Deterministic and stochastic dynamics in spinodal decomposition of a binary system
A model for diffusion and phase separation, which takes into account hyperbolic relaxation of the solute diffusion flux, is developed. Such a ‘hyperbolic model’ provides analysis of ‘hyperbolic evolution’ of patterns in spinodal decomposition in systems supercooled below critical temperature. Analyt...
Збережено в:
| Опубліковано в: : | Успехи физики металлов |
|---|---|
| Дата: | 2009 |
| Автори: | , , |
| Формат: | Стаття |
| Мова: | English |
| Опубліковано: |
Інститут металофізики ім. Г.В. Курдюмова НАН України
2009
|
| Онлайн доступ: | https://nasplib.isofts.kiev.ua/handle/123456789/98091 |
| Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
| Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| Цитувати: | Deterministic and stochastic dynamics in spinodal decomposition of a binary system / D.O. Kharchenko, P.K. Galenko, V.G. Lebedev // Успехи физики металлов. — 2009. — Т. 10, № 1. — С. 27-102. — Бібліогр.: 92 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraine| id |
nasplib_isofts_kiev_ua-123456789-98091 |
|---|---|
| record_format |
dspace |
| spelling |
Kharchenko, D.O. Galenko, P.K. Lebedev, V.G. 2016-04-08T20:11:11Z 2016-04-08T20:11:11Z 2009 Deterministic and stochastic dynamics in spinodal decomposition of a binary system / D.O. Kharchenko, P.K. Galenko, V.G. Lebedev // Успехи физики металлов. — 2009. — Т. 10, № 1. — С. 27-102. — Бібліогр.: 92 назв. — англ. 1608-1021 PACS numbers: 05.40.-a, 05.45.-a, 05.70.Fh, 05.70.Ln, 64.60.-i, 64.75.Nx, 81.30.-t https://nasplib.isofts.kiev.ua/handle/123456789/98091 A model for diffusion and phase separation, which takes into account hyperbolic relaxation of the solute diffusion flux, is developed. Such a ‘hyperbolic model’ provides analysis of ‘hyperbolic evolution’ of patterns in spinodal decomposition in systems supercooled below critical temperature. Analytical results for the hyperbolic model of spinodal decomposition are summarized in comparison with outcomes of classic Cahn−Hilliard theory. Numeric modelling shows that the hyperbolic evolution leads to sharper boundary between two structures of a decomposed system in comparison with prediction of parabolic equation given by the theory of Cahn and Hilliard. Considering phase separation processes in stochastic systems with a field-dependent mobility and an internal multiplicative noise, we study dynamics of spinodal decomposition for parabolic and hyperbolic models separately. It is that the domain growth law is generalized when internal fluctuations are introduced into the model. A mean field approach is carried out in order to obtain the stationary probability, bifurcation and phase diagrams displaying re-entrant phase transitions. We relate our approach to entropy-driven phase-transitions theory. Розвинуто модель дифузії та фазового розшарування, який враховує гіперболічну релаксацію дифузійного потоку. Такий «гіперболічний модель» призводить до «гіперболічного» рівнання щодо формування модульованих структур при спинодальнім розпаді в системах, охолоджених нижче критичної температури. Аналітичні результати для гіперболічного моделю спинодального розпаду порівнюються із відповідними результатами, що випливають з класичної теорії Кана—Хіллярда. За допомогою чисельного моделювання показано, що еволюція системи в гіперболічнім моделю призводить до різкої міжфазної межі у порівнянні з обчисленнями за параболічним модельом Кана−Хіллярда. З розглядом процесів фазового розшарування в стохастичних системах із залежною від поля концентрації рухливістю та внутрішнім мультиплікативним шумом вивчається динаміка спинодального розпаду для параболічного та гіперболічного моделів. Показано, що закон зростання розмірів зерен може бути узагальнений введенням у розгляд внутрішніх флюктуацій, залежних від поля концентрації. Для дослідження стаціонарної картини (функції розподілу, біфуркаційних та фазових діяграм) розвинуто теорію середнього поля, в рамках якої встановлено, що відповідні перетворення носять реверсивний характер. Показано, що опис процесу фазового розшарування у стохастичних системах із внутрішнім шумом забезпечується використанням теорії ентропійнокерованих фазових переходів. В работе развита модель для описания диффузии и фазового расслоения, которая учитывает гиперболическую релаксацию диффузионного потока. Такая «гиперболическая модель» приводит к гиперболическому уравнению описания формирования модулированных структур при спинодальном распаде в системах, охлажденных ниже критической температуры. Аналитические результаты для гиперболической модели спинодального распада сравниваются с соответствующими результатами, следующими из классической теории Кана—Хилларда. С помощью численного моделирования показано, что эволюция системы в гиперболической модели приводит к резким межфазным границам в сравнении с вычислениями согласно параболической модели Кана—Хилларда. При рассмотрении процессов фазового расслоения в стохастических системах с зависимой от поля концентрации подвижностью и внутренним мультипликативным шумом изучена динамика спинодального распада для параболической и гиперболической моделей. Показано, что закон роста размеров зерен может быть обобщен введением в рассмотрение внутренних флуктуаций, зависимых от поля концентрации. Для исследования стационарной картины (функции распределения, бифуркационных и фазовых диаграмм) развита теория среднего поля, в рамках которой установлено, что соответствующие превращения носят реверсивный характер. Показано, что описание процесса фазового расслоения в стохастических системах с внутренним шумом обеспечивается использованием теории энтропийноуправляемых фазовых переходов. We thank David Jou and Alexander Olemskoi for fruitful discussions and useful exchanges. Dmitrii Kharchenko acknowledges financial support from the Fundamental Research State Fund of Ukraine (No. GP/F26/0010). Peter Galenko acknowledges financial support from the German Research Foundation (DFG) under the Project No. HE 160/19 and DLR Agency under contract 50WM0736. Vladimir Lebedev acknowledges financial support from the Russian Foundation of Basic Research (RFBR) under the Project No. 08-02-91957. en Інститут металофізики ім. Г.В. Курдюмова НАН України Успехи физики металлов Deterministic and stochastic dynamics in spinodal decomposition of a binary system Детерміністична і стохастична динаміка в спинодальнім розпаді бінарної системи Article published earlier |
| institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| collection |
DSpace DC |
| title |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| spellingShingle |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system Kharchenko, D.O. Galenko, P.K. Lebedev, V.G. |
| title_short |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| title_full |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| title_fullStr |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| title_full_unstemmed |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| title_sort |
deterministic and stochastic dynamics in spinodal decomposition of a binary system |
| author |
Kharchenko, D.O. Galenko, P.K. Lebedev, V.G. |
| author_facet |
Kharchenko, D.O. Galenko, P.K. Lebedev, V.G. |
| publishDate |
2009 |
| language |
English |
| container_title |
Успехи физики металлов |
| publisher |
Інститут металофізики ім. Г.В. Курдюмова НАН України |
| format |
Article |
| title_alt |
Детерміністична і стохастична динаміка в спинодальнім розпаді бінарної системи |
| description |
A model for diffusion and phase separation, which takes into account hyperbolic relaxation of the solute diffusion flux, is developed. Such a ‘hyperbolic model’ provides analysis of ‘hyperbolic evolution’ of patterns in spinodal decomposition in systems supercooled below critical temperature. Analytical results for the hyperbolic model of spinodal decomposition are summarized in comparison with outcomes of classic Cahn−Hilliard theory. Numeric modelling shows that the hyperbolic evolution leads to sharper boundary between two structures of a decomposed system in comparison with prediction of parabolic equation given by the theory of Cahn and Hilliard. Considering phase separation processes in stochastic systems with a field-dependent mobility and an internal multiplicative noise, we study dynamics of spinodal decomposition for parabolic and hyperbolic models separately. It is that the domain growth law is generalized when internal fluctuations are introduced into the model. A mean field approach is carried out in order to obtain the stationary probability, bifurcation and phase diagrams displaying re-entrant phase transitions. We relate our approach to entropy-driven phase-transitions theory.
Розвинуто модель дифузії та фазового розшарування, який враховує гіперболічну релаксацію дифузійного потоку. Такий «гіперболічний модель» призводить до «гіперболічного» рівнання щодо формування модульованих структур при спинодальнім розпаді в системах, охолоджених нижче критичної температури. Аналітичні результати для гіперболічного моделю спинодального розпаду порівнюються із відповідними результатами, що випливають з класичної теорії Кана—Хіллярда. За допомогою чисельного моделювання показано, що еволюція системи в гіперболічнім моделю призводить до різкої міжфазної межі у порівнянні з обчисленнями за параболічним модельом Кана−Хіллярда. З розглядом процесів фазового розшарування в стохастичних системах із залежною від поля концентрації рухливістю та внутрішнім мультиплікативним шумом вивчається динаміка спинодального розпаду для параболічного та гіперболічного моделів. Показано, що закон зростання розмірів зерен може бути узагальнений введенням у розгляд внутрішніх флюктуацій, залежних від поля концентрації. Для дослідження стаціонарної картини (функції розподілу, біфуркаційних та фазових діяграм) розвинуто теорію середнього поля, в рамках якої встановлено, що відповідні перетворення носять реверсивний характер. Показано, що опис процесу фазового розшарування у стохастичних системах із внутрішнім шумом забезпечується використанням теорії ентропійнокерованих фазових переходів.
В работе развита модель для описания диффузии и фазового расслоения, которая учитывает гиперболическую релаксацию диффузионного потока. Такая «гиперболическая модель» приводит к гиперболическому уравнению описания формирования модулированных структур при спинодальном распаде в системах, охлажденных ниже критической температуры. Аналитические результаты для гиперболической модели спинодального распада сравниваются с соответствующими результатами, следующими из классической теории Кана—Хилларда. С помощью численного моделирования показано, что эволюция системы в гиперболической модели приводит к резким межфазным границам в сравнении с вычислениями согласно параболической модели Кана—Хилларда. При рассмотрении процессов фазового расслоения в стохастических системах с зависимой от поля концентрации подвижностью и внутренним мультипликативным шумом изучена динамика спинодального распада для параболической и гиперболической моделей. Показано, что закон роста размеров зерен может быть обобщен введением в рассмотрение внутренних флуктуаций, зависимых от поля концентрации. Для исследования стационарной картины (функции распределения, бифуркационных и фазовых диаграмм) развита теория среднего поля, в рамках которой установлено, что соответствующие превращения носят реверсивный характер. Показано, что описание процесса фазового расслоения в стохастических системах с внутренним шумом обеспечивается использованием теории энтропийноуправляемых фазовых переходов.
|
| issn |
1608-1021 |
| url |
https://nasplib.isofts.kiev.ua/handle/123456789/98091 |
| citation_txt |
Deterministic and stochastic dynamics in spinodal decomposition of a binary system / D.O. Kharchenko, P.K. Galenko, V.G. Lebedev // Успехи физики металлов. — 2009. — Т. 10, № 1. — С. 27-102. — Бібліогр.: 92 назв. — англ. |
| work_keys_str_mv |
AT kharchenkodo deterministicandstochasticdynamicsinspinodaldecompositionofabinarysystem AT galenkopk deterministicandstochasticdynamicsinspinodaldecompositionofabinarysystem AT lebedevvg deterministicandstochasticdynamicsinspinodaldecompositionofabinarysystem AT kharchenkodo determínístičnaístohastičnadinamíkavspinodalʹnímrozpadíbínarnoísistemi AT galenkopk determínístičnaístohastičnadinamíkavspinodalʹnímrozpadíbínarnoísistemi AT lebedevvg determínístičnaístohastičnadinamíkavspinodalʹnímrozpadíbínarnoísistemi |
| first_indexed |
2025-11-25T20:25:41Z |
| last_indexed |
2025-11-25T20:25:41Z |
| _version_ |
1850521116029747200 |
| fulltext |
27
PACS numbers: 05.40.-a, 05.45.-a, 05.70.Fh, 05.70.Ln, 64.60.-i, 64.75.Nx, 81.30.-t
Deterministic and Stochastic Dynamics
in Spinodal Decomposition of a Binary System
D. O. Kharchenko, P. K. Galenko*,**, and V. G. Lebedev***
Institute of Applied Physics, N.A.S. of the Ukraine,
58 Petropavlivs’ka Str.,
40030 Sumy, Ukraine
*Institut für Materialphysik im Weltraum,
Deutsches Zentrum für Luft- und Raumfahrt (DLR),
51170 Köln, Germany
**Institut für Festkörperphysik, Ruhr-Universität Bochum,
44780 Bochum, Germany
***Udmurt State University, Department of Theoretical Physics,
426034 Izhevsk, Russia
A model for diffusion and phase separation, which takes into account hy-
perbolic relaxation of the solute diffusion flux, is developed. Such a ‘hy-
perbolic model’ provides analysis of ‘hyperbolic evolution’ of patterns in
spinodal decomposition in systems supercooled below critical temperature.
Analytical results for the hyperbolic model of spinodal decomposition are
summarized in comparison with outcomes of classic Cahn−Hilliard theory.
Numeric modelling shows that the hyperbolic evolution leads to sharper
boundary between two structures of a decomposed system in comparison
with prediction of parabolic equation given by the theory of Cahn and Hil-
liard. Considering phase separation processes in stochastic systems with a
field-dependent mobility and an internal multiplicative noise, we study dy-
namics of spinodal decomposition for parabolic and hyperbolic models sepa-
rately. It is that the domain growth law is generalized when internal fluc-
tuations are introduced into the model. A mean field approach is carried out
in order to obtain the stationary probability, bifurcation and phase dia-
grams displaying re-entrant phase transitions. We relate our approach to
entropy-driven phase-transitions theory.
Розвинуто модель дифузії та фазового розшарування, який враховує гі-
перболічну релаксацію дифузійного потоку. Такий «гіперболічний мо-
дель» призводить до «гіперболічного» рівнання щодо формування моду-
льованих структур при спинодальнім розпаді в системах, охолоджених
нижче критичної температури. Аналітичні результати для гіперболічно-
го моделю спинодального розпаду порівнюються із відповідними резуль-
Успехи физ. мет. / Usp. Fiz. Met. 2009, т. 10, сс. 27—102
Оттиски доступны непосредственно от издателя
Фотокопирование разрешено только
в соответствии с лицензией
© 2009 ИМФ (Институт металлофизики
им. Г. В. Курдюмова НАН Украины)
Напечатано в Украине.
28 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
татами, що випливають з класичної теорії Кана—Хіллярда. За допомогою
чисельного моделювання показано, що еволюція системи в гіперболічнім
моделю призводить до різкої міжфазної межі у порівнянні з обчислення-
ми за параболічним модельом Кана−Хіллярда. З розглядом процесів фа-
зового розшарування в стохастичних системах із залежною від поля кон-
центрації рухливістю та внутрішнім мультиплікативним шумом вивча-
ється динаміка спинодального розпаду для параболічного та гіперболіч-
ного моделів. Показано, що закон зростання розмірів зерен може бути
узагальнений введенням у розгляд внутрішніх флюктуацій, залежних
від поля концентрації. Для дослідження стаціонарної картини (функції
розподілу, біфуркаційних та фазових діяграм) розвинуто теорію серед-
нього поля, в рамках якої встановлено, що відповідні перетворення но-
сять реверсивний характер. Показано, що опис процесу фазового розша-
рування у стохастичних системах із внутрішнім шумом забезпечується
використанням теорії ентропійнокерованих фазових переходів.
В работе развита модель для описания диффузии и фазового расслоения,
которая учитывает гиперболическую релаксацию диффузионного пото-
ка. Такая «гиперболическая модель» приводит к гиперболическому
уравнению описания формирования модулированных структур при спи-
нодальном распаде в системах, охлажденных ниже критической темпе-
ратуры. Аналитические результаты для гиперболической модели спино-
дального распада сравниваются с соответствующими результатами, сле-
дующими из классической теории Кана—Хилларда. С помощью числен-
ного моделирования показано, что эволюция системы в гиперболической
модели приводит к резким межфазным границам в сравнении с вычисле-
ниями согласно параболической модели Кана—Хилларда. При рассмот-
рении процессов фазового расслоения в стохастических системах с зави-
симой от поля концентрации подвижностью и внутренним мультиплика-
тивным шумом изучена динамика спинодального распада для параболи-
ческой и гиперболической моделей. Показано, что закон роста размеров
зерен может быть обобщен введением в рассмотрение внутренних флук-
туаций, зависимых от поля концентрации. Для исследования стацио-
нарной картины (функции распределения, бифуркационных и фазовых
диаграмм) развита теория среднего поля, в рамках которой установлено,
что соответствующие превращения носят реверсивный характер. Пока-
зано, что описание процесса фазового расслоения в стохастических сис-
темах с внутренним шумом обеспечивается использованием теории эн-
тропийноуправляемых фазовых переходов.
Key words: spinodal, diffusion, relaxation, model, liquid, structure factor,
stochastic systems.
(Received March 6, 2009; in final version March 26, 2009)
CONTENTS
1. Introduction
2. Hyperbolic model for spinodal decomposition
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 29
2.1. Hyperbolic transport
2.1.1. Equilibrium fluctuations
2.1.2. Power spectra of density and flux fluctuations
2.2. Hyperbolic spinodal decomposition
2.3. Dispersion relation and speeds for atomic diffusion
2.4. Critical parameters for hyperbolic decomposition
2.4.1. Critical wavelength for decomposition
2.4.2. Amplification rate of decomposition
2.4.3. Critical time for instability
2.4.4. Analysis of a structure function
2.5. Comparison with experimental data
3. Modelling of spinodal decomposition
3.1. 1D modelling
3.2. 3D modelling
4. Stochastic models of spinodal decomposition
4.1. Stochastic parabolic model for spinodal decomposition
4.1.1. An early stage of evolution
4.1.2. A late stage
4.1.3. Stationary case
4.1.4. Influence of external and internal noise sources
4.2. Stochasticity in hyperbolic transport
4.3. Stochastic hyperbolic model for spinodal decomposition
4.3.1. Early stages analysis
4.3.2. The effective Fokker−Planck equation for the
hyperbolic model
5. Conclusions
Acknowledgments
References
1. INTRODUCTION
Consider a process of phase separation evolving through spontaneous
growth of fluctuations, e.g., through fluctuations of concentration as
in liquid−liquid systems or fluctuations of density as in gas—liquid sys-
tems. This process is known as a spinodal decomposition, in which, be-
cause of spontaneous fluctuations growth, both phases have equivalent
symmetry but they differ only in composition. It was observed in many
experiments on polymeric mixtures [1], liquid solutions [2, 3], organic
systems [4], and metallic systems [5, 6]. This transformation has been
widely investigated by using theoretical methods as well [7—10]. Phe-
nomenological theory for decomposing phases has been constructed by
Ginzburg and Landau [11]. They described magnetic domains in tran-
sition from the normal to superconducting phase using non-conserved
order parameter. This theory has been successfully advanced by Cahn
and Hilliard for using conserved order parameter for description of
30 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
spinodal decomposition in binary liquids and solids [12]. As it has been
further derived by Cahn [13, 14], kinetics of decomposition is defined
by the growth of unstable fluctuations, and the mean size of a new
phase can be given by the most rapidly growing fluctuation.
In parallel with detailed analysis [10, 15] and tests against experi-
mental data [3, 4], the theory of Cahn and Hilliard has been further ex-
plored and developed. In particular, it has been demonstrated in com-
putational modelling [16] that the rapidly quenched liquid mixtures
under decomposition exhibit unusual non-equilibrium patterns, which
do not consistent with predictions of the Cahn and Hilliard’s theory.
These inconsistencies might be associated with the phase segregation
kinetics induced by hydrodynamic interactions following a rapid
quench below spinodal [16]. They also might be attributed to the spi-
nodal decomposition upon inhomogeneous quenching [17]. In both
cases, there is a boundary for the critical quenching above which the
classic Cahn and Hilliard’s approach has to be extended to the case of
strongly nonequilibrium decomposition provided by deep supercooling
into the spinodal region of a phase diagram. Therefore, earliest stages
and periods of decomposition under large supercooling can provide
pattern’s dynamics different from those predicted by the Cahn and
Hilliard’s theory.
A few advancements were made for strongly non-equilibrium phase
separation. Binder, Frish, and Jäckle [18] generalized the linearized
Cahn—Hilliard’s theory to the case of existence of a slowly relaxing
variable. Their calculations showed that the instability of the system is
not of the standard diffusive type, but rather it is controlled by the re-
laxation of the slow structural variable. Recently, a hyperbolic diffu-
sion equation with phase separation was derived in Refs [19, 20] from
the formalism of extended irreversible thermodynamics [21]. It has
been proposed that the hyperbolic equation is able to describe process
of rapidly quenched decomposition for short periods of time, large
composition gradients or deep supercoolings within a system. Finally,
Grasselli et al. [22] mathematically analyzed extended Cahn—Hilliard’s
equation with hyperbolic relaxation of the diffusion flux. Their treat-
ments have been devoted to one-, two-, and three-dimensional cases of
hyperbolic spinodal decomposition [23—25] to establish existence of the
global and exponential attractors for different phase spaces. These in-
vestigations [18—20, 22—25] show that evolution of phase separation in
deeply supercooled or rapidly quenched systems might be analyzed us-
ing predictions of hyperbolic transport equation.
It is known that considering the phase separation processes one need
to take into account corresponding fluctuations, which lead to memory
effects in the system dynamics. Memory effects in generalized trans-
port equations play a relevant role at high frequency or high speed of
perturbations. The influence of the non-vanishing relaxation time of
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 31
the diffusion flux on the propagation of fast crystallization fronts has
been studied [26, 27] in consistency with extended thermodynamics
[21]. The memory effects play an important role in the propagation of
phase interfaces during fast phase transitions [20].
