cond-mat0603330/Neu.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % authors: J. Neu and L.L. Bonilla 
3: %
4: % title:  Coarsening of binary alloys 
5: %
6: %
7: %
8: % Journal: Phys. Rev. E
9: %
10: % remarks: All correspondence should be sent to L.L. Bonilla
11: %
12: % manuscript number:
13: %
14: % address: Departamento de Matematicas
15: %
16: % Escuela Politecnica Superior
17: %          Universidad Carlos III
18: %          Avda. Universidad, 30
19: %          28911 Leganes (Madrid), Spain
20: %
21: % Internet: bonilla@ing.uc3m.es
22: %
23: % FAX:     34-91-624-9129
24: %
25: % Tel:     34-91-624-9445
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
28: 
29: %% Draft (18 July 2001)
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31: 
32: %
33: %
34: % For preprintstyle the commands \begin{multicols}{2} and
35: %                     \end{multicols}   must be deleted
36: %\documentstyle[multicol,aps,epsfig]{revtex} 
37: % for multicolumn
38: %
39: %\documentstyle[preprint,aps,epsfig]{revtex} 
40:  % for preprint form
41: % (RevTex 3.0)
42: %\documentstyle[prb,aps]{revtex}  % for camera-ready manuscript
43: %(RevTex 3.0)
44: %\def\baselinestretch{2}
45: %\newtheorem{Lemma}{Lemma}
46: %\voffset=-2truecm
47: \batchmode
48: \makeatletter
49: %\def\input@path{{/users/davidr/polimeros//}}
50: %\makeatother
51: \documentclass{article}
52: \usepackage{graphics}
53: 
54: %\makeatletter
55: 
56: 
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
58: \providecommand{\LyX}{L\kern-.1667em\lower.25em\hbox{Y}\kern-.125emX\@}
59: 
60: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Textclass specific LaTeX commands.
61: \newcommand{\lyxaddress}[1]{
62:   \par {\raggedright #1 
63:   \vspace{1.4em}
64:   \noindent\par}
65: }
66: 
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
68: %\usepackage{springer}
69: %\makeatother
70: 
71: \begin{document}
72: 
73: %%%%% USER-DEFINED MACROS HERE %%%%%
74: 
75: \def\RR{\hbox{{\rm I}\kern-.2em\hbox{\rm R}}}
76: \def\pRR{\hbox{{\tiny \rm I}\kern-.1em\hbox{{\tiny \rm R}}}}
77: \def\negro{\hspace*{\fill}$\blacksquare$}
78: \def\div{\mbox{{\rm div}}}
79: \def\rot{\mbox{{\rm rot}}}
80: \def\NN{\hbox{I\kern-.2em\hbox{N}}}
81: \def\eg{{e.g.\ }}
82: \def \l{\lambda}
83: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
84: 
85: %\hyphenation{super-lat-tice semi-con-ductor}
86: 
87: %\tighten       %Gives single-space (RevTex 3.0)
88: 
89: 
90: %\begin{document}
91: %\draft %prints PACS numbers in
92: 
93: \title{Kinetic theory of nucleation and coarsening}
94: %\renewcommand{\thesection}{\Roman{section}}
95: %\renewcommand{\thesubsection}{\Roman{section}.\Alph{subsection}}
96: %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
97: 
98: %\date{\today %May 15, 2000}
99: \maketitle
100: 
101: \author{ J. C. Neu}
102: \lyxaddress{Department of Mathematics, 
103: University of California at Berkeley, Berkeley, 
104: CA 94720, USA}
105: 
106: \author{ L.L. Bonilla}
107: \lyxaddress{Escuela Polit\'{e}cnica Superior, 
108: Universidad Carlos III de Madrid, Avda.\
109: Universidad 30, 28911 Legan{\'e}s, Spain}
110: 
111: \begin{abstract}
112: Classical theory of nucleation based on Becker-Doering equations and coarsening for 
113: a binary alloy. 
114: \end{abstract}
115: 
116: %\pacs{PACS numbers: %05.45.+b, 05.20.-y, 05.40.+j
117: %}
118: 
119: %\begin{multicols}{2}
120: %\narrowtext
121: 
122: \setcounter{equation}{0}
123: \section{Introduction}
124: \label{sec-introduction}
125: The purpose of this chapter is to explain the classical kinetic
126: theory of nucleation in a context simpler than polymer
127: crystallization. Many theories start by assuming that polymer
128: crystallization is an activated process involving crossing of a
129: free energy barrier \cite{zia68}. The latter separates two
130: accessible stable states of the system such as monomer solution
131: and crystal. This general setting for activated processes can be
132: used to describe the formation of a crystal from a liquid
133: cooled below its freezing point \cite{mar95}, precipitation \index{precipitation} and
134: coarsening \index{coarsening} of binary alloys \cite{XH}, colloidal crystallization
135: \cite{gas01}, chemical reactions \cite{pag97}, polymer
136: crystallization \cite{zia68,reg01,zia01}, etc. In all these
137: cases, the theory of homogeneous isothermal nucleation provides
138: a framework to study the processes of formation of nucleii from
139: density fluctuations, and their growth until different nucleii
140: impinge upon each other. In the early stages of these
141: processes, nucleii of solid phase are formed and grow by
142: incorporating particles from the surrounding liquid phase.
143: There is a critical value for the radius of a nucleus that
144: depends on a chemical drive potential, which is proportional to
145: the supersaturation \index{supersaturation} for small values thereof. In this limit,
146: the critical radius \index{critical radius} is inversely proportional to the
147: supersaturation. At the beginning of the nucleation process,
148: nucleii have small critical radius and new clusters are being
149: created at a non-negligible rate. As the size of existing
150: clusters increases, there are less particles in the liquid
151: phase, the supersaturation decreases and the critical radius
152: increases. Then it is harder for new clusters to spontaneously
153: appear from density fluctuations. What happens is that
154: supercritical clusters (whose radii are larger than the
155: critical one) keep growing at the expense of subcritical
156: clusters, that in turn keep losing particles. The size of the
157: nucleii is still small compared to the average distance between
158: them, so that impingement processes (in which two or more
159: clusters touch and interaction between them dominates their
160: growth) can be ignored. This stage of free deterministic growth
161: is called {\em coarsening} \cite{LS}. 
162: 
163: A convenient framework to describe nucleation and coarsening is
164: the classical Becker-D\"oring \index{Becker-D\"oring} kinetic theory. We assume that the
165: dominant processes for nucleus growth or shrinking are addition
166: or subtraction of one particle. Nucleation is thus treated as a
167: chain reaction whereby nucleii of $n$ particles are created by
168: adding one particle to a nucleus of $n-1$ particles, or 
169: subtracting one particle from a nucleus with $n+1$ particles.
170: We can then write rate equations for the number density of
171: nucleii of $n$ particles by using the law of mass action. The
172: kinetic rate constants for the processes of addition and
173: depletion have to be determined by using specific information
174: from the physical process we are trying to model. Typically we
175: impose detailed balance which implies that the ratio of rate
176: constants is proportional to the exponential of the free energy
177: cost of adding one particle to a nucleus of $n$ particles (in
178: units of $k_B T$). This leaves one undetermined rate constant.
179: There are different ways of finding the missing constant. One
180: way is to postulate a microscopic theory for particle
181: interaction and use Statistical Mechanics to determine the free
182: energy of a cluster \cite{leb77}. A different point of view is
183: to impose that our rate constants should provide a description
184: of coarsening compatible with the macroscopic description in
185: terms of balance equations. We shall illustrate this second
186: point of view and be led to a Smoluchowski \index{Smoluchowski} equation from which
187: the Lifshitz-Slyozov \index{Lifshitz-Slyozov} coarsening theory follows \cite{juanjo}. 
188: 
189: 
190: The structure of this paper is as follows. We present the
191: Becker-D\"oring kinetic equations for cluster with $n$
192: particles in Section \ref{sec-kclusters}. One relation between
193: the two rate constants of this theory follows from detailed
194: balance. The other rate constant has to be determined by
195: comparison with the known macroscopic equation for the growth of
196: cluster radii. In the small supersaturation limit, the
197: Becker-D\"oring equations can be approximated by a Smoluchowski
198: equation for the distribution function of cluster radii. Its
199: drift term yields the growth of cluster radii in terms of the
200: missing rate constant. To compare with experimental data, we
201: consider the case of coarsening of a binary alloy \cite{XH}. In
202: Sections \ref{sec-phaseeq} to \ref{sec-quasistatic}, we review
203: phase equilibria, macroscopic kinetics of precipitate and
204: matrix atoms and the quasistatic \index{quasistatic} limit of the kinetic
205: equations, respectively. As a result, we find the growth of the
206: radius of a supercritical cluster in terms of macroscopic
207: parameters. Comparison with the results in Section
208: \ref{sec-kclusters} yields the sought rate constant; see
209: Section \ref{sec-quasistatic}. Numerical values for all the
210: parameters involved in our theories can be calculated from
211: experimental data as explained in Section \ref{sec-XH}. A
212: discussion of our results constitutes the last Section.
213: 
214: \setcounter{equation}{0}
215: \section{Kinetics of clusters}
216: \label{sec-kclusters}
217: Let us assume that we have two stable phases
218: characterized by different values, $c_1$ and $c_2$ of the
219: number density $c$. Phase 1 is solution and Phase 2
220: precipitate. Or Phase 1 is the liquid and Phase 2 the
221: crystal phase. Initially all precipitate particles are in Phase
222: 1. Classic Becker-D\"oring (BD) kinetics treats nucleation as a
223: {\em chain reaction} whereby nucleii (assumed to be {\em
224: spherical}) of
225: $n$ precipitate particles are created by adding one particle to
226: a nucleus with $n-1$ particles, or subtracting one particle
227: from a nucleus with $n+1$ particles. This chain reaction
228: scheme is natural for the situation that BD has
229: in mind, in which bulk precipitate phase consists
230: only of precipitate particles so distinction
231: between particles in nucleus or in solution is
232: clear. 
233: 
234: Let $\rho_n$ be the number density of nucleii of
235: $n$ particles. The monomer density $\rho_1$ represents
236: the concentration of precipitate particles in
237: solution and as such it will be identified with the
238: concentration $c_\infty$ of the macroscopic theory in Section
239: \ref{sec-quasistatic}. Consider the reaction
240: $$ n+1 \rightleftharpoons (n+1) .$$
241: The forward reaction proceeds at a rate proportional
242: to $\rho_1\,\rho_n$ with some rate constant $k_a$.
243: The backward reaction proceeds at a rate proportional
244: to $\rho_{n+1}$ with rate constant $k_d$. Hence the
245: net rate of creation of $(n+1)$-clusters from $n$-clusters per
246: unit volume is the {\em flux} 
247: \begin{eqnarray}
248: j_{n}\equiv k_{a,n}\rho_1\rho_n - k_{d,n+1}
249: \rho_{n+1}.   \label{5.1}
250: \end{eqnarray}
251: The fact that the rate constants depend on cluster
252: size has been explicitly indicated in (\ref{5.1}).
253: Net rate of creation of $n$-clusters is due to their creation
254: from $(n-1)$-clusters minus the rate of creation of
255: $(n+1)$-clusters from $n$-clusters, 
256: \begin{eqnarray}
257: \dot{\rho}_{n} = j_{n-1}- j_n\equiv - D_-\, j_n,
258: \quad n\geq 2.    \label{5.2}
259: \end{eqnarray}
260: This formulation specifies the evolutions of
261: $\rho_2$, $\rho_3$, \ldots with $\rho_1 = c_\infty$
262: {\em given}. The number density of precipitate
263: particles in $n$-clusters is $n\rho_n$ and the density
264: $c$ (equal to the initial concentration of precipitate particles
265: in the solution) of {\em all} precipitate particles is 
266: \begin{eqnarray}
267: c \equiv \sum_{n=1}^{\infty} n\,\rho_{n},   
268: \label{5.3}
269: \end{eqnarray}
270: or equivalently, 
271: \begin{eqnarray}
272: c - c_1 = \gamma_{\infty} + \sum_{n=2}^{\infty}
273: n\,\rho_{n}, \label{5.3b}
274: \end{eqnarray}
275: where we have defined $\gamma = c -c_1$, $\gamma_{\infty} =
276: c_{\infty} -c_1$. There is conservation of all precipitate
277: particles so $c$ or $\gamma_0 = c-c_1$ are constant.
