decrease. Any
fluctuation with λ > λcr shall then survive to grow and any pffiffiffi other with λ < λcr die
away. You will learn shortly that the fluctuation of λ ¼ 2λcr growspfastest,
ffiffiffi thus
leading to phase separation with rather a regular interphase spacing of 2λcr. This is
what is going on with the spinodal decomposition. Let us now explore the theory
quantitatively.

5.4.6 Cahn’s Theory on Spinodal Decomposition

Free Energy Change with Composition Fluctuations


Suppose that minor composition fluctuations set in an otherwise homogeneous
solution of composition co (¼NB,o/Vm) within the spinodal at a given temperature;
see Fig. 5.12b (Early). Assuming that the volume V of our system is kept constant,
we may then write the Helmholtz free energy of the system, F, as4

4
J. W. Cahn and J. E. Hilliard [18] have first derived Eq. (5.52) in a really elegant way:
When composition fluctuates, the local free energy should be a function not only of the local
composition c itself but also of the composition distribution in the immediate surrounding, ∇c, ∇2c,
etc., or f ¼ f(c, ∇c, ∇2c,. . . .). Expanding this function into the Taylor series, one obtains a function
of the form, f ¼ f(c)+κ0∇c+κ1∇2c+κ2(∇c)2 + . . . . Here, κj(j ¼ 0,1,2,. . .) denotes the corresponding
Taylor series coefficient involving the partial derivatives of f with respect to each variable, and
κ00 since f, a scalar, must be invariant with respect to the direction of ∇c. Taking only the leading
terms, noting that, due to the divergence theorem,
Z Z Z
dκ1
κ1 ∇2 cdV ¼ κ1 ∇c  b
ndA  ð∇cÞ2 dV: ðaÞ
dc
V A V

with b
n being the unit normal vector to the surface element dA enclosing V and neglecting
R the surface
integral assuming ∇c  b n ¼ 0 at the surface, one may immediately recognize that F ¼ fdV leads to
V
κκ2–(dκ1/dc). If you are puzzled by Eq. (a), try to carry out the surface integral on its right-hand
side, by taking into account that κ1¼κ1(c), a function of c, for a small volume element
ΔV¼ΔxΔyΔz with the use of the method of converting a surface integral to the volume integral;
see Eq. (1.17) in Chap. 1.
200 5 Kinetics of Phase Transformation: Initial Stage

Z h i
F¼ f ðcÞ þ κð∇cÞ2 dV: ð5:52Þ
V

That is, the free energy of a small volume of nonuniform solution can be
expressed as the sum of two contributions, one being the free energy that this volume
would have in a homogeneous solution at a composition c, f(c), and the other the
gradient energy, κ(∇c)2, which is a function of the local composition. Expanding
f(c) into the Taylor series around f(co), the free energy density of the homogeneous
solution of the initial composition co, you obtain

1
f ðcÞ ¼ f ðco Þ þ f 0 ðco Þðc  co Þ þ f 00 ðco Þðc  co Þ2 þ   : ð5:53Þ
2

Noting that
Z
ðc  co ÞdV ¼ 0 ð5:54Þ
V

due to the mass conservation, the free energy change ΔF due to composition
fluctuations (c–co) may, thus, be written as
Z Z h i
1 00
ΔF ¼ F  f ðco ÞdV ¼ f ðco Þðc  co Þ2 þ κð∇cÞ2 dV ð5:55Þ
2
V V

by taking only the leading 2nd-order term in Eq. (5.53), that is, in the very initial
stage of decomposition where

j c  co j
<< 1: ð5:56Þ
co

If ΔF > 0, our system is stable so that such fluctuations die out; if ΔF < 0, our
system is unstable so that such fluctuations continue to grow. The condition for the
instability (ΔF < 0) may, thus, be written as

1 00
f ðco Þðc  co Þ2 þ κð∇cÞ2 < 0 ð5:57aÞ
2

which means that

1 00
f ðco Þðc  co Þ2 < κð∇cÞ2 < 0: ð5:57bÞ
2
5.4 Spinodal Decomposition 201

as κ > 0. Therefore, the necessity for our system to be unstable with respect to minor
fluctuations is f00 (co) < 0 in agreement with Fig. 5.11. Otherwise, our system would
remain stable, and hence, minor fluctuations are to die out. For an infinite wave
length fluctuation (i.e., ∇c ! 0), the limit of metastability is, thus, f00 (co) ¼ 0 or the
spinodal.
If there arises a coherence strain with composition fluctuations, the strain
energy should also be taken into account in Eq. (5.55). Letting η denote the strain
ε per unit composition change or

ε
η , ð5:58Þ
c  co

the total free energy change may then be written as


Z 
1 00 Eη2
ΔF ¼ f ðco Þðc  co Þ2 þ κð∇cÞ2 þ ðc  co Þ2 dV ð5:59Þ
2 1ν
V

for, e.g., the isotropic case. Here, E stands for Young’s modulus and ν Poisson’s
ratio. For more in-depth treatment, you are referred to the original paper by Cahn
[19]. For the system to be unstable with respect to minor fluctuations or for ΔF < 0,
thus,

1 00 Eη2
f ðco Þðc  co Þ2 þ κð∇cÞ2 þ ð c  co Þ 2 < 0 ð5:60aÞ
2 1ν

or

1 00 Eη2
f ðco Þðc  co Þ2 þ ðc  co Þ2 < κð∇cÞ2 < 0: ð5:60bÞ
2 1ν

The necessity for the instability is now

1 00 Eη2
f ð co Þ ð c  co Þ 2 þ ð c  co Þ 2 < 0 ð5:61Þ
2 1ν

and hence, the limit of metastability (for an infinite wave length fluctuation) is

1 00 Eη2
f ð co Þ þ ¼ 0: ð5:62Þ
2 1ν

The spinodal has earlier been defined as the loci of the point f00 ¼ 0. When the
coherence strain comes into the scene, the spinodal should then be the loci of the
point as defined as Eq. (5.62) or
202 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.14 Incoherent


miscibility gap with the
chemical spinodal
(f00 (co) ¼ 0) and coherent
miscibility gap with the
coherent spinodal
(f00 (co) + 2Eη2/(1ν) ¼ 0)

2Eη2
f 00 ðco Þ ¼  : ð5:62aÞ
1ν

This is called the coherent spinodal, while the earlier f00 (co) ¼ 0 is called the
chemical spinodal. These are compared in Fig. 5.14. You can recognize that the
coherence strain lowers the consolute temperature. Think about why it should be.
Composition Fluctuations and Critical Wave Length
In general, any three-dimensional composition fluctuation c(r,t)–co at position r and
time t may be represented in terms of the Fourier series as
X ! !
c  co ¼ A n ei β n  r : ð5:63Þ
n

where An and βn are the amplitude and wave vector, respectively, of the
nth-component fluctuation wave. For the sake of mathematical simplicity, let us
consider one-dimensional fluctuation in the x-direction. A wave component may
then take the form


c  co ¼ A cos βx; β  : ð5:64Þ
λ

By substituting this equation into Eq. (5.55), you may obtain


Z L 
1 00
ΔF ¼ A2 f ðco Þ cos 2 βx þ κβ2 sin 2 βx dx ð5:65Þ
0 2

across the length of the system L per unit cross-sectional area (i.e., V ¼ Lx1). Upon
integration by using the periodic boundary condition, c(x ¼ 0) ¼ c(x ¼ L),
5.4 Spinodal Decomposition 203

A2 L 00
ΔF ¼ f ðco Þ þ 2κβ2 , ð5:65aÞ
4

which leads to the critical wave vector βcr or wave length λcr for ΔF ¼ 0 as
 1=2
2π f 00 ðco Þ
βcr ¼ ¼ : ð5:66Þ
λcr 2κ

If β > βcr or λ < λcr, ΔF > 0 because the gradient energy predominates (i.e.,
κ(∇c)2 > |f00 (c–co)2|), and hence, such fluctuations die out. If β < βcr or λ > λcr, ΔF < 0
because the volume free energy predominates (i.e., κ(∇c)2 < |f00 (c–co)2|), and hence,
such fluctuations grow. How fast do they grow?
Decomposition Kinetics in the Initial Stage
When composition fluctuates, the flux of each component i (¼1,2) that is observable
in the laboratory frame of reference may be written as

Ji ¼ ci Bi ∇μi þ ci v ð5:67Þ

where Bi denotes the intrinsic mechanical mobility of i and v the drift velocity of
the matrix. Assuming the constant molar volume Vm of the system for the sake of
simplicity (i.e., the volume-fixed or Fick’s frame of reference5),

J1 þ J2 ¼ 0: ð5:68Þ

You may then calculate the drift velocity v and put it back to Eq. (5.67) to obtain
the observable flux of the solute 2 as6

J2 ð¼ J1 Þ ¼ co M∇ðμ2  μ1 Þ ð5:69Þ

where

1
co ¼ c1 þ c2 ¼ ; M ¼ N1 N2 ðN1 B2 þ N2 B1 Þ: ð5:70Þ
Vm

What is the driving force —(μ2–μ1) in the present system involving the volume
free energy and gradient energy as well? It is noted that the free energy density f is
given in any case as

5
Remember that the observable flux Ji in Eq. (5.67) was written as FJi in Chap. 4.
6
Note that
B2∇μ2  B1∇μ1 ¼ (N1B2 + N2B1)∇(μ2  μ1) + (B2  B1)(N2∇μ2 + N1∇μ1), where N2∇μ2
+ N1∇μ1 ¼ 0 due to the Gibbs-Duhem equation.
204 5 Kinetics of Phase Transformation: Initial Stage

f ¼ c1 μ1 þ c2 μ2 ð5:71Þ

and hence,

∂f
¼ μ2  μ1 : ð5:72Þ
∂c2

In order to evaluate the right-hand side of this equation for the present case, let us
calculate the variation of the total free energy δF (¼δΔF) induced by the compo-
sition variation, δc (¼δc2), in Eq. (5.55):
Z
δF ¼ δðΔFÞ ¼ ½f 00 ðco Þðc  co Þδc þ 2κð∇cÞδð∇cÞ dV ð5:73Þ
V

which, if one dimension, would take the form

ZL     
dc dc
δF ¼ δðΔFÞ ¼ f 00 ðco Þðc  co Þδc þ 2κ δ dx: ð5:73aÞ
dx dx
0

By noting that δ∇c ¼ ∇δc,7 you can integrate by parts the 2nd term inside the
integral to obtain

ZL      L ZL  2 
dc d δc dc dc
2κ dx ¼ 2κ δc  2κ δcdx, ð5:74Þ
dx dx dx 0 dx2
0 0

where the first term on the right-hand side is rendered to be 0 again due to the
periodic boundary condition.
Translating Eq. (5.74) back into three dimensions and substituting into
Eq. (5.73a), you get
Z
δðΔFÞ
¼ f 00 ðco Þðc  co Þ  2κ∇2 c dV: ð5:75Þ
δc
V

The integrand should then correspond to the first derivative of the free energy
density f with respect to composition c or

7
Letting δc ¼ c(2)–c(1), ∇δc¼∇(c(2)–c(1))¼δ∇c.
5.4 Spinodal Decomposition 205

∂f
¼ μ2  μ1 ¼ f 00 ðco Þðc  co Þ  2κ∇2 c: ð5:76Þ
∂c

Subsequently, Eq. (5.69) is rewritten as


 
J ¼ co M f 00 ðco Þ∇c  2κ∇3 c ð5:77Þ

where we omitted the subscript for the component 2 for simplicity sake. Upon
comparison with Fick’s first law, the coefficient of ∇c should then be the chemical
or interdiffusion coefficient or

e ¼ co Mf 00 ðco Þ
D ð5:78Þ

which takes a negative value within the spinodal where f00 (co) < 0, in perfect
agreement with Darken’s equation, Eq. (4.43). You are asked to prove this by
rewriting f00 in terms of the thermodynamic factor.
The decomposition kinetics, the temporal variation of local composition c, may
then be described as

∂c
¼ ∇J ð5:79Þ
∂t

or due to Eqs. (5.77), (5.78), and (5.66),


 
∂c e 1 4
¼D ∇ cþ 2∇ c
2
ð5:79aÞ
∂t βc

where we assume that the kinetic and thermodynamic properties M, f00 and κ, or D e
and βc, remain constant for small fluctuations, i.e., in the very initial stage of
decomposition [see Eq.(5.56)].
Instead of formally solving Eq. (5.79), let us assume that the general solution
should take the form
!!
cðr, tÞ  co ¼ AeRβ t ei β  r ð5:80Þ

or in one dimension for mathematical simplicity

cðx, tÞ  co ¼ AeRβ t cos βx ð5:80aÞ

where Rβ, a function of β, may be called the amplification factor. It is easy to find,
by substituting this trial solution into Eq. (5.79), that for it to be the solution, the
amplification factor should be
206 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.15 Schematic of Rβ vs. β2 (a) and equivalently, Rβ vs. λ (b). Note that Rβ exhibits rather a
sharp maximum Rβ,max against λ

"
 2 #
e β
Rβ ¼ Dβ 1 2
: ð5:81Þ
βc

Remembering that D e < 0 within the spinodal, Rβ varies against β2 or equiva-


lently against λ(¼2π/β) as illustrated in Fig. 5.15.
You can see therefrom:
(i) Rβ > 0, if β < βcr or λ > λcr;
(ii) Rβ < 0, if β > βcr or λ < λcr;
(iii) Rβ ¼ 0, if β ¼ βcr or λ ¼ λcr.
That is, any fluctuation with λ > λcr grows with time because Rβ > 0 but that with
λ < λcr dies away with time because Rβ < 0, as we have already predicted on the basis
of Eq. (5.51).
Furthermore, you can see that Rβ exhibits a maximum Rβ,max at (β/βcr)2 ¼ 1/2 or
pffiffiffi
λ ¼ 2λcr (i.e., ∂Rβ/∂β ¼ 0) such that

1e 2
Rβ, max ¼  Dβ ð5:82Þ
4 cr

or due to Eq. (5.66)


 
1 e f 00 ðco Þ
Rβ, max ¼  D : ð5:82aÞ
4 2κ

Equation (5.81) may then be rewritten as


Spinodal Decomposition 207

 2 "  2 #
Rβ β β
¼4 1 : ð5:83Þ
Rβ, max βcr βcr

Morphology Expected
Finally, the composition fluctuations may be represented in one dimension as
X
c  co ¼ Aβ eRβ t cos βx ð5:84aÞ
β

or in three dimensions as
X !!
c  co ¼ A β e Rβ t e i β  r ð5:84bÞ
β

where the summation runs for all possible β or wave length λ (¼2π/β).
pffiffiffi We nowp know
ffiffiffi
that among these wave components, the one with βm ¼ βcr = 2 or λm ¼ 2λcr
grows fastest, thus eventually to dominate the microstructure. You may p then
ffiffiffi easily
expect a stratum-like microstructure with a regular spacing λm ¼ 2λcr for
one-dimensional decomposition as illustrated in Fig. 5.16. Figure 5.17 shows the
equally spaced cross-sectional views of numerical simulation results in three dimen-
sions. Note that the separated phases are each connected while keeping a character-
istic length between the two phases.

Fig. 5.16 Microstructure


expected for a
one-dimensional spinodal
decomposition into α and β

An actual microstructure [21] is shown in Fig. 5.18, which has been formed upon
cooling a homogeneous BaO-SiO2 glass with a composition (0.1BaO + 0.9SiO2)
close to the center of the metastable miscibility gap of the system. Indeed, the
structure appears to consist of two separated phases, each of which formed a three-
dimensional structure, continuously interconnected throughout the glass.
208 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.17 A succession of computed sections through the 50:50 two-phase structure for 100 random
sine waves. Note that all particles are interconnected. The spacing between sections is 1.25/βm. Top:
Z ¼ 1.25/βm (left) and Z ¼ 2.5/βm (right). Bottom: Z ¼ 3.75/βm (left) and Z ¼ 5/βm (right).
(Reproduced from Cahn [20])

Fig. 5.18 Transmission


electron micrograph of 10%
BaO silicate glass cooled
from the melt. This
composition is close to the
top of the metastable liquid-
liquid immiscibility dome.
(From Seward et al. [21])
Problems 209

Problems

1. A metal melts at 1000 K; the latent heat of fusion is 2000 cal/mol. The heat
capacity of solid and liquid are 6 and 10 cal/mol-K, respectively.
(a) What are the Gibbs free energy changes for the system and entropy changes
for system and reservoir for isothermal solidification at 800 K, 880 K, 990 K,
999 K, and 1001 K?
(b) What are the corresponding free energy and entropy changes for complete
solidification in an adiabatic system?
(c) Which of the changes are possible? Which leads to equilibrium?
(d) The metal transforms to another solid at 600 K with a latent heat of 400 cal/
mol-K. What is the metastable melting point of the low-temperature
modification?
2. For the solidification of molten (i) cesium and (ii) nickel by homogeneous
nucleation, construct plots of ΔG and r each as functions of the degree of
undercooling, ΔT.
Data:

Cs Ni
Melting point ( C) 28.44 1453
Surface energy (J/m2) 0.069 1.778
Density (g/cm3) 1.854 7.905
Enthalpy of fusion (kJ/mol) 2.09 17.71
Atomic mass (g/mol) 132.90 58.71

3. (a) Pure metal, M, undergoes a transformation from α-solid by nucleation of


rod-shaped precipitates which can be modelled as cylinders with a radius r
and length ℓ. In this system embryos and nuclei form only with an aspect ratio
(radius/length) which minimizes surface energy. As shown in Fig. 5.19, this
last property is anisotropic: the surface energy is γl across the ends of the
cylinder and γr across the wall of the cylinder. Calculate the size of the critical
nucleus for an undercooling of 50 K.
Data:
Equilibrium transition temperature, Ttr ¼ 400 K
Enthalpy of transition (α ! β), ΔHtr ¼ 1.25 108 J/m3
Axial solid-solid surface energy, γl ¼ 0.2 J/m2
Radial solid-solid surface energy, γr ¼ 0.1 J/m2
(b) Pure metal, N, undergoes a transformation from α-solid by nucleation of
spherical precipitates which change the strain energy of the system. To a first
approximation in this case, the strain energy is proportional to the volume of
the precipitate. Estimate then nucleation rate, I, for an undercooling of 50 K.
210 5 Kinetics of Phase Transformation: Initial Stage

Fig. 5.19 A rod-shaped


nucleus with anisotropic
surface energies

Data:
Equilibrium transition temperature, Ttr ¼ 400 K
Enthalpy of transition (α ! β), ΔHtr ¼ 5 109 J/m3
Axial solid-solid surface energy, γα/β ¼ 0.15 J/m2
Solid-solid strain energy, Eα/β ¼ 2.25 108 J/m3
4. In molten aluminum containing 0.01% titanium, nucleation is observed as near as
0.01 K to the equilibrium transformation temperature. For a nucleation rate of
107 s1 cm3, calculate the height and radius of a spherical cap for heterogeneous
nucleation. Assume that the critical radius for homogeneous nucleation in this
system is about 1 μm. The equilibrium transformation temperature is 932 K, the
enthalpy of fusion is 10.47 kJmol1, and the atomic mass of aluminum is 26.98
gmol1.
5. The rate of nucleation of solid Ge from its liquid has been determined at various
temperatures by rapidly supercooling a large number (1010) of small Ge droplets
and measuring the change in volume of the sample as a function of time. Using
bulk values of the free energy of transformation, △Gv, the investigators were able
to show that the theory of homogeneous nucleation agreed well with their work
and that assumptions of a spherical nucleus and the applicability of the bulk value
of the surface energy γ to nucleus formation were correct within the accuracy of
their experiment. The measurable nucleation rate Iv ¼ 0.1/droplet-sec was
observed for 20 μm diameter droplets when undercooled by 227 K.
(a) Calculate the crystal-liquid surface energy for Ge.
(b) Calculate the critical radius for the nucleation of solid Ge.
Problems 211

Data for Ge:

Tm ¼ 1231 K; ΔHf ¼ 8:3 kcal=mol; ρ ¼ 5:35 g=cm3

6. Liquid, L, wets solid, S, with a contact angle, θ ¼ 20 . Vapor of pure liquid, L, at
1 atm pressure is exposed to a surface of foreign solid, S, which has a site density,
NS ¼ 1015 cm2.
(i) If the surface of solid, S, is perfectly flat, how much undercooling is required
to achieve a nucleation rate of I ¼ 104 cm2 s1?
(ii) If the surface of solid, S, is machined to the shape shown below, how much
undercooling is required to achieve a nucleation rate of I ¼ 104 cm2 s1?

R


6ROLG6

Data:
For species L
ΔS evaporation ¼ 10R
Boiling point ¼1400 K
γvapor/liquid ¼ 0.8 J/m2
Molar volume Vm ¼ 12 cm3/mol
Atomic mass ¼ 39 g/mol
7. Pound and La Mer [J. Am. Ceram. Soc., 47 (1952) 2323] determined the rate of
nucleation of solid tin from its liquid at various temperatures below the melting
point by rapidly supercooling a large number (1010) of small tin droplets,
separated by an oxide film, and measuring the change in volume of the sample
as a function of time. Using the values of transformation free energy per unit
volume, ΔGv, from measurements in bulk transformations, they were able to
show that the theory of homogeneous nucleation agreed well with their work and
that assumptions of a spherical nucleus and the applicability of the bulk value of
the surface energy γ to nucleus formation were correct within their accuracy of
measurement. Calculate the following quantities from their values of nucleation
rate at a constant volume fraction solidified and at 113 K:
(a) The liquid-solid surface energy for tin.
(b) The critical radius for the nucleation of solid tin.
(c) The number of tin atoms in nucleus of critical size.
212 5 Kinetics of Phase Transformation: Initial Stage

Data
(i) Taking into account the change in transformation free energy, ΔGv, with
temperature, the slope of ln (nucleation rate) vs. 1/T curve at 113 K was found
to be:

ðslopeÞ ¼ 23:8 103 K

for any constant fraction of liquid transformed.


(ii) ΔGv ¼ 104 J/m3 at 113  C.
(iii) Radius of tin atom ¼ 1.5 1010 m.
8. (a) Graphite may be converted to synthetic diamonds by application of very high
pressure and temperature. Lower temperatures may be used if small amounts of
nickel are present. Explain. It has been found that the relative lattice parameters of
the graphite and the nickel impurity are very important. Why? What other effects
may be important in this instance.
(b) Suggest another material which might promote the transformation from
graphite to diamond.
9. The system A–B exhibits regular solution behavior in the solid state where the
relative integral molar free energy of mixing, ΔGM, is given as

ΔGM ¼ RTðNA ln NA þ NB ln NB Þ þ ΩNA NB

where NA and NB denote the mole fraction of A and B, respectively. Answer parts
(a) through (e), each when η, the linear strain per unit composition difference, is
equal to (i) η ¼ 0 and (ii) η ¼ 0.05.
(a) Calculate the consolute temperature for solid miscibility.
(b) What is the temperature of the spinodal for the solutions of composition
XA ¼ 0.20?
(c) What is the critical wave length at T ¼ 775 K?
(d) What is the fastest-growing wave length at T ¼ 775 K anywhere in the A–B
system?
(e) What is the maximum value of the amplification factor, R(β), at 775 K
anywhere in the A–B system?
Data:
Regular solution interaction parameter, Ω ¼ 25 kJ/mol
Gradient energy coefficient, κ ¼ 1010 J/m
Young’s modulus, E ¼ 1011 Pa
Poisson’s ratio, ν ¼ 0.3
Self-diffusion coefficient, DA ¼ DB ¼ 104exp(125 kJ/RT) m2/s
Atomic masses, MA ¼ 195 g/mol; MB ¼ 197 g/mol
Densities, ρA ¼ 21.5 g/cm3; ρB ¼ 19.3 g/cm3
Problems 213

10. An Ag-38 at.% Au alloy at 510 K is a single-phase solid solution at equilibrium.


A multilayer thin-film Ag-Au diffusion couple is prepared by evaporation. The
initial composition of the film varies sinusoidally with distance in one dimension
according to

Cðx, tÞ ¼ ð38 at:%AuÞ þ ð12 at:%AuÞ cosβx

where the wave number β ¼ 2π/λ and the wave length λ ¼ 2 109 m.
Estimate the time that it will take to “homogenize” the diffusion couple to the
extent that the maximum composition difference in the sample is 2 at.%
Au. Assume a solution to the diffusion equation having the form

Cðx, tÞ ¼ ð38 at:%AuÞ þ ð12 at:%AuÞ exp ½RðβÞt cos βx

Perform two calculations:


(a) Use Fick’s second law as the diffusion equation:

2
∂c e ∂ c
¼D 2:
∂t ∂x

(b) Use Cahn’s modified diffusion equation:

2 e ∂4 c
∂c e ∂ c 2κ D
¼ D 2  00 :
∂t ∂x f ∂x4

(c) Comment on the difference between your answers to parts (a) and (b). [Note
that the Ag-Au system favors bonds between unlike atoms (“ordering”) and
has a negative gradient energy coefficient.]
Data:
e ¼ 1023 m2 =s; f00 ¼ 5 109 J/m3; κ ¼ 2.6 1011 J/m; λ ¼ 2 109 m
D
11. The rate of development of compositional fluctuations by spinodal decomposi-
tion into zones can also be written as

 2 
e
π Dt
cA ðx, tÞ ¼ cA ðx, 0Þ exp  2
λ

e < 0 since
where cA(x,t) represents the maximum concentration of species A and D
00
G < 0. Answer the following questions to see how changes in t, λ, and T affect
the kinetics of spinodal decomposition.
214 5 Kinetics of Phase Transformation: Initial Stage

(a) What effect does increasing the transition time by a factor of 10 from 10 s to
100 s at room temperature have on the maximum concentration of A when
λ ¼ 0.01 μm and D ¼ 104exp(85 kJ/RT) m2/s?
(b) Compare the maximum concentrations after 100 s at room temperature when
the fluctuation wave length changes by a factor of 10 from 0.1 μm to
0.01 μm.
(c) Compare the maximum concentrations after 100 s for fluctuations of wave
length 0.01 μm in a sample processed at room temperature with those of a
similar sample processed at 100 K above room temperature.
(d) In view of the above calculations, where is the process most sensitive to
diffusion rates?

References

1. D. Turnbull, Phase changes, in Solid State Physics, ed. by D. Turnbull, F. Seitz, vol. 3, (Aca-
demic Press, 1956).
2. K. C. Russel, Nucleation in solids, in Phase Transformations, (ASM, 1970)
3. J. W. Christian, The Theory of Transformations in Metals and Alloys, Part 1, (2nd Ed.,
Pergamon, 1975)
4. D. A. Porter, K. E. Eastering, Phase Transformations in Metals and Alloys, (Van Nostrand
Reinhold(UK) Co., Ltd., 1984)
5. L.S. Darken, R.W. Gurry, Physical Chemistry of Metals (McGraw-Hill Book, Co. Inc., 1953),
p. 163
6. M. Volmer, A. Weber, Z. Physik. Chem. 119, 277 (1925)
7. R. Becker, W. Doering, Ann. Physic 25, 719 (1935)
8. H. Reiss, J. Chem. Phys. 20, 1216 (1952)
9. W.D. Kingery, H.K. Bowen, D.R. Uhlmann, Introduction to Ceramics (Wiley, New York,
1976), p. 182
10. A. Wilm, DRP 244554 (German patent) 1906; Metallurgie 8, 223 (1911)
11. B. Guinier, Structure of age-hardened aluminium-copper alloys. Nature 142, 569–570 (1938)
12. G.D. Preston, Structure of age-hardened aluminium-copper alloys. Nature 142, 570 (1938)
13. M. Hillert, A theory of nucleation for solid metallic solutions, Sc.D. Thesis, M.I.T., 1955
14. J.W. Cahn, Acta Metall. 9, 795 (1961); J. Chem. Phys., 42 (1965) 93; Trans. AIME, 242 (1968)
166
15. J.E. Hilliard, Phase Transformations (ASM, 1970)
16. D.R. Gaskell, Introduction to the Thermodynamics of Materials, 5th edn. (Taylor & Francis,
2008), p. 245
17. R. Becker, Ann. Phys. 32, 128 (1938)
18. J.W. Cahn, J.E. Hilliard, J. Chem. Phys. 28, 258 (1958)
19. J.W. Cahn, On spinodal decomposition in cubic crystals. Acta Metall. 10, 179 (1962)
20. J.W. Cahn, J. Chem. Phys. 42, 93 (1963)
21. T.P. Seward III, D.R. Uhlmann, D. Turnbull, Phase Separation in the System BaO-SiO2. J. Am.
Cearm. Soc. 51, 278–285 (1968)
Chapter 6
Kinetics of Phase Transformation: Later
Stage

6.1 Introduction

A nucleus has just been formed. What will then happen to this nucleus? It starts to
grow via atomic jumps across its boundary or surface. The growth processes
observed in materials may be classified generally into three types:
Type I: Growth without phase change and without composition change, such as
grain growth in a polycrystalline single phase of fixed composition
Type II: Growth with phase change and without composition change, such as
growth of a daughter phase during phase change of a fixed-composition system,
e.g., freezing of water
Type III: Growth with phase change and with composition change, such as
solidification of a binary liquid with solute redistribution or phase separation of
a homogeneous solid solution
Let us learn the old wisdoms dealing with the growth kinetics of each type.

6.2 Type I: Growth Without Phase Change and Without


Composition Change

When heated, an aggregate of fine-grained, strain-free crystals increases continu-


ously in average grain size without change in grain size distribution. This process
is referred to as grain growth. (If strained or deformed, strain-free crystals
nucleate out of the strained matrix and grow at the expense of the strained matrix
with the strain energy as the driving force. This process is called recrystalliza-
tion.) It is experimentally observed that the grain growth is via grain boundary
migration, not via coalescence of neighboring grains as two drops of water

© Springer Nature Switzerland AG 2020 215


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_6
216 6 Kinetics of Phase Transformation: Later Stage

coalesce upon a leaf of lotus or taro, and that the boundaries migrate toward their
centers of curvature. What drives the boundary to migrate?

6.2.1 Driving Force

Let us consider a curved interface or boundary with the principal radii r1 and r2
separating the inner convex side (i) and the outer concave side (o), as illustrated in
Fig. 6.1. The interfacial tension always tends to make the interfacial area shrink,
thus rendering the material in the convex side (i) of the interface to feel a higher
pressure than that in the concave side (o). It is just like that you should apply a higher
pressure inside a rubber balloon in order to keep it ballooned against the external
pressure. How high a pressure do you have to apply to keep a boundary curved?
Let the radii increase virtually by δr, thus causing the volumes of the convex
(i) and concave part (o) to vary by δV(i) and δV(o), respectively, and the interfacial
area by δA. At given temperature and total volume V(i) + V(o), the variation of the
Helmholtz free energy, δF, is then written as

δF ¼ PðiÞ δVðiÞ  PðoÞ δVðoÞ þ γδA ð6:1Þ

where P(i) denotes the pressure inside(i), P(o) the pressure outside(o), and γ the
interfacial energy. From the geometry in Fig. 6.1,

δVðiÞ ¼ δVðoÞ ¼ r1 θ1 r2 θ2 δr;


ð6:2Þ
δA ¼ ðr1 þ δrÞθ1  ðr2 þ δrÞθ2  r1 θ1  r2 θ2  ðr1 þ r2 Þθ1 θ2 δr

for an infinitesimal increase, δr. At equilibrium under given temperature and total
volume, F should be minimum. Thus,

Fig. 6.1 A curved (o)


boundary as defined by the
two principal radii r1 and r2
and the subtending angles θ1
and θ2. When the radii
virtually increase δr
infinitesimally by δr("), the
curved boundary shifts as
delineated by the dotted line
r1
r2
(i)
θ2 θ1
6.2 Type I: Growth Without Phase Change and Without Composition Change 217

   
δF
¼  PðiÞ  PðoÞ r1 r2 θ1 θ2 þ γðr1 þ r2 Þθ1 θ2 ¼ 0, ð6:3Þ
δr T,V

or
 
1 1
ΔP ¼ PðiÞ  PðoÞ ¼ γ þ : ð6:4Þ
r1 r2

If the interface is spherical or r1 ¼ r2 ¼ r, Eq. (6.4) reduces to


Δp ¼ PðiÞ  PðoÞ ¼ : ð6:4aÞ
r

That is, the material in the convex part (i) is subjected to a higher pressure than
that in the concave part (o), and the pressure difference gets larger as r grows smaller.
Equation (6.4) is called the Young-Laplace equation and the pressure difference
ΔP the Laplace pressure.
Remembering the variation of the chemical potential or molar Gibbs free energy
dμ (¼dg) ¼ VmdP at given temperature with Vm being the molar volume, this
pressure difference induces a difference in the chemical potential or molar Gibbs
free energy Δgr (¼Δμ) of the component across the boundary as
 
ðiÞ ðoÞ 1 1
Δgr ¼ μ  μ ¼ Vm γ þ : ð6:5Þ
r1 r2

A consequence is that the atoms in the convex side (i) tend to escape to the
concave side (o), thus making the grain boundary in between migrate toward its
center of curvature. How fast?