Fluctuations for slow (i.e. internal energy, solute density etc.) and
fast variables (e.g., heat flux and atomic diffusion flux) have been con-
sidered frequently. The fluctuations of the heat flux and the viscous
pressure were stressed for the first time by Landau and Lifshitz [28],
who derived the expressions for their correlation. In Refs [29, 30], the
role of rapid fluctuations of the heat flux as a stochastic source has
been considered within the extended thermodynamic formalism. In
these works, a unified description of slow and fast heat fluctuations
has been made [31] for equilibrium and non-equilibrium steady states.
The same idea about separation of slow and fast variables to study fluc-
tuations in a system of particles with inertia has been realized within
the supersymmetric path-integral representation [32]. Besides density
fluctuations, we explore the fluctuations of the diffusion flux and in-
vestigate their role in two different kinds of descriptions: (i) when the
diffusion flux behaves as an independent fluctuating variable; (ii) the
fluctuating part of the flux behaves as a stochastic noise in the evolu-
tion equation for the density.
To study the above-mentioned spatiotemporal phenomena in sto-
chastic analysis, several analytical methods can be used. A linear sta-
bility analysis allows us to set the stability of a homogeneous state with
respect to small perturbations in systems with fluctuating sources
[33]. A fundamental study of noise-induced phase transitions can be
provided by means of dynamic renormalization group theory [34]. In
analytical investigation of noise-induced phenomena, a mean field ap-
proach is widely exploited (see Refs [35—38]). Despite the fact that the
linear stability analysis can be used for a wide class of systems, the re-
normalization group approach cannot be used directly for all models of
stochastic dynamics. The mean field theory has several modifications
for systems with non-conserved and conserved dynamics. Such ap-
proach can be extended to a large number of stochastic systems to give
a qualitative prediction of noise induced ordering and disordering
phase transitions.
The main idea of the present review is to synthesize the previous re-
sults on hyperbolic model of spinodal decomposition and to analyze its
predictions in comparison with outcomes of the parabolic model of
Cahn and Hilliard. Formally, this review can be divided in two parts:
the first one is devoted to study the hyperbolic model in the determi-
nistic case, where we compare it with parabolic model for phase separa-
tion; in the second part, we discuss properties of two above stochastic
models. In Section 2, free energy functional leading to hyperbolic gov-
erning equation for diffusion and phase separation is analyzed. Using
32 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
dispersion relation, the main propagative peculiarities, such as phase
and group speeds, are presented. Critical wavelength and time for in-
stability together with amplification rate for hyperbolic decomposition
are derived. Peculiarities of evolution of patterns are analyzed in Sec-
tion 3 by numerical solution of the hyperbolic and parabolic transport
equations. Here, we discuss a method to define the structure factor and
results obtained in comparison with the outcomes from the Cahn and
Hilliard's theory. Section 4 deals with stochastic approaches related to
study the phase separation in parabolic and hyperbolic models. Start-
ing from a stochastic parabolic model with a concentration dependent
mobility, we introduce internal fluctuations, obeying fluctuation dis-
sipation relation with an intensity reduced to the bath temperature.
We show that at late stages of the system evolution the domain size
growth (Lifshitz—Slyozov) law can be generalized in this model. Study-
ing the stationary case in the mean field approximation, we present
results of re-entrant behaviour of the effective order parameter, when
it takes nontrivial values inside a fixed interval of the system parame-
ters, and prove analytical investigations by computer simulations. In
order to discuss stochastic hyperbolic model, we start with hyperbolic
transport investigations. After, we consider stochastic hyperbolic
model for phase separation and compare results obtained for two above
stochastic models. Finally, in Section 5, a summary for the results is
proposed.
2. HYPERBOLIC MODEL FOR SPINODAL DECOMPOSITION
In this Section, we introduce the hyperbolic model for spinodal de-
composition. Starting from the hyperbolic transport equation, we
analyze equilibrium fluctuations in the system described by two
commensurable variables such as solute concentration and diffusion
flux and discuss spectral properties of these fluctuations (Subsec-
tion 2.1). In Subsection 2.2, we discuss the hyperbolic model for
phase separation in the deterministic case. The detailed study of the
model is presented in Subsections 2.3, where we obtain the disper-
sion relation, group and phase speed, and perform the correspond-
ing analysis, in Subsection 2.4 we discuss critical parameters for
the hyperbolic model and analyze the structure function behaviour.
2.1. Hyperbolic Transport
Let us consider an isothermal and isobaric binary system (both the
temperature T and the pressure P are constants) consisting of atoms
A and B . Following assumptions of Cahn [13], the system is repre-
sented as an isotropic solid solution free from imperfections and with
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 33
the molar volume independent of concentration of A- and B-atoms.
The system under study is described by the particle balance equa-
tion:
= ,
c
t
∂ −∇ ⋅
∂
J (1)
where c is the particle density of a solute in a binary system, J–
the diffusion flux, and t–the time. The diffusion flux is assumed
to be described by the Maxwell—Cattaneo relaxation equation [20,
21, 26, 27],
= ,D D c
t
∂τ + − ∇
∂
J
J (2)
where Dτ and D are the relaxation time and diffusion constant, re-
spectively. The relaxation term is negligible for steady states or
low-frequency perturbations. It becomes dominant at high frequen-
cies or fast speed of propagation.
Density Profiles. Combining Eqs (1) and (2), one gets the following
equation of a hyperbolic type
.= 2
2
2
cD
t
c
t
c
D ∇
∂
∂+
∂
∂τ (3)
Equation (3) predicts the propagation of the density profile with a
sharp front moving with a finite speed DV inside the undisturbed
system. To show this feature of hyperbolic transport, we find an
analytical solution of Eq. (3) for the semi-infinite (one-dimensional)
space by choosing the initial and boundary conditions in the form
( ,0) fc t c= , 0(0, ) ( , )c x c t x c= → ∞ = , (0, ) 0c x t∂ ∂ = (where x is a
spatial coordinate).
Under these conditions, the solution is described by the following
expressions [39, 40]:
behind the diffusion front, 0 < Dx tV≤ ,
0 0 0 = /
( , ) = ( ) exp( / ) ( )( / ) ( , )d ,
t
f a f a t x VD
c t x c c c x l c c x l f t x t+ − − + − ∫
D D
DD
t t x V
f t x I
t x V
⎡ ⎤− τ −
= ⎢ ⎥τ− ⎣ ⎦
2 2 2 1/2
12 2 2 1/2
exp( / 2 ) ( / )
( , )
2( / )
; (4)
at the diffusion front, Dx tV= ,
f a f Dc t x c c c x l c c c t= + − − ≡ + − − τ0 0 0 0( , ) ( ) exp( / ) ( ) exp( / 2 ) ; (5)
ahead of the diffusion front, DtV x< < ∞ ,
34 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
0( , ) .c t x c= (6)
Here,
1/2( / )D DV D= τ and
1/22( )a Dl D= τ are the diffusion speed and
the attenuation distance in the high-frequency limit [41], respectively,
and 1I is a modified Bessel function of the first order.
The concentration profiles described by Eqs (4)—(6) are shown in Fig.
1. In contrast with the concentration profiles described by the para-
bolic differential equation (Fick’s diffusion), the concentration pro-
files in the hyperbolic case have a sharp diffusion front which moves
with the speed DV (Fig. 1). This diffusion front separates the spatial
regions where diffusion occurs ( 0c c> at Dx V t< ; Eq. (4)) and where
diffusion is absent ( 0c c= at Dx V t> ; Eq. (6)). Therefore, the position
of the diffusion front may be examined as a depth, DtV , of density
penetration into a binary system. As it is shown in Fig. 1, the ampli-
tude of the diffusive front at Dx V t= decreases with increasing time
and spatial coordinate, according to Eq. (5).
2.1.1. Equilibrium Fluctuations
Even though solution (4)—(6) describes a smooth profile of density
(with sharp diffusion front) fluctuations always exist in a thermody-
Fig. 1. Profiles of density c at different moments 1 2 3< <t t t as predicted
by solution (4)—(6). Every profile moves with the sharp discontinuity
front, which has the diffusion speed DV . The x-coordinate of this discon-
tinuity front is given by DtV , and the amplitude of the front is decreasing
in time as − τexp[ (2 )]Dt . The density profile at 3 10 Dt t= ≥ τ is matched to
those one described by a partial differential equation of a parabolic type
(i.e., of the form of Eq. (3) with 0Dτ = ).
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 35
namic system. Indeed, techniques of light scattering or neutron scat-
tering allow to explore details of the dynamics of density perturbations
in the system, and it has fostered progress in nonequilibrium statisti-
cal mechanics [42]. Following Ref. [40], we describe features of the
system described by equations (1) and (2) related to c and J fluctua-
tions in equilibrium and to stochastic noise.
The equilibrium second moments of fluctuations of c and J are ob-
tained from the Einstein’s equation for the probability of fluctuations
[28, 43], namely
2
exp ,
2 B
s
Pr
k T
⎡ ⎤δ∝ ⎢ ⎥
⎣ ⎦
(7)
where entropy ( , )s c J is based on the independent thermodynamic vari-
ables c and J. It is known (see, e.g., Ref. [43], Chapter 15) that Ein-
stein’s equation (7) considered as an approximate Gaussian distribu-
tion function predicts the second moments correctly, but it does not
predict third and higher moments accurately. However, since we are
only interested in the second moments, we restrict ourselves in this
Section to the use of the simple Einstein formula (7).
To obtain the second differential
2sδ of entropy in Einstein’s equa-
tion (7), one needs to choose the form of the Gibbs equation for en-
tropy. The generalized Gibbs equation, which incorporates slow and
fast thermodynamic variables, is written as [21]
τμ= − − ⋅1
,Dds du dc d
T T TD
J J (8)
where u is a density of internal energy, D is related to the usual diffu-
sion coefficient D through D D c= ∂μ ∂ , and 1 2μ = μ − μ is the relative
chemical potential of the solute with respect to the one of the solvent.
We focus our attention on the fluctuations of c and J and assume du
negligible for the sake of simplicity. Then, from Eq. (8), we get the sec-
ond differential of the entropy as
2 2 21
( ) ( ) .Ds c J
T c TD
τ∂μδ = − δ − δ
∂
(9)
With the definition (7) and taking the second variation of s from Eq.
(9), the probability of fluctuations is described by
2 2( , ) exp ( ) ( ) ,
2 2
D
B B
vv
Pr c J c J
k T c k TD
⎡ ⎤τ∂μ⎛ ⎞δ δ ∝ − δ − δ⎢ ⎥⎜ ⎟∂⎝ ⎠⎢ ⎥⎣ ⎦
(10)
where v is a small volume in which the fluctuations cδ and δJ occur.
The second moments of fluctuations are given by
36 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
2 2( ) , ( ) .
( / ) ( / )
B B B
T D D T
k T k TD k TD
c J
v c v v c
〈 δ 〉 = 〈 δ 〉 = =
∂μ ∂ τ τ ∂μ ∂
(11)
In what follows, we discuss two important points: (i) the power spectra
of the fluctuations of c and J, (ii) the description of the stochastic
sources in the system (1) and (2).
2.1.2. Power Spectra of Density and Flux Fluctuations
Let us define the correlation functions for the fluctuations of c and J
in the following usual form
( , , , ) ( , ) ( , ) ,
( , , ) ( , ) ( , ) ,
c
J
C t t c t c t
C t t t t
′ ′ ′ ′≡ 〈δ δ 〉
′ ′ ′ ′≡ 〈δ δ 〉
r r r r
r, r J r J r
(12)
where r is the position vector of a point in the system. Since we con-
sider equilibrium (homogeneous, time-invariant state), one has
( , , , ) ( , ),
( , , , ) ( , ),
c c
J J
C t t C t t
C t t C t t
′ ′ ′ ′= − −
′ ′ ′ ′= − −
r r r r
r r r r
(13)
i.e. the correlation functions depend only on relative distances rr ′−
and on the difference in time t t′− . We are interested in the Fourier
transforms of the quantities in Eq. (13), namely
i t i
c c
i t i
J J
S e e C t d dt
S e e C t d dt
ω
ω
ω =
ω =
∫
∫
kr
kr
k r r
k r r
( , ) ( , ) ,
( , ) ( , ) .
(14)
These expressions represent fluctuation spectra and have special theo-
retical and practical interest, as they may be measured by means of
light scattering or neutron scattering techniques [42].
To obtain an explicit form of the fluctuation spectra we first write
Fourier transform (in space) and Laplace transform (in time) of equa-
tions (1) and (2). Using the standard procedure described in Refs [21,
42], we arrive at
D
S c S i J S c
S J S J S i D c S J
δ + δ = δ
τ δ + δ + δ = δ
k k k
k k k k
k
k
( ) ( ) (0),
( ) ( ) ( ) (0),
(15)
where ( )c Sδ k and ( )J Sδ k are the Fourier−Laplace components of cδ
and Jδ , respectively. Then, we have
2
( ) (0)11
.
( ) (0)(1 )
D
D
c S cS i D
J S Ji SS S Dk
δ δ+ τ −⎛ ⎞ ⎛ ⎞⎡ ⎤
=⎜ ⎟ ⎜ ⎟⎢ ⎥δ δ−+ τ + ⎣ ⎦⎝ ⎠ ⎝ ⎠
k k
k k
k
k
(16)
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 37
In equilibrium state (where | | 0→k ), the crossed second moments,
(0) (0)c J〈δ δ 〉k k , vanish because they have opposite time-reversal par-
ity.
Then, from Eq. (16), we have
2
2
2
2
1
( ) (0) | (0)| ,
(1 )
( ) (0) | (0)| .
(1 )
D
D
D
S
c S c c
S S Dk
S
J S J J
S S Dk
+ τ
〈δ δ 〉 = 〈 δ 〉
+ τ +
〈δ δ 〉 = 〈 δ 〉
+ τ +
k k k
k k k
(17)
To obtain the time Fourier transform, one may write
c
J
S k c S i c
S k J S i J
ω = = ℜ 〈δ = ω δ 〉
ω = = ℜ 〈δ = ω δ 〉
k k
k k
k
k
( , | |) 2 [ ( ) (0) ],
( , | |) 2 [ ( ) (0) ].
(18)
Finally, we obtain
c
D D
J
D D
Dk
S k c
D k Dk
S k J
D k Dk
ω = 〈 δ 〉
τ ω + − τ ω +
ωω = 〈 δ 〉
τ ω + − τ ω +
k
k
2
2
2 4 2 2 2 2
2
2
2 4 2 2 2 2
2
( , ) | (0) | ,
(1 2 ) ( )
2
( , ) | (0) | .
(1 2 ) ( )
The corresponding expressions for 2| (0)|c〈 δ 〉k and 2| (0)|J〈 δ 〉k in equi-
librium obtained from Eq. (11) are described by
B B
T D T
k T k T
c J
v c v c
〈 δ 〉 = 〈 δ 〉 =
∂μ ∂ τ ∂μ ∂k k
2 2 2
| (0)| , | (0)| .
( / ) ( / )
(19)
Note that, in Eq. (19), the function ( , )cS kω has a maximum at a
frequency mω given by
1/22 2(2 1) (2 ) .m D DDk⎡ ⎤ω = τ − τ⎣ ⎦ (20)
The fact that the maximum is at 0mω ≠ indicates propagation of
density waves with the speed / mk ω , in contrast with the situation
when the maximum is at 0mω = , which means purely diffusive
transport. It is clear from Eq. (20) that, to observe such a maxi-
mum, i.e., the propagation of density wave, it is needed that
1/2(2 )c Dk k D −> ≡ τ .
Thus, for ck k< , transport is diffusive, and for ck k> , the den-
sity waves may propagate.
This analysis is analogous to the analysis of the transverse veloc-
ity correlation function in generalized thermodynamics for the
Maxwell viscoelastic model [42], which is consistent with the for-
malism of extended irreversible thermodynamics [21].
38 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
2.2. Hyperbolic Spinodal Decomposition
Let us consider a region of phase diagram, in which one phase mix-
ture is unstable with respect to decomposition. This is a spinodal
region where curvature of free energy is negative:
2
2
0,
f
c
∂ <
∂
(21)
and the spinodal itself is defined by
2
2
0,
f
c
∂ =
∂
(22)
where f is the Helmholtz free energy per unit volume and c is the
concentration of B atoms.
For a given temperature, the free energy f is based on the follow-
ing variables: concentration c, gradient of concentration c∇ , and
solute diffusion flux J. Dependence of free energy on concentration
is due to existence of a diffuse interface between appearing phases
in which high concentration gradients may exist. Dependence of
free energy on diffusion flux reflects of the fact that decomposition
may proceed with high rates comparable with the speed
1/2( )D DV D= τ of the front of solute diffusion profile, where D is
the diffusion coefficient and Dτ the time for relaxation of the sol-
ute diffusion flux to its steady-state value. Thus, the selected set of
independent variables { , , }c c∇ J consists of slow conserved variable
c, fast non-conserved variable J, and gradient variable c∇ . Analo-
gous set of variables is generally analyzed within the context of ex-
tended thermodynamics [44] and it is used for models of fast phase
transformations [20].
Free Energy Density. Expanding the dependence of the free energy
density on the concentration gradients and diffusion flux, one gets
[45]
2 2
2
2 2
2
( )
( , , ) ( ,0,0) ...
2 ( )
... .
2
c c
f c f
f c c f c c
c c
f J f
J
∇ = ∇ =
= =
⎛ ⎞∂ ∇ ∂⎛ ⎞∇ = + ∇ ⋅ + +⎜ ⎟⎜ ⎟∂∇ ∂ ∇⎝ ⎠ ⎝ ⎠
⎛ ⎞∂ ∂⎛ ⎞+ ⋅ + +⎜ ⎟⎜ ⎟∂ ∂⎝ ⎠ ⎝ ⎠
0 0
J 0 J 0
J
J
J
(23)
The following points regarding Eq. (23) can be accepted. First, we
define ( ,0,0) ( )hf c f c= as the free energy density of a homogeneous
system with no gradients and fluxes. Second, the term ( )c f c∇ ⋅ ∂ ∂ ∇
must be zero because the free energy of the system does not depend
on the sign of the concentration gradient. Third, one can accept
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 39
that the derivative of the free energy with respect to the diffusion
flux is linear by the flux: ( 2)Jf∂ ∂ = αJ J as approximation consis-
tent with an extended thermodynamics [21]. And fourth, the next
terms of expansion (23) can be omitted as a nonlinear terms in flux
J. Therefore, Eq. (23) can be rewritten as
2
2( , , ) ( ) ( ) ,
2 2
c J
h
r
f c c f c c
α
∇ = + ∇ + ⋅J J J (24)
where 2 2 2( ( ) )c cr f c ∇ == ∂ ∂ ∇ 0 and coefficient Jα is a characteristic of
non-Fickian diffusion which assumed to be [21]
D
T constTD c =
τ ∂μ⎛ ⎞α = ⎜ ⎟∂⎝ ⎠
(25)
with a difference μ of the chemical potentials for both chemical
components. Within the limits of instant relaxation, i.e., 0Dτ → ,
the term with fluxes vanishes and Eq. (24) gives the free energy
density ( , )f c c∇ of the standard (Ginzburg−Landau or Cahn−Hil-
liard) form applicable for local equilibrium system.
Interpretation of Free Energy for Local Nonequilibrium States.
Going beyond local equilibrium requires re-examination in depth
such basic and conceptually relevant concepts as entropy, tempera-
ture, pressure or chemical potential under more general circum-
stances [21, 46]. Therefore, free energy density (24) has to be inter-
preted in terms of a local thermodynamic potential [47, 48].
Equation (24) defines thermodynamic potential with both local
equilibrium contribution ( )hf c and purely local nonequilibrium con-
tribution ( / 2)α ⋅J J (under spatial inhomogeneity defined by the
gradient term). Hence, for the local equilibrium part ( )hf c a local
ergodicity (i.e. the system needs to sample the phase space) is true.
However, as soon as we postulated diffusion flux with a finite re-
laxation time, this means that the local nonequilibrium contribution
α ⋅J J reflects the existence of a slow physical process, which is the
jump of solute atoms [40]. Considering ergodicity of a phase space
for nonequilibrium situation, one may well refer to statistical ef-
fects in fast spinodal decomposition due to existence of many parti-
cles (atoms and molecules) within local volumes. Since the liquid
demixing proceeds very fast, the particles have no time enough to
sample all the phase space. Thus, the number of microstates acces-
sible to each of them will be lower than in equilibrium. This will
imply an increasing in the free energy with respect to the local
equilibrium contribution ( )hf c . This is one of the ways to interpret
the nonequilibrium contribution ( / 2)α ⋅J J to the free energy (24)
that is the simplest conceivable way to express such increasing in
the free energy.
40 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
From the phenomenological pint of view, a purely non-equilibrium
contribution ( / 2)α ⋅J J to the free energy (or entropy) density for sys-
tems under spinodal decomposition controlled by atomic diffusion is
explained as a kinetic energy. Thermodynamic interpretation has been
recently made [49] for this contribution in the framework of multi-
component fluids and of dipolar systems having magnetic moments
with non-vanishing inertia. A model, which takes this kinetic contri-
bution, is called ‘hyperbolic’ model of spinodal decomposition, because
it leads to the constitutive equation of hyperbolic type. In the limit of
instantaneous relaxation, i.e. 0Dτ → , the term α ⋅J J vanishes and Eq.
(24) gives the free energy density ( , )hf c c∇ of Cahn−Hilliard’s form
[12, 13] applicable for local equilibrium system.
Free Energy Functional. Taking Eq. (24), the total Helmholtz free en-
ergy as a free energy functional is given by
2
2( , , ) ( ) ( ) ,
2 2
c J
hv
r
F c c f c c dV
⎡ ⎤α
∇ = + ∇ + ⋅⎢ ⎥
⎣ ⎦
∫J J J (26)
where V is a sub-volume of the system. Evolution of ( , , )F c c∇ J with
time t is described by
ex in
,
dF dF dF
dt dt dt
⎛ ⎞ ⎛ ⎞= +⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
(27)
where ex( )dF dt is the external exchange of the free energy and
in( )dF dt is the internal change of the free energy inside the system.