278: (\ref{5.3b}) establishes that the constant initial
279: concentration disturbance
280: $\gamma_0$ is sum of the disturbance of precipitate particles,
281: $\gamma_\infty$, and the number density of particles in all
282: cluster sizes $n\geq 2$. Given the constraint (\ref{5.3}), the
283: evolution of all $\rho_n$, including $\rho_1$, is
284: specified. 
285: 
286: A most essential point of BD kinetics is
287: identification of rate constants $k_a$ and $k_d$.
288: Ideally, this would be based on basic energetics
289: and dynamics at the microscopic level, but a
290: complete realization of this ideal is clearly elusive. Here
291: is what {\em is} done: {\em The ratio is
292: determined by detailed balance}. Equilibrium, if
293: achievable, is described by a zero flux. Setting
294: $j_n =0$ in (\ref{5.1}) implies
295: \begin{eqnarray}
296: {k_{a,n}\over k_{d,n+1}} = {\rho_{n+1}\over\rho_{1}
297: \rho_{n}}\,.      \label{5.4}
298: \end{eqnarray}
299: Standard equilibrium physicochemical theory
300: states that  
301: \begin{eqnarray}
302: {\rho_{n+1}\over\rho_{n}} = e^{-{\mu_{n}\over
303: \tau}}\,,    \label{5.5}
304: \end{eqnarray}
305: where $\mu_n$ is the free energy cost of creating
306: an $(n+1)$-particle nucleus from an $n$-particle
307: nucleus relative to the state of no nucleus. $\tau= k_B T$ is
308: the temperature measured in units of energy. Clearly, $\mu_n =
309: G_{n+1} - G_n$ ($G_n$ is the free energy \index{free energy} of a
310: $n$-cluster), so (\ref{5.5}) becomes 
311: \begin{eqnarray}
312: {\rho_{n+1}\over\rho_{n}} = e^{-{G_{n+1} -
313: G_{n}\over\tau}}\,,    \nonumber
314: \end{eqnarray}
315: and (\ref{5.4}) now reads
316: \begin{eqnarray}
317: \rho_{1} k_{a,n} = e^{-{G_{n+1} -
318: G_{n}\over\tau}}\, k_{d,n+1}.    \nonumber
319: \end{eqnarray}
320: Thus formula (\ref{5.1}) for the flux is now
321: \begin{eqnarray}
322: j_{n} = k_{d,n+1}\,\left\{ e^{-{G_{n+1} -
323: G_{n}\over\tau}}\,\rho_n - \rho_{n+1}\right\} .   
324: \label{5.6}
325: \end{eqnarray}
326: 
327: The equilibrium considered in the detailed
328: balance argument is achievable only if $G_n\to
329: +\infty$ as $n\to\infty$, whereas in a
330: supersaturated solution, $G_n$ achieves a maximum
331: for finite $n$ and then $G_n\to -\infty$ as $n\to
332: \infty$. The determination of the ratio $k_a/k_d$
333: is assumed to hold regardless.
334: 
335: What is known about $G_n$? Microscopic models for
336: $G_n$ (or, equivalently, the cluster partition
337: function $Q_n \equiv \sum_K e^{U\, b(K)/\tau} =
338: e^{-G_{n}/\tau}$, where $U$ is the binding energy
339: per pair of particles in the cluster of $n$
340: particles, $b(K)$ is the number of
341: nearest-neighbor pairs of particles in the cluster
342: $K$, and the sum is over all translationally
343: inequivalent $n$-particle clusters) are described
344: in \cite{leb77,pen83,pen84}. We would like to
345: follow here a simpler approach, consisting of
346: identifying the resulting expressions for large
347: spherical nucleii with known facts about radius growth in
348: the quasistatic approximation. For nucleii of
349: macroscopic size $n\gg 1$, $n$ can be written in terms of the
350: cluster radius $a$ by
351: \begin{eqnarray}
352: n ={4\pi\over 3}\, c_2\, a^3,    \label{5.7}
353: \end{eqnarray}
354: and $G_n\sim G(a)$, where $G(a)$ is the free
355: energy of a nucleus of radius $a$ as determined
356: by continuum theory. For nucleii of only a few
357: particles, this asymptotic correspondence with
358: continuum theory breaks down. But if the critical
359: nucleus \index{critical nucleus} has $n\gg 1$ particles, the continuum
360: approximation works for $n$ on the order of
361: critical cluster size.
362: 
363: In the limit $|(G_{n+1}-G_n)/\tau| \ll 1$, formula (\ref{5.6})
364: for the flux reduces to 
365: \begin{eqnarray}
366: j_{n} = k_{d,n+1}\,\left\{ -{(G_{n+1} -
367: G_{n})\,\rho_{n}\over\tau} + \rho_n - \rho_{n+1}
368: \right\} \nonumber\\
369: = -  k_{d,n+1}\,\left\{ {1\over\tau}\, (D_+G_{n})
370: \,\rho_{n} + D_+ \rho_n \right\},   \label{5.8a}
371: \end{eqnarray}
372: where $D_{\pm} h_n \equiv \pm\, (h_{n\pm 1} -
373: h_n)$. The basic evolution equation (\ref{5.2})
374: now reads 
375: \begin{eqnarray}
376: \dot{\rho}_{n} - D_- \,\left\{ k_{d,n+1}\,\left(
377: {1\over\tau}\, (D_+ G_{n})\,\rho_{n} + D_+ \rho_n
378: \right)\right\} = 0.   \label{5.8b}
379: \end{eqnarray}
380: This equation looks like a spatially discretized
381: Smoluchowski equation. Asymptotic replacement of
382: difference operators $D_+$, $D_-$ by derivatives
383: is justified if the relative changes in $G_n$,
384: $\rho_n$ and $k_d$ when $n$ increases by one are
385: small. Here we will follow the simple procedure of
386: formulating the continuum limit an checking its validity a
387: posteriori. 
388: 
389: The space-like variable in (\ref{5.8b}) is $n$.
390: Experimental data usually contain histograms showing the
391: distribution of nucleii in the space of their radii $a$, so
392: we adopt the radius $a$ as a more convenient
393: space-like variable. The dependent variable
394: should be $\rho = \rho(a,t)$, the distribution of
395: nucleii in space of radius $a$. Thus $\rho(a,t)\,
396: da$ is the number of nucleii per unit volume with
397: radii in $(a,a+da)$. Conversions $n\to a$, $\rho_n
398: \to\rho$ are now determined. From (\ref{5.7}), it
399: follows that the change $da$ in $a$ when $n$
400: increases by 1 is given by
401: \begin{eqnarray}
402: 1 \sim 4\pi\, c_2 a^2\, da.   \label{5.9}
403: \end{eqnarray}
404: In the general continuum theory of nucleii,
405: the concentration of precipitate particles inside a nucleus,
406: $c_2$, is a function of the radius $a$. But for many
407: experiments, deviations of $c_2$ from its equilibrium value for
408: a planar interface are negligible, so in (\ref{5.9})
409: any term arising from $a$-dependence of $c_2$ is dropped.
410: $\rho_n$ is related to $\rho(a,t)$ by 
411: \begin{eqnarray}
412: \rho_n \sim \rho(a,t)\, da\sim {1\over 4\pi
413: c_{2}}\, {\rho\over a^{2}}\,.  \label{5.10}
414: \end{eqnarray}
415: Given any sequence $h_n$ with continuum
416: approximation $h(a)$, 
417: \begin{eqnarray}
418: D_+ h_n \sim D_- h_n\sim {h_{a}\over 4\pi
419: c_{2}\, a^{2}}\,.  \label{5.11}
420: \end{eqnarray}
421: It follows from (\ref{5.10}) and (\ref{5.11}) 
422: that the continuum limit of (\ref{5.8b}) is 
423: \begin{eqnarray}
424: \rho_{t} - {\partial\over\partial a} \left\{ {k_{d}\over
425: (4\pi c_{2}a^{2})^{2}}\, \left( {\rho\over\tau}\,
426: {\partial G\over\partial a} + a^2 {\partial \over\partial a}
427: \left( {\rho\over a^{2}}\right)\right) \right\} = 0.
428: \label{5.12}
429: \end{eqnarray}
430: (Here $k_d$ is a function of $a$, to be specified). The constraint (\ref{5.3b}) can be
431: written as
432: \begin{eqnarray}
433: \gamma_\infty + c_2\, \int_0^\infty {4\pi\over
434: 3}\, a^3\, \rho(a,t)\, da = c-c_1\equiv \gamma_0. 
435: \label{5.3c}
436: \end{eqnarray}
437: 
438: As time elapses, it will be seen that the diffusive term in Eq.\
439: (\ref{5.12}) becomes negligible in comparison with the drift
440: term. The latter yields the following equation for radius
441: growth: 
442: \begin{eqnarray}
443: \dot{a} = - {k_{d}\over (4\pi c_{2}a^{2})^{2}\tau}\,
444: {\partial G\over\partial a}\,. \label{dot-a}
445: \end{eqnarray}
446: We now present a macroscopic theory that gives an explicit
447: expression for $\dot{a}$ that can be compared to experimental
448: data. Then $k_d(a)$ can be determined and this will specify the
449: limit of $k_{d,n}$ for large $n$. 
450: 
451: \section{Phase equilibria \index{phase equilibria} of a binary material} 
452: \label{sec-phaseeq}
453: 
454: \subsection{Phase equilibria of a binary
455: material} 
456: 
457: Let us consider a medium consisting of two
458: different particles. The more abundant type is
459: called ``matrix'' \index{matrix}, the other ``precipitate'' \index{precipitate}.
460: Suppose that we have a uniform mixture at fixed
461: temperature and pressure. Let $\mu$ be the
462: chemical potential \index{chemical potential}, i.e., the free energy cost of
463: adding one precipitate particle to a pre-existing
464: solution. It is a function of the number density
465: of precipitate particles, $c$:
466: \begin{equation}
467: \mu =Ê\mu(c).\label{1.1}
468: \end{equation}
469: Let us now derive the relationship of the
470: chemical potential, $\mu(c)$, to the bulk free
471: energy density, $g(c)$. We shall add one
472: precipitate particle to a solution of total
473: volume $V$. Then the free energy changes from
474: $g(c) V$ to $g(c) V + \mu(c)$, but the volume
475: changes from $V$ to $V' \equiv V +\nu(c)$, where
476: $\nu(c)$ is the specific volume of a precipitate
477: particle in a solution of number density $c$. The
478: number density changes from $c$ to $c' \equiv
479: (cV+1)/(V+\nu)$, and therefore the new free
480: energy is also expressed as $g(c')V'$. Hence we
481: get the identity
482: \begin{equation}
483: g(c)V +Ê\mu(c) = g\left({cV+1\over V+\nu}
484: \right)\, (V+\nu).\label{1.2a}
485: \end{equation}
486: Since $cV\gg 1$, $\nu\ll V$, this identity
487: reduces to 
488: \begin{equation}
489: \mu(c) = g'(c) + \nu(c)\, \{g(c) - c g'(c)\}.