6.2.2 Growth Rate

You may imagine that when an atom or molecule crosses a boundary, it should
surmount a migrational energy barrier, Δgm. Due to Eq. (6.5), the energy land-
scape from the convex side (i) to the concave side (o) across the boundary may, thus,
look as illustrated in Fig. 6.2.
The absolute reaction rate theory stipulates the jump frequency νAB from the
convex side (i) to the concave side (o) of the boundary to be
 
Δg
νAB ¼ νo exp  m ð6:6Þ
RT

and the jump frequency in the reverse direction νBA to be


218 6 Kinetics of Phase Transformation: Later Stage

Fig. 6.2 Free energy


landscape across a grain g
boundary from its convex
side A (i) to concave side B Δg m
(o), where Δgr is the A
curvature-induced free
energy difference (Eq. 6.5) (i)
Δg r (= Δμ)
B
(o)

reaction coordinate

 
Δgm þ Δgr
νBA ¼ νo exp  ð6:7Þ
RT

where νo ¼ kBT/h. You should recognize that the activation barrier from B to A is
higher by the curvature-induced Δgr than from A to B. The net atomic jump
frequency from A to B, νnet, is then
   
Δg Δg
νnet ¼ νAB  νBA ¼ νo exp  m 1  exp  r : ð6:8Þ
RT RT

Denoting the grain boundary thickness as λ which is the same as the atomic jump
distance crossing the boundary, the grain boundary migration velocity, or growth
rate, u may be written as

u ¼ λνnet : ð6:9Þ

As an atom jumps across the boundary once, the boundary migrates toward its
center of curvature by the boundary thickness λ! It sounds somewhat too effective.
Anyway, this kind of growth rate is called the normal growth rate.
The curvature-induced free energy difference Δgr is normally very small com-
pared to RT. Typical values for, e.g., elemental metals are γ  1 J/m2 and
Vm  105 m3/mol, and hence, for a typical particle size of, e.g., r1 ¼ r2 ¼ r  1 μm
at, say, T ¼ 1000 K,

Δgr
 103  1: ð6:10Þ
RT

Noting that ex  1 + x for x  1, Eq. (6.8) is approximated as


6.2 Type I: Growth Without Phase Change and Without Composition Change 219

 
Δg Δgr
νnet  νo exp  m  , ð6:11Þ
RT RT

or due to Eq. (6.5) with the assumption of r1 ¼ r2 ¼ r,


 
2γVm Δg
νnet  νo exp  m : ð6:11aÞ
rRT RT

Then, the normal growth rate, Eq. (6.9), takes the form
 
2γVm Δg
u¼ λνo exp  m : ð6:12Þ
rRT RT

Here, the inventors of the theory, Burke and Turnbull [1], assume that the linear
growth rate, u, is proportional to the time change of the mean grain radius r or

dr
u/ ð6:13Þ
dt

and the radius of curvature of a specific boundary r is also

r / r: ð6:14Þ

Substituting Eqs. (6.13) and (6.14) into Eq. (6.12) and integrating from t ¼ 0 to
t ¼ t, you obtain a parabolic rate law

r2  r2o ¼ kðTÞt ð6:15Þ

where ro is the mean grain size at t ¼ 0 and the rate law constant k(T) is a function
of temperature as

4γVm
kðTÞ ¼ λν eΔgm =RT ð6:16Þ
CRT o

with C being the product of the proportionality constants in Eqs. (6.13) and (6.14)
that may be not much different from 1. Recognizing that the grain boundary
diffusivity Dgb ¼ λ2νAB neglecting the geometric factor and taking C  1, this
rate law constant may be rewritten more compactly as

4γVm Dgb
kðTÞ  : ð6:17Þ
RT λ

The reality, however, not rarely deviates from this parabolic kinetics [2]. You
may wish to consider why it is so.
220 6 Kinetics of Phase Transformation: Later Stage

6.3 Type II: Growth With Phase Change and Without


Composition Change

This is typically the process by which the daughter (β)-phase grows into the mother
(α)-phase matrix upon α ! β phase transition of a fixed-composition or single-
component system below the equilibrium transition temperature Ttr. For the β-phase
to grow, atoms should jump from the α-matrix to the β-phase crossing its boundary.
What drives these atoms or molecules to jump toward the β-phase? We assume that
the phase boundary here is planar.
As we have already explored in nucleation, the chemical potential or molar
Gibbs free energy μ of each phase is schematically shown in Fig. 6.3a. Now, Δgβ/α
(¼μβ  μα) <0 below Ttr, which has already driven the nucleation overriding the
interfacial energy increase. The free energy landscape from the α-matrix to the
β-phase across its (planar) boundary may, thus, look as in Fig. 6.3b: The atoms or
molecules should surmount the migrational energy barrier Δgm from α to β and
Δgm + (Δgβ/α) from β to α. The net jump frequency from α to β, thus, takes a
similar form as Eq. (6.8) or
   
Δg Δgβ=α
νnet ¼ ναβ  νβα ¼ νo exp  m 1  exp : ð6:18Þ
RT RT

The normal growth rate is then written formally as Eq. (6.9).

Fig. 6.3 (a) Molar Gibbs (a)


free energies vs. temperature
of α- and β-phase. (b) Free μ α
energy landscape across an
Δg β /α
α-/β-phase boundary. Note β
that the escaping tendency
of the atoms or molecules
in α is higher than in β
by Δgβ/α
ΔT

T Ttr T
(b)
μ

Δg m
α

−Δgβ / α
β

reaction coordinate
6.4 Type III: Growth With Phase Change and With Composition Change 221

You may remember that assuming the heat capacity difference between the two
phases Δcp ¼ 0,

ΔT
Δgβ=α ¼ Δhoβ=α  TΔsoβ=α ¼ Δhoβ=α ð6:19Þ
Ttr

where ΔT (¼Ttr  T) stands for the supercooling. The two extreme cases of the
normal growth rate may then be distinguished depending on ΔT:
   
Δgβ=α  Δhoβ=α ΔT
(i) If  RT  ¼ RTTtr  1 or for small supercooling such that ΔT
Ttr  1,

   
Δhoβ=α ΔT Δg
u ¼ λνnet  λνo exp  m ð6:20Þ
RTTtr RT
 o 
Dgb Δhβ=α
 ΔT ð6:20aÞ
λ RT2tr

as T ¼ Ttr  ΔT  Ttr. The growth rate is essentially proportional to ΔT.


   
Δgβ=α  Δhoβ=α ΔT
(ii) If  RT  ¼ RTTtr  1 or large supercooling,

 
Δg
u ¼ λνnet  νo exp  m ð6:21Þ
RT
Dgb
 : ð6:21aÞ
λ

The growth rate is independent of the supercooling ΔT.


You may already recognize that such normal growth rate is pretty much
overestimated because once crossed the energy barrier, atoms or molecules are
assumed to be 100% accommodated. To be more practical, thus, you may wish to
correct the growth rate by taking into account the accommodation or steric factor
f(<1) as

u ¼ f λνnet : ð6:22Þ

6.4 Type III: Growth With Phase Change and With


Composition Change

In this type, we will take two examples: (i) normal solidification of a binary liquid
solution and (ii) growth of spherical precipitates from a solid solution.
222 6 Kinetics of Phase Transformation: Later Stage

6.4.1 Normal Solidification of a Binary Liquid

Partition Coefficient
Let us consider a binary eutectic phase diagram such as in Fig. 6.4. The ratio of the
solute concentrations of the conjugate solid (CS) to liquid (CL) in the two-phase field
(L + α) is called the distribution or partition coefficient, k (¼CS/CL). In the present
case, k < 1. Generally the liquidus (CL) and solidus (CS) are curved, and hence k is
not a constant, but here we assume k to be constant only to make the discussion
simpler. This means that the liquidus and solidus are straight lines with different
slopes as shown in Fig. 6.4. Even in reality, this assumption may be not so bad for
not so large a k.
Now, suppose that an entire charge of melt X at temperature To is solidified one
dimensionally (along the x-axis), with a planar front (solid/liquid interface), from
one end of the mold (x ¼ 0) by cooling therefrom. This kind of solidification is
referred to as normal solidification or normal freezing. In this case, the speed of
planar growth front or the growth rate of solid α, R, can be controlled by the cooling
rate. Note that, when a unit volume of liquid of solute concentration CL is solidified,
the solute should be rejected from the emerging solid to the surrounding liquid by
(1  k)CL, while kCL (¼CS) remains in the solid, as the partition coefficient, k < 1.
We are, thus, concerned with solute redistributions in the growing solid and receding
liquid. Can they be simply uniform throughout each phase?
When Complete Mixing Both in Solid and in Liquid
The phase diagram in Fig. 6.4 says: When you cool a liquid X from temperature To,
the first solid appears at T1 with the composition kCo; as temperature is lowered
further, the solute concentrations of the emerging solid and the remaining liquid are
supposed to follow the solidus (CS ¼ kCL) and liquidus (CL), respectively, with the
relative amount of each phase stipulated by the lever rule; final solidification should
occur at T3 with solid of composition Co and final liquid of composition Co/k. That
is, the composition of the growing solid CS should vary from kCo through to Co and

Fig. 6.4 Hypothetical T


eutectic phase diagram with
a constant distribution
coefficient k < 1 To
Tm,A

CL
CS L
T1
T2
L+α
T3
α
TE

kCo kC2 Co C2 Co/k CE C


6.4 Type III: Growth With Phase Change and With Composition Change 223

the remaining liquid CL from Co through to Co/k both spatially homogeneously all
the way. This means that complete mixing will have been occurring both in the
advancing solid and in the retreating liquid all the way.
When No Mixing in Solid and Complete Mixing in Liquid
If the solidification is not infinitely slow (R ! 0), complete mixing in solid,
however, will be hard to achieve in reality because it is absolutely diffusion-
controlled, while liquid may be kept homogeneous, e.g., by stirring mechanically
even if diffusion is not fast enough. What kind of solute distribution shall we have in
the solid when no mixing in solid at all and complete mixing in liquid?
When the solid growth front proceeds from x by dx into the liquid ahead
extending to L, the amount of solute rejected by the growing solid (CL  CS)dx
increases the solute concentration by dCL uniformly in the remaining liquid from x
to L because of the complete mixing in liquid, or, due to mass conservation,

ðCL  CS Þdx ¼ ðL  xÞdCL : ð6:23Þ

Assuming the local thermodynamic equilibrium at the solid/liquid interface as


is usually the case,

CS ¼ kCL : ð6:24Þ

Upon integrating Eq. (6.23) with Eq. (6.24) from x ¼ 0, where CS ¼ kCo, to x,
where CS¼CS, you may obtain

CS ¼ kCo ð1  gÞk1 ð6:25Þ

where g (¼x/L) denotes the fraction solidified. This is called the Scheil equation to
remember the inventor [3].
The solute distribution along the solid bar may look as shown in Fig. 6.5. As the
solidification proceeds, solute concentration increases exponentially with x or g (¼x/
L), thus, dubbed coring. For one thing, Eq. (6.25) indicates that CS ! 1 as g ! 1

Fig. 6.5 Solute distribution C


along the emerging solid
when no mixing in solid and kCE
complete mixing in liquid

Co/k

Co

kCo

0 g(=x/L) 1
224 6 Kinetics of Phase Transformation: Later Stage

(for k < 1), but the solid should end with the eutectic liquid (CS ¼ kCE) and the last
eutectic liquid (CL ¼ CE) solidifies via the eutectic reaction (L ! α + β); see the
phase diagram, Fig. 6.4.
When No Mixing in Solid and No Mixing in Liquid
Is even complete mixing in liquid realistic? Not quite if not stirred thoroughly,
because the diffusion coefficient in liquid metal is limited only to the order of
1–10 cm2/s at best [4]. What would happen to the solute distribution if even liquid is
not mixed at all either as the extreme? W. A. Tiller et al. [5] solved this solute
distribution problem when there is no complete mixing in solid and no complete
mixing in liquid as well. Let us follow their amusing treatment.
In this case, the solute “vomited” by the emerging solid tends to accumulate in
front of the planar growth front (x0 ¼ 00 ) or solid/liquid interface moving at a
speed R, if the vomiting rate is faster than the diffusion rate into the receding liquid
(x0 > 00 ). The situation is as if a wooden board, with some holes through, plowed
upright through a snow field: snow approaching the board partially passes through
the holes, and most accumulates in front of the board. The concentration (or snow)
profiles CL(x0 ) into the liquid in front of the moving interface (x0 ¼ 0) may look as
shown in Fig. 6.6.

Fig. 6.6 Solute C solid liquid


concentration profile CL(x0 )
in the liquid in front of the
growth front (x0 ¼ 00 ) Co+Ca
moving at a speed R relative
to the beginning of the solid
(x ¼ 0). λC denotes the CL(x′)
characteristic length of the
profile CL(x0 ) and Ca the Co
additional concentration 0′ λc x′
over the initial uniform
ȭ5
concentration Co
0 x

Let us first try to calculate the solute distribution profile in the steady state CL(x0 )
which will be developed after passing through the initial transient state. Noting that
the liquid matrix is drifting at a speed of R “relatively toward” the interface (x0 ¼ 0),
the solute flux JL will be written as the superposition of diffusion flux and drift flux in
the moving frame of reference (x0 ) or

∂CL
JL ¼ DL  RCL ð6:26Þ
∂x0

where DL denotes the solute diffusivity in the liquid. Assuming DL and R to be


constant, the profile CL(x0 ,t) should be such that
6.4 Type III: Growth With Phase Change and With Composition Change 225

2
∂CL ∂J ∂ CL ∂C
¼  L0 ¼ DL þ R L0 : ð6:27Þ
∂t ∂x ∂x0 2 ∂x

In the steady state,

∂CL
¼ 0: ð6:28Þ
∂t

Referring to Fig. 6.6, the boundary conditions are

CL ð1, tÞ ¼ Co ; CL ð0, tÞ ¼ Co þ Ca : ð6:29Þ

Equation (6.27) associated with Eq. (6.28) is then solved to yield the steady state
solution
 
0 Rx0
CL ðx Þ ¼ Co þ Ca exp  : ð6:30Þ
DL

Here, you may recognize that DL/R corresponds to a length, called the charac-
teristic length λC (¼DL/R), where Ca reduces to 1/e of itself or 0.37Ca or the
intercept on the x0 -axis of the tangent at x0 ¼ 0; see Fig. 6.6. This characteristic
length gets smaller, or the profile CL(x0 ) steeper, as the growth rate R grows larger
and DL smaller.
What should be the value for Ca in the steady state? To keep the concentration
profile CL(x0 ) stationary with the advancement of the growth front at the velocity R,
the incoming drift flux RCo should keep being consumed completely by the solid.
Imagine the wooden board plowing through the snow field to be totally permeable.
This indicates that the solute concentration of the growing solid must be CS¼Co.
Assuming the local thermodynamic equilibrium at the growth front or interface,
Co + Ca ¼ Co/k or
 
1
Ca ¼ Co 1 : ð6:31Þ
k

Therefore, the steady state solution, Eq. (6.30), may be rewritten as


    
1 Rx0
CL ðx0 Þ ¼ Co 1 þ  1 exp  , ð6:32Þ
k DL

which shifts as a whole at the speed R relative to the beginning of the solid x ¼ 0; see
Fig. 6.6.
The solid has started first with composition kCo at x ¼ 0 and grown gradually
with x to Co in the steady state. The solute distribution CS(x) from the beginning
(x ¼ 0) to the end of the solid (x0 ¼ 0) may, thus, look as shown in Fig. 6.7.
226 6 Kinetics of Phase Transformation: Later Stage

Co/k
CT
CL(x′)
Cα(x′)
Cα(x′)
Co
0′ x′
CS(x)
kCo

0 x1 x2 x

Fig. 6.7 Solute profile CS(x) in the growing solid behind the moving interface (x0 ¼ 00 ), developed
till the steady state. The solute rejected from the solid (corresponding to the shaded area above
CS(x)) accumulates in the liquid (shaded area below CL(x0 )) to make the solute profile CL(x0 ) in the
steady state. Cα(x0 ) represents the initial transient state profile when the moving growth front
(x0 ¼ 00 ) reaches, e.g., x1 and x2. (From Tiller et al. [5])

Could we calculate this CS(x)? You may easily guess that CS(x) will saturate to Co
exponentially with x to take the form

CS ¼ A þ Beαx ð6:33Þ

where A, B, and α are the constants to be determined. This is actually a consequence


of the natural decay in which the decreasing rate of a difference is proportional to
the difference itself or

dðCo  CS Þ
 ¼ αðCo  CS Þ ð6:34Þ
dx

with α being the proportionality constant. By using the two obvious boundary
conditions at x ¼ 0 and as x ! 1,

CS ð0Þ ¼ kCo ; CS ð1Þ ¼ Co , ð6:35Þ

you can determine A and B to obtain the solution

CS ðxÞ ¼ Co  Co ð1  kÞeαx : ð6:36Þ

There still remains the decay constant α to be determined. Referring back to


Fig. 6.7, you may recognize that the solute “vomited” by the entire solid extending
from x ¼ 0 to the growth front (x0 ¼ 0) should remain accumulated in the liquid
ahead or
6.4 Type III: Growth With Phase Change and With Composition Change 227

Z 1 Z 1
ðCo  CS Þdx ¼ ðCL  Co Þdx0 ð6:37Þ
0 0

as indicated by the shaded areas above CS(x) and below CL(x0 ) in the figure. The left-
hand side integral should have been integrated up to the growth front (x0 ¼ 00 )
which is essentially equal to x ! 1 as Co  CS ! 0. Equation (6.37), in association
with Eqs. (6.35) and (6.32), yields

R 1
α¼k ¼k ð6:38Þ
DL λC

implying that the characteristic length in the solid formed is 1/k times larger than
that in the liquid, λC.
The solute distribution in the growing solid CS(x) is now known as Eq. (6.36)
with Eq. (6.38). It follows that the solute concentration in the liquid at the moving
interface (x0 ¼ 00 ), CT (¼Co + Ca) varies from Co to the steady state value Co/k due to
local thermodynamic equilibrium at the solid/liquid interface or
  
CS Co kR
CT ¼ ¼ 1  ð1  kÞ exp  x : ð6:39Þ
k k DL

Letting Cα(x0 ) denote the transient state solute profile in front of the moving
interface (x0 ¼ 00 ) (see Fig. 6.7), you may immediately guess that it will take the
similar form to the steady state one, Eq. (6.30), but probably with a different decay
constant β, or

0
Cα ðx0 Þ ¼ Co þ CT  Co eβx : ð6:40Þ

The decay constant β should then be such that, due to mass conservation again,
Z 1 Z ℓ
0
ðCα  Co Þdx ¼ ðCo  CS Þdx ð6:41Þ
0 0

where ℓ (¼Rt) is the length of the solid formed for a time duration t as indicated by x1
and x2 at two different time durations in Fig. 6.7. Substituting Eq. (6.36) with
Eqs. (6.38), (6.39), and (6.40) into Eq. (6.41) and integrating, one finds that the
decay constant β is just equal to that for the steady state or

R 1
β¼ ¼ : ð6:42Þ
D L λC

Noting that the thickness of the solid ℓ ¼ Rt and x0 ¼ x  ℓ  0, and substituting


Eqs. (6.39) and (6.42) into Eq. (6.40), one finally obtains
228 6 Kinetics of Phase Transformation: Later Stage

    
1k R2 R
Cα ¼ Co þ Co 1  exp k t exp  ðx  RtÞ : ð6:43Þ
k DL DL

As you may visualize from Fig. 6.7, the steady state growth may continue till the
end of the mold x ¼ L, or x0 ¼ L  ℓ ¼ 0 is approached so as to disturb the steady
state profile CL(x0 ) due to the solute reflection by the end wall of the mold (x ¼ L).
The solute reflection may start when the length of the remaining liquid L  ℓ gets
smaller than, say, 4 or 5 times the characteristic length λC. This solute reflection
increases CT over the steady state value Co/k up to the eutectic composition CE
which, in turn, increases CS from its steady state value Co to kCE. The last liquid of
eutectic composition solidifies via eutectic reaction, and solidification will have
finished when temperature leaves the eutectic temperature. The overall solute
profile may, thus, look as sketched in Fig. 6.8. Look! You’ve now obtained a
homogeneous solid over the steady state portion when practically no mixing in
solid and diffusion-controlled mixing in liquid.

Fig. 6.8 Solute distribution C


CS in the solid bar formed
through the initial transient CE
state, steady state, and final
transient state CE/k

CT
Co
CS

kCo

0 g(=x/L) 1

6.4.2 Growth of Spherical Precipitates from Solid Solution [6]

As the second example of the third type of growth, let us consider the growth of
β-precipitates nucleated in α-matrix upon quenching an alloy X, with the solute
concentration Co (outside spinodal), from a temperature To in the single phase field
to T1 in the two-phase region, as illustrated in a hypothetic immiscibility phase
diagram (Fig. 6.9). How they grow was already introduced schematically in
Fig. 5.12a. Here, we are concerned with how fast they grow or their growth kinetics.
We will assume:
(i) Nucleation has just finished so that no further nucleation occurs.
(ii) All nuclei are spherical with the same radius R over the critical size.
(iii) Growth proceeds isothermally.
6.4 Type III: Growth With Phase Change and With Composition Change 229

Fig. 6.9 A hypothetic


immiscibility diagram. An
To
alloy “X” decomposes into
α and β by nucleation and
growth mechanism outside
the spinodal α β

T1
Xβ Xα

Cα Co Cβ

Fig. 6.10 Solute C


distribution around a
β-precipitate of size R. Note Cβ
that the surface
concentration CR is, in α β α
general, different from the
thermodynamic equilibrium
value Cα stipulated by the Co
phase diagram; see Fig. 6.9.
Refer back to Fig. 5.12a CR

R 0 R r

The solute concentration distribution across a β-precipitate with a radius R may


look, in general, as shown in Fig. 6.10. It is easy to imagine that for the β-precipitate
to grow, the solute should first diffuse down its concentration gradient from the
matrix to the precipitate surface and then accommodate themselves into the precip-
itate via surface reaction at the very surface r ¼ R. It is noted that the surface
concentration CR is, in general, different from the thermodynamic equilibrium value
Cα.
Letting Jdiff and Jrxn denote the diffusion rate (flux) and the surface reaction rate,
respectively,

Jdiff ¼ D∇C; Jrxn ¼ kðCR  Cα Þ ð6:44Þ

where D denotes the solute (chemical) diffusion coefficient and k the surface
reaction rate constant assuming the first-order surface (incorporation) reaction.
The growth rate of a specific precipitate of radius R may then be written as

dR
Cβ 4πR2 ¼ 4πR2 Jrxn , ð6:45Þ
dt
230 6 Kinetics of Phase Transformation: Later Stage

due to mass conservation. The instantaneous surface concentration CR is such that,


due to mass conservation again,

dCR
/ Jdiff  Jrxn ð6:46Þ
dt

at r ¼ R, the surface of the precipitate.


We further assume that the precipitates are not densely enough populated and are
highly dispersed to keep the largest possible separation from each other so that the
diffusion field around a spherical precipitate may be regarded as spherically sym-
metric. We may also assume the situation gets in the steady state after the brief
initial transient state. If spherically symmetric, the steady state concentration gradi-
ent at r ¼ R is given, assuming D being constant, as1

Co  CR
∇C ¼ , ð6:47Þ
R

and ∂CR/∂t ¼ 0 or Jdiff + Jrxn ¼ 0 in Eq. (6.46). Equations (6.44) and (6.47), thus,
lead to

DCo þ kRCα
CR ¼ : ð6:48Þ
D þ kR

This equation tells that if D  kR or the overall kinetics is governed by the


surface reaction, CR  Co; if D  kR or governed by diffusion, CR  Cα. You see,
the local thermodynamic equilibrium at the phase boundary can be maintained only
when the surface reaction is fast enough compared to diffusion in this kind of
situation.
OK! Let us suppose that the overall kinetics is diffusion-controlled (D  kR)
and, hence, CR ¼ Cα as stipulated by the equilibrium phase diagram (Fig. 6.10).
Then, Eq. (6.45) may be rewritten as

dR C  Cα
Cβ ¼ Jdiff ¼ D o : ð6:49Þ
dt R

Upon integration from t ¼ 0 to t, one obtains


 
Co  Cα
R  2
R2o ¼ 2D t ð6:50Þ

where Ro is the initial size of the precipitates at t ¼ 0 that is normally R.


Practically, it is difficult to monitor directly the precipitate size R(t) as a function
of time t. As the precipitates grow as in Eq. (6.50) by consuming the solute from the

1
If not clear, please refer to the steady state solution, Eq. (1.34), in Chap. 1.
6.5 Overall Rate of Transformation 231

matrix, however, the mean concentration of the solute in the α-matrix CðtÞ decreases
with time, and this may be measured against time t, with relative ease, by, e.g.,
monitoring a property proportional to CðtÞ, e.g., hardness or electrical conductivity.
Suppose that we have had “n” precipitates to grow per unit volume. Then,
 
4 3
CðtÞ ¼ Co  n πR Cβ : ð6:51Þ
3

By substituting Eq. (6.50) with Ro neglected as Ro  R, Eq. (6.51) finally takes


the form

CðtÞ ¼ Co  bt3=2 ð6:52Þ

where the system constant


pffiffiffi
8 2 1=2
b¼ πnD3=2 ðCo  Cα Þ3=2 Cβ : ð6:52aÞ
3

That is, if one monitors a C-proportional property against time, then the property
would vary as t3/2 if diffusion-controlled.

6.5 Overall Rate of Transformation

When α-phase transforms isothermally to β-phase below the equilibrium transition


temperature Ttr, the volume fraction of β-phase X(t) will increase with time. We are
now concerned with how the overall volume fraction transformed X(t) increases
with time.
Let us assume:
(i) Only phase changes with no composition change as in freezing of water;
(ii) Phase transforms via nucleation and growth;
(iii) Nucleation and growth rate are each constant;
(iv) Nucleation occurs at random;
(v) Spherical products grow radially until impinging against each other.
If nucleation finished altogether at a moment, say, t ¼ 0 and then growth started
altogether for the same time duration t, the transformation kinetics X(t) would be
trivial: If we have n spherical nuclei/unit volume with negligible size at t ¼ 0, the
volume fraction transformed Xe(t) will take the form

4
Xe ¼ πðutÞ3  n ð6:53Þ
3

as the radius of each nucleus r grows as r ¼ ut with the constant growth rate u. It is
really trivial, isn’t it? Furthermore, see that Xe grows even larger than the obvious
limit Xe ¼ 1!
232 6 Kinetics of Phase Transformation: Later Stage

There is nothing like this situation in reality: the two processes, nucleation and
growth, do not occur consecutively, but overlap, i.e., they occur simultaneously.
Thus, the nuclei formed earlier will grow for longer time duration and vice versa.
What is then the overall volume fraction transformed X(t) in time t? We have now to
take into account the effect of this, so to say, chronological nucleation.
If the nucleation rate is “I,” the infinitesimal number of nuclei “dn” nucleated in
the time interval dτ at time τ is

dn ¼ Idτ: ð6:54Þ

These can grow only for a time span t  τ until time t. Therefore, the infinitesimal
volume fraction transformed by these nuclei, dXe(τ), will be

4 4
dXe ðτÞ ¼ π½uðt  τÞ3  dn ¼ π½uðt  τÞ3  Idτ: ð6:55Þ
3 3

By adding up all these differential contributions by the nuclei formed from τ ¼ 0


to t, you obtain the total volume fraction transformed at time t, Xe(t), as
Z t
1
Xe ðtÞ ¼ dXe ðτÞdτ ¼ πu3 It4 : ð6:56Þ
0 3

It turns out that Xe(t) increases infinitely as t4 as illustrated in Fig. 6.11. But, we
know that the volume fraction transformed cannot be larger than 1 by definition!
What is wrong with our formulation then?
You see, Eq. (6.54) indicates that nucleation occurs everywhere in the total
volume at any time while the earlier-formed nuclei have been growing within the
same volume. That is, nucleation occurs not only in the volume not transformed yet
but also even within the volume already transformed; thus Xe grows indefinitely
larger than 1. Such Xe is, thus, referred to as the extended volume fraction.
Nucleation should be allowed only in the volume not transformed yet. If X denotes
the real volume fraction transformed, then

dn ¼ Idτð1  XÞ: ð6:57Þ

Thus,

4 4
dXðτÞ ¼ π½uðt  τÞ3  dn ¼ π½uðt  τÞ3  Ið1  XÞdτ, ð6:58Þ
3 3

or

dX 4
¼ π½uðt  τÞ3  Idτ: ð6:58aÞ
1X 3

Upon integration, we obtain


6.6 Last Stage of Phase Transformation: Precipitate Coarsening 233

Fig. 6.11 Temporal


X(t) Xe(t)
variation of the overall
volume fraction transformed
X(t) in comparison with the 1
extended volume fraction
transformed Xe(t)

0
0 t

1 3 4
 πu It
XðtÞ ¼ 1  e 3 ¼ 1  eXe ðtÞ ð6:59Þ

which is as illustrated, in comparison with the extended volume fraction Xe(t), in


Fig. 6.11. Indeed, X(t) ! 1 as t ! 1, as it should be. You can see that Xe(t)  X(t) in
much earlier stage of transformation or as t ! 0 or X ! 0 but starts to deviate soon.
It is because the β-particles start to impinge against themselves in Xe(t). This
equation is known as the Johnson-Mehl-Avrami equation [7] after the inventors’
names. For more discussions, you may refer to Burke and Turnbull [1].

6.6 Last Stage of Phase Transformation: Precipitate


Coarsening

Referring back to Fig. 6.9, you may imagine that at the completion of phase
separation via, e.g., nucleation and growth, the β-particles of variable sizes are
dispersed in the α-matrix to the equilibrium fraction Xβ. That’s it then? Should there
be no further change to the system as the thermodynamic driving force for α ! α + β
has now been used up?
It is observed that the larger particles or precipitates grow larger at the expense of
the smaller ones while keeping the equilibrium fraction Xβ. As a consequence, very
small particles tend to be absent in the neighborhood of a large particle. Eventually
there will remain, in principle, a single large particle eliminating all the other smaller
ones. This phenomenon of precipitate coarsening was first described by
W. Ostwald [8] in 1896, thus also called the Ostwald ripening. It is general to
any inhomogeneous phase mixture, e.g., salt grains of variable sizes dispersed in
salt-saturated water.
How does it happen? For a precipitate to grow larger at the expense of the
smaller ones, there must be a mass transfer from the smaller to the larger. That is,
the larger, the smaller the escaping tendency, and vice versa: The larger, the
smaller the solubility in the matrix; the smaller, the larger the solubility. There,
234 6 Kinetics of Phase Transformation: Later Stage

thus, builds up a concentration gradient from the smaller to the larger, driving the
solute to diffuse toward the larger ones. Here, we will learn how the phenomenon
“the richer the richer, the poorer the poorer” can be quantified to describe quanti-
tatively the coarsening kinetics again on the basis of an old wisdom due to
Greenwood [9]. Let us first examine the solubility depending on the
precipitate size.

6.6.1 Thomson-Freundlich Equation

Consider a spherical solute particle of radius r. Equation (6.5) indicates that the
chemical potential of the solute inside the curved surface, μ(r), is enhanced com-
pared to that inside a planar surface (r ¼ 1), μ(1) as

2γVm
μðrÞ ¼ μð1Þ þ : ð6:60Þ
r

Here, we assume that the surface tension, γ, and the molar volume of the solute
particle, Vm, are each independent of r.
Letting So denote the thermodynamic true solubility of the solute under a planar
surface (r ¼ 1) and Sr the solubility in equilibrium with the solute particle of size r,
the diffusive equilibrium criterion stipulates that the chemical potential of the solute
in the precipitate should be equal to that in the solvent matrix or

μðrÞ ¼ μoα þ RT ln Sr ð6:61aÞ


μ ð 1Þ ¼ μoα þ RT ln So ð6:61bÞ

assuming the ideal solution. Here μoα denotes the standard chemical potential of the
solute in the solvent or matrix α. By subtracting Eq. (6.61b) from Eq. (6.61a) and
substituting Eq. (6.60), we have

Sr 2γVm
ln ¼ : ð6:62Þ
So RTr

The solute in the convex side has a higher solubility due to the higher escaping
tendency compared with the solute under a planar surface. This equation is called the
Thomson-Freundlich equation, named after its inventors [10].
We already know from Eq. (6.10) that normally

2γVm
 1, ð6:63aÞ
RTr

which means that the solubility increase, Sr  So should not be that large compared
to the thermodynamic true solubility So for r ¼ 1 or
6.6 Last Stage of Phase Transformation: Precipitate Coarsening 235

S r  So
 1: ð6:63bÞ
So

As Sr/So ¼ 1 + (Sr  So)/So and ln(1 + x)  x for |x|  1, Eq. (6.62) may be
approximated as

2γVm So
Sr  So ¼ ð6:64Þ
RTr

whereby the solubility difference between the precipitates i of radius ri and j of radius
rj amounts to
 
2γVm So 1 1
Sri  Srj ¼  : ð6:64aÞ
RT ri rj

You see Sri < Srj if ri > rj, and hence, the solute tends to diffuse from the smaller
particle j to the larger particle i.