The latter is defined as a dissipative function. Using the procedure de-
scribed in Refs [19, 20] and applied to Eq. (26), one can obtain
( )2 2 2
ex
( ) ,c n c c n
dF c
r c f r c J d
dt t
′
∂⎛ ⎞ ⎡ ⎤= ∇ + − + ∇ Ω⎜ ⎟ ⎢ ⎥∂⎝ ⎠ ⎣ ⎦∫ (28)
( )2 2
in
,c c n J
dF
f r c dV
dt t
∂⎛ ⎞ ⎡ ⎤′= ⋅ ∇ − ∇ + α⎜ ⎟ ⎢ ⎥∂⎝ ⎠ ⎣ ⎦∫
J
J (29)
where Ω is the outer surface of sub-volume V, nJ is the diffusion flux
pointed by the normal vector n , and /c hf f c′ = ∂ ∂ . As it follows from
Eq. (29), the dissipative function includes the term J tα ∂ ∂J , which
has a clear physical meaning: far from equilibrium, the diffusion flux
provides additional ordering that is leading to increasing of the dissi-
pation.
Around a steady state, dissipative function (29) must decrease in
time, so that the free energy of the entire system is decreasing. This
condition implies a relation between fluxes and forces, which is, in the
simplest case, assumed to be linear [21]. For Equation (29), it gives the
following evolution equation for the diffusion flux
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 41
( )2 2 ,c c JM f r c M
t
∂′= − ∇ − ∇ − α
∂
J
J (30)
where M is the atomic mobility. Together with the atomic mass balance
,
c
t
∂ = −∇ ⋅
∂
J (31)
Eq. (30) leads to the following governing equation
( )2
2 2
2
,D c c
c c
M f r c
t t
∂ ∂ ⎡ ⎤′τ + = ∇ ⋅ ∇ − ∇⎢ ⎥⎣ ⎦∂ ∂
(32)
which is the same that it has been previously derived from the entropy
functional [19, 20]. Equation (32) is a general partial differential
equation of a hyperbolic type with the decomposition delay described
by the term
2 2/D c tτ ∂ ∂ . It allows for describe both diffusion mecha-
nism and wave propagation of chemical components.
A natural boundary condition, originating from external exchange
of the free energy (28), is given by
( )2 2 2 0,c n c c n
c
c f c J
t
∂ ′ε ∇ − − ε ∇ =
∂
(33)
where nJ and nc∇ are the projections of the diffusion flux and ‘nabla’-
operator, respectively, on the normal vector to the boundary of the
volume 0υ . Equation (33) represents a dynamical boundary condition,
which shows that the product ( ) nc t c∂ ∂ ∇ should be balanced with the
product c nJμ on the boundary of the subvolume V. From this, in par-
ticular, it follows that if the concentration is fixed, constc = , then the
flux is absent, 0nJ = , on the boundary Ω . In the standard parabolic
situation described by the Cahn—Hilliard equation ( 0Dτ → ), one has
proportionality between the flux and concentration gradient,
n nJ c∝ ∇ , and they both can be cancelled from Eq. (33). In this case,
equation (33) transforms, with some scaling constant, into the known
boundary condition analyzed by Miranville and Zelik (see Eqs (2) and
(1.2) from Refs [50, 51], respectively). Hence, Eq. (32), endowed with a
dynamic boundary condition (33), is a general partial differential
equation of hyperbolic type with the decomposition delay described by
the inertial term
2 2/D c tτ ∂ ∂ . Mathematically, the problem of Cahn—
Hilliard equation with the term
2 2/D c tτ ∂ ∂ , endowed with proper
boundary conditions, has been studied in one-, two-, and three-
dimensions [23—25] to establish existence of the global and exponential
attractors for different phase spaces.
Because we focus on the analysis of the initial stages of decomposi-
tion described by Eq. (32) (i.e., when the large concentration gradients
exist and short periods of time are important) one may neglect all
42 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
terms not linear in c. This one yields
2
2 2 4
2
,D cc c
c c
Mf c Mr c
t t
∂ ∂ ′′τ + = ∇ − ∇
∂ ∂
(34)
where
2 2
cc hf f c′′ = ∂ ∂ . As 0Dτ → , Eq. (34) transfers into the classic
Cahn—Hilliard equation [12, 13]. In the present form, Eq. (34) can be
considered as a modified Cahn—Hilliard equation, which is a linearized
partial differential equation of a hyperbolic type. This equation is true
for spinodal decomposition with local nonequilibrium diffusion (diffu-
sion with relaxation of the solute flux). Such type of decomposition is
expected for short periods of time, large characteristic velocities of
process, large concentration gradients, or under deep supercoolings.
2.3. Dispersion Relation and Speeds for Atomic Diffusion
Main characteristics of diffusion can be found from dispersion analy-
sis of the linearized hyperbolic Cahn—Hilliard equation (34). These are
the phase speed that characterizes propagation of a single (selected)
harmonic, the group speed, which is characteristic of a wave packet,
critical wavelength for decomposition, and critical time for instability,
which both characterize developing coherent structure in decomposi-
tion [52].
We consider the elementary exponential solution of Eq. (34) in the
following form
0( , ) exp[ ( ( ) )],kc z t c a i kz k t− = − ω (35)
where the dispersion relation ( )kω is given by
1/2
2 2 2
2
( ) 1
( ) .
2 4
cc c
D D D
Mk f r ki
k
⎛ ⎞′′ +
⎜ ⎟ω = − ± −
⎜ ⎟τ τ τ⎝ ⎠
(36)
The upper and lower signs for ( )kω in Eq. (36) correspond to the
branches, which are responsible for the wave propagation in the posi-
tive and negative z-directions, respectively. Qualitative behaviour for
( )kω is shown in Fig. 2.
It can be seen that the real part of ω begins to exist only from some
critical value, 0k k= (Fig. 2, a). This value defines confluence of two
branches for imaginary part of ω (Fig. 2, b). In addition, one can de-
fine other two critical values for the wave-vector k. The critical value
ck k= defines a point from which ω takes positive values of its imagi-
nary part (Fig. 2, b). For ck k> , solution (35) exponentially grows in
time and decomposition begins to proceed irreversibly. The critical
value mk k= gives a maximal positive value for ω (Fig. 2, b). Fre-
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 43
quency ( )mkω defines the mostly unstable mode with which pattern
evolves during phase decomposition.
Within the local equilibrium limit 0Dτ → , Eq. (36) arrives to
the following approximation
( )2 2 2( ) 1 1 2 ( ) .
2 D cc c
D
i
k Mk f r k⎡ ⎤′′ω ≈ − ± − τ +⎢ ⎥⎣ ⎦τ
(37)
Equation (37) shows that one of the roots is going to −∞ along
imaginary axis by the law ( ) Dk iω ∝ τ . This leads to exponential de-
cay of the solution (35). The second root of Eq. (37) is finite and it
is equivalent to classic Cahn—Hilliard relation
2 2 2( ) ( ).cc ck iMk f r k′′ω ≈ − + (38)
Thus, local equilibrium limit for dispersion relation (36) gives two
different roots: the first one is diverges and the second one ap-
proaches dispersion relation (38) of Cahn and Hilliard.
Phase Speed. The values of the wave vector 0k above which relation
(36) has the real part, Fig. 2, a, is found from condition
2
2 2
0 2
1
( ) .
2
c
cc cc
c D
r
k f f
r M
⎛ ⎞
′′ ′′⎜ ⎟= + −
⎜ ⎟τ⎝ ⎠
(39)
a b
Fig. 2. Dispersion relations for hyperbolic Cahn—Hilliard equation; Eq. (36).
(a) Real part of frequency, ( ( )).kℜ ω (b) Imaginary part of frequency, ( ( )).kℑ ω
44 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
For
2 2
0k k> and real values of ( )kω , one can obtain from Eq. (36) the
phase speed:
( )1/2
1 2 2 2( ) / ( ) ( ) (2 )p D D cc ck M f r k k− −′′υ = ℜ ω ℜ = τ τ + − , (40)
which may propagate in both positive and negative spatial directions.
The speed pυ incorporates a motion for one of separated single har-
monics. It can be compared with the predictions of the partial differen-
tial equation of a hyperbolic type for solute diffusion without phase
separation. Indeed, analysis of dispersion relation for mass transport
equation
2 2 2/ /D c t c t D cτ ∂ ∂ + ∂ ∂ = ∇ of a hyperbolic type leads to the
following expression [53]:
1/2
2 2 1/2
2
.
( )p
D D c
D
−
⎛ ⎞
υ = ⎜ ⎟τ + τ + ω⎝ ⎠
(41)
Taking into account that 0cr = , for the zero spatial atomic correla-
tion, ccMf D′′ = is the diffusion coefficient in Eq. (40), we use the rela-
tion k ∝ ω for high frequency of disturbances’ propagation. Then,
both expressions (40) and (41) lead to the same result
p D DD Vυ = τ = ω → ∞1/2( / ) with . (42)
In Equation (42), the phase speed pυ is equal to solute diffusion speed
DV , which is a maximal speed for propagation of the solute diffusion
disturbance (profile).
Imaginary part of the phase speed, ( ) (2 )p Di kℑ υ = − τ , specifies the
amplification rate for a given harmonic. With 0k k< , harmonics do not
move with possible changing of their own amplitudes. For both real and
imaginary parts of pυ (with 0k k> ), the harmonics move and change
their own amplitudes. The behaviour is shown in Fig. 3 for ( )p kυ .
Group Speed. Concentration disturbances propagating by diffusion
can be considered as an undistorted wave packet moving with the
group speed given by
( )
( ).
k
W k
k
∂ω = ±
∂
(43)
Using Eq. (36), calculation of the group speed W gives
2 2
2 2 2 1/2
2 ( 2 )
( ) .
(4 ( ) 1)
cc c
D cc c
kM f r k
W k
k M f r k
′′ +
=
′′τ + −
(44)
Dependence ( )W k is shown in Fig. 3. It specifies a speed for concentra-
tion profiles envelope. One may see, as for the phase speed pυ , the real
values for W given by Eq. (44) exist only at 0k k> . In contrast with the
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 45
behaviour of pυ , the imaginary part of ( )W k may exist only at 0k k< .
a b
Fig. 3. Phase and group speeds for hyperbolic Cahn—Hilliard equation. (a)
Real part ( )pvℜ of phase speed (solid line) and real part ( )Wℜ of group
speed (dashed-dotted line). (b) Imaginary part ( )pvℑ of phase speed (solid
line), and imaginary part ( )Wℑ of group speed (dashed-dotted line).
TABLE 1. Predictions for characteristic speeds of diffusion.
Equation Phase speed pυ Group speed W
Parabolic
diffusion
equation
ikD− with 2ik Dω = − 2ikD−
Hyperbolic
diffusion
equation
24 1
2
D
D
i Dk
k
− ± τ −
τ
2
2
4 1D
kD
Dkτ −
Parabolic
Cahn—
Hilliard
equation
2 2( )h ciMk f r k′′− + 2 22 ( 2 )h ciMk f r k′′− +
Hyperbolic
Cahn—
Hilliard
equation
2 2 21 1 4 ( )
2 D h c
D
i
Mk f r k
k
⎛ ⎞′′− ± − τ +⎜ ⎟τ ⎝ ⎠
2 2
2 2 2
2 ( 2 )
4 ( ) 1
h c
D h c
kM f r k
k M f r k
′′ +
′′τ + −
46 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
Analysis of standard parabolic and hyperbolic diffusion equations
[53] as well as parabolic and hyperbolic equations for spinodal de-
composition [52], presented in this Subsection for diffusion speeds,
leads to comparison of both approximations summarized in Table 1.
2.4. Critical Parameters for Hyperbolic Decomposition
2.4.1. Critical Wavelength for Decomposition
Cahn [13] has found a critical wavelength cλ , above which infinitesi-
mal sinusoidal fluctuation of concentration is irreversibly grown. Par-
ticularly, he confirmed the concept of Hillert [54] that cλ → ∞ with
approaching the spinodal, at which one has
2 2/ 0f c∂ ∂ = .
To find the critical wavelength for decomposition under local non-
equilibrium diffusion, we expand ( )hf c in Eq. (26) about some concen-
tration 0c that is
0 0
2 2
0
0 0 2
( )
( ) ( ) ( ) ...
2
h h
h
c c c c
c cdf d f
f c f c c c
dc dc= =
⎛ ⎞−⎛ ⎞= + − + +⎜ ⎟⎜ ⎟
⎝ ⎠ ⎝ ⎠
. (45)
The composition is represented along the z-axis by a series of sinusoi-
dal waves with components of the following form
0 cos( ),c zc c a k z− = (46)
where ca is the amplitude and zk –the frequency of concentration
wave.
Substituting Eq. (46) into Eq. (45), we perform integration of the
functional (26) over the volume υ . Then, for the difference of the
Helmholtz free energy, 0( , , ) ( )hF F c c f c dVΔ = ∇ − ∫J , between a system
with concentration (46) and a homogeneous system, respectively, one
gets:
2
2 2 .
4
c
cc c z
aF
f r k
V
Δ ⎡ ⎤′′= +
⎣ ⎦
(47)
For the reasonable cases of the positive surface tension,
2 0cr > , one
can consider two important points.
First, with 0ccf ′′ > the solution is stable against fluctuation of con-
centration of any wavelength: the free energy only increases in this
case, 0FΔ > .
Second, with 0ccf ′′ < the solution is unstable with respect to the
critical wavelength for decomposition, which can merely be found by
taking the zero value for the square bracket in Eq. (47):
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 47
1/2
22
2 .c
c
c cc
r
k f
⎛ ⎞π ⎜ ⎟λ = = π
⎜ ⎟′′−⎝ ⎠
(48)
with the critical value for wave vector given by
2 1/2( ) .c cc ck f r′′= − (49)
Therefore, with 2 2 0hd f dc < and for cλ > λ , the free energy de-
creases, 0FΔ < , and decomposition starts to proceed. Equation (49)
clearly shows that as the composition tends to the values lying in
the spinodal, 2 2 0f c∂ ∂ = , the critical wavelength approaches to in-
finity, cλ → ∞ [13, 54].
2.4.2. Amplification Rate of Decomposition
Consider a real part of the solution (35) in the following form:
0 cos( ) exp( ) cos( ) exp( ).c c a kz t a kz t+ + − −− = ω + ω (50)
In this solution, signs ‘plus’ and ‘minus’ correspond to growing or
decaying solutions, respectively, in time. Substitution of Eq. (35)
into Eq. (34) defines a real part of the frequency as follows
( )( )1/2
1 2 2 2(2 ) 1 1 4 .D D cc ck M f r k−
±
⎡ ⎤′′ω = τ − ± − τ +⎢ ⎥
⎣ ⎦
(51)
After expanding, the square root in Eq. (51) for 24 [D cck M f ′′τ +
2 2] 1cr k+ ≤ one gets in the local equilibrium limit the expression:
( )2 2 2
0
,lim cc c
D
k M f r k+
τ →
′′ω = + (52)
which is the kinetic amplification rate obtained by Cahn [13] for
purely diffusion regime. Therefore, Eq. (51) can be interpreted as
the kinetic amplification rate for both dissipative and propagative
regimes of atomic transport described by Eq. (34).
From the amplification rate +ω of decomposition, the maximum
can be obtained by differentiation of Eq. (51) with respect to zk .
The extremum condition, / 0zk+∂ω ∂ = , gives maximum frequency
( )1/2
1 2( ) (2 ) 1 1 /m m D D cc ck Mf r− ⎡ ⎤′′ω = τ − + + τ⎢ ⎥⎣ ⎦
(53)
at
( )1/2
2/ (2 ) with < 0.m cc c cck f r f′′ ′′= − (54)
48 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
From Eq. (54), there follows that maximum wavelength, 2 /m mkλ = π ,
is equal to
1/2
22
2 .c
m
cc
r
f
⎛ ⎞
⎜ ⎟λ = π
⎜ ⎟′′−⎝ ⎠
(55)
Consequently, maximum amplification rate allows for the wavelength
(55) greater in exactly 2 times the critical wavelength (49) of insta-
bility against fluctuations of concentration. This result coincides with
Cahn and Hilliard’s results for purely diffusion regime.
2.4.3. Critical Time for Instability
Let us evaluate the time of transitive period from the beginning of
instability (with the beginning of growth of infinitesimal perturba-
tion) up to the arriving into the new metastable state. For the fast-
est growth of infinitesimal perturbation, the maximal frequency
( )mkω is responsible. Therefore, substitution of Eq. (48) into disper-
sion relation (36) leads to [52]
( )1/2
2 2( ) 1 1 ( ) / .
2m m D cc c
D
i
k M f r
⎡ ⎤′′ω = − ± + τ⎢ ⎥τ ⎣ ⎦
(56)
Equation (56) adopts both real and imaginary parts for ω . Using
maximal frequency (56), solution (35) can be rewritten as
0( , ) exp( ) exp( ),k cc z t c a ikz t t− = (57)
where
2 2 1/2
2
(1 ( ) / ) 1
D
c
D cc c
t
M f r
τ
=
′′+ τ −
(58)
is the time for developing coherent structure.
Within the local equilibrium limit, 0Dτ → , we expand square root
in Eq. (58) for
2 2( ) / 1D cc cM f r′′τ = . One gets the following approxima-
tion
2
2
4
,
( )
c
c
cc
r
t
M f
≈
′′
(59)
which can be found from the predictions of pure diffusion theory
(parabolic transport equation) of Cahn and Hilliard. As a result, com-
parative analysis for parabolic and hyperbolic equations in spinodal
decomposition is given in Table 2 for dispersion relations, critical
wavelengths and times for instability.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 49
2.4.4. Analysis of a Structure Function
Different evolution of spinodally-decomposed systems exhibits differ-
ent structures (see Fig. 10). Experimentally, a typical structure after
spinodal decomposition is observed as random, interconnected patterns
with a characteristic length-scale related to maximal amplification rate
of decomposition [4]. Experimental observations using scattering show
a broad Bragg-like peaks, from which information about quenched
structure during spinodal decomposition and decomposition rate can be
read off. For such measurements, a main characteristic for the inten-
sity of scattering is a structure factor. Therefore, the structure factor
can be taken as a parameter for characterization of analyzed evolutions
and for verification of the model predictions with experimental data.
Consider the structure factor ( , )S tk , which describes the intensity
of quasi-elastic scattering observed at time t after the quenching from
the initial temperature iT up to the final temperature fT . The function
( , )S tk can be interpreted as the respective correlation function of the
concentration fluctuations, and it is defined as
( , ) ( , ) ( , ) Tf
S t c t c t= 〈δ − δ 〉k k k . (60)
To obtain expression for the time dependent structure factor (60),
TABLE 2. Predictions of parabolic and hyperbolic models.
Expression for
Parabolic Cahn−Hilliard
equation ( 0)Dτ → Hyperbolic Cahn−Hilliard equation
Dispersion
relation, ( )kω ( )2 2 2
cc ciMk f r k′′− +
1/2
2 2 2
2
( ) 1
2 4
cc c
D D D
Mk f r ki ⎛ ⎞′′ +
⎜ ⎟− ± −
⎜ ⎟τ τ τ⎝ ⎠
Critical wave-
length,
2 /c ckλ = π
( )1/2
22 ( )c ccr f ′′π − [13] ( )1/2
22 ( )c ccr f ′′π −
Amplification
rate, +ω ( )2 2 2
cc ck M f r k′′ + [13] ( )( )1/2
2 2 2
2
1 4 1
D
D cc ck M f r k
τ
′′− τ + −
Maximal wave-
length,
2 /m mkλ = π
( )1/2
22 2 ( )c ccr f ′′π − [13] ( )1/2
22 2 ( )c ccr f ′′π −
Critical time for
instability, ct ( )2 24 ( )c ccr M f ′′
[52] ( )1/2
2 2
2
1 ( ) 1
D
D cc cM f r
τ
′′+ τ −
50 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
we linearize Eq. (34) in terms of concentration inhomogeneity
0( , ) ( , )c t c t cδ = −r r . This one yields
2
2 2 4
2
( ) ( )
( ) ( )D cc c
c c
Mf c Mr c
tt
∂ δ ∂ δ ′′τ + = ∇ δ − ∇ δ
∂∂
. (61)
Following the approach to the concentration fluctuations for the
hyperbolic transport [40], Fourier transforms of ( , )c tδ −k and
( , )c tδ k is expressed as
( , ) exp( ) ( , ),
( , ) exp( ) ( , ).
c t d i c t
c t d i c t
δ − = ⋅ δ −
δ = ⋅ δ
∫
∫
k r k r r
k r k r r
(62)
Then, Eq. (61) for the structure factor (60) is given by
( )2
2 2 2
2
( , ) ( , )
( , ).D cc c
d S t dS t
Mk f r k S t
dt dt
′′τ + = − +k k
k (63)
To solve this equation, we multiply LHS and RHS by exp( )i tω ,
where ω is a frequency of the concentration inhomogeneity. After
some algebra, solution of Eq. (63) is given by
0 2 2 2 2
( ,0) / (1 ) ( ,0)
exp( ) ( , )
( )
D D
D cc c
dS dt i S
i t S t dt
i Mk f r k
∞ τ + + τ ω
− ω =
′′ω − τ ω + +∫
k k
k . (64)
Defining the spectral distribution of fluctuations as
0
( , ) 2 exp( ) ( , ) ,S i t S t dt
∞
ω = ℜ − ω∫k k (65)
one can find from Eq. (64) the spectral distribution for concentra-
tion fluctuations.
Equations (64) and (65) might describe modelled structure for
hyperbolic scenario (with finite Dτ ) and parabolic scenario (for in-
stant relaxation with 0Dτ → ) and experimentally observed struc-
ture after quenching in spinodal decomposition. Such a description
can give information about length scale of concentration fluctua-
tions and, as a consequence, about maximal amplification rate of
decomposition for given scenario.