490: \label{1.2b}
491: \end{equation}
492: One can just as easily consider the chemical
493: potential $\overline{\mu}(c)$, which is the free
494: energy cost of adding one matrix particle to the
495: solution. Adding one matrix particle changes the
496: free energy to $g(c)V + \overline{\mu}(c)$. The
497: volume of the solution changes now to $V'=V +
498: \overline{\nu}(c)$, where $\overline{\nu}(c)$ is
499: the specific volume of a matrix particle in
500: solution. The concentration $c$ of the
501: precipitate changes to $c' \equiv cV/(V+
502: \overline{\nu})$, and therefore the new free
503: energy is $g(c')V'$. Hence we get an identity
504: analogous to (\ref{1.2a}),
505: \begin{equation}
506: g(c)V +Ê\overline{\mu}(c) = g\left({cV\over
507: V+\overline{\nu}} \right)\,
508: (V+\overline{\nu}). \label{1.3a}
509: \end{equation}
510: Again the conditions $cV\gg 1$, $\overline{\nu}
511: \ll V$ lead to an asymptotic reduction of
512: (\ref{1.3a}),
513: \begin{equation}
514: \overline{\mu}(c) = \overline{\nu}(c)\,\{g(c) - c
515: g'(c)\}.   \label{1.3b}
516: \end{equation}
517: 
518: \subsection{Phase equilibria in a dilute solution}
519: Suppose that there are two stable phases
520: characterized by different values, $c_1$ and
521: $c_2$ of the number density $c$, and that these
522: phases occupy adjacent half spaces separated by a
523: planar interface. Imagine that one precipitate
524: particle is removed from phase 1 and dropped in
525: phase 2. The free energy cost is
526: $\mu(c_2) - \mu(c_1)$. In equilibrium, the energy
527: cost is to be zero, therefore
528: \begin{equation}
529: [\mu] \equiv \mu(c_2) - \mu(c_1) = 0. \label{1.4a}
530: \end{equation}
531: Similarly, the energy cost of removing a matrix
532: particle from phase 1 and dropping it in phase 2
533: is also to be zero, so
534: \begin{equation}
535: [\overline{\mu}]  = 0. \label{1.4b}
536: \end{equation}
537: The possible existence of multiple phases is
538: determined by the structure of $g(c)$. Consider a
539: {\em dilute} solution with $c$ much smaller than
540: the total atomic density. In this case, the
541: specific volumes $\nu$ and $\overline{\nu}$ of
542: precipitate and matrix particles should be nearly
543: constants independent of $c$: ``crowding'' effects
544: should be insignificant. In this case,
545: (\ref{1.3b}) and (\ref{1.4b}) imply
546: \begin{equation}
547: [g(c) - c \, g'(c)]\equiv [g-cg'] = 0.\label{1.5}
548: \end{equation}
549: Given this result, it now follows from
550: (\ref{1.2b}) and (\ref{1.4a}) that
551: $$[g'] = 0.$$
552: Hence,
553: \begin{equation}
554: g'(c_1) = g'(c_2) = M\quad \mbox{(common
555: value)},  \label{1.5a}
556: \end{equation}
557: and 
558: \begin{equation}
559: [g] - [c] \, M = 0.   \label{1.5b}
560: \end{equation}
561: Given $g(c)$ with $g''(c)<0$ in some interval of
562: $c$, and $g''(c)>0$ outside, one discerns the
563: standard geometrical construction of the solution
564: to (\ref{1.5a}) and (\ref{1.5b}) for $c_1$ and
565: $c_2$. This is depicted in Figure \ref{f1.1}.
566: \begin{figure}
567: \begin{center}
568: {\par\centering \resizebox*{0.6\columnwidth}{!}{\includegraphics{curva1.eps}} \par}
569: %\vspace{0.5 cm}
570: \caption{Geometrical construction of stable phase
571: equilibria. }
572: \label{f1.1}
573: \end{center}
574: \end{figure}
575: 
576: \subsection{Critical nucleus}
577: Consider a spherical nucleus of phase 2
578: surrounded by phase 1. The energy cost of adding
579: one precipitate particle to this nucleus is
580: $$[\mu] + 8\pi\sigma r\, dr.$$
581: Here $r$ is the initial radius of the nucleus,
582: $dr$ is the change in radius due to adding one
583: precipitate particle, and $\sigma$ is the surface
584: tension. One has 
585: $$4\pi r^2 dr = \nu$$
586: so the energy cost can be written as 
587: $$[\mu] + {2\sigma\nu\over r}\,.$$
588: For a nucleus in equilibrium, this energy cost is
589: zero, therefore
590: \begin{equation}
591: [\mu] = - {2\sigma\nu\over r}\,.\label{1.6a}
592: \end{equation}
593: Now we add one matrix particle to the nucleus. No
594: energy cost for this process implies
595: \begin{equation}
596: [\overline{\mu}] = - {2\sigma\overline{\nu}\over
597: r}\,.\label{1.6b}
598: \end{equation}
599: After substituting for $\overline{\mu}$ from
600: (\ref{1.3b}), this equation becomes 
601: \begin{equation}
602: [g - c\, g'] = - {2\sigma\over r}\,. 
603: \label{1.7}
604: \end{equation}
605: Now substitute (\ref{1.2b}) for $\mu$ and
606: (\ref{1.7}) for $2\sigma/r$ in (\ref{1.6a}) to
607: get 
608: \begin{equation}
609: [g'] = 0,\label{1.8}
610: \end{equation}
611: which is the same as in the case of a planar
612: interface, (\ref{1.5a}). Given the concentration
613: $c_1$ of phase 1, this equation determines the
614: concentration $c_2$ inside the nucleus, and then
615: the Gibbs-Thomson \index{Gibbs-Thomson} relation (\ref{1.6a}) determines
616: the radius of the nucleus, $r$. This
617: determination simplifies when the concentrations
618: $c_1$ and $c_2$ are near ``planar'' values and
619: $\gamma_1$ and $\gamma_2$ are deviations from the
620: planar values. (\ref{1.8}) together with
621: (\ref{1.5}) for the planar case imply 
622: \begin{equation}
623: [g''\,\gamma] = 0.\label{1.9}
624: \end{equation}
625: Let us denote the common values of $g''(c_1)\,
626: \gamma_1$ and $g''(c_2)\, \gamma_2$ by $m$. The
627: variation of $[g-c\, g']$ in (\ref{1.7}) is 
628: \begin{equation}
629: - [c\, g''\,\gamma] = -[c]\, m.\nonumber
630: \end{equation}
631: Hence  (\ref{1.7}) gives 
632: \begin{equation}
633: [c]\, m = {2\sigma\over r}\,.\label{1.10}
634: \end{equation}
635: 
636: \section{Macroscopic kinetics}
637: \label{secMkinetics}
638: \subsection{Balance equations and jump conditions}
639: Let $c(x,t)$, $\overline{c}(x,t)$ denote the 
640: macroscopic number densities of precipitate and
641: matrix atoms. Local volume fractions of
642: precipitate and matrix atoms are $\nu c$ and
643: $\overline{\nu}\, \overline{c}$, respectively.
644: Since matrix and precipitate atoms fill space
645: leaving no gaps, we have the {\em space filling
646: condition}
647: \begin{equation}
648: \nu\, c + \overline{\nu}\, \overline{c} = 1. 
649: \label{2.1}
650: \end{equation}
651: In conventional kinetics, the density of
652: precipitate is locally conserved, with a flux
653: proportional to the gradient of the precipitate
654: chemical potential $\mu(c)$,
655: $$ c_t + \nabla\cdot (-\delta\,\nabla \mu) = 0,
656: $$ 
657: or
658: \begin{equation}
659: c_t =  \nabla\cdot (D\,\nabla c),\quad D =
660: \delta(c)\, \mu'(c). 
661: \label{2.2}
662: \end{equation}
663: Here $\delta(c)$ is a positive mobility
664: coefficient and $-\delta\nabla\mu = - D\,\nabla
665: c$ is the flux of $c$. This flux is formally a
666: diffusion with diffusion coefficient $D =
667: \delta(c)\,\mu'(c)$. Given $\mu(c)$ as in
668: (\ref{1.2b}), and provided $\nu$ and
669: $\overline{\nu}$ {\em do not depend on} $c$
670: (dilute solution), 
671: \begin{equation}
672: D = \delta(c)\,\mu'(c) = \delta (1-\nu c)\,
673: g''(c) = \delta\overline{\nu}\,\overline{c}\,
674: g''(c).      \label{2.3a}
675: \end{equation}
676: In the last equality, the space filling condition
677: has been used to replace $1-\nu c$ by
678: $\overline{\nu}\,\overline{c}$. For {\em stable}
679: bulk phases, $D$ must be positive, and
680: (\ref{2.3a}) then implies $g''(c)>0$. The
681: description of matrix transport is essentially
682: the same. The flux of matrix concentration
683: $\overline{c}$ is
684: \begin{equation}
685: -\overline{\delta}\,\nabla\overline{\mu} =
686: -\overline{\delta}\,\overline{\mu}'\, \nabla
687: \overline{c} = - \overline{D}\,\nabla\overline{c} 
688: \nonumber
689: \end{equation}
690: where $\overline{\mu}(\overline{c})$ is the
691: chemical potential of the matrix atoms as a
692: function of the matrix concentration
693: $\overline{c}$, $\overline{\delta}(c)$ is the
694: mobility of the matrix atoms, and $\overline{D}$
695: is the matrix diffusion coefficient given by 
696: $$
697: \overline{D} =\overline{\delta}\,\overline{\mu}'.
698: $$
699: Let $\overline{g}(\overline{c})$ be the free
700: energy density {\em as a function of} 
701: $\overline{c}$. The space filling condition and
702: $g(c) =\overline{g}(\overline{c})$ imply that
703: $$\overline{\mu}= \overline{g}'(\overline{c}) + 
704: \overline{\nu}\, \{\overline{g}(\overline{c}) -
705: \overline{c}\,\overline{g}'(\overline{c})\},$$ 
706: which is totally symmetric to (\ref{1.2b}). Then
707: the diffusion coefficient $\overline{D}$ is
708: related to $\overline{g}(\overline{c})$ by a
709: formula symmetric to (\ref{2.3a}), 
710: \begin{equation}
711: \overline{D} = \overline{\delta}\nu c
712: \overline{g}''(\overline{c}) . \label{2.3b}
713: \end{equation}
714: The space filling condition leads to a relation
715: between the mobilities $\delta$ and $\overline
716: \delta$. The linear combination $\nu c+\overline{
717: \nu}\overline{c}$ is locally conserved with flux
718: $-\nu D\nabla c - \overline{\nu} \overline{D}
719: \nabla \overline{c}$. But $\nu c+\overline{
720: \nu}\overline{c}\equiv 1$, so this flux is
721: divergence free. Let $C$ be any closed surface.
722: Use of divergence theorem yields 
723: $$\int_C (\nu\, Dc_n +\overline{\nu}\,\overline{D}
724: \overline{c}_n)\, da = 0.
725: $$
726: The space filling condition implies $\nu c_n = -
727: \overline{\nu}\overline{c}_n$, so that we get 
728: $\int_C \nu\, (D-\overline{D})\, c_n\, da = 0$.
729: This holds for all concentration fields $c$ and
730: closed surfaces $C$. Hence,
731: \begin{equation}
732: D=\overline{D} \label{2.4a}
733: \end{equation}
734: and by (\ref{2.3a}) and (\ref{2.3b}), 
735: \begin{equation}
736: \delta(c)\,\overline{\nu}\, \overline{c}\,
737: g''(c) = \overline{\delta}(\overline{c})\,\nu c\,
738: \overline{g}''(\overline{c}) . \label{2.4b}
739: \end{equation}
740: Now, 
741: $$g(c) = \overline{g}(\overline{c}) =
742: \overline{g}\left({1\over\overline{\nu}}\,
743: (1-c\nu)\right),$$
744: therefore, provided again that $\nu$ and
745: $\overline{\nu}$ {\em do not depend on} $c$,
746: $$g''(c) = \left({\nu\over\overline{\nu}}
747: \right)^2\, \overline{g}''(\overline{c}), $$
748: and (\ref{2.4b}) becomes
749: \begin{equation}
750: \delta\,\nu\,\overline{c} = \overline{\delta}\,
751: \overline{\nu}\, c . \label{2.5}
752: \end{equation}
753: This is the relation between mobilities. 