6.6.2 Coarsening Kinetics: Exact Formulation

Now, consider a system of N spherical β-precipitates of various radii rj (j ¼ 1,2,. . .


N), dispersed in the α-matrix or solvent. The solute gain rate of an arbitrary
precipitate i of radius ri from another j, dQj/dt, due to their solubility difference,
Sri  Srj , may be written, in accord with Fick’s first law, as
 
dQj Aj

¼ D Sr i  Sr j ð6:65Þ
dt xj

or, due to Eq. (6.64a),


   
dQj Aj 2γVm So 1 1
¼D  , ð6:65aÞ
dt xj RT rj ri

by introducing the effective diffusion cross section, Aj, and the effective diffusion
distance, xj, between the i and the j. As the rate of solute gain from all the other
particles makes the i grow, its growth rate dri/dt may be written due to the mass
conservation as

dri X dQj
4πr2i Cβ ¼ ð6:66Þ
dt dt
j6¼i

or, due to Eq. (6.65a),


236 6 Kinetics of Phase Transformation: Later Stage

  
dri 2DVm γSo X Aj 1 1
4πr2i Cβ ¼  ð6:66aÞ
dt RT xj rj ri
j6¼i

where Cβ stands for the solute concentration of the β-particles. Here, of course, we
have assumed that the growth kinetics is totally diffusion-controlled.
In principle, by solving this differential equation, Eq. (6.66a), you can get the
growth rate equation of each precipitate, thence the overall growth rate. But, alas!
There is no way to evaluate the diffusional geometric factor Aj/xj for the jth
precipitate where j runs from 1 to N except for i. The growth rate should be
dependent on the local configuration of the precipitates.

6.6.3 Coarsening Kinetics: Highly Dispersed Case

If the volume fraction of the precipitates is small compared to the matrix volume and
they are highly dispersed or distributed keeping the maximum possible separation
distance from each other, however, the solubility distribution around each precipitate
may then be regarded as spherically symmetric as illustrated in Fig. 6.12. The
solubility right at the surface of a particle of radius ri takes the value Sri , as given
in Eq. (6.64), and gradually varies toward a certain uniform value S1 far away from
the surface r ¼ ri or essentially as radial distance r ! 1. The latter, S1, is obviously
higher than the true solubility So because each and every spherical precipitate
dissolves more than So.
The growth rate of a specific particle i is then written, due to the mass conser-
vation, as
 
dri dr
Cβ 4πr2i ¼ 4πr2i Ji jr¼ri or Cβ i ¼ Ji jr¼ri ð6:67Þ
dt dt

where Ji jr¼ri is the influx of solute right at the surface (r ¼ ri) of precipitate i or

∂S
Ji jr¼ri ¼ D : ð6:68Þ
∂r r¼ri

Assuming the steady state in the spherical symmetry, the solubility as a function
of r, S(r) is2

B
Sð r Þ ¼ A þ ð6:69Þ
r

where A and B are to be determined by the two boundary conditions

2
If not obvious to you, see the steady state solution in spherical symmetry, Eq. (1.34), in Chap. 1.
6.6 Last Stage of Phase Transformation: Precipitate Coarsening 237

Fig. 6.12 Spherically S


symmetric solute
distribution around a Sri
β-precipitate i of radius ri.
The solubility takes a value
Sri , enhanced over the
thermodynamic true S∞
solubility So, right at the So
precipitate surface r ¼ ri and
S1 far away from the β
surface or as ri ! 1. (From
0 ri r
Greenwood [9])

S ¼ Sri at r ¼ ri ; S ¼ S1 as ri ! 1; ð6:70Þ

see Fig. 6.12. It follows that

ri ðSri  S1 Þ
SðrÞ ¼ S1 þ ð6:71Þ
r

and, due to Eq. (6.68),

S r i  S1
Ji jr¼ri ¼ D : ð6:72Þ
ri

The growth rate equation, Eq. (6.67), thus, takes the form

dri S  S1
Cβ ¼ D ri , ð6:73Þ
dt ri

where the solubility afar in the matrix, S1, is yet to be determined. As the mass or
total volume of the β-phase is fixed,

X
N
dri X N
S  S1
Cβ 4πr2i ¼0¼ 4πr2i D ri , ð6:74Þ
i¼1
dt i¼1
ri

and hence,

X
N X
N
ri Sri  S1 ri ¼ 0: ð6:75Þ
i¼1 i¼1

Defining the mean particle size, rm, as

1 X
N
rm r, ð6:76Þ
N i¼1 i
238 6 Kinetics of Phase Transformation: Later Stage

Fig. 6.13 Growth rate of an


dri
arbitrary particle i vs. its size
ri. Note that the precipitates dt
smaller than the mean size
rm grow smaller (dri/dt < 0)
and those larger than rm
grow larger (dri/dt > 0)
with the maximum rate at
ri ¼ 2rm. (From
Greenwood [9])
0 rm 2rm ri

Eq. (6.75) in association with Eq. (6.64) gives

1 X
N
2γVm So
S1 ¼ r S ¼ So þ : ð6:77Þ
Nrm i¼1 i ri RTrm

It turns out that the solute concentration afar into the α-matrix, S1, corresponds to
the equilibrium solubility of the mean size (rm) precipitate as expected intuitively.
By substituting Eqs. (6.64) for Sr and (6.77) for S1 into Eq. (6.73), the growth
rate of an arbitrary particle i is rewritten as
 
dri 2D γVm So 1 1 1
¼   : ð6:78Þ
dt Cβ RT ri rm ri

The growth rate varies against precipitate size ri as illustrated in Fig. 6.13. You
see, particles larger than the mean size (ri > rm) grow larger (dri/dt > 0), and vice
versa. The maximum growth rate falls at ri ¼ 2rm, that is, the precipitates two times
larger than the mean size grow the fastest.
How fast do these fastest-growing precipitates grow? Substituting ri ¼ 2rm in
Eq. (6.78) and integrating from t ¼ 0 when rm ¼ rm(0) to t when rm ¼ rm(t), you end
up with a cubic rate law:

3D γVm So
rm ðtÞ3  rm ð0Þ3 ¼ KðTÞt; KðTÞ ¼ ð6:79Þ
4Cβ RT

If the initial mean size rm(0) is negligibly small, then rm/t1/3. Note that K(T) is a
function of temperature mostly via D.
Despite a number of ideality assumptions employed, this cubic rate law may be
regarded as a reasonable estimation of the reality within the order of magnitude. For
this to be valid, the particle size distribution of a system should be wide enough to
include r ¼ 2rm.
Problems 239

Problems

1. If pores of 5 micron diameter are sealed off containing nitrogen at a pressure of


0.8 atm in a glass having a surface tension of 0.30 J/m2 and a relative density of
0.85 (with respect to water), what will be the pore size at which the gas pressure
just balances the negative pressure due to surface tension? What will be the
relative density at this point?
2. Isothermal grain growth may occur in pure metals. The change in size of small
grains adjacent to much larger grains is described by

r2  r2o ¼ 2Kt ðaÞ

where r is the grain radius at time t and r0 is the initial grain radius; see figure
below. When an atom jumps from grain A to grain B, the free energy change of
the system is ΔG ¼ 2γVm/NAr, where γ is interfacial free energy, Vm is molar
volume, and NA is the Avogadro constant.
By deriving Eq. (a), give an expression for the constant, K.

A
r

3. (a) Metal, M, undergoes an allotropic phase transformation between α and β at


Teq ¼ 1000 K. A sample of M containing 107 cm3 small oxide inclusions
which act as effective nucleation sites even for small undercooling is rapidly
quenched from 1100 to 800 K. Estimate the grain size at the end of the
transformation from α to β. Assume that the growth may be described by
the normal growth mechanism.
(b) Alternatively suppose that the sample containing the same number of oxide
inclusions had been very slowly cooled from 1100 to 800 K. What would be
the grain size after completion of the transformation?
Data
Enthalpy of transformation, ΔHtr ¼ 0.5RTeq
Molar volume of α, Vm,α ¼ 10 cm3 mol1
240 6 Kinetics of Phase Transformation: Later Stage

Lattice parameter of α, ao ¼ 3
108cm
Surface energy along α  β, γαβ ¼ 0.05 Jm2
Lattice diffusivities at 800 K, D ¼ 1010 cm2 s1
Diffusivity across α-/β-interface at 800 K, Dgb ¼ 109 cm2 s1

4. A solid-solid transformation produces cubic precipitates. The surface energies are


γcoh ¼ 0.04 J/m2 for coherent precipitates and γincoh ¼ 0.40 J/m2 for incoherent
precipitates. The volume free energy change with transformation and the equi-
librium transition temperature are, respectively,

2
108 ΔT
ΔGv =J  m3 ¼ ; Teq ¼ 1000 K
Teq

and strain energy/unit volume coherent precipitates ΔGε ¼ 5


106J  m3.
ΔGε ¼ 0 for incoherent precipitates.
(a) Which type of precipitate forms at (i) small ΔT and (ii) large ΔT,
respectively?
(b) At what temperature will incoherent and coherent precipitates both be able to
form?
(c) Under conditions of a constant nucleation rate and normal growth, express the
time at which the total transformation rate is a maximum for a given degree of
undercooling.
(d) In a real transformation, the assumptions of a constant nucleation rate and
normal growth are not strictly valid. In each of the following, assess whether
or not the equations in part (c) remain good approximations, and suggest
modifications that will keep the treatment valid.
(i) Nucleation rate is a function of time. Normal growth at a constant rate.
(ii) Heterogeneous nucleation on insoluble impurities. Normal growth at a
constant rate.
(iii) Growth is anisotropic, although for a given direction in the matrix the
growth rate is constant.
(iv) Enthalpy changes are so large that the transformation does not proceed
at constant temperature.
5. Various solid oxide inclusions nucleate and grow in a steel melt which is being
held at constant temperature in a ladle. The growth kinetics and the shape of the
inclusion for three different metal oxide species are illustrated in the sketches
below. For each example suggest a growth mechanism which would be consistent
with the observations. Justify. B and C are spherical.
Problems 241

r r
l

l r r

t t t1/2
Species A Species B Species C

6. Consider the phase diagram (Fig. 6.14) and the time-temperature cycle (Fig. 6.15)
for a sample which was initially of homogeneous composition C0 at temperature
T0. For the annealing schedule shown (Fig. 6.15), indicate the assumptions you
would make for describing the volume fraction transformed after 105 s; that is,
what kind of model (equation) would you use and why?

Fig. 6.14 Phase diagram


T liq

a liq+a liq+B
T0
T1
a+B
T2

C 1 C0
AC 2
XB B

Fig. 6.15 Heat treatment


schedule T

T0
105 sec
T1

T2

0 103 time
242 6 Kinetics of Phase Transformation: Later Stage

7. Livingston [Trans. AIME, 215 (1959) 566] measured the rate of coarsening of
cobalt precipitates in copper using a very sensitive magnetic technique. At 600 C
the kinetics obeyed the Greenwood-Wagner model r3 ¼ r3o þ t=A . The initial
mean radius ro was very small such that r3  r3o  r3. After 106 s, r ¼ 10 nm. The
solubility limit of cobalt in copper at 600 C is 0.378 atom per cent.
(a) Calculate the interfacial energy based on the Greenwood-Wagner model and
the observed kinetics.
(b) Compare your results with the measured value of 0.2 J/m2, and comment on
the applicability of our assumed model.

Copper Cobalt
1
Data: Atomic mass (gmol ) 63.55 58.93
Density (gcm3) 8.93 8.90
Diffusivity in copper (cm2 s1) 2
1012 1012

8. In a pure gold quenched from 700 C, it is thought that the supersaturation of


vacancies is relieved by adsorption of vacancies at dislocation lines.
(a) Considering the dislocation lines to be fixed cylindrical sinks of constant
radius ro, derive an equation given the time dependence of the ratio of the
average vacancy concentration cðtÞ to the initial concentration co (for
0.8 < c=co < 1), which could be used to check this hypothesis.
(b) Derive an equation for the case in which planar grain boundaries act as sinks
for the vacancies.
9. A Ge-Ga ingot containing 10 ppm Ga is solidified at R ¼ 8
103 cm/s with
negligible convection. Show schematically the composition along the length of
the fully solidified ingot, giving the initial composition and lengths of the initial
and final transients. Assume DL ¼ 5
105 cm/s and k ¼ 0.1.
10. When large grains of TiO2 are surrounded by small-sized powders of SrCO3, the
reaction product SrTiO3 forms inward from the surface of the TiO2 grains, with
CO2 gas released at the surface of the product SrTiO3. Letting the volume
fraction of SrTiO3 formed be x assuming TiO2 particles are spherical with a
mean radius ro, the reaction kinetics is found to be

h i2 2k t
p
1  ð1  xÞ1=3 ¼ 2
ro

when diffusion-controlled, where kp is the rate constant. This equation is called


Jander’s equation which has a surprisingly wide applicability.
(a) Derive Jander’s equation. Clearly state any assumptions employed.
(b) For the present reaction system, the activation energy of kp has been found to
be 300 kJ/mol. What is this activation for?
Problems 243

11. A 1 nm diameter particle of gold nucleates in a glass matrix containing 1%


gold at 1000 C. The precipitate particle is essentially pure gold, and the
equilibrium concentration of gold in the glass at 1000 C is 0.1%. Assuming
that the growth of the particle is controlled by diffusion, and the diffusion
coefficient of gold in glass at 1000 C is 1010 cm2/s, use the steady state
approximation for diffusion to a spherical particle to calculate how big the
particle will be after 1 h.
12. Precipitation from solution and sol-gel techniques for preparing ceramic pow-
ders yield spherical particles when appropriate experimental measures are taken.
(a) Assuming that the following expression applies to the solubility of ceramic
powder, determine whether small particles or large particles are more resis-
tant to re-solubilization. Justify your answer semi-quantitatively by deter-
mining the sensitivity of Sr to r.

 
2γVm
Sr ¼ S1 exp
RTr

where
Sr ¼ solubility of particle of radius r (nm)
S1 ¼ solubility of flat particle (r ! 1)
γ ¼ interfacial surface free energy (erg cm2)
Vm ¼ molar volume of solid (cm3)
R ¼ gas constant
T ¼ temperature (K)
r ¼ radius of curvature (cm)
(b) Calculate the value of γ for amorphous silica (SiO2) in water at 25 C if S1
is 25 ppm and Sr is 200 ppm for particles of diameter 2 nm. Molecular
weight and density of amorphous SiO2 are 60 g and 2.2 g/cm3,
respectively.
13. (a) The transformation kinetics in the baking of pizza can be modeled by an
Avrami-type equation. In an SNU laboratory, an ingenious technique has
been developed to measure the transformation rate of the dough. At very
short times the data indicate that dx/dt ¼ 4.2
103 t1.5 min1, where x
represents the fraction transformed. How long will it take to bake the dough
to 95% of completion?
(b) Suppose the same dough in part (a) were used in the baking of cupcakes.
Estimate the value of n, the exponent of the time variable in the Avrami
equation. Justify.
14. The Al–Cu phase diagram is similar to that shown in Fig. 6.16 with
Tm(Al) ¼ 660 C, TE ¼ 548 C, Xmax ¼ 5.65 wt%, and XE ¼ 33 wt%Cu.
The diffusion coefficient for the liquid DL ¼ 3
109 m2 s1. If an Al-0.5 wt%
Cu alloy is solidified with no convection and a planar solid/liquid interface at
5 μm/s:
244 6 Kinetics of Phase Transformation: Later Stage

Fig. 6.16 A hypothetical


phase diagram. k ¼ XS/XL is
constant

(a) What is the interface temperature in the steady state?


(b) What is the thickness of the diffusion layer?
(c) What temperature gradient will be required to maintain a planar interface?

15. A Ge–Ga ingot containing 10 ppm Ga is solidified at R ¼ 8


103 cm/s with
negligible convection. Show schematically the composition along the length of
the fully solidified ingot, giving the initial composition and lengths of the initial
and final transients. Assume DL ¼ 5
105 cm/s and k ¼ 0.1.
16. The rate of solidification of the above ingot is increased abruptly by a factor or
5 and held constant at this new velocity. Sketch the composition variation along
the resulting band, giving the maximum composition and approximate thickness
of the band.
17. Pure metal M is cold-worked at room temperature. For strain relief the material
is annealed at a temperature of 800 K. Under these conditions, new strain-free
grains nucleate and grow. This process is called recrystallization. The driving
force for this process is the reduction in strain energy left in the sample by
cold-working. Nucleation can be heterogeneous, and there are 1019 m3 ran-
domly distributed sites which are suitable heterogeneous nuclei. You may
assume that normal growth occurs.
(i) Calculate the time required for 99% completion of recrystallization. State
your assumptions and justify them where the data permit.
(ii) At 99% recrystallization, what will be the average grain size?
References 245

Data:
Strain energy of cold work, ΔGε ¼ 2.0
106 J/m3
Interfacial energy between transformed and untransformed grains,
γ ¼ 0.20 J/m2
Atomic weight of M, AW ¼ 64 g/mol
Density of M, ρ ¼ 9.0 g/cm3
Lattice constant of M, a ¼ 3.6
1010 m
Diffusion coefficient of M across the grain boundary between the
transformed and untransformed grains, Dgb ¼ 1010 m2/s
18. In a precipitation problem, preferred nucleation occurs on dislocations. The
result is that after a short time, the precipitates appear as infinitely long rods;
and further growth occurs only in the radial direction. Develop an expression for
the rate of precipitate growth or the rate of depletion of solute from the matrix.
Express your answer in terms of the initial concentration, ρo; the equilibrium
concentrations in the matrix, ρe; the precipitate, ρp; and the diffusivity, D.
Indicate the range of times over which your results may be useful.

References

1. J.E. Burke, D. Turnbull, Recrystallization and grain growth. Prog. Met. Phys. 3, 220–292
(1952)
2. H.V. Atkinson, Theories of normal grain growth in pure single phase systems. Acta Metall. 36,
469–491 (1988)
3. E. Scheil, Z. Metallik. 34, 70 (1942)
4. J. Frenkel, Kinetic Theory of Liquids (Oxford University Press, Oxford, 1946), p. 201; W. Jost,
Diffusion (Academic Press, New York, 1952), p. 479
5. W.A. Tiller, K.A. Jackson, J.W. Rutter, B. Chalmers, Acta Metall. 1, 428 (1953)
6. C. Zener, Theory of growth of spherical precipitates from solid solution. J. Appl. Phys. 20,
950 (1949)
7. W.A. Johnson, R.F. Mehl, Reaction kinetics in processes of nucleation and growth. Trans.
AIME 135, 416 (1939).; M. Avrami, Kinetics of phase change. J. Chem. Phys. 7, 1103 (1939);
8, 212 (1940); 9, 177 (1941)
8. W. Ostwald, Lehrbuch der Allgemeinen Chemie, vol. 2, part 1 (W. Engelmann, Leipzig, 1896);
W. Ostwald, Z. Phys. Chem. 22, 289–330 (1987)
9. G.W. Greenwood, The growth of dispersed precipitates in solutions. Acta Metall. 4, 243–248
(1956)
10. W. Thomson, On the equilibrium of vapor at a curved surface of liquid. Philos. Mag. Ser. 4 42
(282), 448–452 (1871)
Chapter 7
Diffusion in Ionic Solids

7.1 Introduction: System and Mobile Species

So far, we have dealt with systems in which the atomic constituents may be regarded
as electrically neutral as in metals. If you apply a voltage across a piece of metal,
electrical conduction arises due to electrons, normally with no atomic migration,
and hence, no physicochemical change across the system and at any electrode as
well. If the current level increases, the constituent atoms may migrate, which is,
however, only a secondary effect called electromigration or electrotransport
(cf. Eq. (1.10) in Chap. 1).
What about ionic solids? Let us consider an ionic compound, MaXb, where “a”
and “b” denote the stoichiometric coefficients. If you apply a voltage across such a
compound, there may arise, of course depending on the system, not only a charge
transfer but also a mass transfer causing physicochemical changes1 not only across
the compound matrix, if polycrystalline, but also in the vicinity of each electrode
depending on its nature. This clearly demonstrates that the mobile atomic constitu-
ents are electrically charged: Metal M normally tends to be transferred to the
cathode and is thus called a cation; metalloid X is transferred to the anode and is
thus an anion [2]. Letting z1(>0) and z2(<0) denote the formal valence of the
cations Mz1 and the anions Xz2 , respectively,

z1 a þ z2 b ¼ 0 ð7:1Þ

for the system should remain electrically neutral.


This stoichiometric composition MaXb, however, is only a special case. Ther-
modynamically, the compound should exist in a range of composition from

1
Typical changes are microstructural changes including grain growth and densification if polycrys-
talline and crystal shift or phase decomposition depending on the nature of the electrodes; see Yoo
and Lee [1].

© Springer Nature Switzerland AG 2020 247


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_7
248 7 Diffusion in Ionic Solids

M-saturated to X-saturated one. The width of this existence, stability, or homo-


geneity range, of course, depends on the system itself. Generally, the more widely
variable the valence of metal M, z1 is, the wider the existence range like transition
metal oxides, e.g., FeO and MnO. It is, thus, more appropriate to represent the
compound as MaXb+δ (or equivalently, Ma2δXb) by taking into account the varying
composition across its stability range. The deviation δ from the stoichiometric
composition (δ ¼ 0) is called the nonstoichiometry. The homogeneity range of
MaXb+δ, then, normally comprises the three regimes: (i) δ < 0 or hypostoichiometry
regime; (ii) δ  0 or near-stoichiometry regime; (iii) δ > 0 or hyperstoichiometry
regime. If δ < 0, the composition is metal-excess or metalloid-deficit: Excess
cations (or missing anions) with a corresponding excess of electrons (e) are
present. If δ > 0, the composition is metal-deficit or metalloid-excess: Cations
and corresponding electrons are missing, or excess anions and corresponding elec-
tron deficits or holes are present.
The compound MaXb+δ (or MaδXb) is a binary comprising two (neutral)
chemical components M and X. By taking these two neutral components as the
mobile species, one may describe the mass transport phenomena, but the accompa-
nying electrical phenomena cannot be described. It is, therefore, more appropriate to
treat the system as comprising three mobile charged components, cations Mz1
(k¼1), anions Xz2 (k¼2), and electrons e (k¼3) with formal valence z3 ¼ 1.
Of course, the charge neutrality should be observed in any case:

X
3
zk ck ¼ 0 ð7:2Þ
k¼1

where ck denotes the concentration of the charged component k.


The ionic solids thereby serve as the major playground of solid state ionics and
solid state electrochemistry. We will learn here how diffusion phenomena in ionic
solids comprising charged components can be described, again on the basis of the
old wisdom.2 We will discuss with a generic binary MaXb+δ.

7.2 Mobile Species and Diffusion Mechanisms

The flow of a neutral chemical component M in MaXb+δ is equivalent to a


combined or coupled flow of charged components, an Mz1 ion and z1 electrons
(e) in the same direction or

M Ð M z1 þ z1 e : ð7:3Þ

2
C. Wagner [3]. (For the English translation of the main part of this pioneering work, see Appendix
II at the end of this book.)
7.2 Mobile Species and Diffusion Mechanisms 249

(a)
Chemical A O
components

Charged ⎧ A2+
⎧ O2-
components ⎨ 2e- ⎨ 2e-
coupled
⎩ ⎩

(b)
Charged A2+ O2- e-
components

Vehicle
defects

Fig. 7.1 (a) Flow of a chemical component of AO1+δ corresponds to the coupled or ambipolar flow
of charged components, its ion and electrons, to keep the charge neutrality. (b) Each charged
component is rendered mobile through its corresponding defects, typically vacancies and intersti-
tials. Arrows denote the directions of flow

Similarly, the flow of a neutral component X is equivalent to a coupled flow of


charged components, an Xz2 ion (z2 < 0) and |z2|e in counter directions or

X Ð X z2 þ z2 e : ð7:4Þ

These are actually nothing but the local ionization equilibria.


The charged components fMz1 , Xz2 , e g are each rendered mobile by their
corresponding point defects, usually vacancies and interstitials (as in metals), but
here they are effectively charged.3 See the summary in Fig. 7.1 for, e.g., a simple
oxide AO1+δ with two chemical components {A,O} or three charged components
{A2+, O2, e}. The cations A2+ via their own vacancies V00A and interstitials Ai ; the
anions O2 via VO and O00i ; the electrons e via holes h, the vacancy-analogues in
the filled valence band and free electrons e0 , the interstitial analogues in the empty
conduction band. Consequently, the self-diffusivity, Dk, of the k-type charged

3
Here, ionic defects are denoted in the Kröger-Vink system of notation, SCL , meaning the species
“S” sitting on the site or locus “L” with an effective charge “C.” The effective charge of a point
defect is defined as the actual charge of the species S less the actual charge that would be at the site L
in perfect ionic crystal and represented by the corresponding number of dots (•) if positive, by
primes (0 ) if negative, and by a cross (x) if neutral. For more detailed treatment, see Appendix I:
“Defect Chemistry of Solid State Ionic Compounds.”
250 7 Diffusion in Ionic Solids

component (k¼1,2,3) is related to the defect diffusivity, Ddk , of the k-corresponding


defects dk via the conservation of mass, sites, and electronic states, respectively, as
X
ck D k ¼ cd k D d k ð7:5Þ
dk

where cs denotes the concentration of species s (¼k, dk) in appropriate units. This is
the very link between the component self-diffusivities and defect structure of a
system compound.
Therefore, any mass/charge transport phenomenon in an ionic solid may be
described in terms of the mobile charged components, cations, anions, and
electrons, or equivalently in terms of mobile defects. The former may be called
the charged-component-level description and the latter the defect-level descrip-
tion. Here, we will mostly stay with the former because this is independent of defect
model.

7.3 Flux Equation and a Few Definitions

We have learned earlier in Chap. 4 that the flux Jk of a mobile species k, whether
component or defect, may be written in the linear regime as

Jk ¼ ck vk ¼ ck Bk Xk ð7:6Þ

where vk is the drift velocity, Bk the mechanical mobility, and Xk the thermody-
namic force acting on particle k. Invoking the Einstein relation from Chap. 4,

Dk
Bk ¼ ð7:7Þ
RT

where Dk is the self-diffusivity as a measure of the total successful jump fre-


quency, Γk of k with the given hopping distance αk in uniform composition

1
Dk ¼ Γk α2k : ð7:8Þ
6

Under the isothermal and isobaric condition, a charged particle k, whether


effectively or actually charged, may be driven not only by its chemical force,
∇μk induced by its own concentration gradient ∇ck as in a neutral particles system,
but also by the electrical force, zkF∇φ because it is charged, where φ stands for the
electrostatic potential or Galvani potential and F the Faraday constant. The sum
7.3 Flux Equation and a Few Definitions 251

of the chemical potential μk and the electrostatic energy zkFφ is called the
electrochemical potential, ηk (¼μk + zkFφ), which is, by definition, the partial
molar Gibbs free energy of k-type particles ((∂G/∂nk)T,P,j(6¼k)). The thermody-
namic force for the k-type particles is, thus, written as

Xk ¼ ∇ηk ¼ ð∇μk þ zk F∇φÞ ð7:9Þ

under isothermal and isobaric conditions. The particle flux equation, Eq. (7.6), thus
takes the form

Jk ¼ ck Bk ∇ηk ¼ ck Bk ð∇μk þ zk F∇φÞ: ð7:10Þ

The mechanical mobility Bk was earlier defined as the drift velocity per unit
force (∇ηk), thus having the units (m/s)/(Newton/mol). When only an electrical
force is applied to a homogeneous composition (∇μk ¼ 0), the drift velocity per unit
electric field strength (∇φ), instead of the electrical force (zkF∇φ), may be
written as

vk
¼ zk FBk  uk : ð7:11Þ
∇φ

Here, uk as such is called the electrochemical mobility and has the units of (m/s)/
(V/m) or m2/Vs.
Whenever charged particles or charge carriers flow, current arises. The current
density, or amount of charge passing through a unit area per unit time, due to the
k-type carriers, ik, is given as

ik ¼ zk FJk ð7:12Þ

when the carrier flux Jk is given as the molar flux.


The partial electrical conductivity of the carrier k, σk, is defined by Ohm’s law
as the current density per unit field strength in homogeneous composition or
∇μk ¼ 0:
 
ik
σk  ð7:13Þ
∇φ ∇μk ¼0

Due to Eqs. (7.10) and (7.12), thus, one may obtain the partial conductivity in
terms of mechanical mobility, Bk, or electrochemical mobility, uk, as

σk ¼ ck ðzk FÞ2 Bk ¼ ck ðzk FÞuk ð7:14Þ

and, further due to Eq. (7.7), in terms of the self-diffusivity as


252 7 Diffusion in Ionic Solids

Dk
σk ¼ ck ðzk FÞ2 : ð7:15Þ
RT

This equation, relating the partial conductivity σk to the self-diffusivity Dk is


called the Nernst-Einstein equation.
The partial conductivity of a charged component k, σk, can then be equated,
via Eq. (7.5), to the k-corresponding defect conductivity σdk as
X
σk ¼ σdk : ð7:16Þ
dk

This is how one can predict the partial conductivities (or self-diffusivities) of an
ionic material from its defect structure and tailor them by tailoring the latter, defect
structure; see Appendix I.
The sum of all the partial conductivities in a given description level makes the
total conductivity, σ, of the system given
X
σ¼ σk ð7:17Þ
k

which should be invariant with respect to the description level for a given system.
Dividing Eq. (7.17) by the total conductivity, one has
X σk X
1¼  tk ð7:18Þ
k
σ k

where tk as such is called the transference number of the k-type carriers.


The total conductivity, σ, of an ionic solid may be divided into the partial ionic
conductivity, σion, the sum of the cationic and anionic conductivities, and the
partial electronic conductivity, σel. For the case of our system MaXb+δ with the
charged mobile components fMz1 , Xz2 , e g,

σ ¼ σion þ σel ; σion ¼ σ1 þ σ2 ; σel ¼ σ3 : ð7:19Þ

It follows that

σion σel
1¼ þ ¼ tion þ tel : ð7:20Þ
σ σ

Any ionic solid may, thus, be regarded, in principle, as a mixed ionic electronic
conductor. Particularly, when σion  σel, thus, tion  1, it is called an ionic
conductor or electrolyte. When σion  σel, thus, tel  1, it is called an electronic
conductor or semiconductor. Note that there is no ionic compound such that tion
(¼1  tel)  1 or 0 in the strict sense.
7.3 Flux Equation and a Few Definitions 253

Finally, the (particle) flux equation, Eq. (7.10), is rewritten in terms of self-
diffusivity or equivalently in terms of partial conductivity due to Eq. (7.15), as4

ck D k σk
Jk ¼  ∇ηk ¼  ∇ηk ð7:21Þ
RT ðzk FÞ2

Equivalently, the partial current or charge flux equation takes the form, due to
Eq. (7.12),

ck D k σ
ik ¼ zk F ∇ηk ¼  k ∇ηk : ð7:22Þ
RT zk F

We will use these particle or charge flux (¼current) equations interchangeably.


Warning: In describing charge transport phenomena in an ionic solid, one should
never dare to intermix the carriers having actual charges (charged components)
with those having effective charges (defects), and one should never leave the
description level chosen at the start of description. You would otherwise likely
generate or annihilate charges!

4
Equation (7.21) is only a limiting case in agreement with Kohlrausch’s law of independent
migration of charge carriers. In reality, however, the charge carriers interfere with each other upon
their transfer, and hence, the general flux equation should be written as
X
Jk ¼  Lkm ∇ηm
m

where m runs for all mobile charged components and the transport coefficients Lkm satisfy the
Onsager reciprocity or

Lkm ¼ Lmk :

For a system with one type of mobile ionic carriers (k ¼ i) and electrons (k ¼ e), the experimental
method to determine the three independent Onsager coefficients (Lii, Lie¼Lei, Lee) is well
established and experimentally practiced upon some mixed conductor systems. It is noted that the
ion-electron interference coefficient Lie (¼Lei) is by no means negligible depending on carrier
concentrations, contrary to the conventional, Kohlrausch’s law of independent migration. The
interested reader may refer to the reports:
1. D.-K. Lee and H.-I. Yoo, “Electron-ion interference and Onsager reciprocity in mixed ionic-
electronic transport in TiO2,” Phys. Rev. Lett., 97 (2006) 255901.
2. T. Lee, H.-S. Kim and H.-I. Yoo, “From Onsager to mixed ionic electronic conductors,” Solid
State Ion., 262 (2014) 2.
254 7 Diffusion in Ionic Solids

7.4 Self-Diffusivities in Stoichiometric Compounds

A stoichiometric compound refers to a compound with negligible homogeneity


range or δ  0 so that its composition may be regarded as fixed. Typical examples
are alkali halides, e.g., NaCl. We will now examine how the ionic self-diffusivities
of such a compound vary against its thermodynamic variables. To this end, we need
to know the equilibrium defect structure5 of the system given or defect concen-
trations (cdk ) vs. thermodynamic variables; see Eq. (7.5). Let us take CdCl2-doped
NaCl as an example.