2.5. Comparison with Experimental Data
The function of the amplification rate predicted by the hyperbolic
model has been compared in Refs [48, 55] with experimental data of
Andreev et al. on phase-separated glasses [56, 57]. The amplifica-
tion rate (51) can be rewritten in the following form
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 51
1/2
1/22 2 2 2 2 2 21 4 (1 ) 1 1 4 (1 ) 1
,
2 2
D c cc D C
D D
D k r k f l k l k
+
⎡ ⎤′′+ τ − − ⎡ ⎤+ − −⎣ ⎦ ⎣ ⎦ω = =
τ τ
(66)
where /C c ccl r f ′′= − is the correlation length, 1/2( )D Dl D= τ is the
diffusion length, and ccD Mf ′′= − is the diffusion constant.
In addition, one can assume that the free energy of a binary sys-
tem can be replaced by
2 4
0 0( , ) ( / 1)( ) ( ) ,h c c cf T c f T T c c B c c⎡ ⎤= − − + −⎣ ⎦ (67)
where cT and cc are the critical temperature and concentration, re-
spectively, cT T< , and 0 0B > . Equation (67) is often used in analy-
sis of kinetics of spinodal decomposition in glasses [58], and the pa-
rameters 0f and 0B are treated as phenomenological input parame-
ters of the theory, which are fitted to experiment [10].
Figure 4 shows data for the relationship ‘ 2/ k+ω versus 2k ’ ex-
tracted from experiments on a binary phase-separated glass [56,
57]. They exhibit non-linear behaviour as predicted by Eqs (66) and
(67) for the following material parameters: 142.3 10D −= ⋅ cm2/s,
117.2 10D
−τ = ⋅ s, 86.2 10cr
−= ⋅ cm⋅ 3/ (mole cm )J ⋅ , / 0.85cT T = ,
0 0.15B = , 4
0 1.88 10f = ⋅ J/(mole ⋅ cm3), and 0.8cc c− = mole frac-
Fig. 4. Dependence
2/ k+ω upon
2k given by the hyperbolic model (solid line;
Eqs (66) and (67)) and scattering data of visible light (points, Refs [56, 57]).
Experimental points were obtained on phase-separated SiO2−12 wt.% Na2O
glass at T = 803 K.
52 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
tion.
In Figure 4, it is shown that good agreement is achieved between
theory and experiment. This result is due to the fact that both
lengths, correlation length Cl and diffusion length Dl , appear in the
theory [48, 55] that is also shown by Eq. (66). The interplay be-
tween these two lengths, i.e. the ratio /D Cl l , governs the transition
as follows (in contrast with the linear Cahn−Hilliard−Cook model
[12, 59] in which only correlation length Cl is important in spinodal
decomposition). With the increase of the correlation length Cl in
comparison with the diffusion length Cl , spinodal decomposition has
the Cahn−Hilliard’s scenario (described by linear or non-linear para-
bolic diffusion equation). With D Cl l≈ , one can accept long-range
interaction within the system and the Cahn−Hilliard’s scenario
takes effect. With D Cl l>> (namely, with 2 2D Cl l≥ [52]), short-
range interaction has effect and local nonequilibrium effect (such as
relaxation of the diffusion flux to its steady state) plays dominant
role in selection of the mode for decomposition. Thus, existence of
these two length, existence of these two length, Dl and Cl , makes
the theory flexible enough to predict non-linear behaviour for am-
plification rate typically observed in experiments and to quantita-
tively describe experiments (Fig. 4).
3. MODELLING OF SPINODAL DECOMPOSITION
In this Section, we present main numeric procedures to be used in
simulations of the spinodal decomposition in deterministic models
of the hyperbolic type. Numerical approaches related for 1D and 3D
simulations are presented in Subsections 3.1 and 3.2, respectively.
Features of hyperbolic spinodal decomposition can be observed in
computational dynamics.
To model decomposition, Eqs (26), (30), and (31) are taken. From
this system, the following dimensionless form of equations is as fol-
lows
2
2 ,c
c
D
rF
f c
c l′
⎛ ⎞δ = − + ∇⎜ ⎟δ ⎝ ⎠
(68)
,
F
t c
∂ δ⎛ ⎞+ = ∇ ⎜ ⎟∂ δ⎝ ⎠
J
J (69)
c
t
∂ = −∇ ⋅
∂
J . (70)
In these equations, the following scales are introduced: / ccM D f ′′=
is the mobility, /D Dl D V= the characteristic spatial length, Dτ the
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 53
time scale as the diffusion relaxation time, and DV the solute diffu-
sion speed as the scale for diffusion flux. To complete this system,
the free energy density is chosen as a double-well potential
2 2
0( ) (1 ) 4,hf c f c c= − (71)
which has minima at 0c = and 1c = , and it has a maximum at 0.5c = .
Boundary conditions are established from the previously obtained
expression for the external exchange ex( / )dF dt of the free energy:
expression in square brackets of Eq. (28) must be zero. This condi-
tion implies the following equalities 0nc∇ = and 0nJ = on the
boundary of the calculated domain. As a result, we arrive to a set
of hyperbolic equations (68)−(71), which describes evolution by the
hyperbolic model described in previous Section (in comparison with
the parabolic model of Cahn and Hilliard using diffusion equation
of a parabolic type).
3.1. 1D Modelling
For integration of equations (68)−(71), we use implicit Euler method
of second order 2( )O τ . Let us introduce the following notations
1 1
2 2
, ,
n nt t t t
c J
p s
t t= + τ = + τ
∂ ∂= =
∂ ∂
(72)
where τ is the time step. Then one can define the system of equa-
tions for the time iterations in the following form
+ + += + τ = + τ ≡ = + τ%( 1) ( ) ( 1/2) ( ) ( 1/2) ( )1 1
, , .
2 2
n n n n n nc c p J J s c c c p (73)
Now, Eqs (68)−(71) can be rewritten as
⎧ ∂ ⎛ ⎞+ + τ =⎪ ⎜ ⎟∂ ⎝ ⎠⎪
⎪ ∂⎛ ⎞⎪ + τ + = −⎜ ⎟⎨ ∂⎝ ⎠⎪
⎪ ⎛ ⎞ ⎛ ⎞∂ τ ∂′⎪ = − −⎜ ⎟ ⎜ ⎟∂ ∂⎪ ⎝ ⎠ ⎝ ⎠⎩
%
%
( )
( )
2 22 ( ) 2
2 2
1
0,
2
1
1 ( ) ,
2
,
2
n
n
n
c c
c
D D
p J s
x
s J M c W
x
r rc p
W f
l x l x
(74)
where /c Dr l is the ratio of correlation parameter cr and diffusion
length /D Dl D V= . Excluding s from (74), one can obtain the final sys-
tem of equations. This one yields
54 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
⎧ ⎛ ⎞ ⎛ ⎞τ ∂ ∂′⎪ + = −⎜ ⎟ ⎜ ⎟∂ ∂⎪ ⎝ ⎠ ⎝ ⎠⎨
⎪ ∂ ∂ ∂⎛ ⎞ ⎛ ⎞− + =⎜ ⎟ ⎜ ⎟⎪∂ ∂ τ τ ∂⎝ ⎠ ⎝ ⎠⎩
%
%
2 22 2 ( )
2 2
( )
,
2
2 2
( ) 1 .
n
c c
c
D D
n
r rp c
W f
l x l x
J
M c W p
x x x
(75)
Equations (75) represent an elliptic set of equations. This system al-
lows us to numerically solve Eqs (68)−(71) relatively the functions p
and W using the following algorithm. Taking initial data for concen-
tration
0
0( ,0) ( )k kc x c x c= = and diffusion flux
0( ,0) 0 kJ x J= = , vari-
ables p and W are obtained from the system (75). Then, new data for
the concentration and the flux are found from the system (73) for the
new time level. These are used for obtaining p and W from Eqs (75).
This procedure is iterated in time to compute concentrations and dif-
fusion fluxes in spinodal decomposition.
Note that the system (75) can be also used for solution of the Cahn
and Hilliard’s parabolic equation for spinodal decomposition. This
procedure has to assume the steady-state diffusion flux, i.e. the flux
needs infinite time for its time changing. Therefore, let τ → ∞ in the
second equation of Eqs (75). This excludes from the numerical proce-
dure the time dependence of the flux and the second order derivative of
concentration with respect to time. In addition, the second equation in
Eq. (73) has to be divided on τ with the further taking the same limit
τ → ∞ . This leads to equality 0s J t= ∂ ∂ = that allows us to exclude
the time dependence of flux from the suggested numeric algorithm.
Choosing a finite difference method and using approximation of the
second order for coordinate x , elliptic system (75) can be tested
against its computational stability. A linearized transfer-matrix
( , )T k n
)
is obtained for deviations cδ and Jδ from exact solutions of
Eqs (75). This one yields
( 1) ( )
( 1) ( )
( , ) ,
n n
n n
c c
T k n
J J
+
+
⎛ ⎞ ⎛ ⎞δ δ
=⎜ ⎟ ⎜ ⎟
δ δ⎝ ⎠ ⎝ ⎠
(76)
where
( )( )
2 2
0 2
2 2
1 1 1
1 1
2 2 2
( , ) ,
1 1
1 1 1 1 3
2 4
ML k i k
T k n N
iMLk MLk
⎛ ⎞⎛ ⎞ ⎛ ⎞− τ + τ − τ + τ⎜ ⎟⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠⎜ ⎟= ⎜ ⎟⎛ ⎞⎜ ⎟− τ + τ − τ + + τ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠
(77)
where k is the wave vector, and factors 0N and L are obtained as
( ) ( )1 12 2
0 1 2 4 1 2 ,N MLk
− −= + τ + τ + τ
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 55
and
2( , ) ( / ) .cc c DL n k f r l k′′= +% % The stability condition requires that, for
given wave vector k, the eigenvalues for transport matrix (77) have to
be not greater than unity. Qualitative dependence of modulus of eigen-
values for transfer-matrix ( , )T k n
)
from the wave vector k is present in
Fig. 5. This dependence is shown at different values of parameters of
the time step τ and relation 0 /c Dr lα = between the correlation pa-
rameter cr and diffusion length /D Dl D V= . One can see that numeri-
cal scheme is conditionally stable. The condition of stability is formed
mainly by a largest possible value of wave vector and weakly depends
on other parameters, particularly, from the time step τ .
Dynamics of spinodal decomposition is presented in Figs. 6 and 7 for
the material and computational parameters summarized in Table 3.
The dynamics is shown in spatial changing of concentration profiles
for a given time step (Figs. 6, 7, a—d). After formation of quasi-
sinusoidal profile from an initially random distribution, this distribu-
tion becomes unstable in further separation due to up-hill diffusion
between decomposing phases (Fig. 6, a—b). This unstable situation
evolves much more faster for the system described by Cahn−Hilliard
equation than for the local nonequilibrium system described by hyper-
bolic equation (Fig. 6, b—c). It occurs, generally, due to propagation of
concentration disturbance with infinite diffusion speed in the
Cahn−Hilliard’s system. Hyperbolic system has a delay described by
the second time derivative in Eqs (32) and (34). As a result, concentra-
tion disturbance in the hyperbolic system propagates with the finite
speed and the instability realizes with the delay relatively to the
Cahn−Hilliard’s system.
Fig. 5. Typical dependence of modulus of eigenvalues for transfer-matrix
( , )T k n
)
on dimensionless wave vector k with the scale of 2 Dlπ . Region of com-
putational stability lies below unity for the eigenvalues of transfer matrix.
Dependence is shown at: (a) different values 0 /c Dr lα = that is the ratio of cor-
relation parameter cr and diffusion length /D Dl D V= , and (b) different τ .
56 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
Finally, the new stable (or metastable) profile is formed which is the
same for both parabolic Cahn−Hilliard’s system and hyperbolic system.
Fig. 6. Dynamics of spinodal decomposition of a binary liquid in 1D case with
the constant mobility 0 constM M= = : (a) metastable state; (b) beginning of
the transition from unstable to metastable state; (c) finishing of the transition
from unstable to metastable state; (d) new metastable state. The first metasta-
ble state (a) and the following metastable state (d) are equivalent for both clas-
sic and modified Cahn−Hilliard equations. The dynamics of transition between
two metastable states is much more faster for classic Cahn−Hilliard equation.
Fig. 7. Dynamics of spinodal decomposition of a binary liquid in 1D case with
the concentration dependence of mobility 0 (1 )M M c c= − : (a) metastable
state; (b) beginning of the transition from unstable to metastable state; (c)
transient period for the transition from unstable to metastable state; (d) new
metastable state. The first metastable state (a) and the following metastable
state (d) are equivalent for both classic and modified Cahn−Hilliard equa-
tions. The dynamics of transition between two metastable states is much more
faster for classic Cahn−Hilliard equation.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 57
The described dynamical scenario qualitatively does not depend on
the atomic mobility of decomposing phases. Comparative dynamics
shown in Figs. 6 and 7 for constant mobility and concentration depend-
ent mobility, respectively, demonstrates that the both dynamics quali-
tatively remain the same. The only difference in dynamics shown in
Figs. 6 and 7 is that the transition from the one metastable state to the
next metastable state proceeds during various periods of time. Conse-
quently, the Cahn and Hilliard model predicts much more faster dy-
namics of spinodal decomposition, which can be resulted in much more
diffuse boundaries between two separated phases.
3.2. 3D Modelling
For 3D numeric solution of Eqs (68)−(71), a cube with linear size N is
taken. The cube is approximated by the numerical grid with equal dis-
tances xΔ between nodes along Cartesian axes. In such a case, coordi-
nates of the nodes are given as x i x= Δ , y i x= Δ , and z i x= Δ , where
1,...,i N= , 1,...,j N= , and 1,...,k N= , respectively.
A random distribution around average concentration 0c within the
cube is accepted for initial time step 0n = . Then for every time step
t n= τ , the following explicit numerical scheme is used:
0 (1 )(1 2 )
2
n
n n n
ijk ijk ijk
ijk
fF
c c c
c
δ⎛ ⎞ = − − − +⎜ ⎟δ⎝ ⎠
TABLE 3. Parameters for a binary system used in numeric computations.
Parameter
Value and unit for
1D modelling
Value and unit for
3D modelling
Initial concentration, 0c 0.5 atomic fraction 0.5 atomic fraction
Height of the free energy, 0f 0.5 0.5
Time for diffusion relaxation, Dτ 111.5 10−⋅ s
111.5 10−⋅ s
Diffusion coefficient, D 95.0 10−⋅ m
2/s 95.0 10−⋅ m
2/s
Bulk diffusion speed,
1/2( / )D DV D= τ 18.26 m/s 18.26 m/s
Spatial diffusion length, /D Dl D V= 90.27 10−⋅ m 90.27 10−⋅ m
Ratio /c Dr l 0.90 0.29
Quantity of computational nodes, N 80 3500
Dimensionless spatial step, xΔ 0.56 0.88
Dimensionless time step, τ 25.13 10−⋅ 35.0 10−⋅
58 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
2
1 1 1 1 1 1( 6 ),n n n n n n nc
i jk i jk ij k ij k ijk ijk ijk
D
r
c c c c c c c
l x + − + − + −
⎛ ⎞
+ + + + + + −⎜ ⎟Δ⎝ ⎠
(78)
1
1 1
(1 ) (
2
n n
n n
ijk ijk
i jk i jk
F F
J J
x c c
+
+ −
τ δ δ⎛ ⎞ ⎛ ⎞= − τ + − +⎜ ⎟ ⎜ ⎟Δ δ δ⎝ ⎠ ⎝ ⎠
1 1 1 1
,
n n n n
ij k ij k ijk ijk
F F F F
c c c c+ − + −
δ δ δ δ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞+ − + −⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟δ δ δ δ⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
(79)
1
1 1 1 1 1 1( ).
2
n n n n n n n n
ijk ijk i jk i jk ij k ij k ijk ijkc c J J J J J J
x
+
+ − + − + −
τ= − − + − + −
Δ
(80)
A set of hyperbolic equations (78)—(80) allows to model decomposi-
tion with conserved function of concentration c. It has been resolved
numerically with material and computational parameters summarized
in Table 3. To test conservation of c, average concentration of a whole
Fig. 8. 3D modelling of spinodal decomposition in undercooled binary liquid:
(a) 0t = , (b) 150t = τ , (c)
31.5 10t = ⋅ τ , (d)
41.5 10t = ⋅ τ . For every time mo-
ment, isoconcentration patterns within the cube with size of
3500 computa-
tional nodes is shown.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 59
computational domain has been compared with the initial value c0. An
error of computations 0 0| | / 100%c c c− ⋅ was found not higher than
3.8%. That confirms correctness of the present system (78)—(80) to
model the spinodal decomposition with conserved concentration c
within computational domain.
Figure 8 shows evolution of concentration inside the cube. It is seen
that initially random parts of distribution with equal concentration
(Fig. 8, a) create isoconcentration surface (Fig. 8, b—d] during decom-
position. Figure 9 presents snapshots of patterns evolving in local non-
equilibrium spinodal decomposition. This sequence exhibits hyperbolic
evolution with a sharp boundary between two types of decomposed liq-
uid especially at the first moments of decomposition (see Fig. 9, b). The
sharp boundary between two liquids follows from the fact that descrip-
tion of diffusion in system (78)—(80) is given by hyperbolic type of
Fig. 9. Evolution of spinodal decomposition: (a) 0t = , (b) 50t = τ , (c)
310t = τ ,
(d)
43 10t = ⋅ τ . For every time moment, cross section for a cube with the size of
3500 computational nodes is shown.
60 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
equation (69). Together with conservation law (70), Eq. (69) predicts
with a finite speed and, as a consequence, with sharp diffusion fronts
between two separating phases.
To compare patterns originating in spinodal decomposition, results
for both hyperbolic evolution and parabolic evolution were extracted
from solution of Eqs (78)—(80). Predictions of complete system (78)—
(80) were taken as for hyperbolic evolution. Predictions of the system
(78)—(80) without relaxation of the solute diffusion flux, i.e., with the
Fick’s diffusion flux
,
F
c
δ⎛ ⎞= ∇ ⎜ ⎟δ⎝ ⎠
J (81)
and its numerical approximation
1 1
1 1 1 1
1
(
2
,
n n
n
ijk
i jk i jk
n n n n
ij k ij k ijk ijk
F F
J
x c c
F F F F
c c c c
+ −
+ − + −
δ δ⎛ ⎞ ⎛ ⎞= − +⎜ ⎟ ⎜ ⎟Δ δ δ⎝ ⎠ ⎝ ⎠
δ δ δ δ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞+ − + −⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟δ δ δ δ⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
(82)
instead of Eqs (69) and (79), respectively, were taken as for para-
bolic evolution.
Results of modelling for patterns in both evolutions are shown in
Fig. 10. Evolution of patterns has been spied upon the complete sepa-
Fig. 10. Comparison of patterns in spinodal decomposition described by (a—d)
hyperbolic equation and (e—h) parabolic equation. Here, (a, e) 10t = τ , (b, f)
40t = τ , (c, g)
22 10t = ⋅ τ , (d, h)
52 10t = ⋅ τ . For every time moment, patterns
are shown for a small cube of
350 computational nodes extracted from central
part of the cube with the size of 3500 computational nodes.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 61
ration of liquids (see Fig. 10, d and h). It is seen that the boundaries
between two demixing liquid phases are sharper for hyperbolic evolu-
tion than for the parabolic evolution. This feature of the hyperbolic
evolution is a result of solute diffusion with the finite speed that leads
to instability realized with the delay relatively to the Cahn—Hilliard’s
system (see comparative dynamics in Figs. 6 and 7).
4. STOCHASTIC MODELS OF SPINODAL DECOMPOSITION
In this Section, we consider to above models for spinodal decomposi-
tion and illustrate in what a manner concentration fluctuations, con-
sidered as internal noise, obeying fluctuation—dissipation relation can
affect on the system dynamics and stationary states.
It is known that contrary to the naive predictions, assuming that
fluctuations lead to disordering effects, phenomena such as noise-
induced transitions in zero-dimensional systems [60—64] (when a sto-
chastic variable x is a function of the time t only, ( )x x t= ), stochastic
resonance [65, 66], noise-induced ordered and disordered phase transi-
tions in extended systems [67—70] (when ( , )x x t= r ), noise-induced pat-
tern formation processes [71, 72], noise-induced effects in excitable sys-
tems [73], and a lot of others are manifestations of the constructive role
of fluctuating environment. An increasing interest in the noise-induced
phenomena in extended systems results in the discovery of new nonequi-
librium universality classes [74, 75] and new types of self-organization
processes such as entropy driven phase transitions [76, 77].
In most problems of noise-induced phenomena in extended systems
external fluctuating sources are considered; their primary role in self-
organizational processes is stated [67, 78]. Recently, a new type of en-
tropy driven phase transitions was discovered [76]. Within this type of
transitions, it was shown that internal fluctuations with intensity re-
lated to a field-dependent kinetic coefficient (mobility) play a principle
role in ordering dynamics. Particularly, it was found that the internal
multiplicative noise leads to the effective entropy dependence on the
stochastic quantity in a functional form but does not change the corre-
sponding free energy functional. As a result, noise-induced effects can
be understood with a help of the entropy mechanism, which follows
from the thermodynamics. Considering parabolic and hyperbolic mod-
els for spinodal decomposition within the frame of the linear stability
analysis and the mean field theory we compare behaviour of above two
models.
The aim of this Section is to perform a somewhat detailed study of
entropy driven phase transitions mechanisms in phase separation
processes. We analyze early and late stages of evolution numerically.
For the stationary picture, we extend the mean field approach to the
systems with the field dependent mobility and investigate mechanisms
62 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
of entropy driven phase transitions in binary stochastic systems. We
generalize the well-known results for early stages of evolution, study
late stages and consider re-entrant ordering effects in a stationary
case. Our analytical results are compared with computer simulations.