754: 
755: The integral form of Equation (\ref{2.2}) informs
756: the upcoming discussion of boundary conditions on
757: a phase interface. Let $R=R(t)$ be a time
758: sequence of closed regions in which $c=c(x,t)$ is
759: a smooth solution of (\ref{2.2}). The number of
760: precipitate particles in $R$ is 
761: $$N = \int_R c\, dx, $$
762: and the time rate of change is 
763: $$\dot{N} = \int_R c_t\, dx + \int_{\partial R} U
764: c\, da, $$
765: where $U$ is the outward normal velocity of
766: $\partial R$. Using (\ref{2.2}) to substitute for
767: $c_t$ above, and then the divergence theorem, we
768: obtain
769: $$\dot{N} = \int_{\partial R} (U\, c+ D\, c_n)\,
770: da.$$
771: The interpretation of this equation is clear:
772: The influx of precipitate atoms per unit area on
773: $\partial R$ is 
774: \begin{equation}
775: U\, c + D\, c_n . \label{2.6}
776: \end{equation}
777: 
778: Suppose now that there is a region $R_1$ of
779: matrix surrounding a region $R_2$ of precipitate.
780: The influx of precipitate atoms per unit area on
781: interface $C$ is given by (\ref{2.6}) with $c$,
782: $c_n$ evaluated on the precipitate side of $C$.
783: The outflux of precipitate atoms from the
784: surrounding matrix into $R_2$ is also given by
785: (\ref{2.6}) with $c$, $c_n$ evaluated on the {\em
786: matrix} side of $C$. Since precipitate atoms do
787: not accumulate on $C$ to form a surface density,
788: the following jump condition holds 
789: \begin{equation}
790: [U\, c + D\, c_n] = 0\quad\mbox{on $C$}.
791: \label{2.7}
792: \end{equation}
793: 
794:  In summary, conservation of precipitate is
795: expressed by the diffusion equation (\ref{2.2})
796: and the associated jump condition (\ref{2.7}).
797: The matrix density $\overline{c}$ satisfies the
798: diffusion equation and jump condition with the
799: same diffusion coefficient $D$. The space filling
800: condition (\ref{2.1}) is automatically upheld by
801: this kinetics. In addition there are
802: ``thermodynamic'' jump conditions
803: \begin{eqnarray}
804: [g'] = 0, \label{2.8a}\\
805: \left. [ g - c g' ] = - 2\sigma\kappa, \right.
806: \label{2.8b}
807: \end{eqnarray}
808: expressing local equilibrium about the phase
809: interface. These are in fact Equations
810: (\ref{1.7}), (\ref{1.8}) with $1/r$ replaced by
811: the mean curvature $\kappa$. Equations
812: (\ref{2.2}), (\ref{2.7}), (\ref{2.8a}) and
813: (\ref{2.8b}) constitute a free boundary problem
814: for the evolution of precipitate concentration
815: $c$ and the phase interfaces. 
816: 
817: \subsection{Time evolution of Gibbs energy}
818: The total Gibbs free energy is 
819: \begin{eqnarray}
820: G = G_1 + G_2 + \sigma S,  \label{2.9a}
821: \end{eqnarray}
822: where 
823: \begin{eqnarray}
824: G_1 \equiv  \int_{R_{1}} g\, dx, \quad 
825: G_2\equiv  \int_{R_{2}} g\, dx \label{2.9b}
826: \end{eqnarray}
827: are free energies of bulk matrix and precipitate
828: phases, and $S$ is the surface area of the phase
829: interface, and $\sigma$ the surface tension. The
830: time evolution of $G$ under the kinetics of the
831: free boundary problem (\ref{2.2}), (\ref{2.7}),
832: (\ref{2.8a}) and (\ref{2.8b}) is examined here.
833: To compute the rates of change of $G_1$ and
834: $G_2$, it is useful to formulate a transport
835: equation for the free energy density $g(c)$: 
836: $$ g_t = g'\, c_t = g'\, \nabla\cdot (D\nabla c)
837: = \nabla\cdot (D\nabla g) - D\, g''\,|\nabla c|^2
838: $$
839: or
840: \begin{eqnarray}
841: g_t - \nabla\cdot (D\nabla g) = - D\, g''|\nabla
842: c|^2 . \label{2.10}
843: \end{eqnarray}
844: 
845: In stable bulk phases, $g''>0$, therefore the
846: source density in (\ref{2.10}) is generally
847: negative. Let us now look at the rate of $G_2$. 
848: \begin{eqnarray}
849: \dot{G}_2 = \int_{R_{2}} g_t\, dx + \int_C
850: U\, g\, da, \label{2.11}
851: \end{eqnarray}
852: where $U$ is the normal velocity of the phase
853: interface, positive if outward from precipitate. 
854: Inserting $g_t$ from (\ref{2.10}) and using the
855: divergence theorem, (\ref{2.11}) becomes 
856: \begin{eqnarray}
857: \dot{G}_2 = \int_C (U\, g+D\, g_n)\big|_2 \, da -
858: \int_{R_{2}} D g'' |\nabla c|^2 \, dx. 
859: \label{2.12a}
860: \end{eqnarray}
861: In the surface integral, the subscript 2 means
862: evaluation on precipitate side of interface.
863: Similarly, 
864: \begin{eqnarray}
865: \dot{G}_1 = -\int_C (U\, g+D\, g_n)\big|_1\, da -
866: \int_{R_{1}} D g'' |\nabla c|^2 \, dx,
867: \label{2.12b}
868: \end{eqnarray}
869: where subscript 1 means evaluation on matrix side
870: of interface. The rate of change of the surface
871: energy $\sigma S$ in (\ref{2.9a}) is given by the
872: standard formula of differential geometry, 
873: \begin{eqnarray}
874: \sigma\dot{S} = 2\sigma\int_C \kappa\, U\, da.
875: \label{2.12c}
876: \end{eqnarray}
877: Adding Equations (\ref{2.12a}) to (\ref{2.12c}),
878: we obtain the rate of change of the total free
879: energy
880: \begin{eqnarray}
881: \dot{G} = &-& \int_{R_{1}+R_{2}} D g'' |\nabla
882: c|^2\, dx \nonumber\\
883: &+& \int_C \{ [U\, g+D\, g_n] +
884: 2\sigma\kappa U\}\, da,     \label{2.13}
885: \end{eqnarray}
886: Here the jump $[\dots ]$ denotes values on
887: precipitate side minus values on matrix side.
888: Using the continuity condition $[g']=0$, it
889: follows that
890: $$ [Ug + Dg_n] = U\, [g] + g'\, [D\, c_n].$$
891: In the right hand side, $g'$ denotes a well
892: defined value on the phase interface. By
893: conservation jump condition (\ref{2.7}), $[D\,
894: c_n] = - U\, [c]$, hence
895: \begin{eqnarray}
896: [Ug + Dg_n] = U \, [g-c g']. \label{2.14}
897: \end{eqnarray}
898: By the thermodynamic jump condition (\ref{2.8b}),
899: $[g - cg'] = -2\sigma\kappa$, so finally,
900: $$[Ug + Dg_n] = - 2\sigma U \kappa,
901: $$
902: and the energy rate formula (\ref{2.13}) reduces
903: to
904: \begin{eqnarray}
905: \dot{G} = -\int_{R_{1}+R_{2}} D g'' |\nabla c|^2
906: \, dx.     \label{2.15}
907: \end{eqnarray}
908: The integral is negative definite. Notice that
909: surface integral contributions to $\dot{G}$
910: cancel. This point is examined by direct physical
911: argument to see what it really means. 
912: 
913: Recall that influx of precipitate atoms into
914: precipitate phase per unit area is 
915: $$ U\, c + D\, c_n.$$
916: The free energy of each precipitate atom changes
917: by an amount $[\mu] = - 2\sigma\kappa\nu$,
918: according to (\ref{1.6a}), as it crosses from
919: matrix to precipitate. Hence there is a
920: contribution to $\dot{G}$ of 
921: \begin{eqnarray}
922: - 2\sigma\, (U\nu c + D\nu c_n)\, \kappa    
923: \label{2.16a}
924: \end{eqnarray}
925: per unit area of phase interface due to crossing
926: of precipitate atoms. Similarly, influx of matrix
927: atoms into precipitate phase per unit area is
928: $$ U\, \overline{c} + D\, \overline{c}_n$$
929: and change in free energy for each matrix atom
930: crossing into precipitate is $[\overline{\mu}] = -
931: 2\sigma\kappa\overline{\nu}$, according to
932: (\ref{1.6b}). Hence, crossing of matrix atoms
933: gives another surface contribution to $\dot{G}$,
934: of
935: \begin{eqnarray}
936: - 2\sigma\, (U\overline{\nu}\, \overline{c} + D
937: \overline{\nu}\, \overline{c}_n)\,\kappa    
938: \label{2.16b}
939: \end{eqnarray}
940: per unit area. Adding (\ref{2.16a}) and 
941: (\ref{2.16b}) yields surface contribution to
942: $\dot{G}$ due to crossing of both types of atoms, 
943: $$- 2\sigma\, \{ U\, (\nu c +\overline{\nu}\,
944: \overline{c}) + D\, (\nu c_n + \overline{\nu}\,
945: \overline{c}_n)\}\,\kappa = - 2\sigma U\kappa
946: $$
947: per unit area. Here the space-filling constraint
948: has been used. From (\ref{2.12c}) it is seen that
949: $2\sigma U\kappa$ can be identified as a rate of
950: change of surface energy per unit area. Hence,
951: the total rate of free energy production per unit
952: area of phase interface is
953: $$- 2\sigma U\kappa + 2\sigma U\kappa = 0.
954: $$
955: 
956: \section{Quasistatic nuclei}
957: \label{sec-quasistatic}
958: 
959: An isolated region $R_2$ of precipitate phase,
960: called a {\em nucleus} is assumed spherical, and
961: concentration field $c$ is assumed spherically
962: symmetric. The kinetics is quasistatic if the
963: time derivative in the diffusion equation
964: (\ref{2.2}) is negligible. Here kinetics is
965: analyzed under the quasistatic assumption and
966: regimes of validity are determined a posteriori
967: by the criterion
968: \begin{eqnarray}
969: {Ua\over D} = {a\dot{a}\over D}\ll 1.   
970: \label{3.1}
971: \end{eqnarray}
972: Here $a$ is the radius of the nucleus, and the
973: normal velocity $U=\dot{a}$. $a^2/D$ represents
974: the characteristic time of diffusive transport in
975: the matrix phase surrounding the nucleus. The
976: characteristic time associated with the kinetics
977: of the radius is $a/U=a/\dot{a}$. Kinetics is
978: quasistatic if the time scale of the radius is
979: much longer than the diffusion time in the
980: surrounding matrix, as in (\ref{3.1}). 
981: 
982: Under assumptions of radial symmetry and
983: quasistatic kinetics, the diffusion equation
984: (\ref{2.2}) reduces to
985: \begin{eqnarray}
986: \partial_r (r^{2}D c_{r}) = 0. \label{3.2}
987: \end{eqnarray}
988: The conservation jump condition (\ref{2.7}) reads
989: \begin{eqnarray}
990: \dot{a}\, [c] = - [Dc_{r}]. \label{3.3}
991: \end{eqnarray}
992: The thermodynamic jump conditions (\ref{2.8a})
993: and (\ref{2.8b}) read
994: \begin{eqnarray}
995: [g'] = 0, \label{3.4a}\\
996: \, [g-cg'] = - {2\sigma\over a}\,. \label{3.4b}
997: \end{eqnarray}
998: Given the value $c_\infty$ of $c$ as
999: $r\to\infty$, Equations (\ref{3.2}) to (\ref{3.4b}) determine
1000: an ordinary differential equation (ODE) for the nuclear radius
1001: $a(t)$. The first integral of (\ref{3.2}) is 
1002: \begin{eqnarray}
1003: r^{2}D c_{r} = Q, \quad\mbox{or}\quad D c_{r} =
1004: {Q\over r^{2}}.   \label{3.5}
1005: \end{eqnarray}
1006: Here $Q$ is a function of time on $r<a$ or $r>a$,
1007: but with a possible jump at $r=a$. Regularity of
1008: $c$ at $r=0$ forces $Q\equiv 0$ on $r<a$. From
1009: (\ref{3.5}) it is now evident that
1010: \begin{eqnarray}
1011: [D c_{r}] = - {Q\over a^{2}}\,,  \nonumber
1012: \end{eqnarray}
1013: where $Q$ now refers to the value on $r>a$. The
1014: conservation jump condition (\ref{3.3}) now reads 
1015: \begin{eqnarray}
1016: \dot{a} [c] = {Q\over a^{2}}\,.   \label{3.6}
1017: \end{eqnarray}
1018: This is a differential equation for $a(t)$, once
1019: the dependence of $[c]$ and $Q$ upon the radius
1020: are determined. The two thermodynamic jump
1021: conditions (\ref{3.4a}) and (\ref{3.4b})
1022: determine $c(a-)$ and $c(a+)$, hence $[c]$ as a
1023: function of $a$. Figure \ref{f3.1} shows the
1024: graphical construction of (\ref{3.4a}) and
1025: (\ref{3.4b}). 
