Evolution of pattern complexity in the Cahn–Hilliard theory of ... Acta Met...The first model for...

12
Evolution of pattern complexity in the Cahn–Hilliard theory of phase separation Marcio Gameiro a , Konstantin Mischaikow b , Thomas Wanner c, * a School of Mathematics, Georgia Institute of Technology, Atlanta, GA 30332, USA b Center for Dynamical Systems and Nonlinear Studies, Georgia Institute of Technology, Atlanta, GA 30332, USA c Department of Mathematical Sciences, George Mason University, 4400 University Drive, MS 3F2, Fairfax, VA 22030, USA Received 21 September 2004; accepted 14 October 2004 Available online 11 November 2004 Abstract Phase separation processes in compound materials can produce intriguing and complicated patterns. Yet, characterizing the geometry of these patterns quantitatively can be quite challenging. In this paper we propose the use of computational algebraic topology to obtain such a characterization. Our method is illustrated for the complex microstructures observed during spinodal decomposition and early coarsening in both the deterministic Cahn–Hilliard theory, as well as in the stochastic Cahn–Hilliard–Cook model. While both models produce microstructures that are qualitatively similar to the ones observed experimentally, our topolog- ical characterization points to significant differences. One particular aspect of our method is its ability to quantify boundary effects in finite size systems. Ó 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. Keywords: Dynamic phenomena; Microstructure; Phase field models; Spinodal decomposition; Coarsening 1. Introduction The kinetics of phase separation in alloys has drawn considerable interest in recent years. Most alloys of commercial interest owe their properties to specific microstructures which are generated through special processing techniques, such as phase separation mecha- nisms. In quenched binary alloys, for example, one typ- ically observes phase separation due to a nucleation and growth process, or alternatively, due to spinodal decom- position [4,7,8]. While the former process involves a thermally activated nucleation step, spinodal decompo- sition can be observed if the alloy is quenched into the unstable region of the phase diagram. The resulting inherent instability leads to composition fluctuations, and thus to instantaneous phase separation. Common to both mechanisms is the fact that the generated micro- structures usually are thermodynamically unstable and will change in the course of time – thereby affecting the material properties. In order to describe these and other processes, one generally relies on models given by nonlinear evolution equations. Many of these models are phenomenological in nature, and it is therefore of fundamental interest to study how well the equations agree with experimental observations. In this paper, we propose a comparison method based on the geometric properties of the microstruc- tures. The method will be illustrated for the microstruc- tures generated during spinodal decomposition. These structures are fine-grained and snake-like, as shown for example in Fig. 1. The microstructures are computed using two different evolution equations which have been proposed as models for spinodal decomposition: The 1359-6454/$30.00 Ó 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actamat.2004.10.022 * Corresponding author. Tel.: +1 703 993 1472; fax: +1 703 993 1491. E-mail address: [email protected] (T. Wanner). Acta Materialia 53 (2005) 693–704 www.actamat-journals.com

Transcript of Evolution of pattern complexity in the Cahn–Hilliard theory of ... Acta Met...The first model for...

  • Acta Materialia 53 (2005) 693–704

    www.actamat-journals.com

    Evolution of pattern complexity in the Cahn–Hilliard theory ofphase separation

    Marcio Gameiro a, Konstantin Mischaikow b, Thomas Wanner c,*

    a School of Mathematics, Georgia Institute of Technology, Atlanta, GA 30332, USAb Center for Dynamical Systems and Nonlinear Studies, Georgia Institute of Technology, Atlanta, GA 30332, USA

    c Department of Mathematical Sciences, George Mason University, 4400 University Drive, MS 3F2, Fairfax, VA 22030, USA

    Received 21 September 2004; accepted 14 October 2004

    Available online 11 November 2004

    Abstract

    Phase separation processes in compound materials can produce intriguing and complicated patterns. Yet, characterizing the

    geometry of these patterns quantitatively can be quite challenging. In this paper we propose the use of computational algebraic

    topology to obtain such a characterization. Our method is illustrated for the complex microstructures observed during spinodal

    decomposition and early coarsening in both the deterministic Cahn–Hilliard theory, as well as in the stochastic Cahn–Hilliard–Cook

    model. While both models produce microstructures that are qualitatively similar to the ones observed experimentally, our topolog-

    ical characterization points to significant differences. One particular aspect of our method is its ability to quantify boundary effects in

    finite size systems.

    � 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

    Keywords: Dynamic phenomena; Microstructure; Phase field models; Spinodal decomposition; Coarsening

    1. Introduction

    The kinetics of phase separation in alloys has drawn

    considerable interest in recent years. Most alloys of

    commercial interest owe their properties to specific

    microstructures which are generated through special

    processing techniques, such as phase separation mecha-

    nisms. In quenched binary alloys, for example, one typ-

    ically observes phase separation due to a nucleation andgrowth process, or alternatively, due to spinodal decom-

    position [4,7,8]. While the former process involves a

    thermally activated nucleation step, spinodal decompo-

    sition can be observed if the alloy is quenched into the

    unstable region of the phase diagram. The resulting

    1359-6454/$30.00 � 2004 Acta Materialia Inc. Published by Elsevier Ltd. Adoi:10.1016/j.actamat.2004.10.022

    * Corresponding author. Tel.: +1 703 993 1472; fax: +1 703 993

    1491.