The outline of this Section is the following. In Subsection 4.1, we
introduce the general stochastic system with conserved dynamics and
obtain the corresponding Fokker−Planck equation that can be used in
the mean field analysis for both parabolic and hyperbolic models. Con-
sidering parabolic model, we discuss effects of the internal multiplica-
tive noise influence at early and late stages of spinodal decomposition
and consider phase transitions by means of the mean field theory. In
Subsection 4.1.4, we discuss effect of a combined effect of both inter-
nal and external stochastic sources. In Subsection 4.2, we study sto-
chasticity of the hyperbolic transport. Subsection 4.3 is devoted to
study of stochastic hyperbolic model for spinodal decomposition.
4.1. Stochastic Parabolic Model for Spinodal Decomposition
In this subsection, we investigate an influence of the conserved-field
(concentration) dependent mobility and the corresponding internal
noise on properties of a phase separation scenario in the parabolic model.
Formally, in a case of a binary system with concentrations Aρ and
Bρ of the components A and B, respectively, the density difference
A Bx = ρ − ρ can be introduced. In a phase separation scenario, the
quantity x obeys a conservation law ( , )d constx t =∫ r r . In our study, we
use the field x to describe the system under consideration and investi-
gate a corresponding dynamics and a stationary picture.
To define a principle model let us start with a continuity equation
for the field ( , )x tr in a d-dimensional space in the form
,x
t
∂ = −∇ ⋅
∂
J (83)
where J is the flux. The deterministic part of the flux is of the form
det
x
M
x
δ= − ∇
δ
J
[ ]F
. (84)
Here, M is the mobility, the Helmholtz coarse-grained free energy
functional F is of the standard (Ginzburg−Landau or Cahn−Hilliard)
form applicable for a local equilibrium system
2[ ] ( ) ( ) ,
2
D
x dV f x x
⎧ ⎫= + ∇⎨ ⎬
⎩ ⎭∫F (85)
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 63
where D is the spatial coupling intensity related to the correlation ra-
dius in the form
2
cD r= with
2 2 2
0( / ( ) ) |c xr x ∇ == δ δ ∇F ; ( )f x is the free
energy density of the bulk. The mobility we assume in a functional
form, i.e. ( )M M x= .
To investigate the influence of the flux fluctuations, let us intro-
duce fluctuation source ζ into the right hand side of Eq. (84). It yields
an expression for the flux in the form as follows:
det ( ; , ).x t= + ζJ J r (86)
Formally, the stochastic part ζ is assumed to be Gaussian, and gener-
ally, it can be a function of the field x. Assuming the x-dependent mo-
bility ( )M M x= and using fluctuation dissipation relation, one gets
the following definition for the averages
2( ; , ) 0, ( ; , ) ( ; , ) 2 ( ) ( ; ).x t x t x t M x C t t′ ′ ′ ′〈ζ 〉 = 〈ζ ζ 〉 = σ − −r r r r r (87)
In the simplest case, we assume the correlation function C in the form
of ( ; ) ( ) ( )C t t t t′ ′ ′ ′− − → δ − δ −r r r r , which allows to consider the white
noise in space and time. Here, we note that the fluctuation dissipation
relation holds that yields an interpretation of the noise intensity
2σ as
an effective temperature of the bath.
In further study of stochastic dynamics, we consider a general case
when control parameters of the system and the noise intensity
2σ are
independent quantities.
Using conditions corresponding to an equilibrium situation, one
gets the flux J as follows [ ] / ( ) ( , )M x x g x t= − ∇δ δ + ξJ rF , where
( ) ( )g x M x= , ( , ) 0t〈ξ 〉 =r ,
2( , ) ( , ) 2 ( ) ( )t t t t′ ′ ′ ′〈ξ ξ 〉 = σ δ − δ −r r r r . Sub-
stituting the generalized flux into Eq. (83), we get the stochastic con-
tinuity equation in the form
[ ]
( ) ( ) ( , ).
x
x M x g x t
t x
∂ δ⎛ ⎞= ∇ ⋅ ∇ + ∇ ξ⎜ ⎟∂ δ⎝ ⎠
r
F
(88)
To study statistical properties of the system one needs to find the
probability density ([ ], )x t=P P . To this end, we represent the system
on a regular d-dimension lattice with a mesh size l . Then, the partial
differential equation (88) is reduced to a set of usual differential equa-
tions written for an every cell i on a grid in the form
( ) ( ) ( ) ( ),i
L ij j R jl L ij j j
l
dx F
M g t
dt x
∂= ∇ ∇ + ∇ ξ
∂
(89)
where index i labels cells: 1, , di N= K ; the discrete left and right op-
erators are introduced ( 2d = ) as follows:
64 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
, 1, 1, ,
, 1 , , 12
1 1
( ) ( ), ( ) ( ),
1
( ) ( ) , ( ) ( ) ( 2 ),
L ij i j i j R ij i j i j
L ij R ji L ij R jl il i l i l i l
− +
+ −
∇ = δ − δ ∇ = δ − δ
∇ = − ∇ ∇ ∇ = Δ = δ − δ + δ
l l
l
(90)
where ijδ is the Kronecker symbol. For stochastic sources, the discrete
correlator is of the form
2 2( ) ( ) 2 ( )i j ijt t t t− ′〈ξ ξ 〉 = σ δ δ −l . In the following
analysis, we use the Stratonovich interpretation of the Langevin equa-
tions (89).
Next, to obtain above distribution let us introduce standard defini-
tions
=1
([ ], ) ( ( ) ) ( ) ,
dN
i i
i
x t x t x t= 〈 δ − 〉 ≡ 〈ρ 〉∏P (91)
where K and 〈 〉K are averages over initial conditions and noise, re-
spectively. To obtain the corresponding Fokker−Planck equation, we
use the standard technique and exploit the stochastic Liouville equa-
tion
( ).i
i i
x
t x
∂ ∂ρ = − ρ
∂ ∂∑ & (92)
Inserting the expression for the time derivative from Eq. (89) and av-
eraging over noise, we get
( ) ( ) ( ) ( ) .L ij j R jl L ij j j
i iji l i
F
M g t
t x x x
⎛ ⎞∂ ∂ ∂ ∂
〈ρ〉 = − ∇ ∇ 〈ρ〉 − ∇ 〈 ξ ρ〉⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠
∑ ∑ (93)
The correlator in the second term can be calculated by means of No-
vikov’s theorem that gives [79]
δ ρ
′ ′〈 ξ ρ〉 = σ δ δ − 〈 〉
′δξ∑ ∫2
0
( )
( ) 2 ( ) ,
( )
t j
j j jk
k k
g t
g t dt t t
t
(94)
where for the last multiplier one has
( ) ( )
( ).
( ) ( )
j l
j
k l k t t
g t x t
g t
t x t ′=
δ ρ δ∂= − ρ
′ ′δξ ∂ δξ
(95)
In a formal solution of the Langevin equation, the response function
takes the form
( )
( )
( )
l
L lk k
k t t
x t
g
t ′=
δ
= ∇
′δξ
. (96)
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 65
After some algebra, we obtain the Fokker−Planck equation for the
total distribution P in the discrete space
2( ) ( ) ( ) ( ) ,L ij j R jl L ij j R ji i
i iji l i j
F
M g g
t x x x x
⎛ ⎞∂ ∂ ∂ ∂ ∂= − ∇ ∇ − σ ∇ ∇⎜ ⎟∂ ∂ ∂ ∂ ∂⎝ ⎠
∑ ∑P P P (97)
where relations between left and right gradient operators are used. The
one-point probability density is defined as follows:
( ) ([ ], )i mm i
P t x t dx
≠
= ∏∫ P .
The above equation can be rewritten in the standard form (in our case,
it has only formal form) for the functional ({ ( )}, )x t= rP P in the con-
tinuum space as follows:
2
2
2
2
( ) ( ) 2 ( )
( ) .
( )
M
d M
t x x x
d M
x
⎡ ⎤∂ δ δ σ δ⎛ ⎞= − ∇ ∇ − ∇ ∇ −⎢ ⎥⎜ ⎟∂ δ δ δ⎝ ⎠⎣ ⎦
δ−σ ∇ ∇
δ
∫
∫
r
r r r
r
r
F
P P
P
(98)
The term proportional to
2σ in the first addendum is the noise-induced
drift. The derived Fokker−Planck equation allows us to write down the
corresponding Langevin equation
2
( , ) ( , ),
2 W
M
x t M g t
t x x
∂ δ σ δ⎛ ⎞= ∇ ∇ − ∇ ∇ + ∇ ξ⎜ ⎟∂ δ δ⎝ ⎠
r r
F
(99)
with a process Wξ , which has a strong mathematical definition through
the Wiener process ( )W t : ( ) ( ) /W t dW t dtξ = ,
2( ) :dW dt [60].
The stationary solution of the Fokker−Planck equation takes the form
2
2
1
[ ] exp [ ] ln ( ) .
2
x x d M x
⎧ ⎫⎛ ⎞σ⎪ ⎪∝ − +⎨ ⎬⎜ ⎟σ⎪ ⎪⎝ ⎠⎩ ⎭
∫ rP F (100)
Exploiting standard thermodynamic definition of the effective inter-
nal energy eff ef effT= +U F S and assuming a quasi-Gibbs form for the
stationary distribution, we can identify the effective entropy
eff (1 / 2) ln ( )d M x= ∫ rS
and the effective temperature
2
efT = σ .
It is principally important that the stationary distribution is exact
and is described not only by the initial functional [ ]xF . Here, the en-
tropy contribution modifies the form of the probability density. De-
spite, we consider the internal multiplicative noise, its action leads to
66 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
the entropy variability. Thus, controlling the noise intensity or parame-
ters in the mobility ( )M x we can govern an ordering process. Therefore,
the system studied below is considered as a nonequilibrium one. The
same situation is observed in the entropy driven phase transitions in
systems with nonconserved dynamics [76, 80−82], where the stationary
distribution is obtained exactly and phase transitions picture is con-
trolled by the effective entropy variations. Below, we relate the entropy
driven phase transitions formalism developed for systems with noncon-
served dynamics to the systems with conserved dynamics. The main at-
tention will be paid on effects where the entropy variability is principle.
Using Langevin or Fokker−Planck equation, one can derive equation
for the first statistical moment directly:
2
( , ) .
( ) 2 ( )
M
x t M
t x x
∂ δ σ δ
〈 〉 = ∇〈 ∇ 〉 − ∇〈∇ 〉
∂ δ δ
r
r r
F
(101)
This equation can be used to analyze the influence of the internal
multiplicative noise on the stability of the null phase in the linear ap-
proximation.
Since the dynamics is conserved,
V
( , ) constx t d =∫ r r ,
V is the system volume), so dynamics of the phase separation can be
considered with a help of the Fourier transform of correlation function
2( , ) (1 / ) ( , ) ( , ) ( , )G t V x t x t d x t′ ′ ′= 〈 + 〉 − 〈 〉∫r r r r r r .
The corresponding structure function is given as
V
( , ) ( , ) iS t G t e d= ∫ krk r r .
In practice, it is convenient to use a spherical average of the correla-
tion and structure functions, which are as follows:
( , ) ( , )
r
g r t G t d
Ω
= Ω∫ r , ( , ) ( , )
k
S k t S t d
Ω
= Ω∫ k .
Here, rΩ and kΩ are spherical shells of radius r and k , respectively.
The above values allow us to extract a mean characteristic size of do-
mains at time t, ( )R t , using scaling relations: ( , ) ( / ( ))g r t r R t= ϕ ,
( , ) ( ) ( ( ))dS k t R t kR t= ϕ . The expected domains growth law is ( ) zR t t∝ ,
where z is the domain-growth exponent.
The Model. In our consideration, we use a model for a binary system,
which is described by the free-energy density, ( )f x . Single-phase equi-
librium ( )f x has a stable single-well structure. In a two-phase region,
( )f x is of a double-well structure; the corresponding model has the form
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 67
2 4( ) 2 4,f x x x= − ε + (102)
where ε is a dimensionless phenomenological constant, playing the
role of a control parameter.
The mobility is used in the functional form
2 1( ) (1 ) , 0.M x x −= + α α ≥ (103)
Variations in the parameter α allow us to consider additive ( 0α = ) or
multiplicative ( 0α ≠ ) noises, separately (the difference in such a clas-
sification is reduced to the following: if the noise term appears in the
evolution equation with a constant multiplier, constg = , then the
noise is additive, else, ( )g g x= , it is multiplicative). The model func-
tion (103) assumes that fluctuations are large in the case where 0x = ,
whereas fluctuations are small in cases where 0x ≠ . Formally, assum-
ing α to be small, an approximate definition is
2( ) 1M x x≈ − α . As
previous studies show, the quantity α can be expressed through the
relation between bulk bD and surface sD diffusion constants, i.e.
1 /b sD Dα ≈ − (see Ref. [83]). Further, we are looking for changes in
system behaviour when α ranges.
In such kind of stochastic models, a possible scenario of phase sepa-
ration depends on the initial conditions: at ( ,0) 0x〈 〉 =r , the system
evolves by spinodal decomposition scenario (see Fig. 11, a), whereas at
( ,0) 0x〈 〉 ≠r , a nucleation process is realized. (Fig. 11, b).
a
b
Fig. 11. Typical spatial patterns: (a) spinodal decomposition ( ,0) 0,x〈 〉 =r (b)
nucleation ( ,0) 0.2x〈 〉 =r . Other parameters are as follows: 120 120,N N× = ×
4,D = 1,ε = 0.5,α =
2 0.2.σ =
68 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
4.1.1. An Early Stage of Evolution
At the first, we investigate the internal multiplicative noise influence
on instability of the homogeneous phase 0x〈 〉 = . Using Fourier trans-
formation for the scalar field
( ) (2 ) ( , )d ix t d x t e− −= π ∫ kr
k r r
in d-dimensional space, in linear approximation one gets
( )2 2 2 .
d x
k Dk x
dt
〈 〉
= − − ε + ασ 〈 〉k
k (104)
It is principally important that the noise contribution denoted as
2 xασ 〈 〉k stabilizes the homogeneous state. The same result was ob-
served in the case of entropy-driven phase transitions with noncon-
served dynamics [76, 81]: as it follows, one can await the similar be-
haviour of the stochastic systems with conserved dynamics where the
entropy driven phase transition formalism can be generalized.
More information of the system behaviour provides the knowledge
of the structure function ( , ) ( ) ( )S t x t x t−= 〈 〉k kk . Following the standard
approach, a linear evolution equation for the spherically averaged
structure function can be derived in the form [84]
2 2 2 2 2 2 2( , ) 1
( ) ( , ) 2 2 ( , ).
(2 )d
dS k t
k Dk S k t k k d S q t
dt
= − − ε + ασ + σ − ασ
π ∫ q (105)
It is seen at 0α = that corresponds to the additive noise case, one ar-
rives at the well-known Cahn−Hilliard−Cook equation for the structure
function [12, 59]. From exponential solutions of Eqs (104, 105), one
can see that only modes with
2( ) /ck k D< = ε − ασ are unstable and
grow at early stages of evolution. With an increase in α or
2σ , the size
of the unstable modes domain ck k< decreases. Modes with ck k> re-
main stable during the linear regime. Note that unstable modes cannot
be realized at condition
2ε < ασ . As it follows, the domain growth
should be different for additive and multiplicative noise.
In Figure 12, we present solutions of the evolution equation (105) at
different values of the parameter α . It can be seen that an increase in
α leads to a shift of the peak position toward smaller values of k . The
peak of ( )S k is less pronounced in the multiplicative noise case than in
the case of the additive noise. It follows that, if the multiplicative noise
is considered, then the dynamic is slowed. A decrease in the peak height
means that an interface is more diffuse in the case of multiplicative
noise (see insertions in Fig. 12). We compare analytical results with
computer simulations at the same time t on the two-dimensional lattice.
In the insertions, a typical patterns and images of spherically averaged
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 69
structure functions are shown. It is seen that in the multiplicative noise
case the pattern has more diffuse interface and the resonance ring in
S(k)-dependence is less pronounced than at the additive noise.
4.1.2. A Late Stage
To get more information about internal multiplicative noise influence,
we simulate the system behaviour and study influence of the parameter
α at the late stage of the evolution. All simulations were done in two-
dimensional lattice with periodic boundary conditions with 120N = . A
criterion for the phase separation in the model under consideration is
the growth of the averaged second moment of x in the real space. We
use the standard definition of the corresponding order parameter
2
2 2
2
=1
( ) ( ) .
N
i
i
M t N x t−= 〈 〉∑ (106)
An alternative formula is 2( ) ( )kk
M t S t= ∑ . In phase separation scenario
during the long time evolution, the quantity 2( )M t grows to the sta-
tionary value 2M when the system tends to the nonzero stationary state.
Fig. 12. Evolution of the structure function at early stage ( 10t = ) at 4D = ,
1ε = ,
2 0.3σ = . Different values of the parameter α are used to compare in-
fluence of additive 0α = and multiplicative 0.9α = noises (solid and dashed
lines, respectively). Insertion shows typical patterns and corresponded images
of spherically averaged structure functions at the same time obtained from
numerical solution of Eq. (99) at 3 0x = with the initial condition ( ,0) 0x〈 〉 =r .
70 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
Figure 13 displays the evolution of the order parameter 2M at different
values of the parameter α and stationary values 2M versus α . It is seen
that an increase in α delays slightly the evolution of 2M (Fig. 13, a) at
small times and suppresses the stationary values (see Fig. 13, b).
Despite the fact that the quantity 2M represents an integral effect,
more information about the system behaviour can be found in the
structure function ( )S k . A convenient quantity is the spherically av-
eraged structure function defined on a circle as follows:
a
b
Fig. 13. Order parameter evolution (a) and its stationary values (b) at dif-
ferent values of α and 1.0ε = , 4D = , 2 0.2σ = .
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 71
1
( , ) ( , ).
k k kk
S k t S t
N ≤ ≤ +Δ
= ∑
k
k (107)
In Figure 14, we display the evolution of the structure function for two
values of the parameter α at fixed noise intensity
2σ . Comparing Fig.
14, a, b, one can see that, in the case under consideration, the peaks of
the structure function are less pronounced when α increases that cor-
responds to the case of the field-dependent mobility case studied in
Ref. [85] and relates to the linear stability analysis. As follows from
Figs. 14, a, b, the positions of the peaks are the same with an increase
in α at equal times. It follows from the linear stability analysis and
corresponds to the fact that in the case of multiplicative noise the dy-
namics is slowed [85]. This result is different from the deterministic
case where at large α peaks are located at higher values of k . The typi-
cal behaviour of the Fourier images in Fig. 14, c, shows the diffuse in-
terface between two phases and change in the peaks position. As Figure
a b
c
Fig. 14. Evolution of the structure function at ε = 1.0, D = 4.0, σ2 = 0.2:
(a) α = 0.01, (b) α = 0.85. The times represented are t = 250, t = 1000,
t = 3000. (c) Fourier images of structure function evolution at α = 0.85
and times t = 250, t = 1000, t = 3000.
72 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
15 shows, the change in peaks height means that with an increase in α
the interface becomes more diffuse.
To investigate the domain growth dynamics we use the standard
formulas for a relevant length:
k k
R t k k S k t kdk S k t dk−= 〈 〉 〈 〉 = ∫ ∫max max1
0 0
( ) , ( , ) ( , ) . (108)
The power law behaviour of the function ( ) zR t t∝ is verified at differ-
ent values of the parameter α , where the domain growth exponent de-
pends on α , i.e. ( )z z= α (see Fig. 16). It is seen that, in the case of ad-
ditive noise ( 0α = ), the exponent 1 / 3z ∝ , whereas at 1.0α = , we ob-
tain 1 / 4z ∝ . Therefore, with an increase in α a crossover of dynami-
cal regimes is observed. Our results are in good correspondence with
deterministic and stochastic approaches, which indicate that an in-
crease in the parameter α delays the dynamics [86−88].
4.1.3. Stationary Case
To investigate the steady states, we can use an extension of the mean
field theory developed for the systems with conserved dynamics [37].
In the framework of this theory, one can use thermodynamic supposi-
tions for the deterministic dynamics and after apply it to the stochastic
one.
Fig. 15. Structure function behaviour at different values of the parameter
α at 1.0ε = , 2 0.2σ = , 4.0D = , 3000t = .
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 73
At first, let us define transition and critical points [89]. Considering
the deterministic case, we use the model tx M x∂ = ∇ ⋅ ∇ δ δF , where
the restriction
0 ( , )
V
x d x t= ∫ r r
is taken into account, 0x is fixed by the initial conditions. For such a
system, the transition point is 0( )T xε : at 0( )T xε < ε the homogeneous
state 0x is stable; at 0( )T xε > ε the system separates in bulk phases, 1x
a
b
Fig. 16. Power law for domains size growth: (a) log−log plot of the evolu-
tion of ( )R t at different values of the parameter α (insertion shows uni-
versal behaviour of the function ( )R t at large times indicated in the rec-
tangle); (b) dependence of the power law exponent z versus parameter α .
Other parameters are as follows: 1.0ε = , 4.0D = , 2 0.2σ = .
74 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
and 2x , with 0x x〈 〉 = . The transition point coincides with critical one
for 0 0x = only, i.e. (0)T cε = ε .
The corresponding steady state solutions are given as solutions of
the equation 0M x∇ ⋅ ∇ δ δ =F . If no flux condition is applied, then
stationary solutions can be obtained from the equation
2 0x∇ δ δ =F ,
due to the mobility M does not affect on the number and extremes po-
sitions of the functional F ; the mobility leads to the change in the dy-
namics of the phase transition only. Hence, the bounded solution is
x hδ δ =F , where h is a constant effective field of the system, in equi-
librium systems h is a chemical potential. In the homogeneous case,
the value h does depend on the initial conditions 0x . Above the transi-
tion point, the steady state is not globally homogeneous, here the sys-
tem separates into two bulk phases with values 1x and 2x . The fraction
u of the system can be defined by the lever rule: 1 2 0(1 )ux u x x+ − = . In
the case of the symmetric form of the free energy functional where two
phases with 1 2x x= − are realized, we get 0h = [37]. Hence, if the field
h becomes trivial, then the transition point can be defined.