1026: \begin{figure}
1027: \begin{center}
1028: {\par\centering \resizebox*{0.6\columnwidth}{!}{\includegraphics{curva2.eps}} \par}
1029: %\fbox{
1030: %\epsfig{%file=curva2.eps,width=7cm
1031: %}}
1032: \vspace{0.5 cm}
1033: \caption{Geometrical construction of $c(a-)$ and
1034: $c(a+)$. }
1035: \label{f3.1}
1036: \end{center}
1037: \end{figure}
1038: 
1039: To determine $Q$, integration of (\ref{3.5}) in
1040: $r>a$ is required. Let $h(c)$ be a function such
1041: that $h'(c) = D(c)$. (\ref{3.5}) now reads
1042: \begin{eqnarray}
1043: \partial_{r} h(c) = {Q\over r^{2}}\,,  \nonumber
1044: \end{eqnarray}
1045: and integration from $r=a$ to $r=\infty$ gives
1046: \begin{eqnarray}
1047: h(c_{\infty}) - h(c_+) = {Q\over a}\quad \mbox{or}
1048: \quad Q = a\,\{h(c_{\infty}) - h(c_+)\}\,.
1049: \nonumber
1050: \end{eqnarray}
1051: Here $c_+$ means $c(a+)$. (\ref{3.6}) now reads
1052: \begin{eqnarray}
1053: \dot{a} = {h(c_{\infty}) - h(c_{+})\over a\,
1054: [c]}\,.   \label{3.7}
1055: \end{eqnarray}
1056: Since $c_+$ and $[c]$ are definite functions of
1057: $a$ as determined by the thermodynamic jump
1058: conditions (\ref{3.4a}) and (\ref{3.4b}),
1059: (\ref{3.7}) is the required ODE for $a(t)$.
1060: 
1061: \subsection{Small supersaturation}
1062: A standard limit called {\em small
1063: supersaturation} is realized when the
1064: concentration is close to planar equilibrium
1065: values $c=c_1$ in matrix phase $r>a$ and $c=c_2$
1066: in precipitate phase $r<a$. Let $\gamma$ denote
1067: disturbance of $c$ from planar equilibrium
1068: values: $\gamma=c-c_1$ in matrix phase,
1069: $\gamma=c-c_2$ in precipitate phase. For
1070: $|\gamma| \ll c_1, c_2$, (\ref{3.7}) reduces to
1071: \begin{eqnarray}
1072: \dot{a} \sim {D_{1}\, (\gamma_{\infty} -
1073: \gamma_{+})\over a\, [c]}\,.   \label{3.8}
1074: \end{eqnarray}
1075: Here $D_1\equiv D(c_1)$ and $[c] = c_2 - c_1$.
1076: Also $\gamma_{\infty}=\gamma(r=\infty)$ and
1077: $\gamma_+ = \gamma(a+)$. $\gamma_+$ is determined
1078: from asymptotic limit of (\ref{3.4a}) and
1079: (\ref{3.4b}): Reduction of (\ref{3.4a}) is 
1080: \begin{eqnarray}
1081: [g'' \gamma] =0\Longrightarrow g''(c_1) \gamma_+
1082: = g''(c_2)\gamma_{-} = m\,\mbox{(common value),}  
1083: \label{3.9a}
1084: \end{eqnarray}
1085: and given $m$, reduction of (\ref{3.4b}) is 
1086: \begin{eqnarray}
1087: [c]\, m = {2 \sigma\over a}\quad\mbox{or}\quad 
1088: g''_1\,Ê[c]\,\gamma_{+} = {2 \sigma\over a}\,.
1089: \label{3.9b}
1090: \end{eqnarray}
1091: Substituting this result for $\gamma_+$ into
1092: (\ref{3.8}) gives the reduced ODE
1093: \begin{eqnarray}
1094: \dot{a} \sim {D_{1}\over [c]\, g''_{1}}\, {1\over
1095: a}\,\left(g''_1\, \gamma_{\infty} - {2 \sigma
1096: \over a\, [c]}\right)\,. \label{3.10}
1097: \end{eqnarray}
1098: This equation indicates that clusters whose radii are smaller
1099: than the critical value
1100: \begin{eqnarray}
1101: a_c \equiv  {2\sigma\over g''_{1}\,\gamma_{\infty}
1102: \, [c]} \,, \label{3.11}
1103: \end{eqnarray}
1104: shrink and disappear. Supercritical clusters of radius larger
1105: than the critical radius $a=a_c$ grow steadily according to
1106: (\ref{3.11}).
1107: 
1108: Is this small supersaturation kinetics consistent
1109: with the quasistatic criterion (\ref{3.1})?
1110: Natural unit of $a$ is $a_c$ given by (\ref{3.11}),
1111: which is the standard formula for critical radius
1112: in small supersaturation limit. Given $a=O(a_c)$,
1113: an order of magnitude estimate of $\dot a$ based
1114: on (\ref{3.10}) is 
1115: \begin{eqnarray}
1116: \dot{a} = O\left( {D_{1}\gamma_{\infty}\over
1117: a_{c}\, [c]}\right)\nonumber
1118: \end{eqnarray}
1119: which can be rearranged as
1120: \begin{eqnarray}
1121: {a\dot{a}\over D_{1}} = O\left( {\gamma_{\infty} 
1122: \over [c]}\right)\label{3.12}
1123: \end{eqnarray}
1124: The analysis here is based on $|\gamma|\ll c_1 ,
1125: c_2$. But for quasistatic kinetics, we need 
1126: \begin{eqnarray}
1127: |\gamma_{\infty}|\ll [c]= c_2 - c_1.\label{3.13}
1128: \end{eqnarray}
1129: 
1130: The reduced ODE (\ref{3.10}) indicates natural
1131: units of $\gamma$, space and time. The limit
1132: (\ref{3.13}) is embodied by measuring $\gamma
1133: \equiv c-c_e$ in units of $\epsilon\, c_e$, where
1134: $\epsilon>0$ is a gauge parameter and limit
1135: $\epsilon\to 0+$ is considered. Given this unit
1136: of $\gamma$, the order of magnitude of $a_c$ in
1137: (\ref{3.11}) is $l/\epsilon$, $l\equiv
1138: \sigma/([c] g''_1 c_e)$. $l/\epsilon$ is adopted
1139: as unit of length. Finally, the unit of time
1140: which gives $\dot a$ as in (\ref{3.12}) is
1141: $\tau_e/\epsilon^3$, $\tau_e\equiv (l^2/D_1)\,
1142: ([c]/c_e)$. This system of units is summarized in
1143: the following Table:\\
1144: 
1145: 
1146: \begin{tabular}{||c|c||cr||}    
1147: \hline Variable& Unit\\
1148: \hline $\gamma\equiv c - c_e$ & $\epsilon c_e$  
1149: \\
1150: \hline $x$& ${l\over\epsilon}$, $l\equiv {\sigma
1151: \over [c] g''_{1} c_{e}}$
1152: \\
1153: \hline $t$ & ${\tau_{e}\over\epsilon^{3}}$,
1154: $\tau_e \equiv {l^{2} [c]\over D_{1} c_{e}}$\\
1155: \hline $G$ & ${\sigma l^{2}\over \epsilon^{2}}$\\
1156: \hline
1157: \end{tabular}
1158: \\
1159: \\
1160: {\em Scaling Table for small supersaturation.}
1161: \bigskip
1162: 
1163: Given these units, the nondimensional version of
1164: (\ref{3.10}) is 
1165: \begin{eqnarray}
1166: \dot{a} = {1\over a}\,\left(\gamma_{\infty} -
1167: {2\over a}\right)\,.\label{3.14}
1168: \end{eqnarray}
1169: 
1170: \subsection{Quasistatic energetics}
1171: Given evolution of precipitate concentration
1172: field $c$, corresponding changes in total free
1173: energy of medium are quantified by the rate
1174: formula (\ref{2.15}). Now suppose concentration
1175: field $c$ corresponds to a nucleus undergoing
1176: quasistatic evolution as set by criterion
1177: (\ref{3.1}). It seems reasonable to approximate
1178: the integral in (\ref{2.15}) using the
1179: quasistatic approximation to $c$ which satisfies
1180: Equations (\ref{3.2}) to (\ref{3.4b}). The rate
1181: formula (\ref{2.15}) reduces to 
1182: \begin{eqnarray}
1183: \dot{G} = -\int_{0}^{\infty} D g'' c_r^2 4\pi r^2
1184: \, dr.\nonumber
1185: \end{eqnarray}
1186: >From (\ref{3.5}), $r^2 D c_r =Q$ for $r>a$, and
1187: $r^2 D c_r=0$ for $r<a$, so this reduces to
1188: \begin{eqnarray}
1189: \dot{G} = - 4\pi Q\int_{a+}^{\infty} g'' c_r\,
1190: dr = - 4\pi Q\,\{ g'(c_{\infty}) - g'(c_+)\}.
1191: \nonumber
1192: \end{eqnarray}
1193: Substituting for $Q$ from (\ref{3.6}), 
1194: \begin{eqnarray}
1195: \dot{G} = - [c]\,\{ g'(c_{\infty}) -
1196: g'(c_+)\} (4\pi a^{2} \dot{a}). \label{3.15}
1197: \end{eqnarray}
1198: In the right hand side, $[c]$ and $g'(c_+)$ are
1199: definite functions of radius $a$, determined by
1200: thermodynamic jump conditions (\ref{3.4a}) and
1201: (\ref{3.4b}). Hence, in quasistatic evolution,
1202: $G$ is effectively a function of the radius $a$.
1203: >From (\ref{3.15}), it is seen that $G(a)$ obeys
1204: the differential relation
1205: \begin{eqnarray}
1206: dG = - [c]\,\{ g'(c_{\infty}) -
1207: g'(c_+)\} (4\pi a^{2} da). \label{3.16}
1208: \end{eqnarray}
1209: In conventional descriptions of nucleus
1210: energetics, the Gibbs free energy cost of a
1211: nucleus regarded as a function of the instantaneous
1212: radius $a$, independent of past history. The
1213: specific formula is
1214: \begin{eqnarray}
1215: G = 4\pi\sigma a^2 - g_{b}{4\pi\over 3}\,
1216: a^{3}  \label{3.17a}
1217: \end{eqnarray}
1218: or in differential form, 
1219: \begin{eqnarray}
1220: dG = 8\pi \sigma a da - g_b 4\pi a^{2} da.