7.4.1 Defect Structure

It is known for the host lattice NaCl (rock-salt structure) that the majority
 ionic

disorder type is the Schottky disorder, i.e., cation-anion vacancy pair V0Na , VCl ,
and electronic disorder (e0, h) is negligible because of a large bandgap. It is further
known that Cd substitutes Na (CdNa), and hence, CdCl2 is incorporated into NaCl as

CdCl2 ! CdNa þ V0Na þ 2ClxCl :


NaCl
ð7:23Þ

The lattice molecule may, thus, be represented as Na12xCdxCl where x denotes


the doping level. Then, the point defects of interest are the set
 
V0Na , VCl , CdNa , ð7:24Þ

and the thermodynamic variables are temperature T and composition x under


atmospheric pressure. Letting [dk] denote the defect concentration (instead of cdk )
following
the
defect

chemistry
jargon,

there are three concentrations to be deter-
mined, V0Na , VCl , and CdNa , and according to the logic of defect chemistry,
there should be exactly three constraints:
(i) Internal or thermal equilibrium: the Schottky defect equilibrium should
prevail, that is, for the Schottky defect formation reaction,

0 Ð V0Na þ VCl ; ð7:25aÞ

5
For a succinct treatment of defect chemistry, see Appendix I, “Defect Chemistry of Solid State
Ionic Compounds,” at the end of this book.
7.4 Self-Diffusivities in Stoichiometric Compounds 255

 


ΔgS
V0Na VCl ¼ KS ¼ exp  ð7:25bÞ
kB T

expressing the defect concentration [dk] as a site fraction in the relevant


sublattice of the NaCl structure. Here, KS and Δgs stand for the Schottky
defect (formation reaction) equilibrium constant and the standard reaction
Gibbs energy of Eq. (7.25a) or Schottky defect formation Gibbs free energy,
respectively.
(ii) Mass conservation: the dopant concentration is fixed, that is,

CdNa ¼ x: ð7:26Þ

(iii) Charge neutrality: our system should remain electrically neutral in any case,
that is,

V0Na ¼ VCl þ CdNa : ð7:27Þ

One may solve these simultaneous equations, Eqs. (7.25b), (7.26), and (7.27),
analytically for the three defect concentrations in the present simple case, but even
more simply by approximating the charge neutrality condition, Eq. (7.27), to the
two limiting or piecewise neutrality conditions as
0

 

VNa  VCl  CaNa ; ð7:28aÞ
0

 

VNa  CaNa  VCl : ð7:28bÞ

These are called the Brouwer approximations, named after the proponent of the
method [4]. One may then get the piecewise solutions for each piecewise neutrality
regime even by inspection as follows:

1=2
(i) Intrinsic regime KS  x : from Eqs. (7.28a), (7.26), and (7.25b),

0


VNa  VCl  KS  CdNa ¼ x:


1=2
ð7:29Þ

1=2
The intrinsic defect fraction (KS ) is much larger than the impurity fraction
(x), and hence, our system is practically “pure.” Noting that Ks is a function of
1=2
temperature

from Eq. (7.25b), this high-temperature regime such that KS 

CdNa is called the intrinsic regime.

1=2
(ii) Extrinsic regime KS  x : from Eqs. (7.28b), (7.26), and (7.25b),

0


K
VNa  CdNa ¼ x  VCl  S : ð7:30Þ
x
256 7 Diffusion in Ionic Solids

ln[S]

Intrinsic region Extrinsic region

1/T
Fig. 7.2 Schematic of defect fractions [S] in logarithmic scale vs. reciprocal temperature for given
x. The exact solutions would be the smooth connection across the boundary of the two piecewise
solutions as delineated by the light solid lines

The impurity is now overwhelming the intrinsic defect concentration or


CdNa  KS ; thus such temperature region is called the extrinsic regime.


1=2

By combining these two sets of piecewise solutions, Eqs. (7.29) and (7.30), one
can get the defect structure of the present system as illustrated schematically in
Fig. 7.2 for a given impurity level x.

7.4.2 Self-Diffusivities

Noting that cdk =ck  ½dk   1, Eq. (7.5) dictates the cation and anion self-diffusivity
DNaþ and DCl , respectively, as

DNaþ ¼ V0Na DV0Na ; DCl ¼ VCl DVCl ð7:31Þ

where the defect diffusivities take the form as in metals (see Chap. 2) as
 
1 Δg þ
DV0Na ¼ ΓV0Na α2 ¼ a2o νo exp  m ; ð7:32aÞ
6 kB T
7.4 Self-Diffusivities in Stoichiometric Compounds 257

 
1  2 Δgm
D VCl ¼ ΓVCl α ¼ ao νo exp 
2
ð7:32bÞ
6 kB T

as each sublattice
pffiffiffi is fcc and, hence, the jump distance is the same for both defects
(α ¼ ao = 2 ). Here, ao is the lattice constant, νo(kBT/h) the thermal vibration
frequency, and Δgm the migrational free energy for cations (m+) and anions
(m).
Now, by combining Eqs. (7.31) and (7.32) with the defect structure, Eq. (7.29) in
the intrinsic regime, and Eq. (7.30) in the extrinsic regime, you can get the analytic
expression for the self-diffusivities as below:
(i) Intrinsic regime

 
Δg =2 þ Δgmþ
DNaþ ¼ a2o νo exp  s ð7:33Þ
kB T

or noting that Δg ¼ Δh  TΔs


 
Δh =2 þ Δhmþ
DNaþ ¼ DNaþ ,0 exp  s ð7:34Þ
kB T

where the pre-exponential factor takes the form


 
Δss =2 þ Δsmþ
DNaþ ,0 ¼ a2o νo exp ð7:34aÞ
kB

Similarly, the self-diffusivity of the anions is written as


 
Δh =2 þ Δhm
DCl ¼ DCl ,0 exp  s ð7:35Þ
kB T

with the pre-exponential factor


 
Δss =2 þ Δsm
DCl 
,0 ¼ a2o νo exp : ð7:35aÞ
kB

(ii) Extrinsic regime

   
Δsmþ Δh þ
DNaþ ¼ xa2o νo exp exp  m
kB kB T
  ð7:36Þ
Δh þ
¼ DNaþ ,0 exp  m
kB T
258 7 Diffusion in Ionic Solids

   
1 2 Δss þ Δsm Δhs þ Δhm
DCl  ¼ ao νo exp exp 
x kB kB T
  ð7:37Þ
Δh þ Δhm
¼ DCl ,0 exp  s
kB T

These self-diffusivities are schematically illustrated in the logarithmic


scale against reciprocal temperature in Fig. 7.3. Upon comparison with the
defect structure in Fig. 7.2, you may recognize that each corresponding line
segment in Fig. 7.3 is rendered steeper by as much as the corresponding
migration enthalpy Δhm . Also note that the Arrhenian slope takes different
values in the two regimes, and hence, by comparing the two different slopes of
each self-diffusivity, you can evaluate the migration enthalpy Δhm and the
Schottky defect formation enthalpy Δhs.
Figure 7.4 shows an as-measured example of cation self-diffusivity of NaCl.
Even if the crystal is not doped intentionally, it is still exhibiting an extrinsic
regime at low temperatures because of background impurities. In comparison
with Fig. 7.3, these impurities are effectively CdNa -like.

Intrinsic region Extrinsic region

ln DK

1/T
Fig. 7.3 Schematic Arrhenian plots of the self-diffusivities of Na+ and Cl in Cd-doped NaCl. The
small triangles upon the line segments denote the corresponding slope

values.
Note

that
the
slope of
each Arrhenian plot has different values in the intrinsic regime ( V0Na  VCl  CdNa ) and in


the extrinsic regime ( V0Na  CdNa  VCl ). The exact solution would be as depicted by the
light solid lines. Compare with the defect structure in Fig. 7.2
7.5 Self-Diffusivities in Nonstoichiometric Compounds 259

Fig. 7.4 The as-measured


self-diffusivity of Na+ (open
circles and solid line) and
from the electrical
conductivity data (closed
circles) vs. 1/T of NaCl.
Even the nominally pure
NaCl exhibits an extrinsic
regime at low temperatures.
(Reproduced from Y.-M.
Chiang et al. [5]. The raw
data, originally from
Mapother et al. [6])

7.5 Self-Diffusivities in Nonstoichiometric Compounds

Unlike a stoichiometric compound, the composition of a nonstoichiometric com-


pound, e.g., AO1+δ, is not fixed, but varies across its stability range, say, in principle,
from the hypostoichiometric regime (δ < 0) through the stoichiometric regime
(δ  0) to the hyperstoichiometric regime (δ > 0) with increasing oxygen activity
in the surroundings at a given temperature. For the composition to vary, the system
should exchange its (neutral) chemical components with the surrounding, thus,
generating electronic defects. What is exchanged is normally the more volatile
component, e.g., O, for the case of oxide AO1+δ. Let us now examine how the
component self-diffusivities of AO1+δ vary against oxygen activity at given tem-
perature. To this end, again we should first understand the defect structure of the
system,6 and then, the correlation Eq. (7.5) will guide you to the end.
We will take a hypothetical, “pure,” nonstoichiometric oxide AO1+δ with
Frenkel disorder (Ai , V00A ) as the dominant ionic disorder and electronic disorder

6
For more detailed treatment, see Appendix I, Defect Chemistry of Solid State Ionic Compounds,
at the end of this book.
260 7 Diffusion in Ionic Solids

(e0, h) induced thermally and/or by nonstoichiometry. Then, the defects of interest
of the present system are the set
 
Ai , V00A ; e0 , h : ð7:38Þ

7.5.1 Defect Structure

The defect chemical logic to calculate the equilibrium defect structure of AO1+δ is
that the system should be in thermal or internal equilibrium and in external or
redox equilibrium with respect to component exchange and electrically neutral.
(i) Internal equilibria: The Frenkel and electronic defects are thermally induced
or of entropic origin. Thus, for the Frenkel defect formation reaction,

AxA Ð Ai þ V00A , ð7:39aÞ

the mass action law requires


 


Δg
Ai V00A ¼ KF ¼ exp  F ð7:39bÞ
kB T

where KF is the Frenkel defect (formation reaction) equilibrium constant


and ΔgF (¼ΔhF  TΔsF) the Frenkel defect formation Gibbs free energy.
For the intrinsic electronic excitation reaction across the bandgap,

0 Ð e 0 þ h , ð7:40aÞ

the mass action law formally reads


 
0  Δgi
½e ½h  ¼ Ki ¼ exp  ð7:40bÞ
kB T

where Ki is the intrinsic electronic excitation reaction equilibrium constant


and Δgi the corresponding Gibbs free energy. Here, the site fractions [e0 ] and
[h] are relative to the effective density of states at the conduction band edge
(NC) and at the valence band edge (Nv), respectively, and Δgi  Eg the
bandgap in semiconductor-physics jargon.
(ii) External equilibrium or redox equilibrium: Our system should be in
equilibrium with the surrounding with respect to component exchange. The
exchange reaction may be written as
7.5 Self-Diffusivities in Nonstoichiometric Compounds 261

1
O Ð OxO þ V00A þ 2h : ð7:41aÞ
2 2ðgÞ

The corresponding mass action law reads


00
 2  
V A ½h  Δg
¼ KOx ¼ exp  Ox ð7:41bÞ
1=2
aO 2 kB T

where KOx is the (oxidation) reaction equilibrium constant and ΔgOx the
standard (oxidation) reaction free energy of Eq. (7.41a). This reaction in
rightward direction is oxidation and in leftward direction reduction, thus
often called the “redox” reaction, and Eq. (7.41b) the redox equilibrium.
(iii) Charge neutrality: The sum of the positive charges should be balanced by the
sum of the negative charges or

2 Ai þ ½h  ¼ 2 V00A þ ½e0 , ð7:42Þ

with the understanding that each defect concentration, in site fraction above, is
now appropriately converted to number per lattice molecule or per unit volume.
We have now the four equations, Eqs. (7.39), (7.40), (7.41), and (7.42), for
the four unknowns in Eq. (7.38). In principle, one can, thus, solve for each defect
concentration as a function of Kj’s (j ¼ F, i, Ox), which are functions only of
temperature or Kj(T) and oxygen activity aO2 . As is seen, however, analytic
solution is extremely hard if not impossible.7 The Brouwer method saves us.
The charge neutrality condition, Eq. (7.42), leads to 2
2 or four possible
piecewise neutrality conditions or majority disorder types as listed in the
matrix form below:

+ − ⬙]
[VA [ e⬘ ]

[A··i ] » [VA
⬙] 2[ A··i ] » [ e⬘ ]
[A··i ] (7.43)
( d » 0) ( d < 0)

[h · ] » 2[VA⬙ ] [h · ] » [ e⬘ ]
[ h·]
( d > 0) ( d » 0)

7
The numerical solution is straightforward with a bit coding, but the Brouwer method allows us to
quickly get analytical insights.
262 7 Diffusion in Ionic Solids

Now, by using each of these piecewise charge neutrality conditions together


with Eqs. (7.39), (7.40), and (7.41), one can solve for the four defect
concentrations analytically with ease for each majority disorder regime in the matrix,
Eq. (7.43), as

(d»0) (d<0)
⁄ ⁄ −1 ⁄ 3 −1 ⁄ 6
[A··i ] » [VA⬙ ] ≈ KF1⁄ 2 2[ A··i ] » [ e⬘] » 2 1 ⁄ 3 KF1 3 Ki2 3 Kox aO
2
1⁄ 2 1⁄ 4
[ h · ] » KF−1⁄ 4 Kox aO 2 [VA⬙ ] » 2 2 ⁄ 3 KF2 ⁄ 3 Ki−2 ⁄ 3 Kox
1⁄3 1⁄6
aO 2
[ e⬘] » KiKF1⁄ 4 Kox
−1⁄ 2 −1⁄ 4
aO 2 [ h · ] » 2 − 1 ⁄ 3 KF−1 ⁄ 3 Ki1 ⁄ 3 K1ox⁄ 3 aO1 2⁄ 6

(7.44)
(d>0) (d»0)
[h · ] » 2[VA⬙ ] ≈ 2 1 ⁄ 3 Kox
1⁄3 1⁄6
aO 2 [h · ] » [ e⬘] ≈ 1 ⁄ 2
−1 ⁄ 3 −1 ⁄ 6
[Ai·· ] » 2 2 ⁄ 3 KF Kox aO 2 [VA⬙ ] » Ki− 1 Kox aO1 2⁄ 2
−1 ⁄ 3 −1 ⁄ 6 − 1 a− 1 ⁄ 2
[ A··i ] » KF Ki Kox
[ e⬘] » 2 − 1 ⁄ 3 Ki Kox aO2 O2

These solutions for each region is called the piecewise solutions based on the
piecewise neutrality conditions.
By now, you may have been puzzled by the assignments of δ < 0, δ  0, or δ > 0
in each majority disorder regime in Matrices (7.43) and (7.44) above. Note that the
nonstoichiometry δ in AO1+δ can be expressed in terms of ionic defect concentra-
tions in number/lattice molecule as

δ ¼ V00A  Ai ð7:45aÞ

or, due to the charge neutrality condition, Eq. (7.42),

1
δ ¼ ð½h   ½e0 Þ: ð7:45bÞ
2

You can then immediately tell whether δ < 0, δ  0, or δ > 0 for each of the four
majority disorder-type regimes; see Matrix (7.43). We know that as the oxygen
activity increases at a given temperature, the three regimes δ<0
(hypostoichiometry), δ  0 (near-stoichiometry), and δ > 0 (hyperstoichiometry)
appear in turn. So, we can assign

a majority

disorder type to each δ-regime. But
oops! There are two types, Ai  V00A and [h]  [e0], for the near-stoichiometry
regime δ  0! Upon comparison of Eqs. (7.39) and (7.40), it is highly unlikely to
have simultaneously the two types in majority; it is simply a matter of whether
KF  Ki or KF  Ki at given temperature!
(i) If KF  Ki or the Frenkel disorder is much cheaper energetically (ΔhF  Eg),
the sequence with increasing aO2 should be
7.5 Self-Diffusivities in Nonstoichiometric Compounds 263

ao2

d<0 d»0 d>0 (7.46a)


2[Ai••] » [e⬘] [Ai••] » [VA⬙] [h•] » 2[VA⬙]

(ii) If Ki  KF or the electronic defects are much cheaper (Eg  ΔhF), the sequence
in the direction of increasing aO2 should be

ao2
(7.46b)
d<0 d»0 d>0
2[Ai••] » [e⬘] [h••] » [e⬘] [h•] » 2[VA⬙]

Now by smoothly connecting the corresponding piecewise solutions in Matrix


(7.44) in the order of Eq. (7.46a) and of (7.46b), respectively, we can get the overall
solution for KF  Ki and for Ki  KF, respectively, as schematically illustrated
against oxygen activity in Fig. 7.4a, b, respectively. These plots are called the
Brouwer diagrams. Note that the neighboring majority disorder regimes should
have one kind of defects in common, and at the very boundary, the concentrations of
this common defect should be the same. This is called the continuity principle,
which serves as the clamp to combine the piecewise solutions in the neighboring
disorder regimes.

7.5.2 Self-Diffusivities

The self-diffusivity of A2+ in AO1+δ may be written, due to Eq. (7.5), as



DA2þ ¼ DAi Ai þ DV00A V00A ð7:47Þ

where the defect concentrations [S] (S ¼ Ai , V00A ) should be in site fractions or
[S] ¼ cS/cA.
The defect diffusivities Dd (d ¼ Ai , V00A ) are given as in Eq. (7.32) or
 
1 Δg
Dd ¼ Γd α2d ¼ γd a2o νd exp  m,d ð7:48aÞ
6 kB T
 
Δh
¼ Dd,o exp  m,d ð7:48bÞ
kB T
¼ Bd RT ð7:48cÞ
264 7 Diffusion in Ionic Solids


C 9JGP-( -K


D 9JGP-K -(

Fig. 7.5 Defect concentrations [S] vs. oxygen activity both in logarithmic scale or Brouwer
diagrams when KF  Ki (a) and Ki  KF (b). The small triangles represent the characteristic
slopes or oxygen exponents “m” such that ½S / am
O2 ; see the solution Matrix (7.44)

where γd denotes the geometric factor for the structure of AO1+δ, that is
normally close to 1, and Δgm,d (¼Δhm,d  TΔsm,d) the migrational free energy.
Upon comparison, you can identify the constituents of the pre-exponential factor
Dd,o in Eq. (7.48b) and the mechanical mobility Bd of the Einstein equation,
Eq. (7.48c).
It should be noted that Dd and Bd are the functions only of temperature. In order to
get an idea about how the self-diffusivity of A2+, DA2þ , varies against aO2 at a given
temperature, let us just assume that DAi ¼ DV00A ¼ 1. Then, Eq. (7.47) in combination
7.5 Self-Diffusivities in Nonstoichiometric Compounds 265

with the piecewise solutions in Matrix (7.44) gets you the variation as shown
schematically in Fig. 7.6.
The cation self-diffusivity apparently

follows
the
ionic defects in majority. In the
majority disorder regime of ½e0   2 Ai  2 V00A , Eq. (7.47) leads to

1=6
DA2þ  DAi Ai / aO2 : ð7:49Þ

In the majority disorder regime of ½h   2 V00A  Ai ,



þ1=6
DA2þ  DV00A V00A / aO2 : ð7:50Þ

In the majority disorder regime of Ai  V00A ,



DA2þ ¼ DAi Ai þ DV00A V00A / a0O2 : ð7:51Þ

In the majority disorder regime of [e0]  [h], the majority ionic disorder type
shifts from Ai to V00A crossing the stoichiometric composition where [e0]  [h] or
δ  0 (see Eq. 7.45). Consequently, the self-diffusivity shows a V-shaped variation
with an oxygen exponent of m ¼ 1/4 and +1/4 in turn or




1
DA2þ ¼ DAi Ai þ DV00A V00A / cosh ln aO2 : ð7:52Þ
4

This is the story of the cation self-diffusivity for a hypothetical oxide AO1+δ.
What about the anion self-diffusivity DO2 ? The same argument applies, viz.,

DO2 ¼ DO00i O00i þ DVO VO : ð7:53Þ

Then, you should know the anionic defect structure in addition. How to estimate
it? You are here adding two more unknowns to the set of the unknowns, Eq. (7.38),
and hence, you need two more constraints which are two internal equilibria con-
straints involving the two types of defects in Eq. (7.53): the anion-Frenkel defect
equilibrium and the Schottky defect equilibrium:
00


Oi VO ¼ K0F ; V00A VO ¼ KS ð7:54Þ

By solving simultaneously these two equations together with those four equations
above, you can additionally get the two as functions of T and aO2, and then Eq. (7.53)
guides you to estimate DO2 . For a succinct treatment, see Appendix I at the end of
this book.
266 7 Diffusion in Ionic Solids


C 9JGP-( -K


D 9JGP-K -(

Fig. 7.6 Schematic log DA2þ vs. log aO2 at a fixed temperature assuming DAi ¼ DV00A ¼ 1: (a)
when KF  Ki; (b) when Ki  KF. Compare with the corresponding defect structures given in
Fig. 7.4. The piecewise variations are connected smoothly by a light solid curve crossing the
majority disorder-type boundaries

7.5.3 Examples

Co12δO
The as-measured cation tracer diffusivity of Co1δO, D Co, varies against aO2 (¼PO2 /
atm) as shown in Fig. 7.7. It is reminded that the tracer diffusivity D Co is related to
the self-diffusivity as D Co ¼ fDCo where f is the correlation factor that takes a value
f ¼ 0.781 for the present system with the fcc lattice (see Chap. 2).
7.5 Self-Diffusivities in Nonstoichiometric Compounds 267

As is seen,

D Co / am
O2 ð7:55Þ

with m varying from m  +1/6 to +1/4 as aO2 increases from the lower bound of the
phase stability field up to aO2 ¼ 1. Referring to Fig. 7.6 or Matrix (7.44), one may
suspect that the system is hyperstoichiometric (δ > 0) or metal-deficit across the
entire aO2 -range examined.
Indeed, Co1δO is known to be metal-deficit across its entire stability range
intuitively with V00Co , h in majority. If this is the case, the redox equilibrium
will be written as

1 V00Co ½h 2
O ¼ OxO þ V00Co þ 2h ; ¼ KOx ð7:56Þ
2 2ðgÞ 1=2
aO2

and the charge neutrality condition as


2 V00Co ¼ ½h : ð7:57Þ

It follows that

1=3 þ1=6
V00Co ¼ 22=3 KOx aO2 : ð7:58Þ

One can, therefore, expect that



þ1=6
DCo ¼ DV00Co V00Co / aO2 : ð7:59Þ

It appears, however, to be valid only as aO2 decreases to the phase boundary. In


reality, DCo increases with aO2 faster (m  1/4) than expected (m  1/6) in the
greater portion of the aO2 -range examined; see Fig. 7.7.
A plausible explanation is that as aO2 increases, the majority
 ionic disorder
 type
shifts from the fully ionized V00Co to the partially ionized V0Co ¼ V00Co þ h due to the
attractive interaction between the charge compensating defects. If this is the case, the
redox equilibrium should read

1 V0Co ½h 
O ¼ OxO þ V0Co þ h ; ¼ K0Ox ð7:60Þ
2 2ðgÞ 1=2
aO2

followed by the charge neutrality condition


0

VCo  ½h : ð7:61Þ


268 7 Diffusion in Ionic Solids

Fig. 7.7 The as-measured


cation tracer ① 1410oC
diffusivity vs. oxygen -7 ② 1310oC
activity of Co1δO from ③ 1210oC
oxygen atmosphere down to ④ 1108oC
the phase stability limit ⑤ 1006oC
-8
(denoted by vertical dotted-

log (D*Co/ cm s )
line segments) at different

2 -1
temperatures. The two small
triangles denote the ideal ①
-9
slopes or oxygen exponents
② 1
m ¼ 1/4 and 1/6,
4
respectively. (From
Dieckmann [7]) -10 ③


-11
1
⑤ 6

-12 -10 -8 -6 -4 -2 0
log aO2

Therefore,

V0Co ¼ K0 Ox aO2 ,
1=2 1=4
ð7:62Þ

and hence,

DCo ¼ DV0Co V0Co ¼ DV0Co K0 Ox aO2 :


1=2 1=4
ð7:63Þ

Fe32δO4
Figure 7.8 shows the as-measured tracer diffusivity of Fe, D Fe , against oxygen
activity both in the logarithmic scale at six different temperatures. Interestingly
enough, D Fe exhibits a V-shaped variation with the oxygen exponents close to
m  2/3 or algebraically

pffiffiffiffiffiffiffi h i
þ2=3 2=3 2=3 1 A
DFe  AaO2 þ BaO2 ¼ AB cosh ln aO2 þ ln ð7:64Þ
2 B

with A and B being the functions of temperature only. As a consequence, the right-
hand branches (m  +2/3) of the V-shaped D Fe even result in that D Fe appears to
decrease with increasing temperature at a fixed oxygen activity. This is apparently
contrary to the “common sense,” viz., that diffusion is a thermally activated process.
You may wish to elucidate the reason for this apparent contradiction.
7.5 Self-Diffusivities in Nonstoichiometric Compounds 269

Fig. 7.8 As-measured tracer diffusivity of Fe, D Fe, vs. oxygen activity both in the logarithmic scale
at six different temperatures. Note that a diffusivity isotherm exhibits a V-shaped variation leaving a
minimum. (From Schmalzried [8])

In comparison with the V-shaped variation in the near-stoichiometry region


(δ  0) in Fig. 7.6b, it is suggested that the majority disorder type may change
crossing this minimum.
The interpretation of this V-shaped behavior has been put forward by Dieckmann
and Schmalzried [9]. Fe3O4 has a spinel structure having two types of cationic
sites, the tetrahedral sites and octahedral sites, created by the cubic close-packed
oxide ions: only one eighth of the former and a half of the latter are occupied by the
cations. The ideal lattice molecule of magnetite is often represented as
 

Fe3þ Fe2þ Fe3þ O4 ð7:65Þ

where (  ) denotes the tetrahedral sites and [  ] the octahedral sites. Because
of plenty of empty tetrahedral and octahedral interstices, one can immediately
expect that the Frenkel disorder will be in a majority. At temperatures as high as
now in Fig. 7.8, on the other hand, the cation distribution on the two types of sites
will become completely disordered. Fe3+ on the Fe2+ site may then be understood
to be a hole, i.e., Fe3þFe2þ
 Fe2þ
Fe2þ
¼ hþ , and the opposite to be an electron or
Fe2þ
Fe3þ
 Fe3þ
Fe3þ
¼ e in terms of the actual charges. Its equilibrium defect structure
may then be described, neglecting the effective charges of the ionic defects, as
follows:
270 7 Diffusion in Ionic Solids

(i) Internal equilibria

FexFe ¼ Fei þ VFe : ½Fei ½VFe  ¼ KF ð7:66Þ


Fe2þ
Fe2þ
þ Fe3þ
Fe3þ
¼ Fe3þ
Fe2þ
þ Fe2þ
Fe3þ
: ½e ½hþ  ¼ Ki ð7:67Þ

(ii) External equilibrium

1 3 3=4 þ 2 1=2
O þ 2Fe2þ ¼ O2
O þ 4 VFe þ 2FeFe2þ : ½VFe 

½h   KOx aO2 ð7:68Þ
2 2ðgÞ Fe2þ

For a random magnetite,


pffiffiffiffiffi
½ hþ   ½ e    Ki : ð7:69Þ

Then, Eqs. (7.68) and (7.66) lead to

2=3
½VFe  ¼ ðKOx =Ki Þ4=3 aO2 ;
2=3
ð7:70Þ
½Fei  ¼ KF ðKOx =Ki Þ4=3 aO2 :

Equation (7.5) finally stipulates that

þ2=3 2=3
DFe ¼ DVFe ½VFe  þ DFei ½Fei  ¼ AaO2 þ BaO2 ð7:71Þ

where you can easily recognize what constitute A and B, and hence, they should
be only functions of temperature. As expected, the V-shaped variation has
indeed been induced by the majority disorder-type change from interstitials
(Fei) to vacancies (VFe) in the present case. Can you now tell that in the vacancy
branch of Eq. (7.71), A grows smaller as temperature increases?

7.6 Diffusion in Concentration Gradients

Suppose that a binary oxide A1δO is put under an oxygen potential gradient ∇μO
at a given temperature. As δ ¼ δ(T,μO), there will have to be a redistribution of
composition δ. For the composition to change, the chemical components A and O
should diffuse, which is, thus, called the chemical diffusion, corresponding to the
interdiffusion in a metallic alloy. How can chemical diffusion occur?
7.6 Diffusion in Concentration Gradients 271

7.6.1 Component Flux Equations

The Gibbs-Duhem equation stipulates for the system A1δO that

ð1  δÞ∇μA þ ∇μO  ∇μA þ ∇μO ¼ 0 ð7:72Þ

as |δ|  1 normally. The gradient ∇μO applied and ∇μA( ∇μO) induced thereby
induce the chemical component fluxes JA and JO in counter directions as

JA ¼ LA ∇μA ; ð7:73aÞ


JO ¼ LO ∇μO ð7:73bÞ

where LK (K ¼ A,O) is the transport coefficient or LK ¼ cKBk; see Eq. (7.10). The
chemical diffusion rate is determined by these transport coefficients, and we want to
know what constitute LK. We know what are actually moving in an ionic solid
A1δO are charged components {A2+,O2,e}; see Fig. 7.9.
Let us describe the fluxes of the chemical components {A,O} in terms of these
charged components {A2+,O2,e}. Letting 1, 2, and 3 denote the cations (z1 ¼ 2),
anions (z2 ¼ 2), and electrons (z3 ¼ 1), respectively, the charged component
fluxes are written for each charged component, due to Eq. (7.21), as

Fig. 7.9 (a) An oxygen


potential gradient ∇μO A O
applied across A1δO
induces the chemical (a)
potential gradient of A, ∇μA,
in the opposite direction. (b) A1- O
These chemical potential
gradients induce the
chemical components A and
O to flow down the
respective gradients. (c)
(b)
These chemical component
flows are actually realized
by the charged component
flows while keeping the
local charge neutrality

(c)
272 7 Diffusion in Ionic Solids

D 1 c1 σ1
A2þ : J1 ¼  ∇η1 ¼  ∇η1 ð7:74aÞ
RT ð z 1 FÞ 2
D 2 c2 σ2
O2 : J2 ¼  ∇η2 ¼  ∇η2 ð7:74bÞ
RT ð z 2 FÞ 2
D3 c3 σ3
e : J3 ¼  ∇η3 ¼  ∇η3 ð7:74cÞ
RT ðz3 FÞ2

Due to the local ionization equilibria, Eqs. (7.3) and (7.4), the driving forces are
subjected to the local equilibrium constraints by necessity as

A ¼ Az1 þ z1 e : ∇μA ¼ ∇η1 þ z1 ∇η3 ; ð7:75Þ



O ¼ O þ z2 e : ∇μO ¼ ∇η2 þ z2 ∇η3 :
z2
ð7:76Þ

Furthermore, there should be no net current across the system because the circuit
involving the system oxide is open8:

X
3 X
3
i¼ ik ¼ zk FJk ¼ 0 ð7:77aÞ
k¼1 k¼1

or, due to Eq. (7.74),

σ1 σ σ
 ∇η1  2 ∇η2  3 ∇η3 ¼ 0 ð7:77bÞ
z1 F z2 F z3 F

By solving Eqs. (7.75), (7.76), and (7.77), one may obtain

t1 t
∇η3 ¼ ∇μA þ 2 ∇μO ð7:78aÞ
z1 z2

or, due to Gibbs-Duhem equation, Eq. (7.72),

tion t
∇η3 ¼ ∇μA ¼ ion ∇μO ð7:78bÞ
z1 z2

where tion (t1 + t2) is the ionic transference number of the system.
Here, we will digress a bit. You see, integration of Eq. (7.78b) across the
thickness L of the system leads to the electronic electrochemical potential
difference

8
This is a special case of the general charge neutrality condition ∇i ¼ 0. For an extensive detailed
treatment, see Lee and Yoo [10].
7.6 Diffusion in Concentration Gradients 273

Z L
tion
Δη3 ¼ dμ : ð7:79Þ
0 z2 O

A chemical potential difference of a component applied, say, ΔμO, induces an


electronic electrochemical potential difference, Δη3, unless tion  0. This is what you
can measure across oxide A1δO by using a voltmeter, viz., the open-circuit voltage
U (¼ Δη3/F). Conversely, a U applied induces ΔμO ( ΔμA) or a composition
difference Δδ. This is the very principle of the chemical-to-electrical energy or
information conversion and storage. You may recognize that the conversion and
storage become most effective when tion  1, and such a system is called an
electrolyte.
Now, back to the mainstream of our discussion, from Eqs. (7.75), (7.76), and
(7.78b), one obtains

∇η1 ¼ tel ∇μA ; ∇η2 ¼ tel ∇μO : ð7:80Þ

Compare Eqs. (7.78b) and (7.80) to see how a ∇μO ( ∇μA) applied is
distributed to ∇ηk (k ¼ 1,2,3). Also, think about what would happen to them if tel  0
or tion  0. To our relief, however, there is no such an ionic solid!
As we now know ∇ηk (k ¼ 1,2,3) in terms of ∇μO ( ∇μA) that is applied, the
ion fluxes in Eq. (7.74) are rewritten as

D 1 c1 σ1
A2þ : J1 ¼  t ∇μ ¼  tel ∇μA ; ð7:81aÞ
RT el A ðz1 FÞ2
D 2 c2 σ2
O2 : J2 ¼  t ∇μ ¼  tel ∇μO ; ð7:81bÞ
RT el O ðz2 FÞ2

and the electron flux is given, due to Eq. (7.77a), as

D3 c3 tion σ3 tion
e : J3 ¼ z1 J1  z2 J2 ¼  ∇μO ¼  ∇μO : ð7:81cÞ
RT z2 ðz3 FÞ2 z2

It is obvious that J1 ¼ JA ¼ LA∇μA and J2 ¼ JO ¼ LO∇μO. You may now


immediately recognize what constitute LA and LO upon comparison with Eq. (7.81).
By comparing with Fick’s first law for A and O, you may further define their
component diffusivities each involving a thermodynamic factor as we did in an
alloy system; see Chap. 4.