Using the above assumptions, let us move to the stochastic case, fol-
lowing prescription [37]. For the one-point probability density,
( ) ([ ], )i mm i
P t x t dx
≠
= ∏∫ P ,
the standard definition of the nearest-neighbours average
( )
([ ], ) 2 ( )j m ij nn i m i
x t x dx d x P t
∈ ≠
= 〈 〉∑ ∏∫ P
can be applied. It allows to rewrite Eq. (125) in a more useful form
( ) ( ),i ij j i
ji
P t M P t
t x
∂ ∂= − Δ 〈 〉
∂ ∂ ∑ % (109)
where
2
2 .
2j
j j j
F M
M M M
x x x
∂ σ ∂ ∂= − + σ
∂ ∂ ∂
% (110)
With no flux condition, the average jM〈 〉%
satisfies the equation
( ) 0.ij j s i
j
M P xΔ 〈 〉 =∑ % (111)
Taking i j= , dropping subscripts and using results of the determinis-
tic analysis with M h〈 〉 =% , we obtain the mean-field stationary equa-
tion
2
2( ) ( ) 2 ( ) ( ),
2s s
V M
hP x M x dD x x M P x
x x x
⎛ ⎞∂ σ ∂ ∂⎡ ⎤= − 〈 〉 − − + σ⎜ ⎟⎢ ⎥∂ ∂ ∂⎣ ⎦⎝ ⎠
(112)
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 75
where we have used mean-field approximation of the Laplacian
( )
( )
2 2 ( ),ij j nn i i
j nn i
x x dx d x x
⎛ ⎞
Δ ≡ − → 〈 〉 −⎜ ⎟
⎝ ⎠
∑ ∑ (113)
the mean-field value x〈 〉 should be defined self-consistently. A solu-
tion of Eq. (112) takes the form
2
2
2
1
( , , ) exp ( ) ( )
2
ln ( ) .
2 ( )
s
D
P x x h N f x x x
dx
M x h
M x
′⎛ ⎡〈 〉 = − + 〈 〉 − +⎜ ⎢σ ⎣⎝
⎞′ ⎤σ+ − ⎟⎥ ⎟′ ⎦ ⎠
∫
(114)
where 2D dD′ = , next we drop the prime.
In order to determine the unknown quantities h and x〈 〉 , we re-
call that considered mean field approach is local and expresses sP of
a field at a given site of the lattice as a function of the field h and
of the mean field 〉〈x in a neighbourhood of the given cell.
In the homogeneous case (below the threshold), the mean field is the
same everywhere and equals the initial value, i.e. 0x x〈 〉 = . Hence, at
the fixed mean-field value, solving the self-consistency equation,
sx xP x x h dx〈 〉 = 〈 〉∫ ( , , ) , (115)
we obtain the constant effective field h . Above the threshold, the
system is separated into two phases with equality 1 2x x〈 〉 = −〈 〉 , and
h must be the same for these two phases and must be zero. Hence,
above the threshold only x〈 〉 should be defined by solving the self-
consistency equation with ( , ,0)sP x x〈 〉 .
The values of the constant effective field h are obtained as solu-
tions of the self-consistency equation with initial concentration
0 0.2x = and shown in Fig. 17. As seen from Fig. 17, a, the field h
decreases monotonically with an increase in the control parameter
ε . If h becomes trivial, we get the transition point Tε . After this
point, 0h = , and the mean-field value x〈 〉 can be calculated self-
consistently. It is clearly seen that the internal multiplicative noise
shifts the transition point toward negative values of the control pa-
rameter ε . With an increase in D , the same circumstance is observed.
The last effect is well defined: an increase in the correlation scale cr or
spatial coupling intensity D induces the ordering behaviour in the
system. The former is the combined effect of the nonlinearity of the
system, multiplicative character of the noise and the spatial coupling.
The noise induced effects are well seen in dependence
2( )h σ , shown
in Fig. 17, b. Here, at positive values of the control parameter ε (solid
76 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
line), the effective field h increases from the zero value at
2 2
Tσ > σ .
Here one has a stable homogeneous phase 0x x〈 〉 ≠ . In the case of nega-
tive values of the control parameter (dashed line), the constant effec-
tive field h becomes zero inside the domain of the noise intensities
2 2 2
1 2[ , ]T Tσ ∈ σ σ . The value h decreases till the first threshold
2
1Tσ ,
above the second one
2
2Tσ it increases monotonically. From the formal
viewpoint, the corresponding mean field value x〈 〉 obtained as a solu-
tion of the self-consistency equation (115) should be nontrivial inside
the domain
2 2 2
1 2[ , ]T Tσ ∈ σ σ where the phase separation with 1 2x x〈 〉 = −〈 〉
occurs.
Let us discuss the mean field x〈 〉 behaviour. Here, we solve the self-
consistency equation, setting 0h = . As Figure 18, a, shows the mean
field value changes, its value critically from zero if the parameter ε
increases. The critical point cε is defined as a bifurcation point when
a
b
Fig. 17. Constant effective field h versus control parameter ε (a) and
noise intensity (b) at fixed initial value 0 0.2x = . Other parameters are
shown in legends.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 77
nontrivial values 1 2x x〈 〉 = −〈 〉 appear. The transition point Tε corre-
sponds to the case when 0x x〈 〉 = . With an increase in both the noise
intensity
2σ and the parameter α , the critical point cε is shifted to-
ward negative values. The dependence of the mean field value versus
noise intensity is shown in Fig. 18, b. Here, one can see re-entrant
phase transitions at negative values of the control parameter at large
spatial coupling intensity. With an increase in ε , the first threshold
2
1cσ is shifted toward small values whereas the second one
2
2cσ becomes
larger. Transition points
2
1Tσ and
2
2Tσ are related to the condition
0x x〈 〉 = . With an increase in the noise intensity at 0ε > , the disorder-
ing phase transition is observed.
The above calculations of the effective field h and the mean field
value x〈 〉 allow us to obtain the corresponding phase diagrams. If the
initial condition 0 0x ≠ is fixed, then one can obtain the transition
lines (dash-dotted lines) which correspond to values of the control pa-
a b
c d
Fig. 18. Mean field value obtained as solution of Eq. (115) at 0h = : (a)
dependence x〈 〉 versus ε at different values of the spatial coupling inten-
sity, noise intensity and the parameter α ; (b) dependence x〈 〉 versus 2σ at
10D = and 0.8α = and different values of the control parameter ε .
78 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
rameter Tε , the noise intensity
2
Tσ and the spatial coupling TD shown
in Figs. 17, 18. The lines of critical points (solid and dashed lines) are
obtained under condition when the bifurcations of the mean field x〈 〉
occur at 0h = (see notes in Fig. 18).
As seen from Fig. 19, a, an increase in the parameter α , yielding the
concentration dependent mobility, leads to a decrease in the values
0cε > at small spatial coupling intensity. If α increases, then the
threshold for the noise intensity grows. It results that the bulk states
with 1 2x x〈 〉 = −〈 〉 exist at large noise intensities only if the mobility
( )M x decreases more abruptly. An interesting situation can be seen
from Fig. 19, b, where the intensity of the spatial coupling is large.
Here, at negative values of the control parameter ε , the mean field
value should behave in a re-entrant manner with variation in the noise
intensity. Indeed, at small and large
2σ the e system is in a homogene-
ous state. Inside the bounded domain of the noise intensity
2σ , the sys-
a b
c d
Fig. 19. Mean field diagrams at different values of the parameter α (solid and
dashed lines of critical points plotted at 0 0x = , 0h = ) and initial conditions
0x (transition dash-dotted lines are plotted at 0 0.6x = , 0h = ). Figures (a)
and (b) correspond to 2.0D = and 10.0D = , respectively. Figures (c) and (d)
correspond to 0.2ε = and 0.2ε = − , respectively.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 79
tem is in inhomogeneous state with 0h = and 0x〈 〉 ≠ . An increase in
the parameter α decreases values ε at which the ordered state with
0x〈 〉 ≠ is observed and extends the corresponding domain of the noise
intensities. Let us consider the diagram in the plane (
2,Dσ ) shown in
Fig. 19, c. Here, at positive values of the control parameter ε , an in-
crease in the noise intensity destroys the state 0x〈 〉 ≠ , as usual. At
negative ε , we get the re-entrant behaviour of the mean field x〈 〉 ,
where with α growth transition values for the spatial coupling inten-
sity decrease, and the domain of the noise intensity with the re-entrant
behaviour extends.
Let us compare our results with computer simulations. To this end, we
have computed averaged value 2M〈 〉 , the moment
2
2M〈 〉 , and the value
2 2
2 2
2
M M〈 〉 − 〈 〉χ =
σ
(116)
that can be understood as a generalized susceptibility or variance. Av-
eraging was done over 7 experiments in a stationary limit ( ∞→t ) at
the time interval of
410t = to
42.3 10t = ⋅ . Obtained results are shown
in Fig. 20. From Figure 20, a, one can see a nonmonotonic behaviour of
the order parameter 2M〈 〉 versus noise intensity. The variance χ in
Fig. 20, b, shows two peaks on observable values of the noise intensity
2σ in the corresponding interval. Peaks in
2( )χ σ dependence are re-
lated to two thresholds of the re-entrant phase transition. Finally, to
show the entropy-driven phase transitions mechanism in the system
under consideration, let us study a topology change of the distribution
st ( ; , 0)P x x h〈 〉 = at fixed mean-field values 0x〈 〉 > . In the ordered (in-
homogeneous) state, the number of extrema of the above distribution
is changed. Indeed, in the domain of point A (see Fig. 21), the stochas-
tic distribution has one peak shifted toward positive values of x due to
a b
Fig. 20. Order parameter 2M (a) and generalized susceptibility χ (b) ver-
sus noise intensity 2σ at 0.5ε = − , 10D = , 0.8α = .
80 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
the fact that 0x〈 〉 > . With an increase in the noise intensity
2σ , mov-
ing through the dashed line toward the point B, an additional peak in
the distribution function appears. The dashed line corresponds to the
system parameters when a double degenerated point of the stochastic
distribution appears. Therefore, here we get the generalization of
noise-induced transitions for extended systems.
4.1.4. Influence of External and Internal Noise Sources
Now, let us assume the presence of the nonequilibrium medium,
which sets external fluctuations. Since the influence intensity of
the medium is determined by the control parameter ε , we may con-
sider an assumption about its fluctuations to be suitable for the de-
scription of real situations: 0 ( , )tε → ε + ζ r . We endow the Langevin
source ( , )tζ r by the Gauss properties
2( , ) 0, ( , ) ( , ) ( ) ( )t t t C t t′ ′ ′ ′〈ζ 〉 = 〈ζ ζ 〉 = σ − δ −r r r r r% (117)
with the spatial correlation function
( )
2
2
| |
( ) 2 exp
2
d
C
− ′⎛ ⎞−′− = λ π −⎜ ⎟λ⎝ ⎠
r r
r r . (118)
Fig. 21. Phase diagram of phase transitions showing the change of the sto-
chastic distribution ( ; 0, 0)stP x x h〈 〉 > = extrema. Insertions display forms
of distributions of stochastic field x . Other parameters are as follows:
0.2ε = − , 0.8α = ; the mean-field value is calculated according to the val-
ues of D and 2σ for points A and B, respectively.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 81
Here, λ is the correlation length of the external noise ζ ;
2σ% is the in-
tensity.
As a result, we arrive at the Langevin equation in the form
[ ]
( ) ( , ) ( ) ( , ),
x x
M x x t g x t
t x
⎛ ⎞∂ δ⎡ ⎤= ∇ ⋅ ∇ + ζ + ∇ ξ⎜ ⎟⎢ ⎥∂ δ⎣ ⎦⎝ ⎠
r r
F
(119)
where noises ξ and ζ are assumed to be independent.
In what follows, we again pass to a discrete space, representing the
continual equation (119) in the form
( ) ( ) ( ) ( ) ( ).i
L ij j R jl l l L ij j j
l
dx F
M x t g t
dt x
⎡ ⎤∂= ∇ ∇ + ζ + ∇ ξ⎢ ⎥∂⎣ ⎦
(120)
Here, left- and right-differences, ( )L ij∇ and ( )R ij∇ , respectively, are
consistent with definitions (90). Equation (120) can be rewritten in a
more convenient form for a further analysis
( ) ( ) ( ) ( ) ( ) ( ),i
L ij j R jl L ij j j ij j j
l
dx F
M g t g t
dt x
⎡ ⎤∂= ∇ ∇ + ∇ ξ + Δ ζ⎢ ⎥∂⎣ ⎦
% (121)
where j j jg M x=% ; the external noise ( )j tζ obeys Gaussian properties:
( ) 0i t〈ζ 〉 = ,
2
| |( ) ( ) 2 ( )i j i jt t C t t−′ ′〈ζ ζ 〉 = σ δ −% , where | |i jC − is a discrete repre-
sentation of the spatial correlation function (| |)C ′−r r .
Early Stage of the System Evolution with Two Noises. As done in pre-
vious sections, we will study the instability of the state ( , ) 0x t =r , tak-
ing only into account the linear terms in Eq. (120). In this case, the dy-
namical equation for the structure function takes the form
2 2
2 2 2 2
( , )
( ) ( , ) 2
2 2
( , ) ( ) ( , ),
(2 ) (2 )d d
dS k t
k S k t k
dt
k k
d S q t d G q S q t
= −ω + σ −
ασ σ− +
π π∫ ∫q q
%
(122)
where ( )G q is the Fourier transform of the external-noise correlation
function ( )C ′−r r . The dispersion relation reads [84]
( )2 2 2 2 2
1 0 1( ) 2 ( ) ,k k D C k d C C⎡ ⎤ω = − σ − ε + σ − σ −⎣ ⎦% % (123)
where σ% is the intensity as in Eq. (117). The dispersion relation indi-
cates that for ( ) 0kω > , the homogeneous null state is stable. This oc-
curs for
2 2 2
0[ (| |)] 0rC =−ε + ασ + σ ∇ >r% , so we can define an effective
control parameter
2 2 2
ef 0[ (| |)] 0,rC =ε = −ε + ασ − σ ∇ >r% (124)
82 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
such that the homogeneous null state is stable for ef 0ε > . Therefore,
the onset of stability is now given by
2 2 2
0[ (| |)]t rC =ε = ασ − σ ∇ r% , which in
a discrete space is
2 2
0 12 ( )t d C Cε = ασ + σ −% . Moreover, the expression
for ( )kω has a noise dependent term that can be considered as a modifi-
cation of the spatial coupling parameter. Thus, we can define an effec-
tive spatial coupling parameter
2
ef 1D D C= − σ% [90].
Mean Field Analysis of the Stationary Case. Let us consider the sta-
tionary case. To this end, we can use the standard procedure described
above to find the stationary distribution as a solution of the corre-
sponding Fokker−Planck equation. The total probability density func-
tional obeys the equation [37, 61, 62, 67]
2 2
| |
,
.
ij j jr r
ij ri j
j j j mn nj n
m nj n
V
M D x
t x x
g g g C g
x x −
⎛ ⎡ ⎤∂ ∂ ∂= Δ − + Δ −⎜ ⎢ ⎥⎜∂ ∂ ∂⎢ ⎥⎣ ⎦⎝
⎞∂ ∂−σ + σ Δ ⎟⎟∂ ∂ ⎠
∑ ∑
∑% %%
P
P
(125)
To perform mean-field calculations, we exploit the one-point probabil-
ity density satisfying the equation
( )
( ),i
ij j i
ji
P t
M P t
t x
∂ ∂= Δ 〈 〉
∂ ∂ ∑ % (126)
where
2 2
| |
,
.j j jr r j j j mn nj n
r m nj j n
V
M M D x g g g C g
x x x −
⎡ ⎤∂ ∂ ∂= − + Δ − σ + σ Δ⎢ ⎥
∂ ∂ ∂⎢ ⎥⎣ ⎦
∑ ∑% % %% (127)
Assuming that a stationary distribution ( )s iP x can be obtained under
no flux conditions, the quantity jM〈 〉%
should obey the equation
( ) 0.ij j s ij
M P xΔ 〈 〉 =∑ % (128)
Following the standard procedure, one can find that the mean-field
distribution function can be computed directly from the stationary
equation
2
2
1 0
( ) ( ) 2 ( ) ( ) ( )
2 ( ) ( ) ( ) ( ),
s
s
V
hP x M x dD x x g x g x
x x
d g x C g x C g x P x
x x
⎛ ∂ ∂⎡ ⎤− = − + 〈 〉 − − σ +⎜ ⎢ ⎥∂ ∂⎣ ⎦⎝
⎞∂ ∂⎡ ⎤+ σ 〈 〉 − ⎟⎢ ⎥∂ ∂⎣ ⎦ ⎠
% % %%
(129)
where, according to the mean-field approximation, one can put
( ) ; ( )g x g x〈 〉 〈 〉 [37], and drop the prime for 2D dD′ = . Therefore, for the
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 83
stationary distribution one has
( ; ; )
( , , ) exp ,
( ; )s
x x h
P x x h N dx
x x
′Ω 〈 〉⎛ ⎞′〈 〉 = ⎜ ⎟′Θ 〈 〉⎝ ⎠
∫ (130)
where
f x M x g x
x x h M x D x x d C h
x x x
∂ σ ∂ ∂⎡ ⎤Ω 〈 〉 = − + 〈 〉 − − − σ +⎢ ⎥∂ ∂ ∂⎣ ⎦
2 2
2
0
( ) ( ) ( )
( ; ; ) ( ) ( ) ,
2
%
%
(131)
a
b
Fig. 22. Dependence of the mean field versus the noise intensities (internal and
external) at 0.4α = , 0.2ε = − , 10D = (a) and the phase diagram at different
values of the correlation radius λ , spatial coupling parameter D and α (b)
(curves 1−3 correspond to 0λ = , whereas curves 1′−3′ relate to 1λ = : 1 and 1′–
8.3D = , 0.4α = ; 2 and 2′– 10D = , 0.4α = ,3 and 3′– 8.3D = , 0.6α = ).
84 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
2 2
0 1( ; ) ( ) 2 ( )( ( ) ( )).x x M x d g x C g x C g xΘ 〈 〉 = σ + σ − 〈 〉% % %%
Unknown field h and x〈 〉 can be obtained form the self-consistency
equation described above.
Let us consider an influence of spatial correlations λ of external
noise on a position of the critical points. Corresponding phase dia-
grams are shown in Fig. 22 at different values of D and α (curves 1−3)
and at different values of spatial correlation radius λ .
As figure shows an increase in D leads to decrease in critical noise
intensity magnitudes
2σ% and promotes re-entrance in phase transition
picture (see curves 1, 2). The same situation is observed when the pa-
rameter α increases (see curves 1, 3). An increase in the correlation
radius of the external noise results in increase in its critical values (see
curves 1′−3′). It leads to the fact that area of re-entrant behaviour of
the order parameter shrinks.
Strong Coupling Limit. Considering a strong coupling limit, we as-
sume D → ∞ and neglect all possible correlations, i.e. ( ) ; ( )x x〈ϕ 〉 ϕ 〈 〉 .
Hence, the stationary distribution functions are given by the mean
field approach for each phase and have the form ( , ) ( )sP x x x x〈 〉 = δ − 〈 〉 .
Next, to obtain an equation for the effective field h , we integrate Eq.
(112) and find
2( ) ( ) ( 2) ( ),h M x V x M x′ ′= 〈 〉 〈 〉 − σ 〈 〉 (132)
where prime denotes derivative with respect to the argument.
In the homogeneous case, Tε < ε , we have 0x x〈 〉 = , and h becomes a
function of the initial conditions. If the value 0x is fixed, then the
field h decreases with an increase in ε until it reaches the null value,
and increases from the null value with an increase in
2σ . As it follows
from the Eq. (132) and mean field analysis, no re-entrance can be found
in strong coupling limit. Therefore, a re-entrant behaviour of the
mean-field value is realized only at finite magnitudes of the spatial
coupling intensity D .
In the case of Tε > ε , we have 0h = . Solutions of Eq. (132) give val-
ues for the bulk phases
( )1/2
2 2 2
1,2
2
1 (1 ) 4 .
2
x
α
〈 〉 = ± − − αε + − αε + α σ
α
(133)
The corresponding transition lines are defined by condition 1 0x x〈 〉 =
that leads to
2 2 2 2
0 0
2
0
(1 )
.
1T
x x
x
α σ − + α
ε =
+ α
(134)
The critical point (for 0 0x = ) is
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 85
2 2,cε = α σ (135)
which coincides with linear stability analysis. (One should not be con-
fused comparing our results with results in Refs [76, 84]. We consider
the internal multiplicative noise influence, whereas in Refs [76, 84]
systems with an external multiplicative noise and conserved dynamics
were studied. In above papers, it was noted that the critical value cε
differs from the obtained in Eq. (135) by the multiplier 2d . Such mul-
tiplier arises only if an external noise is introduced. If we follow the
procedure described in Ref. [76] and consider external fluctuations,
supposing 0 ( , )tε → ε + ση r , where ( , )tη r is the Gaussian noise, then
we will recover results with the multiplier 2d .) It follows that the
strong coupling limit sets the critical values for both the control pa-
rameter and the noise intensity, which correspond to 0ck = . In other
words, at
2ε > ασ , unstable modes start to growth. Moreover, we can
define values cα < α = ε/σ2
at which phase separation exhibits spatial
patterns. Due to [ 1, 1]ε ∈ − and
2 0σ > , 0α ≥ , one gets that cα de-
creases with an increase in the noise intensity
2σ . At negative values of
the control parameter, we get the phase separation with no patterning.