1221: \label{3.17b}
1222: \end{eqnarray}
1223: Here $\sigma$ is surface tension and $4\pi\sigma
1224: a^2$ represents surface energy. $g_b$ is a
1225: constant with units of energy density, sometimes
1226: called ``chemical driving force''. It is the free
1227: energy released per unit volume of increase of
1228: precipitate phase. In its present form
1229: (\ref{3.16}) does not have obvious correspondence
1230: to (\ref{3.17b}). A correspondence is brought out
1231: by reformulation of (\ref{3.16}) with help of
1232: thermodynamic jump conditions (\ref{3.4a}) and
1233: (\ref{3.4b}). By (\ref{3.4a}), it follows that
1234: $g'(c_-) = g'(c_+)=$ common value $m$, and hence 
1235: (\ref{3.4b}) gives 
1236: $$[g] - [c]\, m = - {2\sigma\over a}\,,$$
1237: or, equivalently,
1238: $$[c]\, g'(c_+) = [g] + {2\sigma\over a}\,.$$
1239: Hence (\ref{3.16}) becomes
1240: \begin{eqnarray}
1241: dG = 8\pi\sigma\, a\, da + \{ [g] - [c]\,
1242: g'(c_{\infty})\} (4\pi a^{2} da). \label{3.18}
1243: \end{eqnarray}
1244: The first term on RHS is differential of surface
1245: energy, same as in (\ref{3.17b}). An apparent
1246: correspondence between (\ref{3.17b}) and
1247: (\ref{3.18}) is completed by identifying the
1248: chemical driving force $g_b$,
1249: \begin{eqnarray}
1250: g_b \equiv - [g] + [c]\, g'(c_{\infty}).
1251: \label{3.19}
1252: \end{eqnarray}
1253: Recalling that $[g]$ and $[c]$ are functions of
1254: radius $a$ as determined by thermodynamic jump
1255: conditions, it is evident that the chemical
1256: driving force (\ref{3.19}) is generally a
1257: function of cluster radius, and {\em not a
1258: constant} as in conventional nucleus energetics.
1259: 
1260: We can now write Eq.\ (\ref{3.6}) for $\dot{a}$ in terms of $G_a
1261: = - \{ g'(c_{\infty}) - g'(c_+)\}\, 4\pi a^{2}[c]$ given by
1262: (\ref{3.16}). The result is
1263: \begin{eqnarray}
1264: \dot{a} = - {\int_{c_{+}}^{c_{\infty}} D(c)\, dc\over
1265: g'(c_{\infty}) - g'(c_+)}\, {G_{a}\over 4\pi [c]^{2} a^{3} }.
1266: \label{3.16a}
1267: \end{eqnarray}
1268: 
1269: 
1270: \subsection{Energetics at small supersaturation}
1271: In the small supersaturation limit, $\int_{c_{+}}^{c_{\infty}}
1272: D(c)\, dc \sim D(c_1)\, (c_{+}-c_{\infty}) = D_1\, (\gamma_{+}-
1273: \gamma_{\infty})$, $g'(c_{\infty}) - g'(c_+)\sim g''_1\, (
1274: \gamma_{+} - \gamma_{\infty})$ and (\ref{3.16a}) reduces to
1275: \begin{eqnarray}
1276: \dot{a} = - {G_{a} D_1\over 4\pi g''_1 [c]^{2} a^{3} } .
1277: \label{3.16b}
1278: \end{eqnarray}
1279: Inserting the approximate chemical driving force $g_b = [c]
1280: g'(c_\infty) -[g] \sim g''_1 \gamma_{\infty}$ from (\ref{3.19})
1281: into the free energy (\ref{3.17a}), we obtain 
1282: \begin{eqnarray}
1283: G \sim 4\pi\sigma a^2 - [c]\, g''_1\,
1284: \gamma_{\infty} \, \left({4\pi\over 3}\, a^3
1285: \right)\,.   \label{3.21}
1286: \end{eqnarray}
1287: Notice that the asymptotic chemical driving force, 
1288: \begin{eqnarray}
1289: g_b \sim [c]\, g''_1\,\gamma_{\infty},
1290: \label{3.22a}
1291: \end{eqnarray}
1292: is in fact a constant independent of radius $a$, as in
1293: the conventional wisdom. An alternative expression
1294: for $g_b$ is useful: In certain experiments
1295: surface tension $\sigma$ and critical radius
1296: $a_c$ are measured observables, so it is
1297: convenient to represent $g_b$ in terms of
1298: $\sigma$ and $a_c$, 
1299: \begin{eqnarray}
1300: g_b = {2\sigma\over a_{c}}\,.  \label{3.22b}
1301: \end{eqnarray}
1302: This formula follows from the condition  $G'(a_c)
1303: = 0$. The natural unit of free energy is $\sigma
1304: l^2$. This is entered into the last column of the
1305: scaling table. Then the dimensionless version of the
1306: free energy formula (\ref{3.21}) is 
1307: \begin{eqnarray}
1308: G = 4\pi\, a^2 - \gamma_{\infty}\, {4\pi\over
1309: 3}\, a^3 .   \label{3.23}
1310: \end{eqnarray}
1311: 
1312: \subsection{Identification of rate constant in BD kinetics}
1313: We can now compare Eq.\ (\ref{dot-a}) for the growth of a
1314: (large) cluster radius in BD kinetics with the corresponding
1315: equations (\ref{3.16a}) and (\ref{3.16b}) obtained from our
1316: macroscopic description. We find 
1317: \begin{eqnarray}
1318: k_d &=& {\tau\, \int_{c_{+}}^{c_{\infty}} D(c) dc\over
1319: g'(c_{\infty}) - g'(c_{+})}\, \left( {c_{2}\over
1320: [c]}\right)^{2}\, 4\pi a\label{5.15a}\\ 
1321: & \sim & {D_{1}\tau\over g''_{1}}\, \left( {c_{2}\over
1322: [c]}\right)^{2}\, 4\pi a ,   \label{5.15}
1323: \end{eqnarray}
1324: in the small supersaturation limit.
1325: 
1326: With this determination of $k_d$, the continuum
1327: limit equation (\ref{5.12}) of cluster kinetics
1328: is completely specified. It is easy to check that
1329: the corresponding microscopic rate constant
1330: determined by Penrose et al \cite{pen83,pen84}
1331: is also proportional to $a$ for large clusters.  
1332: One might wonder about the micromolecular basis of
1333: (\ref{5.15}). While that requires further work,
1334: here is a curious observation which might become
1335: relevant to this question: Recall the {\em
1336: mobility} $\delta$ defined in the formulation of
1337: the macroscopic transport theory of Section
1338: \ref{secMkinetics}. It is related to the diffusion
1339: $D$ by (\ref{2.3a}), 
1340: $$ D = \delta\, (1-\nu c)\, g''.$$
1341: Hence the ratio $D_1/g''_1$ in (\ref{5.15}) is
1342: given by
1343: \begin{eqnarray}
1344: {D_{1}\over g''_{1}} = \delta_1\, (1-\nu c_1).   
1345: \label{5.16}
1346: \end{eqnarray}
1347: In the next Section, we show that the volume fraction $\nu
1348: c_1$ of precipitate in matrix phase is small, $\nu c_1
1349: \sim (0.04521$ nm$^3)\, (2.24$ nm$^{-3})\approx
1350: 0.10$, for coarsening experimental data in binary alloys
1351: \cite{XH}. Then $D_1/g''_1\approx
1352: \delta_1$ and formula (\ref{5.15}) for $k_d$ reduces to
1353: \begin{eqnarray}
1354: k_d \approx \delta_{1}\tau\left({c_{2}\over
1355: [c]}\right)^{2}\, 4\pi a .  \nonumber
1356: \end{eqnarray}
1357: 
1358: Let us notice that this rate constant $k_d$ is linear in the
1359: cluster radius so that it scales as $n^{{1\over 3}}$ with
1360: cluster size. This contrasts with the usual Turnbull-Fisher
1361: rate constant that scales as $n^{{2\over 3}}$ \cite{reg01}, but
1362: it agrees with the microscopic considerations of Penrose et al
1363: \cite{pen83,pen84}. The scaling $n^{{1\over 3}}$ has been
1364: shown to yield the Lifshitz-Slyozov distribution function for
1365: cluster radii \cite{juanjo}. The latter is a roughly adequate
1366: description of coarsening \cite{XH,gas01}. 
1367: 
1368: %\setcounter{equation}{0}
1369: \section{Material and energy parameters of
1370: kinetic theory determined from Xiao-Haasen data}
1371: \label{sec-XH}
1372: 
1373: \subsection{Small supersaturation}
1374: The nucleation in Xiao-Haasen's (XH) paper \cite{XH} takes
1375: place under low supersaturation: The initial
1376: sample has uniform composition, with mole
1377: fraction $\chi\equiv 0.12$, or 12 \% of Al in a
1378: Ni matrix. {\em Equilibrium} mole fraction of Al
1379: in matrix phase at annealing temperature of 773 K
1380: is $\chi_1 = 0.101$, or 10.1 \%. Equilibrium mole
1381: fraction of Al in precipitate phase is $\chi_2 =
1382: 0.230$, or 23 \%. Hence, {\em supersaturation} as
1383: a function of equilibrium concentration has
1384: initial value
1385: $${\chi-\chi_{1}\over \chi_{1}} \approx {0.120
1386: -0.101\over 0.101} \approx 0.19.$$
1387: Mole fractions are converted into number
1388: densities: XH report a molar volume of
1389: precipitate phase $V_m \approx 27.16\times
1390: 10^{-6}$ m$^3$. Conversion to an atomic volume by
1391: Avogadro's number gives
1392: $$ \nu_m\equiv {V_{m}\over N_{A}}\approx
1393: 4.51\times 10^{-29} \,\mbox{m}^3$$
1394: or $\nu_m\approx 0.0451$ nm$^3$. 
1395: 
1396: XH also report a lattice constant of $a\approx
1397: 0.356$ nm for the precipitate phase, and atomic
1398: volume corresponding to this lattice constant is
1399: $\nu_m = a^3 \approx 0.0451$ nm$^3$. It is clear
1400: that the molar volume $V_m$ was derived from the
1401: lattice constant. A lattice constant for the
1402: matrix phase is not reported explicitly, so it is
1403: presumably close to the value $a\approx 0.356$ nm
1404: of the precipitate phase. It seems there is an
1405: implicit assumption: Local structure of alloy in
1406: a lattice, with sites that can be occupied by Al
1407: or Ni atoms. In this case, atomic volumes of Al
1408: and Ni are de-facto the same, i.e., 
1409: $$
1410: \nu_m =\overline{\nu}_m \approx
1411: 0.0451\,Ê\mbox{nm}^3 .$$ 
1412: Now number densities of Al and Ni easily follow.
1413: For instance, $c_1$, the equilibrium number
1414: density of Al in matrix phase is 
1415: $$c_1 ={\chi_{1}\over \nu} \approx {0.101\over
1416: 0.0451}\,\mbox{nm}^{-3}\approx 2.24\,
1417: \mbox{nm}^{-3} .$$
1418: Table 1 gives initial concentration of
1419: Al in matrix phase, and equilibrium
1420: concentrations $c_1$ and $c_2$ of Al in matrix
1421: and precipitate phases. 
1422: \bigskip
1423: 
1424: \begin{tabular}{||c|c|c||cr||}    
1425: \hline $c$& $c_1$& $c_2$\\
1426: \hline
1427: 2.66 & 2.24 & 5.10   \\ \hline
1428: \end{tabular}
1429: \\
1430: 
1431: \noindent {\bf Table 1}: Number densities
1432: (nm$^{-3}$).