7.6.2 Nonstoichiometry Flux

For a binary system like A1δO, there is a single composition variable δ, which is
experimentally measurable by, e.g., thermogravimetry or coulometric titrometry.
274 7 Diffusion in Ionic Solids

Finally, let us examine how the composition or nonstoichiometry δ varies under a


∇μO applied.
For the system A1δO,

1δ 1
cA ¼ c1 ¼ ; cO ¼ c2 ¼ : ð7:82aÞ
Vm Vm

Thus,

δ
cO  cA ¼ ð7:82bÞ
Vm

where Vm denotes the molar volume of the system oxide and is assumed to be
constant.
We know that

∂cA ∂cO
¼ ∇JA ¼ ∇J1 ; ¼ ∇JO ¼ ∇J2 ð7:83Þ
∂t ∂t

for the component A and O are each conserved. Rate of the nonstoichiometry change
then takes the form
 
∂ δ ∂c ∂c
¼ O  A ¼ ∇ðJO  JA Þ  ∇Jδ : ð7:84Þ
∂t Vm ∂t ∂t

Here, we may call Jδ(JO –JA) the nonstoichiometry flux. By combining the
component fluxes, Eq. (7.81), we finally end up with Jδ in terms of the component
self-diffusivities or equivalently in terms of partial ionic conductivities, due to the
Nernst-Einstein equation, as

ð D 1 c1 þ D 2 c2 Þ
Jδ ¼ JO  JA ¼  tel ∇μO ð7:85aÞ
RT
ðσ1 þ σ2 Þ
¼ tel ∇μO ð7:85bÞ
ðz2 FÞ2

where we used the relation z1 ¼ |z2| and we know z2 ¼ 2 for AO.

7.6.3 Chemical Diffusion Coefficient

Fick’s first law defines the diffusion coefficient in the given situation. The first law
for δ of the system A1δO stipulates
7.6 Diffusion in Concentration Gradients 275

 
e δ
Jδ ¼ D∇ ð7:86Þ
Vm

where D e is the chemical diffusion coefficient controlling the thermodynamic


equilibration kinetics. Upon comparison with Eq. (7.85), the chemical diffusivity
takes the form, in terms of the component self-diffusivities and in terms of the
partial ionic conductivities, respectively, as
 ∂μ 
e ¼ D1 c1 þ D2 c2 tel
D O
ð7:87aÞ
RT ∂ðδ=Vm Þ
   
σ1 þ σ2 ∂μO
¼ t 3 ð7:87bÞ
4F2 ∂ðδ=Vm Þ

by setting z1 ¼ z2 ¼ 2 for A1δO. Keep in mind that σion ¼ σ1 + σ2, tel ¼ t3 and
furthermore, σiontel ¼ tionσel.
The last factor involving only the thermodynamic quantities is called the ther-
modynamic factor as is done in metallic interdiffusion systems. If the oxide is in
external equilibrium with respect to oxygen exchange,

1 1 o
μO ¼ μO2 ðgÞ ¼ μO2 þ RT ln aO2 ð7:88Þ
2 2

assuming the gas oxygen in the surrounding is ideal. Here, μoO2 is the standard
chemical potential of gas oxygen. By substituting Eq. (7.88) into the thermodynamic
factor in Eq. (7.87), one obtains
 
e 1 ∂ ln aO2
D ¼ ðD1 c1 þ D2 c2 Þtel  ð7:89aÞ
2 ∂ðδ=Vm Þ
 
RT ðσ1 þ σ2 Þσ3 1 ∂ ln aO2
¼ 2  : ð7:89bÞ
4F ðσ1 þ σ2 þ σ3 Þ 2 ∂ðδ=Vm Þ

In the latter, we employed the definition of tel (¼σ3/σ) to show the configuration
of the partial conductivities. Remember that we have already encountered the
primitive functions, e.g., Eq. (7.58), of this kind of thermodynamic factor involving
a defect concentration in establishing a Brouwer diagram above.
The combinations of the component self-diffusivities or ionic conductivities via
the charge neutrality as in Eqs. (7.87) and (7.89) is called the Nernst-Planck-type
diffusion coefficient. You may remember the Darken-type diffusion coefficient,
which is a consequence of the flux coupling to conserve the local lattice. In ionic
solids, the component fluxes are coupled to keep the local charge neutrality, thus
resulting in the chemical diffusivity as derived up to now. As is seen, the chemical
e is governed by the faster of the cations and anions (σ1 + σ2) and by the
diffusivity D
slower of the electrons and ions (σionσel/(σion + σel)).
276 7 Diffusion in Ionic Solids

You are strongly recommended to derive De for a system of general composition,


e.g., A1δOν, where ν is the stoichiometric coefficient such that ν ¼ z1/|z2| and
|δ|  1 as is the case with not-weird, ordinary nonstoichiometric compounds. By
following the same flow of logic, you will end up with the expression, for the
nonstoichiometry flux,

1 σ t
Jδ ¼ JO  JA ¼  ion el 2 ∇μO
ν z1 jz2 jF
  ð7:90Þ
e δ
¼  D∇
Vm

from which you can easily identify the chemical diffusivity by introducing the
thermodynamic factor.
e of the system Co1δO in Fig. 7.7. The oxide is
As an example, let us examine D
known to be a p-type semiconductor across its entire stability range, and hence

tion  ttel  1: ð7:91aÞ

From the defect structure, Eq. (7.58),


δ ¼ V00Co / aO2 :
1=6
ð7:91bÞ

Furthermore, the self-diffusivity of O is known to be normally three to four orders


of magnitude smaller than that of cations in CoO as is the case with other close-
packed oxides or

D1 c1  D2 c2 ð7:91cÞ

where c1  c2 ¼ 1/Vm in molar concentration and, due to Eqs. (7.5) and (7.91b),

δ
D1 c1 ¼ DV cV ¼ DV ð7:91dÞ
Vm

where the subscript V means VCo, the Co-cation vacancy.


The full-fledged chemical diffusivity in Eq. (7.89a) may then be approximated as
     1
e δ 1 ∂ ln aO2 1 ∂ ln ðδ=Vm Þ
D DV  ¼ DV : ð7:92aÞ
Vm 2 ∂ðδ=Vm Þ 2 ∂ ln aO2

Equation (7.91b) gets you the numerical value 6 for the thermodynamic factor,
the last factor in Eq. (7.92a), and hence

e  3DV :
D ð7:92bÞ
7.7 Nonstoichiometry Re-Equilibration Kinetics 277

Fig. 7.10 Initial


nonstoichiometry
t=0
distribution (solid line at
i
t ¼ 0) of an infinite slab of
A1δO with thickness L. As
time goes by, the initial t
nonstoichiometry δi across
L relaxes (dotted lines with
increasing t) symmetrically A1- O
toward the final value δf,
which is fixed by the f
boundary conditions x
0 L

That is, the chemical diffusivity of the system CoO is three times larger than the
cation vacancy diffusivity!
 Obviously,
 however, this is valid over the oxygen
activity region where V00Co , h in majority; thus, with the oxygen
 exponent
 of the
cation self-diffusivity or electrical conductivity, m ¼ 1/6. If V0Co , h in majority or
m ¼ 1/4, then D e  2DV .

7.7 Nonstoichiometry Re-Equilibration Kinetics

Suppose that somehow we let there be a nonstoichiometry distribution for an infinite


slab of, say, A1δO with a thickness L along the x-direction as shown in Fig. 7.10.
The initial and boundary conditions may be written as

δð0 < x < L, t ¼ 0Þ ¼ δi ; ð7:93Þ


δðx ¼ 0, tÞ ¼ δðx ¼ L, tÞ ¼ δf : ð7:94Þ

There then arises a nonstoichiometry flux Jδ as given in Eq. (7.85), which varies
local nonstoichiometry value due to Fick’s second law as
 
∂δ ∂ e ∂δ
¼ ∇Jδ ¼ D ð7:95Þ
∂t ∂x ∂x

where the chemical diffusivity is as given in Eq. (7.87) or (7.89). By solving this
equation, thus, one can obtain the spatial and temporal variation of nonstoichiometry
δ(x,t). For solution, of course, you should know the composition dependence of the
chemical diffusivity or D e ¼D e ðδÞ.
If
 
δi  δf 
 
 δi   1, ð7:96Þ
278 7 Diffusion in Ionic Solids

we may, however, assume D e to be constant in the range of [δi, δf] and will end up
with the “trivial” solution δ(x,t) of infinite trigonometric series which we obtained
by using the separation-of-variables method in Chap. 1:
 
δðx, tÞ  δf 4 X 1 ð2j þ 1Þπx e
ð2j þ 1Þ2 π2 Dt
¼ sin exp  ð7:97Þ
δi  δf π j¼0 ð2j þ 1Þ L L2

or, in terms of the spatial mean of δ(x,t) at time t, δðtÞ, as


 
δð t Þ  δf 8X 1 e
ð2j þ 1Þ2 π2 Dt
¼ 2 exp  ð7:98Þ
δi  δf π j¼0 ð2j þ 1Þ2 L2

In an appropriate time determined by your error tolerance as discussed in Chap. 1,


these solutions may be approximated only to the first term (j ¼ 0) to make the long-
time solutions.
The boundary value of nonstoichiometry δf is often experimentally adjusted by
controlling the oxygen activity in the surrounding of a binary system oxide A1δO at
given temperature as δ ¼ δðaO2 , TÞ. Even if we change the oxygen activity literally
stepwise, the target surface nonstoichiometry value δf normally cannot be
established instantaneously because the redox reaction, e.g., Eq. (7.60), should
occur at the surfaces, and as a consequence, this reaction provides another kinetic
step in series with the solid state chemical-diffusion step. If this is the case, the
boundary conditions should be written differently from Eq. (7.93). The normal
practice is to assume that the surface reaction rate is of the first order or

Jδ ¼ kðδf  δo Þ at x ¼ 0 and L ð7:99Þ

where k is the surface reaction rate constant and δo the instantaneous concentra-
tion at the surface (x ¼ 0,L). Then, one should solve the diffusion equation,
Eq. (7.95), with the same initial condition as before, Eq. (7.93), but with different
boundary condition, Eq. (7.99). The analytic solution due to the Laplace-transform
method is available when the chemical diffusivity is constant.9
Let us here consider how long one should wait to completely equilibrate
A1δO. The first-term approximation of Eq. (7.98) or the long-time solution is
written as

δð t Þ  δf 8 L2
 2 et=τ ; τ ð7:100Þ
δi  δf π e
π2 D

where τ is the relaxation time of the system with a given dimension L. For 99%
equilibration or δ  δf ¼ 0:01ðδi  δf Þ, one needs to wait for a time duration, t  4τ,

9
For the analytic solutions, see Crank [11], and for experimental practices, see, e.g., Song and
Yoo [12].
Problems 279

and for 99.9% equilibration, t  7τ. For Co1δO, for example, D e  3DV  106
cm /s at ca. 1000 K. In order to “fully” re-equilibrate a plate-like specimen with
2

1 mm thickness at 1000 K, thus, you have to wait more than 7τ  1000 s or an hour.

Problems

1. At 723 K the predominant point defect in UO2 is the oxide ion vacancy with an
effective charge of +2. For the oxide ion in this crystal, calculate (i) the electro-
chemical mobility and (ii) the transference number.
Data: at 723 K
Total electrical conductivity of UO2, σ ¼ 0.1 S/cm
Density of UO2, ρ ¼ 10.5 g/cm3
Atomic mass of UO2 ¼ 270 g/mol
Self-diffusion coefficient of O2 ¼ 1013 cm2/s
2. Figure 7.11 shows the dependence of the self-diffusivity of zinc, DZn, in ZnO on
the partial pressure of zinc, pZn, at 800 K.

(a) Write the predominant defect incorporation reaction for ZnO at this
temperature.
(b) Plot the temperature dependence of DZn from room temperature to its melting
point under an atmosphere of zinc vapor fixed at pZn ¼ 105 kPa. Assume the
ZnO to be free of impurities.
Data:
ZnO melts at 2240 K.
For Zn transport through ZnO:

DZn/cm2s–1

10–15

10–14

10–13

0.001 0.01 0.1 1 10


pZn/atm

Fig. 7.11 DZn vs. pZn of ZnO in both logarithmic scales at 800 K
280 7 Diffusion in Ionic Solids

ΔH/eV ΔS/kB
Frenkel disorder 2.6 9.0
Zn incorporation 2.0 2.0
Zn migration 0.7 2.7
γa2ν ¼ 0.1 cm2/s
3. In CaF2 the predominant defects are Frenkel pairs on the anionic sublattice (anti-
Frenkel disorder). The important cation defects are calcium vacancies. There are
virtually no Ca interstitials. The solubility limit of NaF in CaF2 is several percent
at the eutectic temperature.
(a) Do you expect DCa or DF to be greater in pure CaF2?
(b) For the solution of NaF in CaF2, write the solute incorporation reaction and
the equilibrium constant at saturation with respect to NaF.
(c) What effect do you expect that NaF addition will have on DCa and DF?
(d) Do you expect DNa to be greater or lesser than DCa? Justify.
4. NiO, suspended in a thermobalance, loses weight if the oxygen pressure is
suddenly changed from 1 bar to 1010 bar. After 0.442 h, 90% of the total weight
change has occurred. This weight change was 0.214 mg, and the disclike sample
with parallel surface had a weight of 1 g, the surface area being 2.1 cm2. It is
known that NiO is a metal-deficient p-type semiconductor with a molar volume of
11.2 cm3/mol.
(a) Calculate the chemical diffusion coefficient of NiO.
(b) Estimate the self-diffusion coefficient of nickel ions. If one would find that
D Ni > DNi , what would one conclude from this unusual result of a reduction
experiment?
(c) Estimate the ionic conductivity of NiO.
5. NiO is of rock-salt structure type with a lattice parameter of 0.4177 nm and has a
bandgap of Eg ¼ 4.2 eV. We estimate the Schottky defect formation energy to be
~6 eV. Because Ni can be oxidized to the trivalent state, NiO tends to be slightly
cation-deficient (Ni1δO or NiO1+δ).
(a) In the hypothetically pure and stoichiometric state (δ ¼ 0), do you expect NiO
to be an electronic or ionic conductor at a temperature of 1000 K? Show how
you arrive at your answer.
(b) In the nonstoichiometric state (δ > 0), do you think the cation deficiency of
Ni1δO will be accommodated as cation vacancies or as oxygen interstitials?
Why? Write a formally correct defect reaction and mass action equilibrium
constant for the oxidation mechanism of your choice.
(c) Will the nonstoichiometric material be p- or n-type in its electronic conduc-
tivity? Explain. Draw the Brouwer diagram in this regime, and find the Po2
dependence of the electronic conductivity and nonstoichiometry.
(d) Assume cation vacancies to be the predominant defect accommodating
nonstoichiometry. Given a cation vacancy migration energy of 2.4 eV,
Problems 281

estimate the vacancy diffusion coefficient, DV, and the cation self-diffusion
coefficient, DNi, respectively, at 1000 K. Clearly state the assumptions
employed if any. (Note: The as-measured tracer diffusivity is
1.2
1018 m2/s!)
(e) Referring back to Part (d), what concentration of cation vacancies is neces-
sary for stoichiometric NiO to be a mixed conductor (at 1000 K) with an ionic
transference number of 0.5? Assume the electron mobility to be 10 cm2/V-s
and the hole mobility to be 0.1 cm2/V-s. Compare your calculated result with
the Ni concentration of NiO.
(f) Derive an expression for the chemical diffusion coefficient D e of Ni1δO in
terms of vacancy diffusivity and electronic transference number.
(g) On the basis of Parts (c) and (d), evaluate the chemical diffusivity.
(h) Let us now dope our NiO with Li2O up to 10 m/o. Assuming Li+ substitutes
Ni2+, write the incorporation reaction. Remember that Ni1δO is stable only
for δ > 0.
(i) Draw the Brouwer diagram for the 10 m/o Li2O-doped NiO and compare with
that of pure NiO in Part (c).
(j) On the basis of the Brouwer diagram in Part (i), explain figuratively how the
electronic conductivity and ionic conductivity respectively vary against the
oxygen partial pressure.
6. A slab of semiconducting A1δO is placed between two different oxygen partial
pressure atmospheres (P0O2 ¼ 104 atm and P00O2 ¼ 101 atm) as shown in the
figure.

A1−δ O

PO′2 = 10−4 atm PO″2 = 10−1 atm

150 μm

(a) In which direction does the oxide shift? Explain why.


(b) Estimate the shift velocity of the oxide.
(c) Evaluate the flux of gas oxygen (in mole O2/cm2∙s) across the slab with a
thickness of 150 μm. You may assume here that the surface reaction rate is
fast enough.
(d) Assuming that the majority type of disorder in this oxygen partial pressure
e
range is V00 and h•, evaluate the chemical diffusivity of the system, D.
A

Data: Self-diffusion coefficient of A, DA ¼ 109 cm2/s


Molar volume of the oxide, Vm ¼ 10 cm3/mol
282 7 Diffusion in Ionic Solids

7. Calcia-stabilized zirconia (CSZ) has the following properties at 1000 C:


Transference number of O2, tO2 ¼ 0.999
Hole mobility, uh ¼ 104 cm2/Vs
Self-diffusion coefficient of O2, DO2 ¼ 3
108 cm2/s
Total
••
00
σ ¼ 0.4 S/cm

conductivity,
VO  CaZr ¼ 0.12/lattice molecule
Density, ρ ¼ 5.6 g/cm3
DZr, DCa < 1015 cm2/s.
(a) An oxygen potential gradient (1 atm–104 atm) is applied across a
1-mm-thick plate at 1000 C as shown in Fig. 7.12a. Calculate the permeation
flux of oxygen across the plate in mole O2/cm2s, assuming the surface
reaction rate is fast enough.
(b) In addition to the oxygen potential gradient, a voltage of 0.1 V is applied as
shown in Fig. 7.12b. What is the oxygen flux?
(c) Calculate the oxygen flux when the polarity of the voltage applied is reversed.
(d) Calculate the oxygen flux when the circuit in Part (b) is short as shown in
Fig. 7.12c.

Fig. 7.12 A zirconia slab under an oxygen partial pressure difference only (a), together with an
external voltage applied (b), and with the both electrodes short-circuited (c)

8. A crystal of ultrapure KCl is doped with K2O to an oxygen impurity level of


105.
(a) Plot the diffusion coefficients of K and Cl against reciprocal temperature, and
indicate the values of the slopes.
(b) What is the temperature of the intrinsic-to-extrinsic transition?
(c) Estimate the magnitude of DO relative to DK and DCl. What additional
information is required? Can you estimate the magnitudes of these quantities?
References 283

Data Migration Schottky disorder


ΔHm/kJmol1 ΔSm/R ΔHS/kJmol1 ΔSS/R
K+ 67 2.7 250 9.6
Cl 96 4.1
Assume that the solubility of K2O above room temperature is greater than 105
and γa2o ν  0:1 cm2 s1 .
9. PbS has the rock-salt (NaCl-type) crystal structure in which the cations and
anions are on fcc sublattices. The predominant defects in PbS are Frenkel defects
on the Pb-sublattice. Furthermore, PbS frequently contains a small excess of Pb
ions.
(a) Would you expect DPb or DS to be greater? Explain.
(b) How would you expect additions of Ag2S to affect DPb?
(c) How would you expect additions of Bi2S3 to affect DPb?
Ionic radii: r(Pb2+) ¼ 0.120 nm; r(S2) ¼ 0.184 nm
10. For Ca diffusion in NaCl, it is known

that Ca

diffuses by vacancy mechanism on
 0
the Na sublattice and that CaNa  VNa over the particular range of interest.
Certainly one can write Fick’s first law for the flux of Ca. Is it correct to write


2

1 ∂ CaNa ∂ CaNa
¼ ?
DCa ∂t ∂x2

Justify your answer.

References

1. H.-I. Yoo, K.-C. Lee, J. Electrochem. Soc. 145, 4243 (1998)


2. M. Faraday, Experimental Researches in Electricity Art (Taylor and Francis, London, 1839),
p. 1339, quoted from K. Funke, Sci. Technol. Adv. Mater. 14, 043502 (2013)
3. C. Wagner, Z. Phys. Chem. B21, 25 (1933)
4. G. Brouwer, A general asymptotic solution of reaction equations common in solid-state
chemistry. Philips Res. Rep. 9, 366–376 (1954)
5. Y.-M. Chiang, D. Birnie III, W.D. Kingery, Physical Ceramics (John Wiley & Sons, Inc.,
New York, 1997), p. 202
6. D. Mapother, H.N. Crooks, R. Maurer, Self-diffusion of sodium in sodium chloride and sodium
bromide. J. Chem. Phys. 18, 1231 (1950)
7. R. Dieckmann, Z. Phys. Chem. NF 107, 189–210 (1977)
8. H. Schmalzried, Solid State Reactions (Verlag Chemie, Weinheim, 1981), p. 175
9. R. Dieckmann, H. Schmalzried, Ber. Bunsenges. Phys. Chem. 81, 344–347 (1977)
10. K.-C. Lee, H.-I. Yoo, J. Phys. Chem. Solids 60, 911–927 (1999)
11. J. Crank, The Mathematics of Diffusion, 2nd edn. (Clarendon Press, Oxford, 1975), p. 60
12. C.-R. Song, H.-I. Yoo, Solid State Ion. 120, 141–153 (1999)
Chapter 8
Kinetics of Gas/Solid Reaction:
Diffusion-Controlled Case

8.1 Introduction: Tarnishing Reaction

Have you ever seen shined silverware tarnish to black when it is left for long in
ambient air? Or shined brassware tarnish to green? When a metal M is exposed to an
oxidizing gas X2(g) such as Cl2, Br2, I2, O2, S2, etc., the metal tarnishes losing its
luster. It is simply due to the formation of a tarnishing layer or scale MXν upon the
surface of the metal M; see Fig. 8.1. Let us call this MXν “excide.”1 Such a tarnish-
or scale-formation reaction between M and X2 is called the tarnishing reaction or
oxidation.2
There are two kinds of tarnishing scales. One suppresses or prohibits further
oxidation, thus protecting the underlying metal from further oxidation, and the other
does not, thus letting oxidation continue. The former is said to be protective and the
latter non-protective.
Whether a tarnishing layer is protective or non-protective is very important for the
practical usage of metals. If protective, the metal beneath may survive long, and
otherwise, not. People, thus, wanted to know how to judge whether the “excide”
scale of a given metal will turn protective or non-protective.
In 1923, Pilling and Bedworth [1] in the USA proposed an empirical rule to
judge it, that is, the Pilling-Bedworth ratio (PBR) defined as

1
The oxidation products, MaXb in general, are to be generically represented here as the “excide”
MXν where ν¼b/a.
2
Scaling process and tarnishing process are often distinguished in literatures: the former refers to the
oxidation process at high temperatures where thick oxide layers appear and the latter to that at low
or intermediate temperatures where thin oxide layers form.

© Springer Nature Switzerland AG 2020 285


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1_8
286 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

Fig. 8.1 (a) When a metal


M is exposed to an oxidizing
gas X2(g), (b) a tarnishing 0 ; J 0 0;ν ; J
layer MXν grows upon the
metal M, thus physically
separating the reactants M Δ[
and X2(g) across the
thickness Δx
D E

Volume of the}exide} formed ðMXν Þ MMXν =ρMXν


PBR  ¼ ð8:1Þ
Volume of the metal replaced ðMÞ MM =ρM

where Mk and ρk denote the molar weight and density of k(¼MOν, M), respectively:
If PBR < 1 or the volume of the scale MXν formed upon a metal M is smaller than the
volume of M consumed or replaced beneath, then the scale volume is not enough to
completely fill up the volume occupied previously by the metal M consumed or
replaced, and hence the scale should be discontinuous or porous, thus, non-protec-
tive. If PBR > 1, the scale should be compact and continuous and, hence, protective.
If PBR >> 1 or the scale volume is excessively larger than the metal volume
consumed, however, the scale should spall or chip off rendering the scale non-
protective again.
The PBR values for some selected oxides are:
K2O(0.47); Na2O(0.54); Li2O(0.57);
SrO(0.61); CaO(0.64); BaO(0.67); MgO(0.81); PbO(1.28); CoO(1.78); NiO(1.65);
ZnO(1.58); FeO(1.7); SnO(1.33);
Fe3O4(1.90); Fe2O3(2.14); Al2O3(1.28); Cr2O3(2.07);
SiO2(2.15); TiO2(1.73); ZrO2(1.56).
You may now tell which oxide may be protective and which may be not, but do
not forget that this is just an empirical rule based upon the presumption that the
oxidant X transfers across the scale MXν to the metal M.

8.2 Thermodynamics of Oxidation

What drives the oxidation reaction in any case? Thermodynamics teaches us that
the natural, spontaneous, or irreversible direction of change is in the direction in
which the Gibbs free energy of the reaction system decreases at given temperature
T and pressure P. For the oxidation reaction of M to MXν,

ν
M þ X2 Ð MXν , ð8:2Þ
2

the change in Gibbs free energy or reaction Gibbs free energy, ΔG, is written as
8.2 Thermodynamics of Oxidation 287

ν
ΔG ¼ μMXν  μM  μX2 ð8:3Þ
2

where μk denotes the chemical potential of the reacting species k(¼MXν, M, X2).
If ΔG < 0, thus, the reaction, Eq. (8.2), should proceed from left to right
(oxidation); if ΔG > 0, the reaction recedes from right to left (reduction). In other
words, the reaction free energy ΔG drives the reaction forward or backward to
reduce the system free energy. When ΔG ¼ 0, the reaction neither proceeds nor
recedes, and there sets in the three-phase (M/MXν/X2) equilibrium.
The “excide” MXν exists across a composition range, in principle, from the
M-saturated to the X-saturated, and hence, it should be represented more appro-
priately as M1+δXν to take into account the nonstoichiometry or deviation δ from
the stoichiometric composition (δ¼0). By neglecting the δ-dependence of its
chemical potential as |δ| < <1 normally, however, we may take

μMXν  μoMXν , ð8:4Þ

where μoMXν is the standard (state) chemical potential depending only on T and P.
Also neglecting the solubility of X in M as is the case,

μM  μoM ð8:5Þ

for a “pure” metal M in its natural (Raoultian) reference state.


Assuming the ideal behavior of X2 as is the case,

μX2 ¼ μoX2 þ RT ln aX2 ð8:6Þ

where μoX2 and aX2 (¼PX2 =atm ) denote the standard chemical potential and the
activity of gas X2, respectively.
By substituting Eqs. (8.4), (8.5) and (8.6) into Eq. (8.3), we obtain
 
ν ν ν
ΔG ¼ μoMXν  μoM  μoX2  RT ln aX2  ΔGo  RT ln aX2 ð8:7Þ
2 2 2

where ΔGo defined as such is the standard reaction free energy of the oxidation
reaction, Eq. (8.2).
When “pure” M  is in equilibrium
 with MXν, the equilibrium chemical potential
of X(¼X2/2), μeqX ¼ μ eq
X2 =2 , is, due to Eq. (8.3), such that

ΔG ¼ μoMXν  νμeq
X  μM ¼ 0:
o
ð8:8Þ

Thus,
288 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

   
1 eq 1 o
μeq ¼ μ ¼ μ  μ o
M , ð8:8aÞ
X 2 X2 ν MXν

or in terms of the gas X2-activity,


 
2ΔGo
aeq ¼ exp ð8:9Þ
X2 νRT

due to Eq. (8.7). For ΔG < 0 or for the oxidation reaction to proceed, thus,

μX > μeq eq
X or aX2 > aX2 ð8:10Þ

by necessity. Otherwise, already existing MXν, if any, would reduce back to M.

8.3 Kinetic Steps and Rate Laws Observed

The thermodynamic necessity for oxidation of M, Eq. (8.10), is now satisfied.