For the system with two stochastic sources, one can find that critical
value for the control parameter is renormalized as
2 2
0 12 ( ).c d C Cε = ασ − σ −% (136)
From this, it follows that two stochastic sources compete with each
other. Here, we get shift of the critical point with the multiplier 2d
related to the noise intensity
2
0Cσ% and spatial correlations 1C .
4.2. Stochasticity in Hyperbolic Transport
In this Section, we focus on a more conceptual question related to the
fluctuations of the flux J or, more concretely, to their conceptual in-
terpretation. We shall see that, depending on the value of the relaxa-
tion time Dτ and of the observational time scale, such fluctuations can
be interpreted in two different ways described in Ref. [40]. To discuss
these ideas, we must recall some results concerning hydrodynamic sto-
chastic noise.
Hyperbolic Transport with Noise. We discuss now the stochastic noise
in a system,
,Lτα + α = − α + ζ&& & (137)
which may generally represent the hyperbolic transport (3) with noise.
Taking into account both independent variables α and α ≡ β& , Eq.
(137) may be written as
86 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
α βα = β + ζ β = − α − β + ζ
τ τ
% & %&
1
, .
L
(138)
The set of equations (138) represents the second-order equation (137)
as two first-order evolution equations, in a way that the system be-
comes Markovian. In Eq. (138), αζ and βζ are the respective stochastic
sources, whose second-order moments have to be obtained. The correla-
tor of fluctuating terms is defined as follows
2 2
eq eq eq eq
2 2
eq eq eq eq
0 1 0
( ) ( ) 1
1
1
L
t t L
⎡ ⎤−⎡ ⎤ ⎢ ⎥⎡ ⎤ ⎡ ⎤〈α 〉 〈αβ〉 〈α 〉 〈αβ〉 τ⎢ ⎥′〈ζ ζ 〉 = + ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ 〈βα〉 〈β 〉 〈βα〉 〈β 〉 ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ −⎢ ⎥τ τ⎣ ⎦ ⎢ ⎥τ⎣ ⎦
% % . (139)
In Equation (139), we have eq eq 0〈αβ〉 = 〈βα〉 = because α and α& have
opposite time-reversal symmetry. Thus, it is found
α β α β β α〈ζ 〉 = 〈ζ 〉 = 〈β 〉 〈ζ ζ 〉 = 〈ζ ζ 〉 = 〈α 〉 − 〈β 〉
τ τ
% % % % % %2 2 2 2 2
eq eq eq
2
0, , .
L
(140)
To obtain the second moment of equilibrium fluctuations of α and
β , we assume, as in Section 2, that α and β are independent variables.
Then, including both of these variables into the entropy, one can write
2 2
eq
1 1
( , ) ,
2 2
S S A Bα β = − α − β (141)
where only second-order terms have been considered. According to Eq.
(7), the probability of fluctuations is described by
2 2( , ) ~ exp .
2 2B B
A B
Pr
k k
⎡ ⎤
α β − α − β⎢ ⎥
⎣ ⎦
(142)
Note that, in Eq. (142), we have identified the fluctuations δα and δβ
with α and β , respectively, because their equilibrium average values
are zero for both of them due to the form of the evolution equations
(138). Then, in equilibrium, we have
2 2
eq eq, .B Bk k
A B
〈α 〉 = 〈β 〉 = (143)
To evaluate the ratio /A B , we obtain the entropy production corre-
sponding to Eq. (141). This one yields
[ ] 0.dS dt A B A B A B= − αα − ββ = − αα − αα = −α α + α ≥&& & & && & && (144)
From this, and by following the usual methods of nonequilibrium
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 87
thermodynamics [91], it follows the linear relation between thermody-
namic flux −α& and its conjugated force A Bα + α&& . This is
.A Bα + α = −μα&& & (145)
Comparison of Eq. (145) with Eq. (137) yields B = τ , 1μ = , and
A L= . Thus, as follows, the second moments of α and β are related by
2
eq
2
eq
.
B
A L
〈α 〉 τ= =
〈β 〉
(146)
Introducing this relation into Eq. (140) it is found
L
α α β α β β〈ζ 〉 = 〈ζ ζ 〉 = 〈ζ ζ 〉 = 〈ζ 〉 = 〈β 〉 = 〈α 〉
τ τ
2 2 2 2
eq eq2
2 2
0, 0, .% % % % % % (147)
Since
2 0α〈ζ 〉 =% , the first equation in Eq. (138) may be introduced into
the second one. Therefore, we get
,L βτα + α = − α + τζ%&& & (148)
with
2
eq(0) ( ) (0) ( ) 2 .t t Lβ β β β〈ζ ζ 〉 = 〈τζ τζ 〉 = 〈α 〉% % (149)
Thus, the expression for the noise keeps the same form as in the case
with 0τ = . This is in agreement with the ideas of fluctuation-
dissipation, which relate the fluctuations to the dissipative part of the
equation [the term in L in Eq. (148)].
The transition from noisy hyperbolic transport described by equa-
tion (137) to Langevin equation Lα = − α + ζ& might be analyzed by con-
sidering the generalized entropy (141) in the following form
2 2
eq( , ) .
2 2
A A
S S
L
τα β = − α − β (150)
With 0τ → , the last term in
2β disappears together with the term in
α&& in Eq. (137). In this case, the dynamics of α is described by a simple
relaxation with a temporal constant given by
1L− . One interesting
situation may be found when
1 1L−τ << << . In this case, α decays
slowly and β decays fast. Assume, for instance, that 1≈L s
−1
and
310−τ ~ s. The typical relaxation of α will be of the order of 1 second
and β will decay in a millisecond scale. In this case, βζ% describes the
effect of all the variables whose relaxation time is much less than
310−
s, in such a way that they may be assumed to decay instantaneously in
comparison with β .
Fluctuations or Independent Variable. Now, consider a special system
88 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
with two independent variables α and β in such a way that the relaxa-
tion rate of β is characterized by its high but finite value. In this case,
one may write Lδβ ≡ β + α , i.e. = Lβ − α + δβ , where δβ being the inde-
pendent part of β . This part of β is orthogonal to the slow subspace
generated by α . Then, we may write
α = − α + δβ δβ = − δβ
τ
&&
1
, .L (151)
and
| | 2 | |(0) ( ) .t tBLk L
t e e
A
− τ − τ〈δβ δβ 〉 = = 〈α 〉
τ τ
(152)
The following three cases may be outlined in considering Eqs (151)
and (152).
(i) If τ is sufficiently short, δβ acts as a ‘noise’ in the equation for α
(the first equation of system (151)).
(ii) If τ is not completely negligible as compared to
1L− , δβ acts as a
coloured noise.
(iii) In the limit 0τ → , one has
2
eq(0) ( ) 2 ( ).t L t〈δβ δβ 〉 → 〈α 〉 δ (153)
Thus, it is seen that the transition of Eq. (150) and (151) from small
τ to vanishing τ is conceptually interesting. It is illustrative of how
the variable β (i.e., α& ) goes from an independent variable with its own
dynamics to a purely Markovian stochastic noise. In physical terms,
the frontier between small τ and vanishing τ is settled by the time
scale one is able to measure. For instance, if 1L = s
−1
and
810−τ = s,
the system will have two independent variables, α and α& , for an ob-
server which is able to measure picoseconds (
1210−
s). However, it will
have only one independent variable, α , plus a stochastic noise, in a
form of α& , for an observer which is only able to measure milliseconds
( 310−
s). The latter observer will be able to work in the ‘adiabatic’ ap-
proximation with very slow α in comparison with β .
The situation described by Eqs (1) and (2) is interesting from this
perspective. For negligible values of the diffusion relaxation time Dτ ,
one should use the Landau−Lifshitz formalism for the fluctuating hy-
drodynamics [28] and write
2 ,c
c
D c
t
∂ = ∇ + ζ
∂
(154)
where the stochastic noise cζ is interpreted as
cζ = −∇ ⋅ δJ (155)
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 89
with δJ being a fluctuating part of the diffusion flux, i.e.
.D c= − ∇ + δJ J (156)
According to the Landau−Lifshitz approach [28], one has
(0) ( ) 2 ( ).Bt k DT t〈δ δ 〉 = δJ J (157)
Assume, in contrast, that Dτ is small but still measurable. In this
case, we get
, ,D J
c
D c
t t
∂ ∂= −∇ ⋅ τ + = − ∇ + ζ
∂ ∂
J
J J (158)
with c and J being independent variables of the entropy given by
the generalized Gibbs equation (8). The second moments of the fluc-
tuation of c and J are given by Eq. (11). The noise Jζ would cor-
respond to values relaxing in time scales much shorter than Dτ , in
such a way that it may be considered as Markovian noise:
(0) ( ) 2 ( ).J J Bt k DT t〈ζ ζ 〉 = δ (159)
In the limit of vanishing Dτ , the fast part of J becomes a stochas-
tic noise, and we get cζ in Eq. (154) described by Eq. (157).
Still another form to discuss the interpretation of noise in the
context of Eq. (137) is to write Eq. (138) without any added noise,
i.e., in the following form
, .
D D
L βα = β β = − α −
τ τ
&& (160)
This set of equations may be integrated to give
0
( ) ( ) ( ) ( ),
t
t
t M t t t dt t′ ′ ′α = − − α + β∫& (161)
with the memory function
( ) exp ,
D D
L t t
M t t
⎛ ⎞′−′− = −⎜ ⎟τ τ⎝ ⎠
(162)
and the exponential relaxation
0
0( ) ( ) exp .
D
t t
t t
⎛ ⎞−
β = α −⎜ ⎟τ⎝ ⎠
& (163)
In this case, the ‘noise’ is due to the uncertainty in the value of
( )tα& at the initial time 0t .
90 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
Mori’s expression for the fluctuation-dissipation theorem states
that
2
eq( ) ( ) ( ) .t t M t t′ ′〈β β 〉 = − 〈α 〉 (164)
Indeed, if we use the result (146), we obtain that
2 2 2
0 eq 0 eq eq( ) ( ) .
L
t t〈α 〉 ≡ 〈β 〉 = 〈α 〉
τ
& (165)
Combination of Eq. (165) with Eqs (162) and (163) gives Eq. (164).
Again, when τ becomes negligible, ( )M t t′− given in Eq. (162) be-
comes
( ) ( ),M t t L t t′ ′− = δ − (166)
and Eq. (161) becomes
( ) ( ) ,t L t αα = − α + ζ& (167)
with
2 22 .Lα〈ζ 〉 = 〈α 〉 (168)
Note that, in the limit (166), we consider t t′> , whereas in the limit
(168), a factor 2 appears because one considers | | 0t t′− > rather than
0t t′− > , i.e. one considers both 0t t′− > and 0t t′ − > .
4.3. Stochastic Hyperbolic Model for Spinodal Decomposition
Considering the stochastic hyperbolic model, let us start with a set of
equations for the local concentration and the flux in the forms as follow:
,x
t
∂ = −∇ ⋅
∂
J (169)
D
x
M x t
t x
∂ δτ = − − ∇ + ζ
∂ δ
J J r
[ ]
( ; , )
F
. (170)
Here, Dτ is a relaxation time for the flux J . In further analysis, let us
introduce a dimensionless time / xt t′ = τ and operator
1
x x
−′∇ = ∇l ,
where scales xl and xτ are introduced (usually, the time scale for such
transition is defined through the Debye frequency Dω , diffusion en-
ergy diffE and temperature T as
1
diffexp( / )x D E T−τ = ω ). If the diffu-
sion is caused by the vacancy mechanism, then the quantity diffE is en-
ergy for migration of vacancies. Next, let us move to dimensionless
quantities / DJ J V′ = with /D x xV = τl , 0/M M M′ = , 0/′ ′=F F F ,
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 91
DV′ζ = ζ for flux, mobility and free energy, and noise respectively.
Inserting such quantities into Eq. (169), dropping the prime, and as-
suming 0 0 D x xM V= τlF , we reduce the set of two first-order time equa-
tion to the second-order equation in the form of
( )
2
2
( ) ( ) ( , ) ,
x x
M x g x t
t t
∂ ∂δ + = ∇ ⋅ ∇μ + ζ
∂ ∂
r (171)
where /D xδ = τ τ states the ratio of relaxation times for
( ) ( )g x M x= . At 1δ << , one arrives at the parabolic model for the
spinodal decomposition discussed previously in Section 4.1.
In our consideration, it can be shown that two conjugate variables,
as the local concentration and the flux, should be considered as com-
mensurable variables as a special case. Moreover, we will explain that
even the flux is supposed to be fast variable our results leads to the
well-known picture of nonlinear dependence of an amplification rate at
early stages, whereas at late stages (stationary case), the hyperbolicity
of the model does not affect on the system behaviour essentially.
4.3.1. Early Stages Analysis
Let us consider an early stage of the system evolution. As done before,
we can obtain an evolution equation for the spherically averaged struc-
ture function ( , )S k t . In the following, we use the special case of the
white noise assumption, C t t t t′ ′ ′ ′− − → σ δ − δ −r r r r2( , ) ( ) ( ) . After some
algebra, one gets
( )
2
2 2 2
2
2 2 2 2
( , ) ( , )
1
2 2 ( , ).
(2 )d
d d
S k t k Dk S k t
dt dt
k k d S q t
⎛ ⎞
δ + = − − ε + ασ +⎜ ⎟
⎝ ⎠
+ σ − ασ
π ∫ q
(172)
This equation is reduced to Eq. (63) in the deterministic case
( 2 0σ = ), whereas in the case of 1δ << and 0α = one arrives at the
Cahn−Hilliard−Cook equation [12, 59]. It can be seen that the dis-
persion relation now takes the form
2 2 2
2
( ) 1
( ) ,
2 4
i k k D
k
− ε + ασω = − ± −
δ δ δ
(173)
which, in the case of the local equilibrium, 1δ << , is as follows:
1 2 2 2
1
( ) (2 ) 1 2 ( ) .k i k k D−
δ=
⎡ ⎤ω = − δ ± δ − ε + ασ⎣ ⎦ (174)
At ck k< , the imagine part of the frequency (173)
92 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
− ⎡ ⎤ℑ ω = δ ± − δ − ε + ασ
⎣ ⎦
1 2 2 2( ( )) (2 ) 1 1 4 ( )k k k D (175)
can be used to obtain maximal wave-vector amplitude mk .
To obtain a normalized amplification rate, we need to compare dis-
persion relations obtained for both parabolic and hyperbolic stochastic
models. To that end, we use the quantity CHC m( )kω obtained from the
parabolic Cahn−Hilliard−Cook model with multiplicative fluctuations
as a normalization factor,
2
m ( ) / 2k D= ε − ασ (see Subsection 4.1.1).
Acting in such a manner, the normalized amplification rate takes the
form
* 2 2
hyp hyp CHC m( ) / ( ) / ( ) /q q k k qω = ω ω , where / cq k k= , and disper-
sion relation hyp( )kω is taken from Eq. (175) as ω = ℑ ωhyp ( ) ( ( ))k k .
Therefore, for the stochastic hyperbolic model one has
2 2
2 2
* 2
2 2 2
( )
1 4 (1 ) 1
2
( ) .
( )
q q
D Dq q
q
ε − ασ+ δ − −
ω =
δ ε − ασ
(176)
In such a case, both parabolic and hyperbolic normalized amplification
rates take values:
2( ) / 0q qω = at 1q = and
2( ) / 4q qω = at 0q = . At
0δ → (transition to the one slow variable (parabolic) model), one ar-
rives at the linear dependence
2( ) /q qω versus
2q .
The corresponding dependence from Eq. (176) is shown in Fig. 23.
Comparing different curves, one can see that in the nonlinear behav-
iour appears only if
2 2( ) 0δ ε − ασ ≠ . It means that in the deterministic
case
2 0σ = the nonlinearity is caused by 0δ ≠ , whereas the stochastic
contribution leading to renormalization of the control parameter ε is
able to promote essential contribution to the above effect. Indeed, at
large difference between
2ασ and | |ε at fixed α , the nonlinear effect
becomes well pronounced.
To relate the stochastic approach to the deterministic one, let us re-
write the dimensionless dispersion relation in the form
2 2 2( ) 1
( ) ,
2 4
xx ck f r ki
k
′′ +
ω = − ± −
δ δ δ
%
(177)
where the notation
2
xx xxf f′′ ′′= + ασ%
is used, 0|xx xf =′′ = −ε . From this, it fol-
lows that the internal noise influence leads to a change in the barrier
height for the effective free energy f% due to
2
0|xx xf =′′ = −ε + ασ% . Moving
back to dimensional variables, let us interpret our results for the sto-
chastic case. Here, instead of originally exploited correlation and dif-
fusion lengths, /C c xxl r f′′= − and 0D xx Dl M f′′= − τ , respectively, it is
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 93
convenient to use the effective lengths, /C c xxl r f′′= − %%
and
0D xx Dl M f′′= − τ%% . In such a case, we arrive at formula for the amplifica-
tion rate given by Eq. (66). The difference in deterministic and sto-
chastic cases lies only in the renormalization of the barrier height for
the effective free energy f% . An increase in the intensity
2σ of multi-
plicative fluctuations determined by the field dependent mobility de-
creases the length
2
Dl ∝ ε − ασ%
and increases the scale
21 /Cl ∝ ε − ασ% . The local nonequilibrium effect at small
2σ is en-
hanced by the diffusion flux relaxation that is in good correspondence
with results obtained for the deterministic case analysis.
4.3.2. The Effective Fokker−Planck Equation for the Hyperbolic Model
To describe the system states, we need to know the distribution func-
tion of the field x . In order to get it, one should obtain the correspond-
ing Fokker−Planck equation. According to the standard procedure, we
represent our system in a discrete d-dimensional space with
dN cells
with the mesh size l . Then, following Eq. (171), the system dynamics
will be described by a set of equations for every cell of the space:
Fig. 23. Comparison of the function * 2( ) /q qω for hyperbolic (modified
Cahn−Hilliard) model for deterministic and stochastic cases at different
values of the ratio /D xδ = τ τ and different noise intensities 2σ . Other pa-
rameters are as follows: 0.5ε = , 1D = , 0.8α = .
94 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
2
2
.L Ri i
ij j jl j j
l
d x dx F
M g
dt dt x
⎛ ⎞∂δ + = ∇ ∇ + ζ⎜ ⎟∂⎝ ⎠
(178)
Next, let us introduce a new variable, ip , playing a role of an effective
momentum, i ip x= δ & . Then, by definition, the probability density
function is given by the averaging of the density function ( , , )x p tρ
over noise: ( , , ) ( , , )x p t x p t= 〈ρ 〉P . To construct an equation for the
macroscopic density function P , we exploit the conventional device to
proceed from the continuity equation:
0i i
i i i
x p
t x p
⎡ ⎤∂ρ ∂ ∂+ + ρ =⎢ ⎥∂ ∂ ∂⎣ ⎦
∑ & & . (179)
Inserting the momentum definition, we obtain
( ) ,
t
∂ρ = + ζ ρ
∂
L N (180)
where the operators ii
= ∑L L and ii
= ∑N N are defined as follows
1
,L Ri
i ij j jl i
i i l
p F
M p
x p x
⎛ ⎞∂ ∂ ∂≡ − − ∇ ∇ −⎜ ⎟δ ∂ ∂ ∂ δ⎝ ⎠
L (181)
.L
i ij j
i
g
p
∂≡ −∇
∂
N (182)
Within the interaction representation, the density function reads
exp( )t℘ = − ρL that allows to rewrite Eq. (180) as
,
t
∂ ℘ = ℘
∂
R (183)
( , , ) .t t
i i i i
i i
x p t e e−⎡ ⎤= ≡ ζ ⎣ ⎦∑ ∑ L LR R N (184)
The well-known cumulant expansion method serves as standard and
effective device to solve such a type stochastic equation [92]. Neglect-
ing terms of the order
3( )O R , we get the kinetic equation for the aver-
aged quantity ( )t〈℘ 〉 in the form
0
( ) ( ) ( ) ( ).
t
t t t dt t
t
⎡ ⎤∂ ′ ′〈℘〉 = 〈 〉 〈℘〉⎢ ⎥∂ ⎣ ⎦
∫ R R (185)
Within the original representation, the equation for the probability
density reads
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 95
( )
0
( ) ( ) ( ).
t
P t C e e d P t
t
τ − τ⎧ ⎫∂ ⎪ ⎪⎡ ⎤= + τ τ⎨ ⎬⎣ ⎦∂ ⎪ ⎪⎩ ⎭
∫ L LL N N (186)
Due to the physical time much larger than a correlation scale ( )t >> τ ,
one can replace the upper limit of the integration by ∞ .
To proceed this one, we use the procedure proposed in Ref. [69, 70]
to obtain the effective Fokker−Planck equation for the hyperbolic sto-
chastic model. Expanding exponents, we arrive at the equation
( ) ,P Pt
∂ − =∂ L C (187)
where the operator
(1) (2)≡ +L L L (188)
has the components
(1) (2)
1
, .L Ri
ij j jl i
i ii l i i
p F
M p
x x p p
⎛ ⎞∂ ∂ ∂ ∂≡ − + ∇ ∇ ≡⎜ ⎟δ ∂ ∂ ∂ δ ∂⎝ ⎠
∑ ∑L L (189)
The collision operator C is defined as follows:
∞ ∂= = = ≡ − ∇
∂∑ ∑( ) ( ) ( ) ( ) (0)
=0
, ( ), , n n n n L
ij j
n i i
C C g
p
C M NL L N N , (190)
where
( )nL in the collision operator C is defined through the commuta-
tor
( ) ( 1)[ , ]n n−=L L L ; moments of the noise correlation function ( )C τ are
defined as follows:
( ) 1
0
( !) ( ) .n nn C d
∞−= τ τ τ∫M (191)
Substituting all definitions into commutators, one can calculate the
collision operator in the form:
( )
2
(0) (1)
2
2 2
(1)
2
.