1433: 
1434: \subsection{Nucleation energetics}
1435: In XH, the nucleus energy takes the classic form 
1436: \begin{eqnarray}
1437: G = 4\pi\sigma a^2 -g_b\,\left( {4\pi\over 3}\,
1438: a^{3}\right)\,.      \label{4.1}
1439: \end{eqnarray}
1440: Here $\sigma$ is surface tension. A value $\sigma
1441: \approx 0.014$ J m$^{-2} = 1.4\times 10^{-20}$ J
1442: nm$^{-2}$ is deduced from interpreting coarsening
1443: data with the Lifshitz-Slyozov (LS) theory. XH deals
1444: with chemical driving force $g_b$ in two ways: 
1445: 
1446: \noindent (i) {\em ``Experimental''}. The
1447: distribution of nuclei in the space of their
1448: radii goes through a transient phase with two
1449: peaks, separated by a local minimum at about 1.2
1450: nm. XH conjecture that the initial radius $a_c$
1451: is in fact this 1.2 nm. Estimate of $g_b$ now
1452: follows from (\ref{3.22a}), 
1453: $$ g_b \sim {2\sigma\over a_{c}}\approx 2.33
1454: \times 10^{-20}\, \mbox{J\, nm}^{-3} .$$
1455: 
1456: Given an experimental estimate of the critical
1457: radius, $a_c \approx 1.2$ nm at outset, one can
1458: estimate the number of atoms in the critical
1459: nucleus, both Al and entrained Ni:
1460: $$ {1\over \nu}\, {4\pi\over 3}\, a_c^3 \approx
1461: 160.$$
1462: Of these, 23\% are Al, so there are 
1463: $$ n_c= 0.23\times 160 = 37$$
1464: Al atoms in the critical nucleus. It seems that
1465: the critical nucleus is ``just big enough'' so
1466: energetics based on continuum theory applies. 
1467: 
1468: \noindent (ii) {\em ``Theory''}. In standard
1469: theories, $g_b$ is computed from both
1470: thermodynamic properties of precipitate and bulk
1471: phases. As such, it comes out as a constant
1472: independent of nucleus radius $a$. These
1473: derivations do not face up to the fine points of
1474: the real situation, summarized in the formula
1475: (\ref{3.19}) for $g_b$. So our approach is to
1476: stick with the determination of $g_b$ based on
1477: $\sigma$ and $a_c$, 
1478: $$g_b \approx {2\sigma\over a_{c}}\approx 2.33
1479: \times 10^{-20}\,\mbox{J\, nm}^{-3},$$ 
1480: and then see what can be said about $g(c)$. In
1481: (\ref{3.22a}) one can determine $g''_1$ because
1482: all the other quantities are known. In fact, one
1483: gets 
1484: $$g''_1 = {g_{b}\over [c]\gamma_{\infty}} \approx
1485: {2.33\times 10^{-20} \,\mbox{J\, nm}^{-3}\over
1486: (5.10 - 2.24)\, \mbox{nm}^{-3}\, (2.66 - 2.24)\,
1487: \mbox{nm}^{-3} } \approx 1.94\times 10^{-20} \,
1488: \mbox{J\, nm}^{-3} .
1489: $$
1490: The annealing temperature of 773 K defines a
1491: basic unit of energy, 
1492: $$\tau = (773\,\mbox{K})\, ( 1.38\times 10^{-23}
1493: \, \mbox{J/K}) \approx 1.07\times 10^{-20}\,
1494: \mbox{J}.
1495: $$
1496: One now has
1497: $${g''_{1}\over \tau}Ê\approx 1.81\,\mbox{nm}^3.
1498: $$
1499: This is just one number imposed upon free energy
1500: function $g(c)$ by the XH data, but it is
1501: sufficient to establish the \\
1502: \\
1503: {\em Nonideal character of Al solution in matrix
1504: phase.}\\
1505: 
1506: Suppose the solution is ideal. Then the chemical
1507: potential of an Al particle in matrix phase is
1508: given by 
1509: \begin{eqnarray}
1510: \mu(c) = \mu_1 + \tau\ln {c\over c_{1}} ,
1511: \label{4.2}
1512: \end{eqnarray}
1513: where $\mu_1$ is the chemical potential when $c=
1514: c_1$ is the planar solvability. Now the relation
1515: between $\mu(c)$ and $g(c)$ is given by
1516: (\ref{1.2b}), which is repeated here for easy
1517: reference, 
1518: \begin{eqnarray}
1519: \mu(c) = g'(c) + \nu\, (g - c\, g'). \label{4.3}
1520: \end{eqnarray}
1521: >From this equation is evident that $g''(c)$ gives
1522: information about $\mu'(c)$. In fact,
1523: differentiation of (\ref{4.3}) yields 
1524: \begin{eqnarray}
1525: \mu'(c) = (1- \nu c)\, g''(c)\Longrightarrow
1526: \mu'_1 = (1-\nu c_1)\, g''_1 .\label{4.4}
1527: \end{eqnarray}
1528: Numerical value of $\mu'_1/\tau$ based upon
1529: previous value of $g''_1$ turns out to be 
1530: $${\mu'_{1}\over\tau} \approx \{ 1- (0.0451 \,
1531: \mbox{nm}^3)\, (2.24 \,\mbox{nm}^{-3})\}\, (1.81\,
1532: \mbox{nm}^3)\approx 1.63\, \mbox{nm}^3 .
1533: $$
1534: If the ideal solution formula (\ref{4.2}) were
1535: correct, one would get 
1536: $${\mu'_{1}\over\tau} = {1\over c_{1}} \approx {1
1537: \over 2.24 \,\mbox{nm}^{-3} }\approx 0.45\,
1538: \mbox{nm}^3 ,
1539: $$
1540: which is 1/274 of value that follows from XH
1541: parameters. That the solution of Al in Ni phase
1542: is not ideal was already known to XH. They in
1543: fact considered that our chemical driving force
1544: $g_b$ is sum of two terms: (i) a chemical driving
1545: force estimated from the activity of Al component
1546: at the concentrations $\chi$ and $\chi_1$, and
1547: (ii) and the elastic strain energy per unit
1548: volume. With the corresponding expressions, they
1549: obtained a value for the critical radius, $a_c
1550: \approx 1.7$ nm, which is not too far from the
1551: experimental value, $a_c \approx 1.2$ nm
1552: \cite{XH}. 
1553: 
1554: The experimentally derived values of surface
1555: tension $\sigma$ and chemical driving force
1556: $g_b$ in (\ref{4.1}) set important parameters for
1557: macroscopic nucleation theory, namely: Energy
1558: barrier for nucleation, and typical free energy
1559: cost to add one Al particle to a nucleus. 
1560: 
1561: {\em Energy barrier} is given by 
1562: $$G_{\mbox{nuc}} = G(a_c) = {4\pi\over
1563: 3}\,\sigma a_c^2 \approx {4\pi\over 3}\,
1564: (1.4\times 10^{-20}\,\mbox{J\, nm}^{-2})\, (1.2\,
1565: \mbox{nm})^2 \approx 8.4\times 10^{-20}\,\mbox{J}
1566: .  $$
1567: Energy barrier in units of thermal energy is 
1568: $${G_{\mbox{nuc}} \over\tau} \approx 7.8 .
1569: $$
1570: A reasonable looking number. Notice that
1571: exponential 
1572: $$ e^{ - {G_{\mbox{nuc}} \over\tau} } \approx
1573: 3.7\times 10^{-4}, $$
1574: which appears in the nucleation rate is not too
1575: small. This makes anthropomorphic sense: In XH
1576: experiment, nucleation kinetics unfolds in hours
1577: and days time scales, not unduly taxing to
1578: humans. Evidently, the annealing temperature is
1579: tuned so as to achieve a ``reasonable''  nucleation
1580: rate. 
1581: \bigskip
1582: 
1583: {\em Free energy cost to add one particle.}
1584: 
1585: The number $n$ of Al particles in nucleus is
1586: related to radius $a$ by 
1587: \begin{eqnarray}
1588: n = {4 \pi\over 3}\, c_2 a^3 \Longrightarrow a =
1589: \left({3n\over 4\pi c_{2}}\right)^{{1\over 3}} .
1590: \label{4.5}
1591: \end{eqnarray}
1592: Substituting (\ref{4.5}) for $a$ in (\ref{4.1})
1593: gives nucleus energy as a function of $n$,
1594: \begin{eqnarray}
1595: G_n = (36 \pi)^{{1\over 3}}\,\sigma c_2^{-{2\over
1596: 3}} \, n^{ {2\over 3}} - g_b\, c_{2}^{-1}\, n.
1597: \label{4.6}
1598: \end{eqnarray}
1599: Free energy cost to add one particle to nucleus
1600: of $n$ particles, in units of thermal energy
1601: $\tau$, is
1602: $${\mu_{n}\over\tau} \equiv {G_{n+1} - G_{n}
1603: \over\tau}  = (36 \pi)^{{1\over 3}}\,\sigma 
1604: c_2^{-{2\over 3}}\tau^{-1} \,\{ (n+1)^{ {2\over 3}}
1605: -  n^{ {2\over 3}}\} - g_b\, c_{2}^{-1}\,\tau^{-1}.
1606: $$
1607: Since this formula is based on continuum theory,
1608: its validity requires $n\gg 1$, in which case it
1609: reduces to 
1610: \begin{eqnarray}
1611: {\mu_{n}\over\tau} \sim {2\over 3} (36
1612: \pi)^{{1\over 3}}\,\sigma c_2^{-{2\over 3}}\,
1613: \tau^{-1} n^{ -{1\over 3}} - g_b\, c_{2}^{-1}\,
1614: \tau^{-1}.    \label{4.7}
1615: \end{eqnarray}
1616: Substituting XH parameter values in RHS,
1617: \begin{eqnarray}
1618: {\mu_{n}\over\tau} \sim 1.43\, n^{ -{1\over 3}} -
1619: 0.42.    \label{4.8}
1620: \end{eqnarray}
1621: In the limit $n\to\infty$, we get $|\mu_n/\tau|
1622: \sim 0.42$. While less than 1, one would not call
1623: this value ``small compared to one''. For $n=37$,
1624: corresponding to a critical nucleus, of course
1625: one gets $\mu_n/\tau=0$. Hence there will be a
1626: range of $n$ about $n= n_c = 37$ in which $|\mu_n
1627: /\tau|\ll 1$. In particular, $|\mu_n/\tau| < 0.2$
1628: in the rather generous interval $12<n<300$. In
1629: this range of $n$, asymptotic reduction of
1630: discrete kinetic  models such as Becker-D\"oring
1631: to a Smoluchowski partial differential equation (PDE) should be
1632: reasonable. 
1633: 
1634: 
1635: \section{Discussion}
1636: \label{sec-BDasymptotics}
1637: In coarsening experiments, one starts from a
1638: situation of equilibrium at high temperature in
1639: which most clusters are monomers. Then the
1640: temperature is lowered to a value below the
1641: critical temperature, and kept there. Clusters are
1642: nucleated and grow, supersaturation changes with
1643: time so that nucleation of new clusters becomes
1644: unlikely, and the coarsening of clusters proceeds.
1645: As explained by Penrose et al \cite{pen83,pen84},
1646: this process is reasonably well described by the
1647: Becker-D\"oring model (better than by the
1648: Lifshitz-Slyozov distribution function), provided
1649: the volume fraction of precipitate is small. Let us describe
1650: the nucleation and coarsening processes in typical experiments
1651: such as XH and which parts thereof are mathematically
1652: understood. 
1653: 
1654: The nucleation process described by the BD equations starts at
1655: $t=0$ with some initial value of $\rho_1= c_{\infty}$ and no
1656: supercritical clusters. According to the XH data, the energy
1657: barrier corresponding to the initial value of the critical
1658: nucleus is relatively high, $G_{\mbox{nuc}}/\tau
1659: \approx 7.8$, so that we may consider the clusters
1660: below critical size ($n<n_c$) to be in a
1661: quasistationary state. The flux across the energy
1662: barrier is then uniform and it supplies the source
1663: for coarsening of clusters larger than the
1664: critical size. As explained in Section
1665: \ref{sec-XH}, there is a range of sizes (about the
1666: critical size) for which we may approximate the
1667: discrete BD kinetics by a continuum Smoluchowski
1668: equation for the distribution function
1669: $\rho$. The latter will describe the coarsening
1670: process and it should be approximately solved with a
1671: boundary condition obtained by matching to the
1672: solution of the BD equations for $n<n_c$. 
1673: For $t>0$, supercritical clusters are
1674: created at the rate $j$ per unit volume given in Eq.\
1675: (\ref{6.6}) below, and $\rho_1$ starts to decrease. A small
1676: change of $\rho_1$, $O(1/n_c)=O(\epsilon^3 l^3/c_2)$, induces
1677: an $O(1)$ relative change of $j$. 