What kinds of kinetic steps are involved eventually to form the reaction product
MXν? The kinetic necessity for the reaction to proceed is that the reactants M and X
should meet physically each other; see Fig. 8.1b. To this end, X should transfer from
the gas X2 to the M/MXν interface and/or M to the MXν/X2 interface.
For X to meet M at the internal M/MXν interface and to react there, the elemen-
tary kinetic steps involved may be:
(i) Transport of gas X2 to the surface of MXν;
(ii) Adsorption of gas X2 there;
(iii) Dissociation into X (X2!2X);
(iv) Incorporation of X into MXν (X  z2 e ! Xz2 );
(v) Diffusion of X(¼Xz2 þ z2 e ) through MXν to the MXν/M interface;
(vi) Reaction with M there to MXν (νX + M ! MXν).
For M to meet X at the external MXν/X2 interface and to react there, on the other
hand, the likely kinetic steps may be:
(i) M incorporation into MXν (M ! Mz1 þ z1 e );
(ii) M(¼Mz1 þ z1 e ) diffusion through MXν to the MXν/X2 interface;
(iii) Reaction with X adsorbed there to MXν (M + νX ! MXν).
Note that all the kinetic steps in each case occur in series. You may immediately
imagine that there can be a variety of kinetic rate laws depending on which step is
rate-determining in what circumstance. It is indeed so. There have been observed
typically the following rate laws for the oxidation reaction. Letting Δx denote the
thickness of MXν upon M,
8.4 Parabolic Rate Law and Wagner’s Theory 289

(i) Linear rate law: Δx / t


(ii) Parabolic rate law: Δx / t1/2
(iii) Cubic rate law: Δx / t1/3
(iv) Logarithmic rate law: Δx / ln t
Can you guess which kinetic step results in which rate law? Recollecting Chap. 3,
you may easily guess that the linear rate law may be the case for a phase-boundary
reaction being rate-determining and, recollecting Chap. 1, the parabolic rate law for
diffusion being rate-determining. What about the cubic and logarithmic laws? The
cubic rate law may arise when the high-diffusivity paths in the polycrystalline
“excide” are somehow progressively blocked as the “excide” grows [2] or when the
M/MXν-interface coherence strain retards the diffusion [3]. When the “excide” is
thin enough in the very earlier stage of oxidation in particular, electron tunneling
and interfacial space charge may affect the reaction kinetics, possibly leading to the
cubic or logarithmic laws. For the mechanistic interpretations of the kinetic laws,
see, e.g., Hauffe [4] and Kofstad [5]. Particularly for the space charge effect, see
Wagner [6]. In the following, we will learn the diffusion-controlled, parabolic rate
law in detail again from the old wisdom [7].3

8.4 Parabolic Rate Law and Wagner’s Theory

It has earlier been observed that when the “excide” layer is protective (PBR > 1) and
thick (Δx > ~30 nm) [8], the overall kinetics always follows the parabolic rate law:

dðΔxÞ kP
¼ or ðΔxÞ2 ¼ 2kP t: ð8:11Þ
dt Δx

It is normally the case for high-temperature oxidation. Here, the rate law
constant kP is variously called the practical tarnishing rate law constant or, to
honor those who first recognized this kinetics, G. Tamman [9] and several years
later N.B. Pilling and R.E. Bedworth [1], the Tamman constant, or the Pilling and
Bedworth constant.
Sometimes, this law is represented in terms of mass gain Δm instead of thickness
Δx, as

dðΔmÞ k0p
¼ or ðΔmÞ2 ¼ 2k0P t, ð8:12Þ
dt Δm

for the mass gain Δm is normally (actually more easily) measured at high temper-
atures in particular. Noting that

3
For the English translation of the main part of the former of Ref. [7], see Appendix II at the end of
this book.
290 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

Fig. 8.2 (a) Spatial ′ ″


configuration of the (a) M M1+δXν X2(g)
oxidation system,
MjM1+δXνjX2(g), where the μ o
MX ν

two interfaces are denoted 1


μ = μ′M
o
μM μX μ′′X = μ X2 (g)
as 0 and 00 in turn;
M
(b) 2
(b) schematic not-to-scale
distributions of ′
 μ eq
X = μX μ′′M
μMXν  μoMXν , μM, and μX,
where ∇μMXν ¼ δ′
∇μM þ ν∇μX  0; δ
(c)
(c) schematic distribution of
the metal excess δ; (d) flows δ′′
of the chemical components JM
M and X driven by
(d)
∇μX(¼ ∇ μM/ν) imposed JX
in Part (b); (e) flows of the
charged components, MZ1 þ , J Mz1+
XZ2  , and e; (f) oxidation
reactions at the two (e) J e−
interfaces, where the small
arrows beneath the charged J X z2 −
components indicate their
directions of moving in (f) Δx
accord with those in Part (e).
Note that νz2 ¼ z1 M + νX z2 −
M z + + z1e − + νX 2 / 2
1

= MX ν + z 2 e − = MX ν

Δm ¼ ρMXν AΔx ð8:13Þ

with “A” being the exposed surface area of the oxidizing specimen, you may easily
find the interrelationship between the parabolic rate law constants kP and kP0 .
How come the parabolic rate law? Its inner working was elegantly figured out
by C. Wagner in his celebrated classic paper [7] in 1933 which may be regarded as
having opened the door to a new world of solid state ionics and solid state
electrochemistry. This theory starts with the following assumptions:
(i) The scale is dense (PBR > 1) and homogeneous, thus, a diffusion barrier.
(ii) Only a single oxidation product forms.
(iii) Local equilibrium prevails at the metal/scale and scale/gas interfaces and
throughout the thickness of the scale.
The thermodynamic and kinetic situation is schematically illustrated in Fig. 8.2,
where the internal interface (M|MXν) and external interface (MXν|X2(g)) are indi-
cated by 0 and 00 , respectively. Assumptions (i) and (iii) ensure the overall kinetics to
be diffusion-controlled, and Assumption (iii) ensures μ0M ¼ μoM and μ0X ¼ μeq X at the
8.4 Parabolic Rate Law and Wagner’s Theory 291

internal interface (0 ) and μ00X ¼ μX2 ðgÞ =2 at the external interface (00 ). Otherwise,4 e.g.,
μ00X < μX2 ðgÞ =2.
It is noted that, due to Eqs. (8.3), (8.7), and (8.8a),

1 1
μ00X  μ0X ¼ μX2ðgÞ  μeq
X ¼  ν ΔG: ð8:14Þ
2

For a single phase M1+δXν, the Gibbs-Duhem equation stipulates that, noting
ν ¼ z1/|z2|,

∇μM ∇μX
∇μM þ ν∇μX  0 or  ð8:15Þ
z1 z2

insofar as |δ| << 1 at given T and P. This means that upon integrating across MXν in
Fig. 8.2,

1  00 
μ00X  μ0X ¼  μ  μ0M : ð8:16Þ
ν M

You have learned from Chap. 7 that the fluxes of the charged components
fMz1 þ , Xz2  , e g are written respectively as

σ1
M z1 þ : J 1 ¼  ∇η1 ð8:17aÞ
ðz1 FÞ2
σ2
Xz2  : J2 ¼  ∇η2 ð8:17bÞ
ðz2 FÞ2
σ
e : J3 ¼  32 ∇η3 ð8:17cÞ
F

under the isothermal and isobaric condition, where zk, σk, and ηk denote the formal
valence, partial conductivity, and electrochemical potential of the charged com-
ponent k(¼1,2,3), respectively. Remember that z2 < 0, z3 ¼ 1 and ν¼z1/|z2| for the
system MXν.
The local ionization equilibria throughout the “excide” stipulate:

∇η1 ¼ ∇μM  z1 ∇η3 for Mz1 Ð M  z1 e and ð8:18aÞ


z2  
∇η2 ¼ ∇μX  z2 ∇η3 for X Ð X  z2 e ð8:18bÞ

Due to the constraint of charge neutrality or open-circuit,

4
Confer Eq. (6.48) and what follows.
292 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

X
3
zk FJk ¼ 0: ð8:19Þ
k¼1

Substituting Eq. (8.18) into Eq. (8.17) and solving for ∇η3 via Eq. (8.19), you
obtain

t1 t
∇η3 ¼ ∇μM þ 2 ∇μX ð8:20aÞ
z1 z2

which may be rewritten due to the Gibbs-Duhem equation, Eq. (8.15), as

tion t
∇η3 ¼ ∇μM ¼ ion ∇μX : ð8:20bÞ
z1 z2

Here, tion ¼ t1 + t2 is the ionic transference number. This is the very origin of the
open-circuit voltage U(¼Δη3/F) which you can measure across the entire system
in Fig. 8.2a by using the electronic probes,5 say, Pt-wires.
Now, substitute Eq. (8.20b) into Eq. (8.18) and you will get

∇η1 ¼ tel ∇μM ; ∇η2 ¼ tel ∇μX ð8:21Þ

where tel(¼1tion) is the electronic transference number of the excide.


Substitution of Eq. (8.21) into Eqs. (8.17a) and (8.17b), then, yields the (chem-
ical) component fluxes in terms of the chemical driving force ∇μX(¼∇μM/ν)
applied as

σ1 tel
JM ¼ J1 ¼  ∇μM ; ð8:22aÞ
ð z 1 FÞ 2
σ2 tel
JX ¼ J2 ¼  ∇μX : ð8:22bÞ
ðz2 FÞ2

According to Eq. (8.2), one unit of M transferred to the external surface (00 ) yields
one unit of MXν outward into X2(g); one unit of X transferred to the internal surface
(0 ) yields 1/ν unit of MXν inward into M. The total formation rate of MXν per unit
area JMXν may, thus, be written as

5
An electronic probe, e.g., an electronic conductor metal, communicates with a mixed ionic
electronic conductor system MXν by exchanging electrons (e) to achieve the electrochemical
equilibrium with respect to electron exchange at the point of contact, i.e., η3(probe)¼η3(MXν).
Consequently, any pair of the electronic probes in contact with MXν measures the electrochemical
potential difference of electrons Δη3 between them. Likewise, an ionic probe can be equilibrated
with the system with respect to the exchange of common mobile ions.
8.4 Parabolic Rate Law and Wagner’s Theory 293

 
1  1
JMXν ¼ jJM j þ  JX  ¼ JM  JX : ð8:23Þ
ν ν

By substituting Eq. (8.22) into Eq. (8.23), replacing ∇μM ¼ ν∇μX due to Gibbs-
Duhem equation, Eq. (8.15), and ν ¼ z1/|z2| by definition, we obtain

ðσ1 þ σ2 Þtel
JMXν ¼ ∇μX : ð8:24Þ
z1 jz2 jF2

Can you recognize that (σ1+σ2)tel¼σionσel/(σion+σel) ¼ tionσel? This combination


of the partial electronic and ionic conductivities is often called the ambipolar
conductivity, which we met earlier in Chap 7. The diffusion-controlled oxidation
rate JMXν is determined, like the chemical diffusion rate Jδ in Chap. 7, by the faster of
the two kinds of mobile ions (Mz1 , Xz2 ) and the slower of the faster ions and
electrons.
Let us now integrate Eq. (8.24) across the excide thickness Δx in Fig. 8.2 to get

00 a00
ZΔx ZμX ZX2
σion tel RT
JMXν dx ¼ d μX ¼ σion tel d ln aX2 ð8:25Þ
z1 jz2 jF 2
2z1 jz2 jF2
0 μ0X a0X
2

with the use of Eq. (8.6) for μX(¼μX2 =2) in the last equality. Then what?
C. Wagner [7] says, “By the assumption of quasi-steady-state growth, the rate
of transport of material (JMXν ) through the scale is independent of distance (x) along
the direction of diffusion.” Equation (8.25) can then be immediately integrated and
rearranged as
0 1
a00
ZX2
1 B
B RT
C
JMXν ¼ σion tel d ln aX2C
A: ð8:26Þ
Δx @2z1 jz2 jF2
a0X
2

Now, due to the mass conservation,

dðΔxÞ
JMXν ¼ ð8:27Þ
Vm dt

where Vm is the molar volume of MXν.


Therefore, the parabolic rate law constant kp takes the form as
294 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

a00
ZX2
dðΔxÞ kP Vm RT
¼ ; k ¼ σion tel d ln aX2 ð8:28Þ
dt Δx P 2z1 jz2 jF2
a0X
2

What a breathtaking beauty! The parabolic rate law constant kP has turned out
to comprise the so-called ambipolar conductivity, σiontel, of the excide in concern,
which is also the basic ingredient of the chemical diffusivity D e of the excide (see
Eq. (7.88)), thus suggesting an intimate interrelationship as you will see shortly. By
using the Nernst-Einstein equation, Eq. (7.15), relating the partial conductivity to
self-diffusivity, you may rewrite kp equivalently in terms of the self-diffusivities of
the three types of charged components.
Integration of Eq. (8.28) with the initial condition Δx ¼ 0 at t ¼ 0 leads to the
famous parabolic rate law equation, Eq. (8.11).

8.5 Remark on the Quasi-steady State Assumption

The quasi-steady state assumption may be regarded as the most critical one among
others because it saves JMXν out of the integral with respect to the distance x in
Eq. (8.25), thus resulting in the parabolic rate law. How can it be justified and what
does this assumption connote?
The present thermodynamic situation in Fig. 8.2 is exactly that for the chemical
diffusion in Chap. 7. Let us consider the nonstoichiometry variation across the
excide M1+δXν. Its nonstoichiometry δ may be written as

δ 1
¼ CM  CX ð8:29Þ
Vm ν

because CM ¼ C1 ¼ (1 + δ)/Vm and CX ¼ C2 ¼ ν/Vm. It follows that

∂ðδ=Vm Þ ∂CM 1 ∂CX  


1
¼  ¼ ∇ JM  JX  ∇Jδ ð8:30Þ
∂t ∂t ν ∂t ν

as the chemical components M and X are all conserved. In comparison with


Eq. (8.23), you can immediately recognize that

1
JMXν ¼ JM  JX ¼ Jδ ð8:31Þ
ν

Distance independence of JMXν ð¼ Jδ Þ means that


8.6 Telltale Marker Experiment 295

Fig. 8.3 Shift of the inert


markers (xxxxx) with X2(g)
(a) M
oxidation depending on the
relative magnitude of JM and
JX: (a) initial position upon
M; (b) when JM >> |JX|; (c)
when JM << |JX|; (d) when
JM/|JX/ν| ¼ a/b where (b) M MXν X2(g)
a + b ¼ Δx

(c) M MXν X2(g)

(d) M MXν X2(g)

b a

∂ðδ=Vm Þ
∇JMXν ¼ ¼ 0: ð8:32Þ
∂t

As long as |δ| << 1, this will be approximately valid, and hence, the steady state-
growth assumption is justified. With the growth of the excide scale, only the spatial
extent (Δx) of a fixed δ-range, from the M-saturated(δ0 ) to the X-saturated(δ00 )
commensurate with the μ00X applied, is expanded; see Fig. 8.2c.

8.6 Telltale Marker Experiment

Suppose that we have no idea about the relative mobility of M and X in MXν. Could
we know afterward whether the excide growth to a thickness Δx was due to either JM
or JX or both? Here is a telltale simple experiment.
If you put inert markers, say, fine Pt-particles, upon the surface of M before
oxidation (Fig. 8.3a), you may have the three possibilities concerning the markers’
position afterward as illustrated in Figs. 8.3b–d. Depending on the positions of the
inert markers afterward, you may tell the relative contribution of JM and JX to the
overall thickness Δx. Knowing that the mobile components are moving relative to
the inert markers which are sitting on the local lattice reference frame, you may
figure out how they are depending on the relative magnitude of JM and JX as
illustrated.
296 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

8.7 An Example: Oxidation of Co to CoO

Equation (8.28) indicates that when you know the functional dependence of the
partial electronic and ionic conductivities σel and σion upon the oxidizing gas activity
aX2 ð¼ PX2 =atmÞ or partial pressure, PX2 , you can evaluate the parabolic rate law
constant, kP, and such information will be available from the defect chemistry of
the system MXν as we have learned in Chap. 7.
Let us here evaluate the parabolic rate law constant kP of Co1-δO, for example,
which is one of the oxides whose defect structure is the most studied. Co1-δO is
known to be protective (PBR ¼ 1.78), thence its growth kinetics upon Co follows
the parabolic rate law. We like to evaluate the parabolic rate law constant kp as
given in Eq. (8.28) by using the basic mass/charge transport properties of Co1-δO.
We know that Co1-δO is a p-type semiconductor across its entire stability
range, and the self-diffusivity of Co2+ cations is more than four orders of magnitude
larger than that of O2 anions, that is,

σel >> σion ; σ1 >> σ2 : ð8:33Þ

This means that

tel  1 ; σion  σ1 : ð8:33aÞ

The oxide is hyperstoichiometric or metal deficit across its entire stability


range. Referring back to the defect structure

of Co1-δO in Chap. 7, the majority
type disorder is believed to be V00Co , h particularly as the oxygen activity
approaches the Co/CoO boundary or the defect concentrations grow smaller. The
external equilibrium condition (Eq. (7.55))

1
O2ðgÞ ¼ OxO þ V00Co þ 2h ; V00Co ½h 2 ¼ KOx aO2
1=2
ð8:34Þ
2

and the charge neutrality condition (Eq. (7.56)) yield

00 ½h  1=3 þ1=6


VCo ¼ ¼ 22=3 KOx aO2 , ð8:35Þ
2

where the external equilibrium constant KOx is a function of temperature only. The
ionic conductivity, which is essentially equal to the cationic conductivity in the
present case, may then be written as

σion  σ1 ¼ V00Co  2F  uV00Co ¼ σi,o aO2
1=6
ð8:36Þ

where uV00Co is the electrochemical mobility of V00Co and

1=3
σi,o  21=3 FuV00Co KOx : ð8:36aÞ
8.7 An Example: Oxidation of Co to CoO 297

This corresponds to the partial cation conductivity at aO2 ¼ 1, of which, if you want,
uV00Co may be replaced by the vacancy diffusivity DV by using the Nernst-Einstein
equation.
It is now enough for Co1-δO. Let us substitute Eqs. (8.33a) and (8.36), together
with the numerical values z1 ¼ |z2| ¼ 2, into Eq. (8.28) to have

a00
ZX2
Vm RT 3Vm RTσi,o  00 1=6 
a O2  a0 O2 :
1=6 1=6
kP ¼ σi,o aO2 d ln aO2 ¼ ð8:37Þ
8F2 4F 2
a0X
2

 
If a00O2 >> a0O2 ¼ aeq
O2 as is the case for the normal oxidation in ambient air or so,
this equation can better be approximated as

3Vm RTσi,o 00 1=6


kP  a O2 ð8:37aÞ
4F2

or, in terms of the cation self-diffusivity at a00O2 , DCo ða00O2 Þ,

kP  3DCo ða00O2 Þ ð8:37bÞ

due to the local lattice conservation relation

D
V00Co DV ¼ Co : ð8:38Þ
Vm

You are strongly recommended to derive Eq. (8.37b) for yourself by using the
Nernst-Einstein equation.
The story up to now is based on the assumption that theh majority
i disorder type is
00 
00  0 00
VCo , h or 2 VCo ¼ ½h  in the oxygen activity range aO2 , aO2 . It is, however,
known that as oxygen activity increases or the defect

concentration
increases, the
majority disorder type appears to shift to V0Co , h or V0Co  ½h .6 If this were the

6
In this case, the effective charge value (¼1) of V0Co is different from the actual or formal charge
value (¼2) of Co2+. Nevertheless, C1D1 ¼ CVDV for local lattice conservation, which leads to the
Nernst-Einstein equation, σ1  C1 ð2FÞ2 RT D1
¼ CV ð2FÞ2 DRTV  σV . This indicates that the effective
charge units of VCo should be 2 (VCo ) in disagreement with the reality, V0Co ! To our surprise, the
00

conventional Nernst-Einstein equation is failing! This is simply a consequence of the Onsager-type


electron-ion cross effect in a mixed ionic electronic conductor. Interested readers are referred to:
C. Wagner, “Equations for transport in solid oxides and sulfides of transition metals,” Prog. Solid
State Chem., 10 (1975) 3–16; H.-I. Yoo, H. Schmalzried, M. Martin, J. Janek, “Cross effect
between electronic and ionic flows in semiconducting transition metal oxides,” Z. Phys. Chem.
NF 168 (1990) 129–142; and H.-I. Yoo and H.-S. Kim, “Complete representation of isothermal
mass and charge transport properties of mixed ionic electronic conductors,” Solid State Ionics
225 (2012) 166–171.
298 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

h i
case in a0O2 , a00O2 , the redox reaction and its equilibrium constant K0Ox would then
be written as Eq. (7.59) or

1 V0Co ½h 
O ¼ OxO þ V0Co þ h ; ¼ K0Ox ; ð8:39Þ
2 2ðgÞ 1=2
aO2

and hence,

½h   V0Co ¼ K0 Ox aO2 :
1=2 1=4
ð8:39aÞ

It follows that

σion  σ1 ¼ V0Co  F  uV0Co ¼ σ0i,o aO2
1=4
ð8:40Þ

with

σ0i,o  FuV0Co K0 Ox :
1=2
ð8:40aÞ

Accordingly, the parabolic rate law constant is evaluated as

a00
ZX2
Vm RT Vm RTσ0i,o  00 1=4 
σ0i,o aO2 d ln aO2 ¼ 0 1=4
1=4
kP ¼ a O  a O : ð8:41Þ
8F2 2F2 2 2

a0X
2

You may then approximate this equation similarly to Eq. (8.37a) and rewrite in
terms of self-diffusivity by using the approximate Nernst-Einstein equation.

8.8 Internal Oxidation

When a base metal, e.g., Ag, containing a small amount of less noble metal, e.g., Cu,
is exposed to oxygen, something different from above occurs: The oxides of the less
noble Cu form within the alloy bulk, and this oxide precipitation zone grows
inward away from the metal surface, as illustrated in Fig. 8.4. This sort of oxidation
is called the internal oxidation in contradistinction with what we learned above, the
external oxidation.
8.8 Internal Oxidation 299

O2 Ag + δ Cu

0 ξ x

Fig. 8.4 Schematic of internal oxidation: Oxidation of a noble metal, e.g., Ag, containing a small
amount(δ) of less noble metal, e.g., Cu (Ag+δCu). The oxides of the less noble metal precipitate
within the bulk, and the oxide precipitation zone, ξ, grows inward away from the alloy surface

Fig. 8.5 Oxide- 100


o
precipitation-zone thickness 1- 0.42% Mn, 750 C
squared (abscissa) vs. time o
1 2- 0.103% Si, 750 C
(ordinate) of Cu-alloys
80 2 o
3- 0.033% Mn, 750 C
containing a small amount o
of less noble Mn or Si as 4- 0.103% Si, 875 C
o
indicated within the figure. 5- 1.55% Mn, 1000 C
o
60 6- 0.103% Mn, 1000 C
time/ hours

(From Rhines [10])

40
3

20
4
5 6
0
0.000 0.002 0.004 0.006 0.008 0.010 0.012
2 2
thickness / cm

It is observed that the depth of the oxide precipitation zone, ξ, increases


parabolically with time or

ξ / t1=2 ; ð8:42Þ

see Fig. 8.5.


How can such thing happen? You can imagine that, for this thing to happen,
oxygen should diffuse into the alloy to meet and react with the less noble component
inside the alloy. By necessity, then, the base metal matrix, say, Cu in Fig. 8.5, should
have an appreciable solubility of oxygen, and its diffusivity DO in the matrix should
be no smaller than that of less noble component, say, Mn or Si in Fig. 8.5, D2.
Otherwise, the solute would, instead, diffuse out to the surface and react with oxygen
there oppositely to the present internal oxidation, namely, external oxidation
occurs. Furthermore, the minority alloying element, Mn or Si, should have much
higher affinity for oxygen than the host metal, Cu. Otherwise, both would compete
against each other for oxygen.
300 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

Fig. 8.6 (a) The alloy (A,


B) with oxide BOν
(a) (A,B)
precipitated internally to a
zone thickness ξ(t) for a time
duration t. (b)
Corresponding (b) C
concentration distributions JO JB
of O and the less noble CO(s)
component B, CO, and CB, CB
respectively. The oxygen CB(b)
solubility limit in the alloy is CO
CO(s), and the bulk
concentration of B is CB(b).
0 ξ(t) x
(From Schmalzried [13])

Let us now explore the kinetics of internal oxidation, Eq. (8.42), on the basis of
this scenario. For the original treatment, you are referred to C. Wagner [11] for
metallic alloys, and for its extension to oxide alloys, to H. Schmalzried [12]
(internal oxidation to a higher oxide, e.g., B2O3, in a lower oxide matrix, e.g., BO).
Here is a dilute metallic alloy (A,B) with B being the solute. We assume that B is
much less noble compared to the solvent A, that is, the reaction free energies of
oxidation or oxygen affinities of A and B, ΔGAO and ΔGBOν , respectively, are such
that

j ΔGBOν j>>j ΔGAO j : ð8:43Þ

Wherever and whenever oxygen is available, therefore, BOν forms preferentially. It


is also supposed that the oxygen solubility in the base metal A, CO(S) in equilibrium
with oxygen activity in the surrounding, and the diffusivities of oxygen and the less
noble B in A, DO, and DB are given.
With the formation of oxide BOν internally, the concentration distributions of O
and B, CO, and CB, respectively, within the alloy (A,B) may be as illustrated in
Fig. 8.6. Thus, O diffuses inward and B outward down the respective concentration
gradients ∇CO and ∇CB. Then, the precipitate zone front x ¼ ξ(t) is expected to fall
where the oxygen flux JO, coming from the surface to the immediate left of the front
(x ¼ ξ), is met by the B-flux JB, coming from the bulk to the immediate right of the
front (x ¼ ξ+), in such a way that

JO jx¼ξ ¼ νJB jx¼ξþ ð8:44Þ

or, due to Fick’s first law,


8.8 Internal Oxidation 301

   
∂CO ∂CO
lim DO ¼ lim νDB ð8:44aÞ
ε!0 ∂x x¼ξε ε!0 ∂x x¼ξþε

for ε > 0 definite. Here, of course, we have assumed that the chemical reaction
B + νO ! BOν is fast enough to consume all the incoming reactants there so that
CO ¼ 0 ¼ CB at x ¼ ξ(t) in the matrix all the way; see Fig. 8.6b. You see, this is
pretty much like the moving boundary problem that you have learned in Chap. 1,
indeed.
Referring back to the observation, Eq. (8.42), let us tentatively assume that
pffiffiffiffiffiffiffiffi
ξð t Þ ¼ β  2 DO t ð8:45Þ

where β is a dimensionless parameter to be determined.


In order to evaluate the concentration gradients at the zone front x ¼
h i

ξ  lim ðξ  εÞ in Eq. (8.44a), we need to know the transient solutions to
ε!1
Fick’s second law for each of the diffusing reactants CO(x,t) and CB(x,t) with the
boundary conditions

CO ð0, tÞ ¼ COðsÞ ; CO ðx  ξ, tÞ ¼ 0 ð8:46aÞ


CB ð1, tÞ ¼ CBðbÞ ; CB ðx ξ, tÞ ¼ 0; ð8:46bÞ

see Fig. 8.6b.


The particular solutions may take the two-parameter error functional form or
 
x
Ck ¼ Ak þ Bk erf pffiffiffiffiffiffiffi ; ðk ¼ O, BÞ: ð8:47Þ
2 Dk t

Solving for the two parameters, Ak and Bk, by using the two boundary conditions
Eq. (8.46a) for k ¼ O and Eq. (8.46b) for k ¼ B, in association with Eq. (8.45), you
may obtain
  pffiffiffiffiffiffiffiffi
erf x=2 DO t
CO ðx, tÞ ¼ COðsÞ 1 ; ð8:48aÞ
erf ðβÞ

"  pffiffiffiffiffiffiffiffi#
erfc x=2 DB t
CB ðx, tÞ ¼ CBðbÞ 1  pffiffiffiffi ð8:48bÞ
erfc β ψ

where
302 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

DO
ψ : ð8:49Þ
DB

Substituting the solutions, Eq. (8.48) into Eq. (8.44a), and rearranging, you end
up with the equation for β involving the known system parameters:

eγ erf ðβÞ
2
COðsÞ
¼   ð8:50Þ
νC2ðbÞ ψ1=2 eγ2 φ erfc βψ1=2

which allows you to evaluate the dimensionless parameter, β, in Eq. (8.45). You are
urged to derive Eq. (8.48) through Eq. (8.50) for yourself as an exercise to remind
yourself of the moving boundary problem in Chap. 1.
Summing up, the oxide precipitation zone is moving inward parabolically with
time as Eq. (8.45) with β such as in Eq. (8.50). Let us here consider two possible
extreme cases by invoking the first-order approximations of the error function that

ez 1
2
2
erf ðzÞ  pffiffiffi z for z << 1; erfcðzÞ  pffiffiffi  for z >> 1: ð8:51Þ
π π z

pffiffiffiffi
(i) When β << 1 and β ψ << 1,

pffiffiffi  
π COðsÞ pffiffiffiffi
β ψ: ð8:52Þ
2 νCBðbÞ

Thus,
!
π1=2 COðsÞ DO
ξðtÞ ¼ 1=2
 t1=2 : ð8:53Þ
νCBðbÞ DB

pffiffiffiffi
(ii) When β << 1 and β ψ >> 1,

 1=2
COðsÞ
β ð8:54Þ
2νCBðbÞ

and hence, Eq. (8.45) may be rewritten as


 
2COðsÞ DO 1=2 1=2
ξðtÞ  t : ð8:55Þ
νCBðbÞ
Problems 303

Under the present condition, the advancement of the precipitation zone into the
interior of the alloy is determined only by oxygen diffusion. The present condition
together with Eq. (8.52) indicates that

DB COðsÞ
<< << 1, ð8:56Þ
DO 2νCBðbÞ

which is normally the case for internal oxidation problems.


In conclusion, for a binary alloy (A,B) to oxidize internally, the appreciably less
noble component B present in relatively small amount and the matrix of the nobler
metal A must display sufficient oxygen solubility (CO(s)) and mobility (DO) so that
CODO > CBDB.

Problems

1. The oxidation rate of cobalt to CoO at 1148


C is shown in the figure below as a
function of oxygen potential.

⎡1 ⎛ Δm ⎞ 2 ⎤
log ⎢ ⎜ ⎟ ⎥
⎣⎢ t ⎝ A ⎠ ⎦⎥ 4

1
0.5

0.01 0.1 1.0


PO2 / atm
(Unit of argument of log: g2cm-4hr-1)

(i) Evaluate the PO2 dependence of the parabolic rate law constant kp by paying
special attention to the scales of x-axis and y-axis.
(ii) Explain this effect of oxygen activity on the oxidation rate constant in the
light of the defect structure of Co1-δO, which is a metal-deficit p-type
semiconductor.
2. A specimen of metal M exposed to an atmosphere of pure oxygen will form an
oxide with a composition MO on the surface. The oxide layer is pure, continuous,
and adherent to the surface of the metal. Marker experiments show that the oxide
layer grows by diffusion of the anion. Other experiments show the transference
numbers to be tM2+~0, tO2 ¼ 0.2, and tel ¼ 0.8.
304 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

What is the position of the interface between the oxide film and the base metal
relative to the original metal interface after 4 hours at 1000
C?
Data:
For the reaction
M(s) + 12 O2(g) ¼ MO(s), ΔGo/Jmol1 ¼ 235,000 + 85 T

For MO at 1000
C
Electrical conductivity of MO, σ ¼ 0.12 S/cm
Self-diffusivity of oxide, DO ¼ 107 cm2/s
Self-diffusivity of metal, DM ¼ 1010 cm2/s
Density of metal, ρm ¼ 7 g/cm3
Density of MO, ρMO ¼ 5 g/cm3
Atomic mass of metal ¼ 60 g/mol
Atomic mass of oxygen ¼ 16 g/mol
3. The practical parabolic rate constant k such that (Δm)2 ¼ 2kt has been determined
by a weight change with a thermobalance at 1000 K. The surface area of the plane
metal A was 2 1 cm2, and the weight change after one hour was 20 mg.
Oxidation took place by formation of only AO in an atmosphere of 99% N2 and
1% O2 (Ptotal ¼ 1 bar), and the inert markers were found at A/AO interface. It is
known that AO is metal-deficient across its entire stability range and, hence, a
p-type semiconductor.
(a) Calculate the parabolic rate law constant k.
(b) Write the parabolic rate law constant k in terms of the ionic conductivity of
AO, and calculate the mean ionic conductivity in the oxygen partial pressure
range examined.
(c) Is this conductivity due to cations or anions? Justify your answer.
(d) Calculate the corresponding self-diffusivity.
Data:
Standard free energy of formation of AO, ΔGoAO ¼ 200 kJ/mol
Molar volume of AO, Vm ¼ 25 cm3/mol
Molar weight of AO, MAO ¼ 75 g/mol
4. Thin sheet of metal M is exposed on both surfaces to air at 600
C.
(a) Do you expect the thickness of the MO scale at this temperature to be linear or
parabolic with time?
Problems 305

(b) Assuming the rate law which you have identified in part (a), derive the scale
thickness Δx as a function of time to evaluate the rate law constant, klinear or
kparabolic.
(c) It has been experimentally found that the rate law constant turns out to be
Arrhenian or

 
180 kJ=mol
k / exp  :
RT

Explain any discrepancy with your estimate in part (b).


(d) With such a large enthalpy of reaction (see the data below), it is unreasonable
to assume that the thin sheet is able to dissipate the heat fast enough to
maintain isothermal oxidation. Write the differential equation and the bound-
ary conditions which express the oxidation heat balance for this geometry.
Data: M O MO
Atomic mass (g/mol) 24.31 16.00 40.31
Density (g/cm3) 1.74 – 3.60
D ¼ A exp (Q/RT)
1 1
A/ m2 s Q/kJmol
DM2+ in MO 2.5 103 331
DO2 in MO 2.5 102 261
DM in M 1.5 104 136
ΔGoformation(MO) ¼ 601,000 + 124 T J/mol for 298 < T/K < 922
M melts at 922 K.
MO melts at 2915 K.
M is hcp with lattice constants 3.21 1010 m and 5.21 1010 m.
MO has the NaCl-type crystal structure with lattice constant 4.2 1010 m.

5. The practical parabolic rate constant has been determined by a weight change
with a thermobalance at 1000 K. The surface area of the plane metal A was 2 1
cm2, and the weight change after one hour was 20 mg. The oxidation took place
by formation of only AO in an atmosphere of 99% N2 and 1% O2 (Ptotal ¼ 1 bar).
AO is a semiconducting oxide. Calculate the (component) diffusion coefficient
DA and the tracer diffusion coefficient DA by assuming the following data to be
given:
AO crystallizing in the NaCl structure; ΔGAO
0
¼ 200 kJ/mol; Vm ¼ 25 cm3/
mol; MAO ¼ 75 g/mol.
306 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

6. Growth of AgBr on Ag was observed in a bromine gas atmosphere of


PBr2 ¼ 2.3 104 Pa at 400
C as shown in Fig. 8.7a: (I) when short-circuited
between Ag and the Pt-gauze-covered AgBr surface (see Fig. 8.7b);
(II) conventional tarnishing experiment. It has been reported that the partial
conductivities of AgBr vary against the activity of Ag as shown in Fig. 8.7c,
e.g., at 277
C.
(a) Calculate the driving forces for the tarnishing reaction at 400
C for the two
cases, respectively.
(b) Evaluate the parabolic rate law constants kp such that Δm2 ¼ 2kpt (where Δm
being weight increase) for the two cases, respectively, at 400
C.
(c) Explain quantitatively why Case (I) has orders of magnitude higher reaction
rate?
(d) Evaluate the ionic conductivity and electronic conductivity of AgBr, respec-
tively, at 400
C.
The following data may be helpful:
The standard free energy: ΔGo ¼ 79.366 kJ at 400
C for the reaction,
Ag + Br2/2(g) ¼ AgBr.
Density of AgBr: 6.473 g/cm3
Molar weight of AgBr: 187.77

a 2.0
(weight increase)2×10-5/ mg2⋅cm-4

I
1.5

1.0

0.5

normal bromination II
0.0
0 20 40 60 80 100 120 140 160 180 200
time/ min

Fig. 8.7 (a) Parabolic growth of a AgBr product layer on Ag in Br2-gas (T ¼ 400
C,
PBr2 ¼ 2.3 104 Pa). (I) Short circuit between the Ag and a Pt-gauze on the AgBr surface.
(II) Normal growth of a compact AgBr product. (From J. H. Eriksen and K. Hauffe, Z. Phys. Chem.
NF59 (1968) 326, quoted from Schmalzried [13]). (b) Schematic diagram of the experiment. (c) The
electrical partial conductivities σk for ions (k ¼ ion) and for electronic charge carriers (k¼h,e0) in
AgBr at T ¼ 277
C against silver activity aAg. (From Wagner [14])
Problems 307

b c
-2
σion
-4

log (σk/ Ω-1cm-1)


-6 σh.