L R
ij j jl l
i i
R
jl lL R L
ij j jl l ij j i
i i l i i
g g
p
g
g g g p
x p x p p
∂= − − ∇ ∇ +
∂
⎧ ⎫⎛ ⎞∂∇ ⎛ ⎞∂ ∂ ∂⎪ ⎪+ ∇ ∇ − ∇ +⎜ ⎟⎨ ⎬⎜ ⎟⎜ ⎟∂ ∂ ∂ ∂ ∂⎝ ⎠⎪ ⎪⎝ ⎠⎩ ⎭
∑C M M
M
(192)
Noting that, in further consideration, we are interesting in behav-
iour of the distribution ({ }, )ix tP , not the total one, ({ },{ }, )i iP x p t . The
reduced distribution can be obtained according to the moments
( ) ({ }, ) ({ },{ }, ) d ,n n
i i i i i
i
P x t x p t p p⎡ ⎤≡ ⎣ ⎦∏∫P (193)
96 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
where the integration is provided over a set { }ip . Then, performing
corresponding manipulations with Eq. (187), we arrive at the recursive
relations for the moments
( ) ({ }, )n
iP x t , where
(0) ({ }, )iP P x t≡ , the first
moment
(1) ({ }, )iP x t can be considered as a flux J of the probability
density, i.e.
(1)P = J .
Indeed, taking the zeroth moment of momentum, p , over Eq. (187),
we obtain the expected continuity equation
1
.t
i i
P
x
∂∂ = −
δ ∂∑ J (194)
The first-moments’ calculation leads to
1 L R
t ij j jl
i l
F
M P
x
⎡⎛ ⎞∂∂ = − + ∇ ∇ +⎢⎜ ⎟δ ∂⎢⎝ ⎠⎣
∑J J
(2)
(1) 1
R
jl lL R L
ij j jl l ij j
i l i
gP P
g g g P
x x x
⎤⎧ ⎫⎛ ⎞∂∇∂ ∂⎪ ⎪ ⎥∇ ∇ + ∇ −⎜ ⎟⎨ ⎬⎜ ⎟∂ ∂ δ ∂ ⎥⎪ ⎪⎝ ⎠⎩ ⎭ ⎦
+ M . (195)
For the second moment, we obtain
(2)
(1) (0)1
( ) .L R
ij j jl l
i
P
g g P
x
∂ = − ∇ ∇
δ ∂
M M (196)
As a result, the evolution equation for the flux J is of the form
⎡⎛ ⎞⎛ ⎞∂∇∂⎢⎜ ⎟∂ = − + ∇ ∇ − ∇ +⎜ ⎟⎜ ⎟⎜ ⎟δ ∂ ∂⎢ ⎝ ⎠⎝ ⎠⎣
⎤∂+ ∇ ∇ ⎥∂ ⎦
∑ (1)
(0)
1
.
R
jl lL R L
t ij j jl ij j
i l l
L R
ij j jl l
i
gF
M g P
x x
g g P
x
J J M
M
(197)
Therefore, we arrive at the set of two differential equations, Eq. (194)
and Eq. (197). Eliminating the flux J , we, finally, get the Fok-
ker−Planck equation for the hyperbolic stochastic model in the form
⎡ ⎤⎛ ⎞∂∇∂ ∂ ∂ ∂δ + = − ∇ ∇ − ∇ −⎢ ⎥⎜ ⎟⎜ ⎟∂ ∂ ∂ ∂∂ ⎢ ⎥⎝ ⎠⎣ ⎦
∂− ∇ ∇
∂
∑
∑
2
(1)
2
2
(0) .
R
jl lL R L
ij j jl ij j
i i l j
L R
ij j jl l
ij i j
gP P F
M g P
t x x xt
g g P
x x
M
M
(198)
In the continuum space, the obtained effective Fokker−Planck equa-
tion for the probability density functional is as follows:
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 97
⎡ ⎤δ δ δ⎛ ⎞δ∂ + ∂ = − ∇ ∇ − ∇ ∇ −⎢ ⎥⎜ ⎟δ δ δ⎝ ⎠⎣ ⎦
δ− ∇ ∇
δ
∫
∫
(1)
2
2
(0)
2
( ) ( ) 2 ( )
( ) .
( )
tt t
M
P P d M P
x x x
d M P
x
r
r r r
r
r
F M
M
(199)
The stationary probability density can be obtained explicitly under
no flux conditions. Indeed, here, we arrive at the result
( )(0) (1)
(0)
1
[ ] exp [ ] ( ) ln ( )sP x x d M x
⎧ ⎫∝ − + −⎨ ⎬
⎩ ⎭∫ rF M M
M
. (200)
This is similar to that presented in Eq. (100). The main difference is
that in the distribution (200) we have a contribution related to the first
moment of the noise correlation function with respect to the time. It
means that the stochastic source appeared in the hyperbolic model
should have smeared temporal correlation function. In other words, its
frequency spectrum should have some cut-off frequency and the noise
should be considered as a coloured noise in time. It looks absolutely
natural, because as it was shown in previous subsection the stochastic
hyperbolic transport should be characterized with fixed correlation
time, ζτ , which may be small comparing to the relaxation time for the
concentration field. Taking the temporal correlation function in the
exponential form related to the Ornstein−Uhlenbeck process ζ ,
( )1( ) exp | | /C t t t t−
ζ ζ′ ′− = τ − − τ , one can find
(0) 2= σM ,
(1) 2
ζ= τ σM . It is
seen that the temporal correlation radius ζτ makes the renormalization
of the critical values for the noise intensity, whereas main results ob-
tained for the stochastic parabolic model remain the same.
Making use the fluctuation dissipation relation one can relate the
correlation scale with the relaxation time for the flux, ; Dζτ τ . It follows
that, in the white noise limit characterized by 0ζτ → , one gets the
parabolic model for the spinodal decomposition, whereas one arrives at
the hyperbolic model at fixed but small ζτ . This conclusion is in a good
correspondence with the linear stability analysis, where it was shown
that in order to get the nonlinear dependence of the amplification rate
it is necessary to consider two modes: concentration field and flux.
5. CONCLUSIONS
A model for kinetics of fast spinodal decomposition in a binary system
free from imperfections and with the molar volume independent of
concentration is developed. The model takes into account a finiteness
of the diffusion speed DV and assumes that the spinodal decomposition
may proceed with the rate of the order of DV . Such an approach leads to
the description with independent variables of the concentration and
98 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
atomic diffusion flux having different relaxation times to their own
local equilibrium values. It leads to a modified Cahn—Hilliard equa-
tion, describing the spinodal decomposition for both diffusion and
wave propagation of atoms (components of the binary system). In such
a case, the equation describes kinetics of fast separation the rate of
which compatible with the diffusion speed DV .
To describe fast decomposition, the free energy density
e ne( , , )f c c f f∇ = +J includes standard local equilibrium contribution
e ( , )f f c c= ∇ and purely non-equilibrium contribution nef proportional
to the square of the atomic flux, ⋅J J . The use of ( , , )f c c∇ J in such a
form has a statistical basis: there are many particles within every local
volume, and it includes a reduction of the available phase space for
each particle. It is shown that the fast spinodal decomposition is de-
scribed by a hyperbolic type of differential equation. This description
is true if the time scale of the process of phase separation has the order
of the relaxation time Dτ . It occurs for the case of the fast frequency
and short wave-lengths which cannot be neglected in the description of
evolution from an unstable state to a new metastable state in spi-
nodally-decomposing system and, generally, for fast moving inter-
faces. Despite classic Cahn−Hilliard scenario described by parabolic-
diffusion equation predicts much more diffuse boundaries, the hyper-
bolic scenario exhibits evolution with sharper boundaries between two
separating phases. It occurs due to description of hyperbolic evolution
by the equation with a finite diffusion speed and description of the
parabolic evolution by the equation with an infinite diffusion speed.
The provided analysis leads to the obtaining the phase and group
atomic speeds. The real values for speeds define the finite propagation
of a single harmonic (for the phase speed) and a packet of harmonics
(for the group speed). The hyperbolic model is able to give prediction
for scenario from very earliest up to latest of spinodal decomposition.
We considered the fluctuations of the solute density and the solute dif-
fusion flux at the equilibrium steady state. The power spectra of a sol-
ute number density and a solute diffusion flux have been reviewed.
The latter has a non-vanishing relaxation time leading to a hyperbolic
transport equation for the evolution of the density. Several interpreta-
tions of the stochastic source related to fast variables eliminated from
the description have been examined. Particularly, the phase separation
scenario of the system with internal multiplicative noise related to the
field dependent mobility for parabolic (Cahn—Hilliard) and hyperbolic
models is analyzed. Analysis was performed for early and late stages of
the evolution analytically and by computer simulations. The stationary
case is considered with the help of the mean field approach.
For the system undergoing spinodal decomposition with the field
dependent mobility, we have derived the Fokker−Planck equation. It
was found that its stationary solution can be written in exact form.
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 99
Our theoretical approach shows that we dealt with the entropy driven
phase transitions mechanism. As shown, the field-dependent mobility
leads to delays in dynamics at early stages and, therefore, leads to de-
lays in domain growth law at late stages. Stationary values of the order
parameter reduced to the second moment of the stochastic field depend
on the parameter that governs the functional dependence of the mobil-
ity: with an increase in such parameter, the order is suppressed. Con-
sidering stationary states, we extend the mean field approach to the
systems with the field-dependent mobility. It was found that an in-
crease in the parameter that governs the functional dependence of the
mobility, the critical points for the phase transitions are shifted. The
system can undergo a re-entrant phase transitions when the mean field
becomes nontrivial inside the fixed domain of the noise intensity. The
strong coupling regime shows that a position of the critical point de-
pends on the constant governing the field-dependent mobility.
Our results can be applied to investigations of polymer mixtures
where relaxation flows are driven by field-dependent coefficients,
phase separation in binary alloys, and microstructure phase transi-
tions.
ACKNOWLEDGMENTS
We thank David Jou and Alexander Olemskoi for fruitful discussions
and useful exchanges. Dmitrii Kharchenko acknowledges financial
support from the Fundamental Research State Fund of Ukraine (No.
GP/F26/0010). Peter Galenko acknowledges financial support from
the German Research Foundation (DFG) under the Project No. HE
160/19 and DLR Agency under contract 50WM0736. Vladimir Lebe-
dev acknowledges financial support from the Russian Foundation of
Basic Research (RFBR) under the Project No. 08-02-91957.
REFERENCES
1. H. Jinnai, T. Hashimoto, D. Lee, and S.-H. Chen, Macromolecules, 30: 130
(1997).
2. F. Mallamace, N. Micali, and S. Trusso, J. Phys.: Condens. Matter, 8: A81
(1996).
3. N. F. Bunkin, A. V. Lobeev, and G. A. Lyakhov, Usp. Fiz. Nauk, 167: 1070
(1997) (Physics—Uspekhi, 40: 1090 (1997)); Yu. I. Ustinovshchikov and B. E.
Pushkarev, Usp. Fiz. Nauk, 176: 611 (2006) (Physics—Uspekhi, 49: 593
(2006)).
4. D. Beysens, Materials Sciences in Space (Eds. B. Feuerbacher, H. Hamacher,
and R. J. Naumann) (Berlin: Springer: 1986), p. 191.
5. J. Mainville, Y. S. Yang, K. R. Elder, M. Sutton, K. F. Ludwig, and G. B.
Sephenson, Phys. Rev. Lett., 78: 258 (1997).
100 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
6. Yu. I. Ustinovshchikov and B. E. Pushkarev, Usp. Fiz. Nauk, 176: 611
(2006) (Physics—Uspekhi, 49: 593 (2006)).
7. J. S. Langer and M. Baron, Annals of Physics, 78: 421 (1973); J. P. Donley
and J. S. Langer, Phys. Rev. Lett., 71: 1573 (1993).
8. G. F. Mazenko, Phys. Rev. B, 43: 5747 (1991).
9. K. Binder, C. Billotet, and P. Mirold, Zeitschrift für Physik B—Condensed
Matter, 30: 183 (1978); K. Binder, J. of Chemical Physics, 79: 6387 (1983).
10. K. Binder and P. Fratzl, Phase Transformations in Materials (Ed. G. Kos-
torz) (Weinheim: Wiley: 2001), p. 409.
11. V. L. Ginzburg and L. D. Landau, JETP, 20: 1064 (1950).
12. J. W. Cahn and J. E. Hilliard, J. Chem. Phys., 28: 258 (1958).
13. J. W. Cahn, Acta Metall., 9: 795 (1961).
14. J. W. Cahn, Trans. Metall. Soc. AIME, 242: 166 (1968).
15. V. P. Skripov and A. V. Skripov, Usp. Fiz. Nauk, 128: 193 (1979)
(Sov. Phys. Usp., 22: 389 (1979)).
16. S. Bastea and J. L. Lebovitz, Phys. Rev. E, 52: 3821, (1995).
17. B. Liu, H. Zhang, and Y. Yang, J. Chem. Phys., 113: 719, (2000).
18. K. Binder, H. L. Frish, and J. Jäckle, J. Chem Phys., 85: 1505 (1986).
19. P. Galenko, Phys. Lett. A, 287: 190 (2001).
20. P. Galenko and D. Jou, Phys. Rev. E, 71: 046125 (2005).
21. D. Jou, J. Casas-Vazquez, and G. Lebon, Extended Irreversible Thermody-
namics (Berlin: Springer: 2001).
22. M. Grasselli and V. Pata, J. Evolution Equations, 4: 27 (2004).
23. S. Gatti, M. Grasselli, A. Miranville, and V. Pata, J. Math. Anal. Appl., 312:
230 (2005).
24. M. Grasselli, G. Schimperna, and S. Zelik, arXiv:0804.0988v1 [math.AP].
25. A. Segatti, Math. Models Methods Appl. Sci., 17, No. 3: 411 (2007).
26. P. Galenko and S. Sobolev, Phys. Rev. E, 55: 343 (1997).
27. P. K. Galenko and D. A. Danilov, Phys. Lett. A, 235: 271 (1997).
28. L. D. Landau and E. M. Lifshitz, Fluid Mechanics (Addison Wesley: 1958).
29. D. Jou and J. M. Rubi, Phys. Lett. A, 72: 78 (1979).
30. D. Jou and C. Perez-Garsia, Physica A, 104: 3320 (1980).
31. D. Jou, J. E. Llebot, and J. Casas-Vazquez, Physica A, 109: 208 (1981).
32. H. Kleinert and S. V. Shabanov, Phys. Lett. A, 235: 105 (1997).
33. M. Ibanes, J. Garcia-Ojalvo, R. Toral, and J. M. Sancho, Lecture Notes in
Physics (Berlin: Springer-Verlag: 2000), vol. 557, p. 247.
34. J. M. Sancho, J. Garcia-Ojalvo, and H. Guo, Physica D, 113: 331 (1998).
35. C. Van der Broeck, Phys. Rev. Lett., 73: 3395 (1994).
36. C. Van der Broeck, J. M. R. Parrondo, R. Toral, and R. Kawai, Phys. Rev.
E, 55: 4084 (1997).
37. M. Ibanes, J. Garcia-Ojalvo, R. Toral, and J. M. Sancho, Phys. Rev. E, 60:
3597 (1999).
38. J. Buceta, M. Ibanes, J. M. Sancho, and K. Lindenberg, arXiv:cond-
mat/0211181.
39. P. K. Galenko and M. D. Krivilyov, Modelling Simul. Mater. Sci. Eng., 8: 67
(2000).
40. D. Jou and P. Galenko, Physica A, 399: 149 (2006).
41. P. Galenko, Phys. Rev. B, 65: 144103 (2002).
42. J. P. Boon and S. Yip, Molecular Hydrodynamics (New York: McGraw Hill:
DYNAMICS IN SPINODAL DECOMPOSITION OF A BINARY SYSTEM 101
1970).
43. H. B. Callen, Thermodynamics (New York: Wiley: 1960).
44. R. Luzzi, A. R. Vasconcellos, J. Casas-Vazquez, and D. Jou, Physica A, 248:
111 (1998).
45. P. Galenko and D. Jou, Proceedings of Joint European Thermodynamics
Conference IX (Eds. B. Guy and D. Tondeur) (Saint-Etienne: Ecole Nationale
Superieure des Mines: 2007), p. 75.
46. J. Casas-Vazquez and D. Jou, Rep. Prog. Phys., 66: 1937 (2003).
47. P. Galenko and V. Lebedev, Phil. Mag. Lett., 87: 821 (2007).
48. P. Galenko and V. Lebedev, Phys. Lett. A, 372: 985 (2008).
49. P. Galenko and D. Jou, Physica A, 388: 3113 (2009).
50. A. Miranville and S. Zelik, Math. Methods Appl. Sci., 28: 709 (2005).
51. C. Gal, Dyn. Partial Differ. Equ., 5: 39 (2008).
52. P. Galenko and V. Lebedev, The International Journal of Thermodynamics,
11: 21 (2008).
53. P. Galenko, Phys. Rev. B, 65: 144103 (2002).
54. M. Hillert, Acta Metall., 9: 525 (1961).
55. P. Galenko and V. Lebedev, Pis’ma Zh. Eksp. Teor. Phys., 86: 525 (1990)
(JETP Lett., 86: 458 (2007)).
56. N. S. Andreev, G. G. Boiko, and N. A. Bokov, J. Non-Cryst. Solids, 5: 41
(1970).
57. N. S. Andreev and E. A. Porai-Koshits, Discuss. Faraday Soc., 50: 135
(1970).
58. E. P. Feldman and L. I. Stefanovich, Zh. Eksp. Teor. Fiz., 98: 1695 (1990)
(Sov. Phys. JETP, 71: 951 (1990)).
59. H. E. Cook, J. Chem. Solids, 30: 2427 (1969); H. E. Cook, Acta. Metall., 18:
297 (1970).
60. W. Horsthemke and R. Lefever, Noise-Induced Transitions (Berlin:
Springer-Verlag: 1984).
61. H. Risken, The Fokker—Planck Equation (Berlin: Springer-Verlag: 1989).
62. C. W. Gradiner, Handbook of Stochastic Methods (Berlin: Springer-Verlag:
1985).
63. A. I. Olemskoi and D. O. Kharchenko, Physica A, 293: 178 (2001).
64. D. O. Kharchenko and S. V. Kohan, Eur. Phys. J. B, 29: 97 (2002).
65. K. Wiesenfeld and F. Moss, Nature, 373: 33 (1995).
66. L. Gammaitoni, P. Hänggi, P. Jung, and F. Marchesoni, Rev. Mod. Phys.,
70: 223 (1998).
67. J. Garcia-Ojalvo and J. M. Sancho, Noise in Spatially Extended Systems
(New York: Springer-Verlag: 1999).
68. D. O. Kharchenko, Physica A, 308: 113 (2002).
69. D. O. Kharchenko and I. A. Knyaz’, Eur. Phys. J. B, 32: 375 (2003).
70. A. I. Olemskoi, D. O. Kharchenko, and I. A. Knyaz’, Phys. Rev. E, 71:
041101 (2005).
71. J. M. R. Parrondo, C. Van der Broeck, J. Buceta, and F. J. de la Rubia,
Physica A, 224: 153 (1996).
72. F. Sagués, J. M. Sancho, and J. Garcia-Ojalvo, Rev. Mod. Phys., 79: 829
(2007).
73. B. Linder, J. Garcia-Ojalvo, A. Neiman, and L. Shimansky-Geier, Phys. Rep.,
392: 321 (2004).
102 D. O. KHARCHENKO, P. K. GALENKO, and V. G. LEBEDEV
74. Y. Tu, G. Grinstein, and M. A. Munoz, Phys. Rev. Lett., 78: 224 (1997).
75. W. Genovese, M. A. Munoz, and J. M. Sancho, Phys. Rev. E, 57: R2495
(1998).
76. M. Ibanes, J. Garcia-Ojalvo, R. Toral, and J. M. Sancho, Phys. Rev. Lett.,
87: 020601 (2001).
77. B. von Haeften, G. Izús, S. Mangioni, A. D. Sánchez, and H. S. Wio, Phys.
Rev. E, 69: 021107 (2004).
78. D. O. Kharchenko, Usp. Fiz. Met., 6: 1 (2005).
79. E. A. Novikov, Sov. Phys. JETP, 20: 1290 (1965).
80. J. Buceta and K. Lindenberg, Phys. Rev. E, 69: 011102 (2004).
81. D. O. Kharchenko and A. V. Dvornichenko, Eur. Phys. J. B, 61: 95 (2008).
82. D. O. Kharchenko and A. V. Dvornichenko, Physica A, 387: 5342 (2008).
83. S. van Gemmert, G. T. Barkema, and S. Puri, arXiv:cond-mat/0501196.
84. J. Garcia-Ojalvo, A. M. Lacasta, J. M. Sancho, and R. Toral, Europhys.
Lett., 42: 125 (1998).
85. L. Ramirez-Piscina, J. M. Sancho, and A. Hernandez-Machado, Phys. Rev. B,
48: 125 (1993).
86. I. M. Lifshitz and V. V. Slyozov, J. Phys. Chem. Solids, 19: 35 (1961).
87. C. L. Emmott and A. J. Bray, Phys. Rev. E, 59: 213 (1999).
88. J. M. Sancho, A. Hernandez-Machado, L. Ramirez-Piscina, and A. M. La-
casta, Acta Phys. Polonica B, 24: 733 (1993).
89. J. D. Gunton, M. San Miguel, and P. S. Sahni, Phase Transtions and Criti-
cal Phenomena (New York: Academic Press: 1983), vol. 8.
90. D. O. Kharchenko, A. V. Dvornichenko, and I. A. Lysenko, Ukr. Jour. of
Phys., 53, No. 9: 917 (2008).
91. S. De Groot and P. Mazur, Non-Equilibrium Thermodynamics (Amsterdam:
1962).
92. N. G. Van Kampen, Stochastic Processes in Physics and Chemistry (Amster-
dam: 1992).
|