1678: %Hence we adopt the
1679: %nondimensionalization suggested by the Scaling
1680: %Table of Section \ref{secMkinetics}, so that the
1681: %continuum limit of BD is the problem (\ref{5.17})
1682: %to (\ref{5.19}). We would like to describe the
1683: There is a transient situation during which 
1684: $\rho$ becomes a bimodal distribution function with peaks at sub
1685: and supercritical sizes. As time evolves, the
1686: supercritical peak increases at the expense of
1687: the subcritical peak, which disappears given
1688: enough time. Then the resulting unimodal
1689: distribution evolves toward a function with the LS
1690: scaling. 
1691: 
1692: Currently it is known that the LS distribution function
1693: \cite{LS} is a solution of the Smoluchowski equation for a very
1694: special boundary condition at small cluster size \cite{juanjo}.
1695: Although the stability properties of the LS distribution
1696: function are not completely elucidated, it seems clear that the
1697: Smoluchowski equation may have other stable solutions that may
1698: match the quasistationary distribution at small cluster sizes.
1699: The appropriate solution of the Smoluchowski equation should
1700: then describe the transient stage of coarsening. As the time
1701: advances, the peak of the distribution function at subcritical
1702: sizes decreases and disappears while the peak at supercritical
1703: sizes takes over. The latter should have the LS scaling to
1704: explain experimental \cite{XH} and numerical data
1705: \cite{pen83,pen84}. To carry out an asymptotic analysis of
1706: nucleation and coarsening providing the same qualitative
1707: description sketched here is a challenging future task.
1708: 
1709:  \section{Acknowledgements}
1710:  The present work was financed through the Spanish
1711: DGES grant PB98-0142-C04-01 and carried out during J.\ Neu's
1712: sabbatical stay at Universidad Carlos III supported by the
1713: Spanish Ministry of Education. 
1714: 
1715: 
1716: %\appendix
1717: \section{Appendix: BD kinetics for $n<n_c$}
1718: The quasistationary state is a solution of the BD
1719: equations characterized by uniform flux,
1720: \begin{eqnarray}
1721: j_{n} = k_{d,n+1}\,\left\{ e^{-{G_{n+1} -
1722: G_{n}\over\tau}}\,\rho_{n} - \rho_{n+1}
1723: \right\} \equiv j,   \label{6.1}
1724: \end{eqnarray}
1725: for $n<n_c = c_2 4\pi a_c^3/3$. At high
1726: temperature, before the experiment starts, we have
1727: the following equilibrium solution
1728: \begin{eqnarray}
1729: \rho_{eq,n} = {c\, e^{-{G_{n}\over\tau}}\over
1730: \sum_{l=1}^{\infty} l\, e^{-{G_{l}\over\tau}}}
1731: \sim c\, e^{-{G_{n}-G_{1}\over\tau}} . \label{6.2}
1732: \end{eqnarray}
1733: To write the above approximation, we have assumed
1734: that $G_2/\tau\gg G_1/\tau\gg 1$ and that
1735: $G_n$ increases with $n$. After nucleation and
1736: coarsening start, we shall assume that $\rho_n$ is
1737: close to its equilibrium value [given by the
1738: approximate expression (\ref{6.2}) at the correct
1739: temperature $\tau$], as $n\ll n_c$. Thus $\rho_n =
1740: O(c\, e^{-(G_{n}-G_{1})/\tau})$ if $n<n_c$, and
1741: much smaller than this order if $n>n_c$. This
1742: means that $e^{{G_{n}\over\tau}}\,\rho_{n}/c =
1743: O(e^{G_{1}/\tau})$ if $n <n_c$, and that
1744: $e^{{G_{n}\over\tau}}\,\rho_{n}/c = o(e^{G_{1}/
1745: \tau})$ if $n\gg n_c$. 
1746: 
1747: Equation (\ref{6.1}) can be written as
1748: $$ e^{{G_{n+1} \over\tau}}\,\rho_{n+1} - 
1749: e^{{G_{n}\over\tau}}\,\rho_{n} = -
1750: {j\over k_{d,n+1}}\, e^{{G_{n+1}\over\tau}}\,,
1751: $$
1752: and therefore easily integrated under the
1753: condition $e^{{G_{n}\over\tau}}\,\rho_{n} \to 0$
1754: as $n\to\infty$:
1755: \begin{eqnarray}
1756: e^{{G_{n}\over\tau}}\,\rho_{n} = j\,
1757: \sum_{l=n}^\infty { e^{{G_{l+1}\over
1758: \tau}}\over k_{d,l+1}}\,.   \label{6.3}
1759: \end{eqnarray}
1760: The terms in this sum are largest for $l\sim n_c$,
1761: at which $G_l$ is maximum. For such integers, the
1762: continuum approximation holds, and we can write
1763: \begin{eqnarray}
1764: { e^{{G_{l+1}\over\tau}}\over k_{d,l+1}}\sim
1765: {e^{{G_{n_{c}}\over\tau}}\over k_{d,n_{c}}}
1766: \sum_{l=n}^\infty e^{-{4\pi\sigma\over\tau}\,
1767: (a_{l}-a_{c})^{2}}.    \label{6.4}
1768: \end{eqnarray}
1769: We have used $G_{l+1} - G_{n_{c}}\sim - 4\pi\sigma 
1770: (a_l-a_c)^2$, for $l+1$ close to $n_c$. We now
1771: approximate $a_l = a_c + x\sqrt{\sigma/\tau}$ in
1772: (\ref{6.3}), and $1\sim 4\pi c_2 a_c^2 da_l = 4\pi
1773: c_2 a_c^2 \sqrt{\tau/\sigma}\, dx$, so that
1774: (\ref{6.3}) becomes 
1775: \begin{eqnarray}
1776: e^{{G_{n}\over\tau}}\,\rho_{n} \sim {j\over
1777: D_{1}}\, {[c]\over c_{2}}\,{\sqrt{ {\sigma\over
1778: \tau}}\over \gamma_{\infty}}\, e^{{G_{n_{c}}\over
1779: \tau}}\, 2\, \int_{\sqrt{{\sigma\over\tau}}\,
1780: (a-a_{c})}^\infty  e^{-4\pi x^{2}}\, dx.  
1781: \label{6.5}
1782: \end{eqnarray}
1783: The equilibrium solution of the BD equations is 
1784: (\ref{6.2}). If we impose that $\rho_n \sim
1785: \rho_{eq,n}$ as $n\ll n_c$, (\ref{6.2}) and
1786: (\ref{6.5}) yield 
1787: \begin{eqnarray}
1788: j\sim \sqrt{{\tau\over\sigma}}\, {c D_{1}
1789: \gamma_{\infty} c_{2}\over[c]}\,
1790: e^{-{G_{n_{c}}-G_{1}\over\tau}} \,. 
1791: \label{6.6}
1792: \end{eqnarray}
1793: This constant flux is exponentially small because$(G_{n_{c}}
1794: -G_1)/\tau \sim G(a_c)/\tau \gg 1$. Notice that it is also
1795: proportional to the supersaturation $\gamma_\infty$. It is
1796: clear that a small change in the supersaturation,
1797: $\delta\gamma = O(\gamma_{\infty}/n_c)$, produces
1798: an $O(1)$ change in $n_c$ and in $G_{n_{c}}$,
1799: $\delta n_c = - 3 n_c \delta\gamma/
1800: \gamma_{\infty}$ and $\delta G_{n_{c}} = - 2
1801: G_{n_{c}} \delta\gamma/\gamma_{\infty}$, and hence
1802: a significant relative change of $j$ in
1803: (\ref{6.6}):
1804: \begin{eqnarray}
1805: {\delta j\over j}Ê\sim \exp\left({g''_{1} [c]
1806: n_{c}\over\tau}\,\delta\gamma\right)\,.
1807: \label{6.6b}
1808: \end{eqnarray}
1809: 
1810: Notice that, in the continuum limit, the flux
1811: $j_n$ becomes 
1812: \begin{eqnarray}
1813: j_n &\sim & -{D_{1}\tau\over g''_{1}[c]^{2}}\, {e^{-
1814: {G(a)\over\tau}}\over 4\pi a}\, {\partial\over \partial a}
1815: \left( e^{{G(a)\over\tau}}\, {\rho\over a^{2}}\right)\nonumber\\
1816: &=& {D_{1}\tau\over g''_{1}[c]^{2}}\, {1\over 4\pi a}\,
1817: \left( {\rho\over\tau a^{2}}\, {\partial G\over \partial a}
1818: + {\partial\over
1819: \partial a}\left( {\rho\over a^{2}}\right)\right)\,. 
1820: \label{6.6d}
1821: \end{eqnarray}
1822: The relation between drift and diffusion coefficients here is
1823: $G_a/\tau$ in agreement with the formulas provided by
1824: Nonequilibrium Thermodynamics; see Ref.\ \cite{reg01}. 
1825: 
1826: \begin{thebibliography}{10}
1827: \bibitem{zia68} 
1828: A. Ziabicki, {\it Generalized theory of nucleation
1829: kinetics. I. General  formulations}. J. Chem. Phys. {\bf 48},
1830: 4368-4374 (1968). 
1831: 
1832: \bibitem{mar95} 
1833: I. V. Markov, {\it Crystal growth for beginners}. (World Sci.,
1834: Singapore 1995). 
1835: 
1836: \bibitem{XH} 
1837: S. Q. Xiao and P. Haasen, {\it HREM
1838: investigation of homogeneous decomposition  in a
1839: Ni-12 at.\% Al alloy}. Acta metall. mater. {\bf
1840: 39}, 651-659 (1991). 
1841: 
1842: \bibitem{gas01} 
1843: U. Gasser, E.R. Weeks, A. Schofield, P.N. Pursey
1844: and D. A. Weitz, {\it Real-space imaging of nucleation and
1845: growth in colloidal crystallization}. Science {\bf 292},
1846: 258-262 (2001). 
1847: 
1848: \bibitem{pag97} 
1849: I. Pagonabarraga, A. P\'erez-Madrid and J. M.
1850: Rub\'{\i}, {\it Fluctuating  hydrodynamics approach to chemical
1851: reactions}. Physica A {\bf 237}, 205-219  (1997).
1852: 
1853: \bibitem{reg01}
1854: D. Reguera, J. M. Rub\'{\i} and L. L. Bonilla, contribution to
1855: this book. 
1856: 
1857: \bibitem{zia01}
1858: A. Ziabicki, contribution to this book. 
1859: 
1860: \bibitem{leb77} J. Lebowitz and O. Penrose,
1861: {\it Cluster and Percolation inequalities for
1862: lattice systems with interactions}. J. Stat.
1863: Phys. {\bf 16}, 321-337 (1977).
1864: 
1865: \bibitem{LS}
1866: I. M. Lifshitz and V. V. Slyozov, {\it The kinetics
1867: of precipitation from  supersaturated solid
1868: solutions}. J. Phys. Chem. Solids {\bf 19}, 35-50
1869: (1961). 
1870: 
1871: \bibitem{juanjo} J. J. L. Vel\'azquez, {\it The
1872: Becker-D\"oring equations and the 
1873: Lifshitz-Slyozov theory of coarsening}. J. Stat.
1874: Phys. {\bf 92}, 195-236 (1998). 
1875: 
1876: \bibitem{pen83}  O. Penrose and A. Buhagiar,
1877: {\it Kinetics of nucleation in a lattice gas
1878: model: microscopic theory and simulation
1879: compared}. J. Stat. Phys. {\bf 30}, 219-241 
1880: (1983).  
1881: 
1882: \bibitem{pen84} O. Penrose, J. L. Lebowitz, J.
1883: Marro, M. Kalos and J. Tobochnik , {\it Kinetics 
1884: of a first order phase transition: computer
1885: simulations and theory}. J. Stat.  Phys. {\bf
1886: 34}, 399-426 (1984). 
1887: 
1888: \end{thebibliography}
1889: 
1890: %\end{multicols}
1891: \end{document}