-8

-10
σe′
-12
-2 -4 -6 -8
log aAg

Fig. 8.7 (continued)

7. C. Wagner carried out the following tarnishing experiment with the system of
silver (Ag) as shown in Fig. 8.8 (Z. Phys. Chem. B21 (1933) 25): Against the
lower open end of a glass tube, two cylinders, each consisting of compressed
polycrystalline Ag2S with cross section of 0.12 cm2 and height of 0.6 cm, were in
close contact followed by a piece of silver sheet. The tube was filled with molten
sulfur. After one hour at 200
C, he found changes in mass of each of the
cylinders as:
Ag: Δm ¼ 108 mg
Ag2S I: Δm¼ 0 mg
Ag2S II: Δm ¼ +124 mg
(a) What kind of reactions take place at the interfaces between (i) Ag and Ag2S I,
(ii) Ag2S I and Ag2S II, and (iii) Ag2S II and liquid sulfur, respectively? Note
that Ag2S is an electronic conductor.
(b) Calculate the practical parabolic rate constant kP defined by the equation

ðΔxÞ2 ¼ 2kP t

where Δx denotes the thickness of the tarnished layer and t the duration of
tarnishing.
(c) What is the rate-determining species for the tarnishing reaction 2Ag(s) + S
(l) ¼ Ag2S? Calculate the mean diffusivity and conductivity of this species.
Data:
Atomic weight of Ag: 108
Atomic weight of S: 32.1
Density of Ag2S: 7.32 g/cm3
308 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

Fig. 8.8 Reaction system


configuration and mass
changes Δm. Hand-drawn
and handwritten by
C. Wagner. (Reproduced
from the lecture note of
Wagner [15]. Courtesy of
Prof. H. Schmalzried)

Standard free energy of formation of Ag2S at 200


C, ΔGo¼ 44.98 kJ/mol-
Ag2S

8. The figure below shows the rate of oxidation of pure hafnium and two of its
alloys. In each case the temperature was 1000 K, and the oxygen partial pressure
was 100 kPa. All metal phases have hcp crystal structure.
(a) Explain the trend in the data with specific reference to all the relevant defect
incorporation reactions. Assume that the dominant defects in HfO2 are
vacancies of Hf (V0000 
Hf ) or O (VO ) or both and that Al2O3 and Ta2O5 form
substitutional solid solutions with HfO2 under these conditions.
(b) Identify the site of scale formation, i.e., at the gas/oxide interface, at the
oxide/base metal interface, etc. Justify.
(c) On a plot of lnDHf vs. 1/T, show the variation of the diffusion coefficient of
hafnium, DHf, in HfO2 with temperature from the melting point of HfO2 to
room temperature at two different oxygen pressures. Derive expressions for
the slope of your graph in each regime. HfO2 is nonstoichiometric (oxygen
deficient).

:HLJKW +IDW$O
JDLQHG
2
⎛ ȟm ⎞ +I
⎜ ⎟
⎝ A ⎠

+IDW7D

7LPHW

9. The effective diffusivity is a function of grain size because grain boundaries are
high-diffusivity paths. A polycrystalline oxide scale, MO, forms on metal M. At
the temperature of the experiment, the scale coarsens measurably with time such
that at any time

Deff ¼ Dlattice þ fDg:b:


Problems 309

where Dlattice is the diffusivity measured in a single crystal of MO, Dg.b. is the
grain boundary diffusivity, and the coefficient, f, is inversely proportional to the
MO grain size. The diffusivities are independent of oxygen partial pressure PO2
over the range of interest. It is known that tel  1 and that tcation >> tanion so that
the Wagner equation for the Tammann scaling constant, kT, reduces to

00
ZμO
1
kT ¼ Deff d μO :
RT
μ0O

(a) Derive an expression for the scale thickness, Δx, as a function of time
assuming that the kinetics of scale formation is diffusion-controlled and
that the kinetics of grain growth are interface-controlled.
(b) On the same plot of (Δx)2 versus t, show how scale thickness varies with
time
(i) Under the conditions described in part (a)
(ii) In the complete absence of grain growth in the scale
10. A copper alloy containing 1 atomic percent Mg is allowed to oxidize at 1000 K
in two different oxygen partial pressures. One specimen is oxidized at
PO2 ¼ 105 atm (I) and another at PO2 ¼ 1012 atm (II). The thickness of
oxide (η) and of subscale (ξ and ε), see the illustrations below, all depends on
t1/2. Mg forms a substitutional solid solution in copper. MgO forms a substitu-
tional solid solution in Cu2O, but it is not soluble in copper.
(a) What causes the difference in behavior in case I and case II?
(b) If the difference in chemical potential of atomic oxygen (μOg-μOi) across the
external scale (case I) is 105 J/mol, what is the flux (equiv/cm2-sec) across an
external scale (Cu2O) of 102 cm thickness? Neglect the effect of internal
oxidation.
(c) If the equilibrium concentration of the oxygen in the alloy (case II) is given
by

1 CO atom frac:
O Ð O in alloy Kð1000 KÞ ¼ ¼ 11 ,
2 2 1=2
PO1 ðatmÞ1=2

what is the thickness of the internally oxidized zone ε in case II after 105 sec?
310 8 Kinetics of Gas/Solid Reaction: Diffusion-Controlled Case

(I) (II)

PO = 10−5 atm PO = 10−12 atm


2 2

gas &X0JDOOR\ gas &X0JDOOR\

VXEVFDOH VXEVFDOH

h x e

For copper at 1000 K:


DO ¼ 107cm2/s; DCu ¼ 2 1011cm2/s; DMg ¼ 6 1011cm2/s;
ρ ¼ 8.96 g/cm3; M.W. ¼ 63.5 g/mol.

For Cu2O at 1000 K:


σCu2O ¼ 5 102 S/cm; tel1; DCu vacancy ¼ 108 cm2/s; DO < <DCu;
σi ¼ Di 105 S/cm;
ρ¼6 g/cm3; M.W. ¼ 143 g/mol.

Thermodynamic data at 1000 K:


2Cu + O2/2 ¼ Cu2O ΔGo ¼ 105 kJ/mol
Mg + O2/2 ¼ MgO ΔGo ¼ 506 kJ/mol
11. The reduction of iron ore of density ρB ¼ 4.6 g/cm3 and size R ¼ 5 mm by
hydrogen can be approximated by the unreacted core model. With no water
vapor present, the stoichiometry of the reaction is

4H2 þ Fe3 O4 ! 4H2 O þ 3Fe

with rate approximately proportional to the concentration of hydrogen in the gas


stream. The first-order rate constant has been measured by Otake et al. (1957) to
be

kS ¼ 1:93 10þ5 e24,000=RT cm=s:

(a) Taking D ¼ 0.03 cm2/s as the average value of the diffusion coefficient for
hydrogen penetration of the product layer, calculate the time necessary for
complete conversion of a particle from oxide to metal at 600
C.
(b) Does any particular resistance control? If not, what is the relative importance
of the various resistance steps?
References 311

References

1. N.B. Pilling, R.E. Bedworth, The oxidation of metals at high temperatures. J. Inst. Met. 29,
529–591 (1923)
2. G.P. Sabol, S.B. Dalgaard, The origin of the cubic rate law in zirconium alloy
oxidation. J. Electrochem. Soc. 122, 316 (1975)
3. H.-I. Yoo et al., A working hypothesis on oxidation kinetics of zircaloy. J. Nucl. Mater 299,
235 (2001)
4. K. Hauffe, Oxidation of Metals (Plenum Press, New York, 1965)
5. P. Kofstad, High Temperature Oxidation of Metals (Wiley, New York, 1966)
6. C. Wagner, The formation of thin oxide films on metals. Corros. Sci. 13, 23–52 (1973)
7. C. Wagner, “Beitrag zur Theorie des Anlaufvorgangs,” Z. Phys. Chem, B21 (1933) 25;
“Diffusion and high temperature oxidation of metals”, in Atom Movements, (American Society
for Metals, Clevelan, 1951)
8. H. Schmalzried, Chemical Kinetics of Solids, Chap. 7 (VCH Verlagsgesellschaft mbH,
Weinheim, 1995)
9. G. Tamman, Z. Anorg. Allgem. Chem 111, 78 (1920)
10. F.N. Rhines, Gas-metal diffusion and internal oxidation, in Atom Movements, (ASM, Cleve-
land, 1951), pp. 174–191
11. C. Wagner, Z. Elektrochem. 63, 772 (1959)
12. H. Schmalzried, Ber. Bunsen. Phys. Chem 87, 551–558 (1983)
13. H. Schmalzried, Solid State Reactions (Verlag Chemie, Weinheim, 1981), p. 184
14. C. Wagner, Ber. Bunsenges. Phys. Chem. 63, 1027–1030 (1959)
15. C. Wagner, Lecture Note: Kinetics in Metallurgy (MIT, Cambridge, 1955), p. 77
Appendix I: Defect Chemistry of Solid State Ionic
Compounds

Introduction

An ionic compound, e.g., MO, consists of charged components, cations (M2+),


anions (O2), and electrons (e). These charged components are rendered mobile
only via defects. It is thus a prerequisite to understand the defect structure of the
system of interest, that is, types of defects and their concentrations against the
thermodynamic variables of the system.
Defects are defined as whatsoever makes a crystalline solid deviate from its ideal
crystal structure. An ideal crystal refers to an infinite crystal with component ions all
sitting at their respective lattice sites in a periodic array as stipulated by its crystal-
lographic structure, the interstitial sites all empty, electrons all at the valence
band sites, and the conduction band sites all empty; For example, an “infinite”
crystal M2+O2 with M2+ on M2+-sites (M2þ M2þ
); O2 on O2 sites (O2
O2
); electrons
  
e on the filled valence band e sites (ee ); ionic emptiness or vacancy V0 with
actual charge 0 on interstitial sites where there should be no charge (V0I0 ); and
electronic emptiness or holes h0 with actual charge 0 on the empty conduction
band h0 sites(h0h0 ).
We here denote the structure elements constructing the ideal structure MO as Sm Ln ,
indicating the species S(¼M, O, V, e, h), with an actual charge unit m(¼. . .,2, 1,
0, +1, +2,. . .), sitting at the site or locus L(¼M, O, I, e, h) where there would have to
be an actual charge unit n(¼. . .,2, 1, 0, +1, +2,. . .) if ideal. Defining the effective
charge of a structure element as c ¼ mn and denoting it, instead of numerals, as the
same number of dots (•) and primes (0 ) for c > 0 and c < 0, respectively, and a cross
(x) for c ¼ 0, the structure elements constructing the ideal structure MO may be
simplified in the form of ScL as suggested by Kröger and Vink as

MxM , OxO , VxI ; exe , hxh

which are often called the regular structure elements.

© Springer Nature Switzerland AG 2020 313


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1
314 Appendix I: Defect Chemistry of Solid State Ionic Compounds

Anything which makes the structure deviate from the ideal one or defects or
irregular structure elements may then be immediately sorted out as: missing cation or
cation vacancy V00M ; missing anion or anion vacancy VO•• ; interstitial cation MI•• ;
interstitial anion O00I ; misplaced cation MO••••; misplaced anion O0000
M ; electron in empty
conduction band e0h ; missing electron or hole in filled valance-band he• ; impurity or
alien ion substituting cation, AcM, or substituting anion, AcO ; and interstitial impurity
or alien ion, AcI .
Each of these is confined to a site or point; thus, they are called point defects.
There are more: Dislocations disturbing the periodicity of lattice sites; grain bound-
aries and surfaces spatially confining the crystal which would have to be infinite if
ideal; and voids and inclusions that are three-dimensional aggregates of point defects
of a kind. Depending on their geometries, they are often called line defects, planar
defects, and volume defects, respectively.
Formation of a defect always requires work exerted on the system, thus increasing
the energy of the system. Therefore all these defects can only be stabilized by
configurational entropy gain as temperature goes up. In this sense, only point defects
are thermodynamically stable in the normal temperature range of existence of ionic
solids. It means that their concentrations are uniquely determined by the thermody-
namic variables of the system in equilibrium state. For higher-dimensional defects,
the entropy gain can hardly compensate the energy increase, if not impossible.
Here, we will discuss the generation modes of thermodynamically stable point
defects and the defect-chemical logic to calculate the equilibrium defect structure of
a given system. As a stereotype of systems, we will consider only a binary oxide
MO, but the idea and logic can be readily extended to other binary, ternary, and
higher systems with minor modifications.

Intrinsic or Thermal Disorders

For a given system, the ratio of the M-sites to O-sites must remain fixed in any case,
as required by its crystallographic structure. For the case of, e.g., MO,

½M‐sites=½O‐sites  1

where [ ] stands for the concentration of the thing therein. Whenever whatsoever
defects are generated, this site ratio or structure condition must be observed by
necessity. Otherwise, the structure would be no longer the structure of the system
itself.
For pure MO, possible ionic defects are exhaustively: V00M , MI•• , VO•• , O00I , MO•••• ,
and O0000
M . These defects can only be generated in pair involving either M- or
O-sublattice alone or both sublattices in order to maintain the structure as
Appendix I: Defect Chemistry of Solid State Ionic Compounds 315

0 ¼ V00M þ VO•• ð1aÞ


MxM þ OxO ¼ MI•• þ O00I ð2aÞ
MxM þ VxI ¼ MI•• þ V00M ð3aÞ
OxO þ VxI ¼ O00I þ VO•• ð4aÞ
MxM þ OxO ¼ MO•••• þ O0000
M: ð5aÞ

In these generation reactions, one should recognize that in addition to the


structure condition or site (L) conservation, charge (c) and mass (S) are conserved,
as a Kröger-Vink symbol ScL suggests. These three must be conserved in any case.
If the defect concentrations are small enough compared to those of the regular
structure elements as is normally the case, one may apply the mass action law to
write for each case in order as
 00  •• 
VM VO ¼ KS ð1bÞ
 ••  00 
MI OI ¼ KaS ð2bÞ
 ••  00 
MI VM ¼ KF ð3bÞ
 00  •• 
OI VO ¼ KaF ð4bÞ
 ••••  0000 
MO OM ¼ Ka : ð5bÞ

Here, the mass action law constant Kj(j ¼ S, aS, F, aF, a) takes the form
 
Δhj
Kj ¼ Kj,0 exp  ð6Þ
kB T

where Kj,0 and Δhj are the pre-exponential factor and standard enthalpy change of
the generation reaction of j-type and kBT has the usual significance. These defect
pairs are named, in order, Schottky disorder (j ¼ S), anti-Schottky disorder (j ¼ aS),
Frenkel disorder (j ¼ F), anti-Frenkel disorder (j ¼ aF), and anti-structure disorder
(j ¼ a). All these defects are only internally generated for entropic reason and, hence,
called the intrinsic or thermal disorder. The defect pair which is energetically least
costly in a given structure usually overwhelms the rest, thus making the majority
disorder type. The anti-structure disorder, however, will be too costly in ionic solids,
thus, normally out of concern.
In a similar way, electronic disorders are generated internally or thermally as

exe þ hxh ¼ he• þ e0h : ð7aÞ

It is here noted that even in generating the electronic defects, the three (in ScL )
should be conserved, viz., the densities of states in the valence and conduction band,
316 Appendix I: Defect Chemistry of Solid State Ionic Compounds

charge, and mass. By defining free holes and electrons as h •  he•  exe and e0 
e0h  hxh, respectively, this intrinsic electronic excitation reaction is often represented
more succinctly as

0 ¼ e0 þ h • : ð7bÞ

Again applying the mass action law, one may write


 0  • 
eh he ¼ Ki ð7cÞ

where Ki is called the intrinsic


  electronic excitation
  equilibrium constant. In terms of
semiconductor jargons, e0h ¼ ½e0  ¼ n and he• ¼ ½h •  ¼ p in number concentra-
tion and then
 
Eg
np ¼ Ki ¼ Nv Nc exp  ð7dÞ
kB T

where Nv and Nc are the effective density of states at the valence (e sites) and
conduction band edge (h0 sites), respectively, and Eg the band gap.

Impurity-Induced Disorders

Nothing is pure for the thermodynamic reason. An impurity can be incorporated into
the host lattice either substitutionally or interstitially depending mostly on the
relative size of the impurity to host ion. For simplicity sake, we here consider two
types of cation impurities only: one with a lower valence and the other with a higher
valence than the host ion M2+ in MO. It is noted that our system exchanges mass
particles with the surrounding only in electrically neutral forms, because it would
otherwise be electrically charged to energetically prevent further exchange. It is thus
always easier to consider an impurity doping as incorporating the impurity in the
form of an oxide when the host is MO.
Let us first consider the incorporation of A2O in MO. There can be two possibil-
ities, A0M and AI• , which are charge-compensated by generating oppositely charged
native defects, either VO•• or MI•• and either V00M or O00I , respectively, or by themselves
if amphoteric. Incorporation reactions may, thus, be formulated exhaustively as

A2 O ¼ 2A0M þ VO•• þ OxO ð8aÞ


A2 O ¼ 2A0M þ MI•• þ 2OxO ð8bÞ
A2 O ¼ 2AI• þ V00M þ OxO ð8cÞ
Appendix I: Defect Chemistry of Solid State Ionic Compounds 317

A2 O ¼ 2AI• þ O00I ð8dÞ


A2 O ¼ A0M þ AI• þ OxO : ð8eÞ

It is informative to write down the lattice molecular formula for each case. For
xA2O added, they are in order: M1-2xA2xO1-x; M1-xA2xO; M1-xA2xO; MA2xO1+x;
M1-xA2xO. This means that the mechanisms, Eqs. (8b), (8c), and (8e), cannot be
distinguished chemically.
We will next consider the case where the impurity with higher valence E2O3 is
incorporated. Following the same logic as before, we may write the incorporation
equations exhaustively as


E2 O3 ¼ 2EM þ V00M þ 3OxO ð9aÞ

E2 O3 ¼ 2EM þ O00I þ 2OxO ð9bÞ
E2 O3 ¼ 2EI••• þ 3V00M þ 3OxO ð9cÞ
E2 O3 ¼ 2EI••• þ 3O00I : ð9dÞ

The lattice molecular formulae for xE2O3 added are in turn: M1-3xE2xO; M1-
2xE2xO1+x; M1-3xE2xO; and ME2xO1+3x. In this case, they are all chemically distin-
guished from each other.
Which of these multiple possibilities is really responsible in a given structure is
again determined by which is the least costly. Rule of thumb is that the charge
compensating native defect is determined by the majority type of thermal disorder
for the given system.

Redox-Induced Disorders

No compound exists in one and only fixed composition again for the thermodynamic
reason. For example, MO should exist over a range of composition or homogeneity
range between the M-saturated and O-saturated composition. The compound should,
thus, be represented more appropriately as M1-δO or MO1+δ. Its homogeneity range
may then be divided, in general, into three regions: hypostoichiometric (δ < 0), near-
stoichiometric (δ  0), and hyperstoichiometric (δ > 0). In this light, the stoichio-
metric composition MO is nothing but a special composition in the region δ  0 or a
symbolic representation of the compound M1-δO. An extreme example will be
“FeO,” which does not even exist within its homogeneity range (Fe1-δO with δ > 0
always).
For the composition of a system to vary, component particles should be
exchanged with its surrounding, and the exchange is driven by a difference in
component chemical potentials across the boundary. Once component chemical
potential distributions are rendered uniform, particle exchange ceases, and then the
318 Appendix I: Defect Chemistry of Solid State Ionic Compounds

system is said to be in external equilibrium. For the case of binary MO1+δ, the
number of composition variables is only one (δ), and only one of the two component
chemical potentials or activities aM and aO2 can be varied independently at given
temperature T and pressure P due to the Gibbs-Duhem equation:

1
d ln aM þ ð1 þ δÞd ln aO2 ¼ 0: ð10Þ
2

When the external equilibrium condition is disturbed, the system MO may change
its composition by exchange of both components M and O in principle but normally
by the exchange of the more volatile component which is O for the case of oxides.
When the oxygen chemical potential in the surrounding is higher than that in the
solid oxide, component oxygen tends to be incorporated to raise the oxygen content
by creating oxygen interstitials or metal vacancies depending on which are energet-
ically cheaper. This process is called oxidation. In the opposite case, oxygen leaves
the crystal to lower the oxygen content by leaving behind oxygen vacancies or metal
interstitials depending on which are energetically cheaper. This is called reduction.
What is eventually achieved is reduction or oxidation equilibrium, thus often dubbed
redox equilibrium.
It is noted that because only neutral oxygen are exchanged, what are left behind
should also be electrically neutral. A neutral interstitial oxygen may be regarded as a
normal oxygen ion interstitial O00I bearing two free holes 2h• within it and a neutral
metal vacancy as a missing metal ion V00M bearing two free holes 2h• within it or
 x  x
OxI ¼ O00I , 2h • ; VxM ¼ V00M , 2h • : ð11Þ

Likewise
 x  x
MxI ¼ MI•• , 2e0 ; VxO ¼ VO•• , 2e0 : ð12Þ

In this sense, the former two bearing holes (or missing electrons) are like electron
acceptors and the latter two bearing extra electrons electron donors in elemental
semiconductors. They are indeed so: MO1+δ always tends to be of p-type when δ > 0
and of n-type if δ < 0. For the sake of simplicity, we here assume that once these are
generated, they immediately donate all holes (or, equivalently, accept electrons) or
donate all electrons. This is termed “fully ionized” and actually happens at elevated
temperatures.
Oxidation reactions and corresponding mass action laws may then be written as

1  
O2ðgÞ ¼ OxI ¼ O00I þ 2h • : O00I p2 ¼ KOx,1 aO2 ;
1=2
ð13aÞ
2
Appendix I: Defect Chemistry of Solid State Ionic Compounds 319

1  
O2ðgÞ ¼ OxO þ VxM ¼ OxO þ V00M þ 2h • : V00M p2 ¼ KOx,2 aO2
1=2
ð13bÞ
2

Reduction reaction equilibria may be written as

1 1   1=2
OxO ¼ O2ðgÞ þ VxO ¼ O2ðgÞ þ VO•• þ 2e0 : VO•• n2 ¼ K Re,1 aO2 ; ð13cÞ
2 2
1 1   1=2
MxM þ OxO ¼ O2ðgÞ þ MxI ¼ O2ðgÞ þ MI•• þ 2e0 : MI•• n2 ¼ K Re,2 aO2 : ð13dÞ
2 2

These four redox reaction equilibrium constants Kj (j ¼ Ox,1; Ox,2; Re,1; Re,2)
each take the shape as in Eq. (6). The reader, however, should note that these four
Kj’s are not all independent of each other owing to the internal equilibria, Eqs. (1),
(2), (3) and (4) or KOx,1KRe,1 ¼ KaFKi2; KOx,1KRe,2 ¼ KaSKi2; KOx,2KRe,1 ¼ KSKi2;
and KOx,2KRe,2 ¼ KFKi2. It turns out that only one out of the four is independent and,
hence, only one out of the four is enough to describe the redox equilibrium of the
binary system MO. If ternary, there would be two composition variables, and hence,
there would be two external equilibria.

Defect Structure of “Pure” Nonstoichiometric Compound,


MO1+δ

Nothing can be absolutely pure. Here “pure” means that the majority type of intrinsic
disorder overwhelms impurities in concentration. Such a compound is said to be in
its “intrinsic regime.”
In order to calculate the equilibrium defect structure of a compound whether
intrinsic or extrinsic, one should first postulate the most likely defects on the basis of
the structure of MO. Let us suppose that our system MO has the Schottky disorder as
the majority type of ionic disorder. The defects of present interest may then be listed
as

V00M , VO•• ; e0 , h • :

We therefore need four equations to calculate these four concentrations as


functions of the thermodynamic intensive variables of the system MO1+δ, T, P,
and aO2 . These equations are formulated from the requirements that the system has to
meet:
1. 2 internal equilibria: Eq. (1b) and Eq. (7c)
2. 1 external equilibrium: Eq.  (13a)
 or (13c)
 
3. 1 charge neutrality: n þ 2 VM ¼ p þ 2 VO••
00

One may solve, in principle, these four simultaneous equations for each defect
concentration as a function of T, P (via Kj), and aO2 , but analytic solution is
320 Appendix I: Defect Chemistry of Solid State Ionic Compounds

impossible in many cases because of different algebraic character of the charge


neutrality condition compared to the other constraints. If the latter, however, can be
approximated in terms of only one pair of oppositely charged defects or majority
disorder type depending on δ-ranges, the analytic solution will be rendered trivial in
the form of
 c n
SL ¼ ΠKj j am
O2 ð14Þ
j

where nj and m are the exponents. This trick was first proposed by Brouwer, thus
called the Brouwer approximation. It goes as follows:
The charge neutrality condition may be approximated to the 2  2 limiting charge
neutrality conditions or majority disorder types:
       
ðiÞ n  2 VO•• ðδ < 0Þ; ðiiÞ n  p ðδ  0Þ; ðiiiÞ V00M  VO•• ðδ  0Þ; ðivÞ 2 V00M  pðδ > 0Þ
ð15Þ

Problem is then how to allocate these majority disorder types along the axis of
oxygen activity at a given temperature. It is reminded that the homogeneity range of
M1-δO is generally divided into the three regions: δ < 0; δ  0; and δ > 0. The
nonstoichiometry δ defined here as oxygen excess or metal deficit is given in the
present case as

    1
δ ¼ V00M  VO•• ¼ ðp  nÞ ð16Þ
2

due to the charge neutrality condition. By using this relationship, each majority
disorder type can be assigned to one of the three δ-ranges, which is already done in
Eq. (15). Now one can see there are two majority disorder types (ii) and (iii)
simultaneously in the near-stoichiometry region causing a logical conflict in appear-
ance, but it is a matter of whether Ki >> KS or KS >> Ki for the system given. If the
former is the case, sequence of the majority disorder types will be (i)!(ii)!(iv) with
increasing oxygen activity aO2 and if the latter, (i)!(iii)!(iv). The defect structure
of the system across its entire range of existence can finally be constructed by
combining the piecewise solutions, Eq. (14) for each disorder regime, in accord
with the sequence of majority disorder types. Figure 1 shows an example for
Ki >> KS..
Appendix I: Defect Chemistry of Solid State Ionic Compounds 321

Fig. 1 Defect structure of


δ< 0 δ≈0 δ> 0
M1-δO vs. oxygen activity at
fixed temperature, as n≈2[V••O] n≈p ′′]
p≈2[VM
calculated assuming n p
Ki >> KS (not to scale). The 1 1
6 6
triangles represent the
oxygen exponents “m” of 1 1

log [ ]
6 1 6
the piecewise solutions
4

1
4
V′′M VO••

log aO2

Defect Structure of “Impure” Nonstoichiometric Compound,


M1-2xA2xO1-x+δ

Let us next calculate the defect structure of a more general case, A2O-doped
MO. Here we assume that A substitutes M and MO has the anti-Frenkel disorder
as the majority type of ionic disorder. Then we may list the defects of the most
concern as

A0M , VO•• , O00I ; e0 , h • :

We thus need five constraints for these five unknowns, which comprise:
1. 2 internal equilibria: Eq. (4b) and (7c)
2. 1 external equilibrium, Eq. (13a) or (13c);
 
3. 1 mass conservation for the dopants: A0M ¼ x (constant)
     
4. 1 charge neutrality condition: n þ A0M þ 2 O00I ¼ p þ 2 VO••
These are all and exhaustive. The last charge neutrality condition may be approx-
imated to 3  2 limiting conditions or majority disorder types as
       
ðiÞ n  pðδ  0Þ; ðiiÞ n  2 VO•• ðδ < 0Þ; ðiiiÞ A0M  pðδ > 0Þ; ðivÞ A0M  2 VO•• ðδ  0Þ;
 00   00   •• 
ðvÞ 2 OI  pðδ > 0Þ; ðviÞ OI  VO ðδ  0Þ:
ð17Þ

Defining oxygen nonstoichiometry of the present xA2O-doped MO as M1-


2xA2xO1-x+δ,
322 Appendix I: Defect Chemistry of Solid State Ionic Compounds

Fig. 2 As-calculated defect δ< 0 δ≈0 δ> 0


structure of extrinsic M1-
n≈2[VO••] [A′M]≈2[VO••] [A′M]≈p p≈2[O′′I ]
2xA2xO1-x+δ vs. oxygen
activity at fixed temperature, n 6 6 p
not to scale. Note the
1 ′
AM 1
2
continuity principle between 1
nearest-neighboring

log [ ]
1 1
disorder regimes. The 4 4
triangles represent the O ′′I 2
1
V••
O
oxygen exponents “m” of 1 1
6 6
the piecewise solutions
p n

log aO2

    1
δ ¼ O00I  VO•• þ x ¼ ðp  nÞ ð18Þ
2

due to the charge neutrality condition. Each majority disorder type is now assigned
to each of the three δ-regions as indicated in Eq. (17). It should first be noted that
once “impure” or extrinsic, the intrinsic disorder types should not be in majority in
the near-stoichiometry region (δ  0) by definition; thus, possibilities (i) and (vi) are
already ruled out. Then noting that any two contiguous disorder regimes should have
one defect in common, which is termed the “continuity principle,” one may establish
the sequence of majority disorder types with increasing aO2 as (ii)!(iv)!(iii)!(v).
By combining the piecewise solution, Eq. (14) in each region in this sequence, one
finally ends up with the defect map for the present system as shown in Fig. 2.

Defect Structure of Impure “Stoichiometric” Compound MO

Finally, we consider a special case, a “stoichiometric” compound, which normally


refers to a compound with negligible nonstoichiometry. We have seen that for a
nonstoichiometry to be generated, electronic defects are always involved accompa-
nying the component particle exchanges with the surrounding. As Eqs. (11) and (12)
suggest, the range of nonstoichiometry is determined by the variability of the
valences of the component ions from chemistry point of view. This is why transition
metal compounds normally exhibit wider ranges of homogeneity. Therefore the
compounds with almost fixed-valent ions, e.g., alkali halides, alkali earth oxides,
and the like, may be regarded “stoichiometric,” and their electronic defects may also
be neglected compared to the ionic defects. Once stoichiometric, the composition is
fixed, and hence it is as if the system boundary was closed; therefore external
equilibrium is no longer a concern. Then the defect structure will be determined
Appendix I: Defect Chemistry of Solid State Ionic Compounds 323

Fig. 3 Defect structure of Intrinsic Extrinsic


E2O3-doped, [M••
I
] ≈ [VM′′ ] [E•M] ≈ 2[VM′′]
“stoichiometric” MO with
Frenkel disorder as the
-ΔhF /2k
majority, not to scale. Small
VM
′′
triangles denote the slopes.

ln [ ]

Note the intrinsic regime at EM
the high temperatures and
extrinsic regime at the low -Δh F /k
temperatures
M••
I

1/T

only by the internal equilibria or temperature and impurity content no matter how
many components there are.
Suppose that our “stoichiometric” MO has the Frenkel disorder as the majority
and substitutional impurity E2O3 (see Eq. (9a)). Then the defects of concern will be


EM , V00M , MI••

We have now
1. 1 internal equilibrium: Eq. (3b)

2. 1 mass conservation: EM ¼ x(constant)
 •    
3. 1 charge neutrality: EM þ 2 MI•• ¼ 2 V00M .
By solving these three equations, one can get each defect concentration as a
function of T and x independently of the component activities. The present situation
can be solved even analytically,
 •  but for00 consistency
   sake, we will employ the
Brouwer approximations as: EM  2 VM ; MI••  V00M . The former prevails
1=2
at low temperatures such that x >> KS and the latter at high temperatures such that
1=2
x << KS . By combining the piecewise solutions for each temperature region, one
obtains the defect structure as in Fig. 3. Note that in the high-temperature intrinsic
regime, the impurity concentration is overwhelmed by the intrinsic disorder and vice
versa in the low-temperature extrinsic regime.

Concluding Remarks

The very basic logical framework of defect chemistry has thus far been given to a
binary system. This can be readily extended to ternary or higher systems simply by
adding external equilibria each corresponding to additional composition variables in
addition to the exhaustive internal equilibrium conditions, mass conservation con-
dition, and charge neutrality condition. If the nonstoichiometry is negligible, thus,
324 Appendix I: Defect Chemistry of Solid State Ionic Compounds

the composition practically remains fixed; however, one may disregard the external
equilibria.
The defect structure gets more involved as defect concentrations increase so that
the mass action laws are no longer applicable. This basic logical framework can also
be extended by taking into account defect associates or long-range interactions in
terms of activity coefficients of defects.
For more extended treatments, the reader is referred to the general references
listed below.

General References

1. F.A. Kröger, V.J. Vink, Relations between the concentrations of imperfections in crystalline
solids, in Solid State Physics, ed. by F. Seitz, D. Turnbull, vol. 3, (Academic Press, New York,
1956)
2. W.D. Kingery, H.K. Bowen, D.R. Uhlmann, Introduction to Ceramics (Wiley, New York,
1960). (Chap. 4)
3. F.A. Kröger, The chemistry of imperfect crystals, in Imperfection Chemistry of Crystalline
Solids, vol. 2, 2nd edn., (North-Holland, Amsterdam, 1974)
4. H. Schmalzried, Point defect in ternary ionic crystals, in Progress in Solid State Chemistry,
ed. by H. Reiss, vol. 2, (North-Holland, Amsterdam, 1964)
5. D.M. Smyth, The Defect Chemistry of Metal Oxides (Oxford University Press, 2000)
6. H.I. Yoo, Defect structure, nonstoichiometry, and nonstoichiometry relaxation of complex
oxides, in Ceramic Science and Technology, ed. by R. Riedel, I. W. Chen, vol. 2, (Wiley-
VCH, Weinheim, 2010)
Appendix II: English Translation: Carl Wagner,
“Beitrag zur Theorie des Anlaufvorgangs,”
Z. Phys. Chem. B21 (1933) 25

(On the Theory of Scaling Reactions)


In the scaling of metals in oxygen, sulfur, or halogen gases, compact reaction layers
are formed in many cases, so that the diffusion of either oxygen inward or metal
outward through the scale is required for further reaction. For these cases, either an
equivalent amount of cations plus electrons can migrate to the oxide-gas interface or
oxygen anions may diffuse inward, with electrons in an equivalent amount migrating
in the opposite directions. Usually, the diffusion of only one ionic species is
significant, but scale growth is dependent upon “mixed” ionic and electronic migra-
tion. A concentration gradient is required for diffusion, so it is necessary to relate the
ionic concentrations in the scale to the non-metal or metal activities.
The scale (compound on the surface of the metal) is composed of metal cations
(component 1) of valence z1 (+) and metalloid anions (component 2) of valence z2
(). For the stoichiometric scale composition, one has

n1 : n2 ¼j z2 j:j z1 j ð1Þ

where n1 and n2 are the number of moles of ions of type1 and 2.


If the scale is a metal-excess compound, then excess cations with a corresponding
excess of electrons (component 3 with valence z3 ¼ 1) are present. For a metal-
deficient compound, cations and a corresponding number of electrons are missing. In
all cases, the conditions of electrical neutrality must be met, i.e.,

n1j z1 j¼ n2j z2 j þn3 : ð2Þ

With the assumption of independent migration of the individual ions and elec-
trons in the scale, the average migration velocity vi of component i (i ¼ 1, 2, and 3) in
the direction of the ξ-axis may be considered as a superposition of the effects of
forces and a concentration gradient (pure diffusion, vi ¼ viD). Thus

© Springer Nature Switzerland AG 2020 325


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1
326 Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . .

 

vi ¼ vD
i þ Bi z i e þ Fi ð3Þ

where e is the elementary charge, dϕ/dξ is the electrical potential gradient, Bi is the
mobility of species i (velocity per unit force per particle), and Fi is any externally
applied force (magnetic field, etc.).
A local difference in concentration or chemical potential can exist in thermody-
namic equilibrium only if external forces are present. The equilibrium condition
(ΣFi ¼ 0) is

1 dμi dϕ
þ zi e  Fi ¼ 0 ð4Þ
No dξ dξ

where No is Avogadro’s number.


At equilibrium, the total migration velocity must be zero. Then, upon comparing
Eqs. (3) and (4), one obtains the migration velocity for the case in which only a
concentration gradient exists:

Bi dμi
i ¼
vD
No dξ
ð5Þ

In the case of scaling reactions, external forces are not present, so Fi ¼ 0. Then
from Eqs. (3) and (5), one obtains
 
1 dμi dϕ
vi ¼ Bi þ zi e : ð6Þ
No dξ dξ

From this average velocity, one can calculate ηi, the amount of species i
transported through area A in unit time, or
 
1 dμi dϕ
ηi ¼ eci Avi ¼ ABieci þ zi e ð7Þ
No dξ dξ

where ηi has units of equivalents of species i per second and eci is concentration of
species i in equivalents per cm3.
The product Bi eci can be determined directly in electrical conductivity measure-
ments. For a normal electrolysis experiment, wherein no concentration gradient is
present (dμi/dξ ¼ 0), one has
 
dϕ
j ηi j electrolysis
¼ ABieci jzi je : ð8Þ

On the other hand, the current I (in amperes) can be calculated in terms of specific
electrical conductivity (in ohm1 cm1), or
Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . . 327

 
dϕ
I ¼ 300Aσ 

[Note: Wagner used 300|dϕ/dξ| where ϕ is expressed in absolute volts, as is used


in Eq. (8); 300 practical volts ¼ one absolute volt.]
The total charge transport (ions + electrons) in equivalents per second is obtained
by dividing the total current, I, by Faraday’s constant, i.e., I/96,487. The various
components contribute to the total current in proportion to their transference number
ti. Therefore
 
I 300Ati σ dϕ
j ηi j electrolysis
¼ ti ¼ : ð9Þ
96, 487 96, 487  dξ 

Comparing Eqs. (8) and (9)

300ti σ
Bieci ¼
96, 487jzi je

Substituting into Eq. (7) for components 1, 2, and 3


 
300At1 σ dμ dϕ
η1 ¼  1  z1 N o e ð11aÞ
96, 487jz1 jeNo dξ dξ
 
300At2 σ dμ dϕ
η2 ¼  2  z2 N o e ð11bÞ
96, 487jz2 jeNo dξ dξ
 
300At3 σ dμ3 dϕ
η3 ¼   z3 N o e ð11cÞ
96, 487jz3 jeNo dξ dξ

These equations have general validity. For the special case of a scaling reaction,
the sum of the positive charges transported through area A of the scale must equal the
sum of the negative charges transported through area A of the scale, or

η1 ¼ η2 þ η3 : ð12Þ

Substituting Eqs. (11a) to (11c) into Eq. (12) and solving for dϕ/dξ, which cannot
generally be measured, one obtains
 
dϕ 1 t dμ1 t dμ2 t dμ3
¼  1 þ 2 þ 3 : ð13Þ
dξ No e jz1 j dξ jz2 j dξ jz3 j dξ

For the reactions:


328 Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . .

1 metal cation þ jz1 j electrons ¼ 1 metal atom ð14aÞ


1 metalloid anion ¼ 1 metalloid atom þ jz2 j electrons ð14bÞ

one has the following equilibrium conditions:

μ1 þ jz1 jμ3 ¼ μMe ð15aÞ


μ2 ¼ μX þ jz2 jμ3 ð15bÞ

where μMe is the chemical potential of neutral metal atoms and μX is the chemical
potential of neutral metalloid atoms. The latter quantities are related through the
Gibbs-Duhem equation:

nMe dμMe þ nX dμX ¼ 0 ð16aÞ

For small deviations from stoichiometry, one may substitute Eq. (1) in Eq. (16a)
to yield
 
z 
dμX   2 dμMe : ð16bÞ
z1

Upon substitution of Eqs. (13), (15b), and (16b) into Eqs. (11a), (11b), and (11c),
one obtains
 
300At1 t3 σ 1 dμMe
η1 ¼  ð17aÞ
96, 487No e jz1 j dξ
 
300At2 t3 σ 1 dμMe
η2 ¼ ð17bÞ
96, 487No e jz1 j dξ
 
300Aðt1 þ t2 Þt3 σ 1 dμMe
η3 ¼  : ð17cÞ
96, 487No e jz1 j dξ

[Obviously, with the use of Eq. (16b), the term dμMe/|z1|dξ could be replaced by
dμX/|z2|dξ in each of Eqs. (17a), (17b), and (17c).]
The total reaction rate, which is equal to the increase in the compound of the
reaction product in equivalents, e
n, per second, is given by
 
de
n 300Aðt1 þ t2 Þt3 σ 1 dμMe 
¼ jη1 j þ jη2 j ¼ jη3 j ¼
dt 96, 487No e jz1 j  dξ 
 
300Aðt1 þ t2 Þt3 σ 1 dμX 
¼ : ð18Þ
96, 487No e jz2 j  dξ 
Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . . 329

Knowledge of the individual ionic transference numbers t1 and t2 is not needed;


only their sum, which is the ionic contribution to the conductivity, is needed. Notice
that, in principle, the reaction rate could be limited by either negligible (t1 + t2) or t3
because the rate is proportional to the product (t1 + t2)t3; however, because only a
few pure binary compounds are ionic conductors (AgI, NaCl, CuCl, but no oxides or
sulfides), rate control is usually limited by ionic diffusion, and not by electronic
migration.
The conductivity σ and the transference numbers t1, t2, and t3 are in general still
functions of the composition or of the chemical potentials μMe or μX. Correspond-
ingly, the chemical potential gradient in the reaction product is not constant. (By the
assumption of quasi-steady-state growth, the rate of transport of material through the
scale is independent of distance along the direction of diffusion.) Thus, from
Eq. (18), the gradient in the chemical potential is an inverse function of conductivity:
 
dμMe  const
 
 dξ  ¼ ðt1 þ t2 Þt3 σ ð19Þ

In regions of the scale where (t1 + t2)t3σ is large, the gradient is small, and to a first
approximation, this region controls the rate. In regions where (t1 + t2)t3σ is small, a
steeper gradient is present. In any case, the rate of transport of ionic species through
the scale, |η1| + |η2|, is the same throughout the scale.
Equation (18) may be integrated after multiplying both sides of the equation by dξ
(note again that de n/dt is supposed to be independent of ξ):
( Z μð i Þ )
de
n A 300 Me dμMe
¼ ðt þ t2 Þt3 σ
dt Δξ 96, 487No e μðaÞ 1 j z1 j
Me

( Z ðaÞ
)
μX
A 300 dμ
¼ ðt1 þ t2 Þt3 σ X ð20Þ
Δξ 96, 487No e μX
ði Þ jz2 j

Note that if A does not remain constant during oxidation (e.g., oxidation of a wire
rather than a large sheet), then the integration of Eq. (18) must be altered to include
A ¼ A(ξ).
If only poor data are available for conductivity and transference numbers, then
average values of the conductivity and transference numbers may be inserted and
removed from the integral to yield
( )
ðiÞ ð aÞ
de
n A 300ðt1 þ t2 Þt3 σ μMe  μMe
¼ 
dt Δξ 96, 487No e j z1 j
330 Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . .

( )
ðaÞ ðbÞ
A 300ðt1 þ t2 Þt3 σ μX  μX
¼  : ð21Þ
Δξ 96, 487No e j z2 j

The value

ðiÞ ðaÞ ðaÞ ðbÞ


μMe  μMe μX  μX
¼
jz1 j jz2 j

is equal to the Gibbs’ free energy of the reaction per equivalent. This equals the
electrical work of a corresponding galvanic cell (i.e., a pure ionic conductor with the
appropriate chemical potential gradient), which equals NoeEo (Eo is the emf in
absolute units). Calculating the emf in volts, one obtains

ðiÞ ð aÞ
300 μMe  μMe
Eo ðvoltsÞ ¼ 300Eo ¼ : ð22Þ
No e j z1 j

Equation (21) may then be written as


de
n A ðt1 þ t2 Þt3 σ
¼  Eo ðvoltsÞ : ð23Þ
dt Δξ 96, 487No e

For a given atmosphere (PX), the terms in brackets {} in Eqs. (20), (21), and (23)
are constant (independent of time) and may be called the rational rate constant, k(kr),
defined by the equation
 
de
n A A
¼ k ¼ k : ð24Þ
dt Δξ Δξ r

From Eq. (23)

ðt1 þ t2 Þt3 σ
k ¼ kr ¼  E ðvoltsÞ: ð25Þ
96, 487No e o

One way to measure the rate at which a metal oxidizes is to measure the increase
in thickness of the scale. If it is limited by solid state diffusion through the scale
(of ions and electrons), then the rate of growth at any time is inversely proportional
to the instantaneous thickness of the scale (first noted by G. Tamman, Z. anorg.
allgem. Chem. 1920, vol. 111, p. 78, and several years later independently by N. B.
Pilling and R. E. Bedworth, J. Inst. Metals, 1923, vol. 29, p. 529), i.e.,
Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . . 331

dðΔξÞ k0 k
¼ ¼ P ð26aÞ
dt Δξ Δξ

where k0 (kP) is called the practical (or parabolic or Tamman) reaction rate constant.
Through integration one obtains the equivalent expression

ðΔξÞ2
¼ 2k0 ¼ 2kP : ð26bÞ
t

The value of k0 or kP is to be calculated from theory through use of the rational


rate constant. The product ΔξA is the total volume of the scale. After dividing by the
equivalent volume V e (¼equivalent weight/density), one obtains the equivalent
number e n:

Δξ  A
e
n¼ ð27Þ
Ve

Through differentiation,

de
n A dΔξ
¼ : ð28Þ
dt V e dt

Then, from Eqs. (28) and (26a),

de
n A k0 A kP
¼ ¼ : ð29Þ
dt Δξ Ve Δξ V e

Upon comparison of Eqs. (29) and (24), one obtains the conversion formula:

k0 kP
k ¼ kr ¼ ¼ ð30Þ
e
V e
V

In some oxidation experiments, the course of the reaction is determined by


monitoring the change in weight of the specimen, Δm. Then, the reaction rate is
often expressed as


1 Δm 2
¼ k00 : ð31Þ
t A

The increase in weight Δm is equal to the product of the equivalent number e


n and
the equivalent weight of component X (¼atomic weight AX divided by the valence
|z2|). In addition, with Eq. (27) one has
332 Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . .

e
nV e j z2 j
e Δm  V
Δξ ¼ ¼ : ð32Þ
A A  AX

Inserting Eq. (32) into Eq. (26b), one obtains

 e 2
1 Δm 2 jz2 jV
¼ 2k0 ¼ 2kP : ð33Þ
t A AX

Thus, from Eqs. (31) and (33),

 
0
e 2 00
1 j z2 j V
k ¼ kP ¼ k , ð34Þ
2 AX

and from Eqs. (30) and (34),


 2
1 e jz2 j
k ¼ kr ¼ V k00 : ð35Þ
2 AX

Electromotive Forces in Reaction Scales


Consider the placement of electrodes (electronic conductors) on each side of a
reaction scale during its formation. The total emf between the electrodes is com-
posed of the two contact potential differences between the electrodes and the scale
and the potential difference within the scale (“diffusion potential”). The latter
contribution may be obtained through use of Eq. (13):
Z ðiÞ Z ðiÞ  
dϕ 1 t t t
ΔϕðdiffusionÞ ¼ dξ ¼  1 dμ1 þ 2 dμ2 þ 3 dμ3 ð39Þ
ðaÞ dξ No e ðaÞ j z1 j j z2 j j z3 j

The equilibrium condition for electrons between the electrode and the scale is

μ3 þ z3 No eϕ ¼ μE3 þ z3 No eϕE ð40Þ

where μE3 is the chemical potential of electrons in the electrode and ϕE is the
accompanying electrical potential. The corresponding quantities in the scale are
written without a phase index, except for the use of (i) or (a) to distinguish the
inner and outer sides of the scale. The potential difference between the electrode and
the scale (the contact potential) on the inner side is given by

1 ðiÞ
ΔϕðiÞ ¼  μE3  μ3 : ð41Þ
z3 N o e

Correspondingly, the contact potential at the outside electrode is


Appendix II: English Translation: Carl Wagner, “Beitrag zur Theorie des. . . 333


1 ðaÞ
ΔϕðaÞ ¼  μ3  μE3 : ð42Þ
z3 N o e

The sum [Δϕ(i)+ Δϕ(diffusion)+Δϕ(a)] is the emf between the two electrodes;
setting z3 ¼ 1 and t3 ¼ 1t1t2 and using Eqs. (15a) and (15b), one obtains
Z ðiÞ  
1 t1 t
E¼  dμMe þ 2 dμX , ð43Þ
No e ðaÞ j z1 j jz2 j

and with Noe ¼ F and Eq. (16b),


Z ðiÞ
1 dμMe
E¼ ðt 1 þ t 2 Þ
F ð aÞ j z1 j
Z ð aÞ
1 dμX
¼ ðt1 þ t2 Þ : ð44Þ
F ðiÞ j z2 j
Index

A C
Absolute reaction rate theory, 217 Cation, 247
Ambipolar conductivity, 293 Charged-component-level description, 250
Ambipolar diffusion, 249 Charged components, 248, 271
Anti-Frenkel disorder, 321 Charged mobile components, 252
Arrhenian form, 87 Charge neutrality, 248, 249, 255, 261, 262, 267,
Atomic mobility, 1 271, 275, 296
Atomic theory, diffusion Chemical components, 249, 259
correlation effect, 90, 92, 93 Chemical diffusion coefficient, 145, 146, 167,
correlation factor, 93–95 270, 274–277
defect diffusion coefficient, 86, 87 Chemical diffusivity, 158, 167, 275, 277
mechanisms, 80–82 Chemical potential, 317
random walk, 77–80 Chronological nucleation, 232
self-diffusion coefficient, 83–86 Coarsening kinetics
thermodynamic variables, 87–90 diffusional geometric factor, 236
Atomistic, 74 effective diffusion cross section, 235
effective diffusion distance, 235
Fick’s first law, 235
B growth kinetics, 233, 236
Boltzmann-Matano analysis, 145, 147, 150 highly dispersed case, 236–238
error-function solutions, 54 nucleation, 233
Fick’s second law, 52 Ostwald ripening, 233
mass-conservation principle, 55 Thomson-Freundlich equation, 234, 235
Matano interface, 55, 56 Collinear interstitialcy mechanism, 82
Matano-plane, 55 Component diffusivities, 167, 273
semi-infinite source, 55 Component flux equations, 271–273
short-time solutions, 54 Correlated jumps, 92
total differential equations, 53 Correlation effect, 90, 92, 93
Boltzmann transform, 53 Correlation factors, 93–96, 266
Brouwer approximations, 255 Co to CoO oxidation
Brouwer method, 261 electrochemical mobility, 296
Brownian motion, 78 external equilibrium condition, 296

© Springer Nature Switzerland AG 2020 335


H.-I. Yoo, Lectures on Kinetic Processes in Materials,
https://doi.org/10.1007/978-3-030-25950-1
336 Index

Co to CoO oxidation (cont.) point, 314


external equilibrium constant, 296 redox-induced disorders, 317–319
hyperstoichiometric/metal deficit, 296 structure (see Defect structure)
local lattice conservation, 297 structure elements, 313
Nernst-Einstein equation, 297, 298 Defect structure, 250, 252, 254–256, 260–263
parabolic rate law constant, 296, 298 impure nonstoichiometric compound, 321,
p-type semiconductor, 296 322
redox reaction, 298 impure stoichiometric compound, 322, 323
Coulometric titrometry, 273 pure nonstoichiometric compound,
Crowdion mechanism, 82 319–321
Cubic rate law, 238 Diffusion, 148, 149
Curvilinear interstitialcy mechanism, 82 concentration gradients
chemical diffusion coefficient, 270,
274–277
D component flux equations, 271–273
Darken’s analysis nonstoichiometry flux, 273, 274
Boltzmann-Matano analysis, 150 definition, 1, 2
Boltzmann transform, 152 diffusional driving force, 161
component diffusivities, 151 dimensions, 56–58
diffusion, 148, 149 thermodynamic treatment, 154–158
drift, 148, 149 Diffusion coefficient, 2, 75, 79, 229
drift velocity, 151 Diffusion mechanisms
Fick’s first law, 151 collinear interstitialcy mechanism, 82
Fick’s reference frame, 150 curvilinear interstitialcy mechanism, 82
inert-markers, 150 defect structure, 83
interdiffusion coefficient, 151, 152 direct exchange, 81
interdiffusion couple, 150 ideal crystal, 80, 81
intrinsic diffusion coefficients, 153 interstitial, 81, 82
Kirkendall effect, 149 mass conservation constraint, 82
Kirkendall experiment, 151 point defects, 81
laboratory frame of reference, 151 ring, 81
laboratory reference frame, 150 self-diffusivity, 83
Matano analysis, 152 vacancy, 81, 82
Matano-interface, 153 Diffusion medium, 73
mean molar volume, 149 Diffusional driving force, 161
Nernst-Planck type, 151 Direct exchange mechanism, 81
number-fixed reference frame, 153 Divergence operator, 10, 11
partial molar volumes, 149 Drift velocity, 148, 149, 151, 154, 162, 250,
reference frame, 150 251
Smigelskas-Kirkendall experiment, 151 Driving force, 216, 217
tracer diffusivity, 153 Drunkards’ walk, 79, 80
volume-fixed reference frame, 153
Darken’s chips, 149
Darken-type diffusion coefficient, 165 E
Debye/Einstein frequency, 76 Einstein-Smoluchowski relation, 158
Defect diffusion coefficient, 86, 87 Electrical conductivity, 259, 277
Defect diffusivities, 96, 167, 250 Electrochemical mobility, 251
Defect-level description, 250 Electrode, 247
Defects Electrolyte, 273
definition, 313 Electromigration/electrotransport, 247
formation, 314 Electron tunneling, 289
impurity-induced disorders, 316, 317 Electronic conductor, 252
intrinsic/thermal disorders, 314–316 Electronic defects, 260, 315
Index 337

Electrostatic energy, 251 normal solidification, 222–228


Equilibrium chemical potential, 287 spherical precipitates, solid solution,
Error function, 23, 24 228–231
Extended volume fraction, 232, 233 with phase/without composition change,
External equilibrium, 260, 270, 275 220, 221
External oxidation, 299 Growth rate, 217–219, 236, 238

F H
Fick’s first law, 73, 75, 151, 155, 235, 300 Helmholtz free energy, 216
application, 7, 8 Henry’s law constant, 158
definition, 1, 2 High-diffusivity paths, 289
gradient operator, 2, 3 Homogenization time, 47
limitations, 4–6
linear rate laws, 3, 4
magnitudes of D, 6, 7 I
units, 6, 7 Ideal crystal, 80, 81, 313
Fick’s reference frame, 150 Impurity-induced disorders, 316, 317
Fick’s second law, 301 Inert markers, 149, 295
continuity equation, 8–10 Instantaneous line source, 73
c(x,t) evolution with time, 13 Instantaneous planar source, 73
cylindrical symmetry, 12 Instantaneous point source, 73
divergence operator, 10, 11 Interdiffusion coefficient, 145–147, 151, 167
Fourier’s law, heat conduction, 14, 15 Interfacial energy, 216
Laplacian operator, 12 Internal oxidation
mass conservation, 8–10 boundary conditions, 301
spherical symmetry, 12 definition, 298
steady-state solutions, 15, 16 Fick’s first law, 300
transient-state solutions, 16, 17 Fick’s second law, 301
First-term approximation, 46, 47 kinetics, 300
Flux equation moving boundary problem, 301, 302
definitions, 251–253 oxide precipitation zone, 299
Fourier’s law of heat conduction, 3 oxygen solubility, 300
Frenkel disorder, 262, 269 precipitation zone, 302, 303
reaction free energies/oxygen affinities, 300
system parameters, 302
G Interstitial mechanism, 81, 82
Galvanic cell, 330 Intrinsic diffusion coefficients, 165
Gaussian distribution, 18, 21, 24, 30, 32 Intrinsic diffusivity, 149, 170
Gauss Theorem, 9 Intrinsic/thermal disorders, 314–316
Gibbs-Duhem equation, 156, 159, 271, 272, Inverse transform, 50, 51
292, 293, 328 Ionic conductor, 252
Gibbs’ free energy, 286, 330 Ionic defects, 249, 314
Gradient operator, 2, 3 Ionic solids, 247
Grain growth, 215 Ionization equilibria, 291
Growth kinetics, 228 Isotropic crystalline solid, 78
Growth processes
without phase/without composition change
driving force, 216, 217 J
grain size distribution, 215 Johnson-Mehl-Avrami equation, 233
growth rate, 217–219 Jump distance, 76, 79, 84, 86
with phase/with composition change
338 Index

K Matano interface, 55, 56, 150, 153


Kinetic steps and rate laws Matano-plane, 55
cubic rate law, 289 Mean particle size, 237
elementary, 288 Mean square displacement/spread, 73, 78
linear rate law, 289 Mechanical mobility, 154, 155, 251
logarithmic laws, 289 Migration enthalpy, 258
parabolic rate law, 289 Migrational energy barrier, 217
reactants, 288 Mixed ionic electronic conductor, 252
Kirkendall effect Mixing free energy, 160
Boltzmann-Matano analysis, 147 Molar Gibbs free energy, 217, 220
Darken-type diffusion coefficient, 165 Moving boundary problem, 301, 302
drift velocity, 162 diffusivity ratio, 61
edge dislocation climb, 165 error-function solution, 60
interdiffusion coefficient, 147 flux, 59, 60
interdiffusion couple, 165 local thermodynamic equilibrium, 59
intrinsic diffusion coefficients, 162 phase boundary, 59
Kirkendall pores, 167 saturation solubilities, 59
local defect equilibrium, 164, 165 subsolidus phase diagram, 58
Mo-wires, 146, 147
principle of local lattice conservation, 163
repeatable growth sites, 165, 166 N
“such and such effect”, 147 Natural distribution, 18
vacancy concentration, 164 Natural vibration frequency, 76
vacancy flux, 164 Nernst-Einstein equation, 252, 274, 294
vacancy generation rate, 164 Nernst-Planck-type diffusion coefficient, 151,
Kirkendall experiment, 151 275
Kirkendall pores, 167 Net jump frequency, 220
Kohlrausch’s law of independent migration, Neutral chemical component, 248
253 Neutral components, 248
Neutral interstitial oxygen, 318
Neutral metal vacancy, 318
L Neutral oxygen, 318
Laplace transform Nonstoichiometric compound
definition, 48 chemical components, 259
Fick’s second law, 49 defect structure, 260–263
inverse transform, 50, 51 Nonstoichiometry flux, 248, 273,
method, 278 274, 323
thin-film solution, 47 Nonstoichiometry re-equilibration kinetics,
total differential equation, 48 277, 278
Laplacian operator, 12 Normal growth rate, 218–221
Lattice diffusivities, 96 Normal solidification
Lattice parameter, 76 characteristic length, 225, 227
Leak test, 22, 23, 26 coring, 223
Linear growth rate, 219 diffusion coefficient, 224
Linear rate laws, 3–6 distribution/partition coefficient, 222
Local defect equilibrium, 163–166 distribution profile, 224
Local lattice conservation, 297 eutectic composition, 228
Local lattice reference frame, 295 growth front, 222, 224
Long-time solution, 16, 45, 46 local thermodynamic equilibrium, 225, 227
natural decay, 226, 227
Scheil equation, 223
M steady state growth, 228
Mass action law, 261, 316, 318 steady state solution, 225
Matano analysis, 152
Index 339

O diffusion mechanisms, 80
Ohm’s law, 251 diffusivities, 96
Orthogonality theorem, 40–42 displacement, 77
Ostwald ripening, 233 drunkards’ walk, 79, 80
Oxidation, 261, 318 isotropic crystalline solid, 78
Oxidation thermodynamics jump direction, 77
chemical potential, 287 jump distance, 79
Gibbs free energy, 286 mean square displacement, 78, 79
natural (Raoultian) reference state, 287 root mean square displacement, 79
nonstoichiometry/stoichiometric successful jump frequency, 78–80
composition, 287 travel distance, 79
reaction free energy, 287 Raoul’s law, 158
standard reaction free energy, 287 Rate law constant, 219
Oxide precipitation zone, 299, 302 Reaction free energies/oxygen affinities, 300
Oxygen solubility, 300 Redox equilibrium, 267
Redox-induced disorders, 317–319
Redox reaction equilibrium, 319
P Reduction, 318
Parabolic kinetics, 219 Reference frame, 150
Parabolic rate law, 219 Reflection and superposition method, 28, 29
Parabolic rate law and Wagner’s theory Relaxation time, 47, 278
ambipolar conductivity, 293 Root-mean-square displacement, 21
charged components, 291
chemical driving force, 292
electronic transference number, 292 S
Gibbs-Duhem equation, 292, 293 Scale, 325
ionic transference number, 292 Scale growth, 325
ionization equilibria, 291 Scaling reactions theory
Nernst-Einstein equation, 294 chemical potential, 326
practical tarnishing rate law constant, 289 electromotive forces, 332, 333
quasi-steady-state growth, 293 Gibbs’ free energy, 330
solid state electrochemistry, 290 metal-excess, 325
solid state ionics, 290 metal oxidizes, 330
Tamman constant/pilling, 289 metals, 325
thermodynamic and kinetic situation, 290 migration velocity, 326
Partial electrical conductivity, 251 rational rate constant, 330, 331
Partial molar Gibbs free energy, 251 reaction rate constant, 331
Partition coefficient, 222 scale growth, 325
Piecewise solutions, 262, 263 stoichiometric, 325
Pilling-Bedworth ratio (PBR), 285, 286 Scheil equation, 223
Point defects, 249, 254, 314 Schottky defect formation enthalpy, 258
Probability density, 22 Schottky defect formation Gibbs free energy,
255
Schottky disorder, 254
Q Secondary-ion mass spectroscopy (SIMS), 29
Quasi-steady state assumption, 294, 295 Self-diffusion coefficient, 76
attempt frequency, 85
energy barrier, 85
R isotopes, 84
Radioactive isotopes, 29 jump distance, 84
Random walk theory mean square displacement, 84
Brownian motion, 78 self-diffusivity, 86
diffusion coefficient, 79 self-tracer-diffusivity, 84
340 Index

Self-diffusion coefficient (cont.) protective, 286


successful jump frequency, 83, 84 types, 285
total vibration frequency, 85 Telltale marker experiment, 295
tracer diffusivity, 84 Temperature gradient, 3
vacancy formation free energy, 86 Thermal conductivity, 3
Self-diffusivities, 83, 86, 87, 90, 93, 154, 167 Thermal diffusivity, 14
Co1-δO, 266–268 Thermal energy, 14
Fe3-δO4, 268–270 Thermodynamic driving forces, 1
geometric factor, 264 Thermodynamic factor, 273, 275
ionic defects, 265 clustering, 159
majority disorder regime, 265 diffusion enhancement factor, 158
nonstoichiometric compound Gibbs-Duhem equation, 159
(see Nonstoichiometric compound) Henry’s law constant, 158
stoichiometric compounds inflection point, 160
(see Stoichiometric compounds) integral mixing free energy, 160
Self-tracer-diffusivity, 84 mixing free energy, 160
Semi-infinite source regular solution, 160
error function solutions solution thermodynamics, 159
generalization, 35, 36 spinodal decomposition, 160
variation, 33–35 thermodynamic bias, 159
features, 32 uphill-diffusion, 161
Fick’s second law, 30–32 Thermodynamic force, 251
properties of erf(z), 36 Thermodynamic treatment
Short-time solution, 16, 26, 32, 46 activity coefficient, 156
Solid state ionics, 248 chemical diffusivity, 158
Space charge, 289 chemical potential, 154
Stable isotopes, 29 component diffusivity, 155, 158
Standard reaction free energy, 287 directional walk, 157
Steady-state solutions, 7, 15, 16 Einstein-Smoluchowski relation, 158
Stoichiometric coefficients, 247, 276 electrochemical potential, 154
Stoichiometric composition, 247 Fick’s first law, 155
Stoichiometric compounds, 322, 323 Fickian force, 156
cation self-diffusivity, 258 fundamental mobility, 155
defect structure, 254–256 Gibbs-Duhem equation, 156
extrinsic regime, 257 mechanical mobility, 154, 155
intrinsic regime, 257 self-diffusivity, 154, 157, 158
migrational free energy, 257 tracer diffusivity, 154, 157
migration enthalpy, 258 Thermodynamic variables, 73, 254
Schottky defect formation enthalpy, 258 activation energy, 87, 89
thermal vibration frequency, 257 doping, 87
thermodynamic variables, 254 Gibbs-Helmholtz equation, 88
Successful jump frequency, 76, 78, 79, 83, 87, pre-exponential factor, 87
167 self-diffusion coefficient, 88, 90
“Such and such effect”, 147 Thermogravimetry, 273
Supercooling, 221 Thin-film source
Superposition method, 30 application, 29
concentration profiles, 19
error function, 23–25
T Fick’s second law, 17–19
Tarnishing layer, 285 flux, 20
Tarnishing reaction, 285 inflection points, 21
Tarnishing scales leak test, 22, 23, 26
non-protective, 286 local concentration change, 20
Index 341

maximum concentration, 19 Fick’s second law, 37–40


mean displacement, 21, 22 first-term approximation, 46, 47
mean square displacement, 21, 22 homogenization time, 47
reflection and superposition method, 28, 29 long-time solution, 45, 46
variations, 27 orthogonality theorem, 40–42
Thomson-Freundlich equation, 234, 235 short-time solution, 37
Tracer diffusion, 29, 93 trivial initial condition, 42–44
Tracer diffusivities, 29, 46, 84, 93, 96, 146, 167
Transference number, 252
Transformation rate U
extended volume fraction, 232, 233 Unit normal vector, 9
and growth, 232 Uphill-diffusion, 161
Johnson-Mehl-Avrami equation, 233
nucleation, 231, 232
Transient-state solutions, 15–17 V
Travel distance, 79 Vacancy diffusivity, 87
Trigonometric infinite series solution Vacancy flux, 164
concentration distribution c(x,t), 37 Vacancy generation rate, 164
error-functional solutions, 36 Vacancy mechanism, 81, 82, 84
features, 44, 45

You might also like