    E-mail address: [email protected] (T. Wanner).

    inherent instability leads to composition fluctuations,and thus to instantaneous phase separation. Common

    to both mechanisms is the fact that the generated micro-

    structures usually are thermodynamically unstable and

    will change in the course of time – thereby affecting

    the material properties. In order to describe these and

    other processes, one generally relies on models given

    by nonlinear evolution equations. Many of these models

    are phenomenological in nature, and it is therefore offundamental interest to study how well the equations

    agree with experimental observations.

    In this paper, we propose a comparison method

    based on the geometric properties of the microstruc-

    tures. The method will be illustrated for the microstruc-

    tures generated during spinodal decomposition. These

    structures are fine-grained and snake-like, as shown

    for example in Fig. 1. The microstructures are computedusing two different evolution equations which have been

    proposed as models for spinodal decomposition: The

    ll rights reserved.

    mailto:[email protected]

  • Fig. 1. Microstructures obtained from the Cahn–Hilliard–Cook model (4). The top row shows solution snapshots for the deterministic case r = 0,the bottom row is for noise intensity r = 0.01. In both cases we used c1/2 = 0.005 and m = 0. The set X+(t) defined in (5) is shown in dark blue. (Forinterpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

    694 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    seminal Cahn–Hilliard model, as well as its stochastic

    extension due to Cook.

    The first model for spinodal decomposition in binary

    alloys is due to Cahn and Hilliard [4,7]; see also the sur-veys [6,16]. Their mean field approach leads to a nonlin-

    ear evolution equation for the relative concentration

    difference u = .A � .B, where .A and .B denote the rel-ative concentrations of the two components, i.e.,

    .A + .B = 1. The Ginzburg–Landau free energy is givenby

    EcðuÞ ¼ZX

    WðuÞ þ c2jruj2

    � �dx; ð1Þ

    where X is a bounded domain, and the positive param-eter c is related to the root mean square effective interac-tion distance. The bulk free energy, W, is a double wellpotential, typically

    WðuÞ ¼ 14u2 � 1� �2

    : ð2Þ

    Taking the variational derivative dEc/du of the Ginz-burg–Landau free energy (1) with respect to the concen-tration variable u, we obtain the chemical potential

    l ¼ �cDuþ oWou

    ðuÞ;

    and thus the Cahn–Hilliard equation ou/ot = Dl, i.e.,

    ouot

    ¼ �D cDu� oWou

    ðuÞ� �

    ; ð3Þ

    subject to no-flux boundary conditions for both l and u.Due to these boundary conditions, any mass flux

    through the boundary is prohibited, and therefore mass

    is conserved. We generally consider initial conditions for

    (3) which are small-amplitude random perturbations of

    a spatially homogeneous state, i.e., u(0,x) = m + e(x),where e denotes a small-amplitude perturbation with to-tal mass �Xe = 0. In order to observe spinodal decompo-sition, the initial mass m has to satisfy (o2W/ou2)(m) < 0.

    One drawback of the deterministic partial differential

    equation (3) is that it completely ignores thermal fluctu-

    ations. To remedy this, Cook [10] extended the model by

    adding a random fluctuation term n, i.e., he consideredthe stochastic Cahn–Hilliard–Cook model

    ouot

    ¼ �D cDu� oWou

    ðuÞ� �

    þ r � n; ð4Þ

    where

    hnðt; xÞi ¼ 0; hnðt1; x1Þnðt2; x2Þi ¼ dðt1 � t2Þqðx1 � x2Þ:

    Here, r > 0 is a measure for the intensity of the fluctua-tion and q describes the spatial correlation of the noise.

    In other words, the noise is uncorrelated in time. For the

    special case q = d we obtain space-time white noise, butin general the noise will exhibit spatial correlations. See

    also [23].

    Both the deterministic and the stochastic model pro-

    duce patterns which are qualitatively similar to themicrostructures observed during spinodal decomposi-

    tion [5]. Recent mathematical results for (3) and (4) have

    identified the observed microstructures as certain ran-

    dom superpositions of eigenfunctions of the Laplacian,

    and were able to explain the dynamics of the decompo-

    sition process in more detail [3,27,28,33,34,38].

    Even though both models are based on deep physi-

    cal insight, they are phenomenological models. Many

  • M. Gameiro et al. / Acta Materialia 53 (2005) 693–704 695

    researchers have therefore studied how well the models

    agree with experimental observations. See for example

    [2,13,17,30], as well as the references therein. Most of

    these studies have been restricted to testing scaling

    laws. Exceptions include for example Ujihara and

    Osamura [37], who perform a quantitative evaluationof the kinetics by analyzing the temporal evolution of

    the scattering intensity of small angle neuron scattering

    in Fe–Cr alloys. Another approch was pursued by

    Hyde et al. [18], who consider spinodal decomposition

    in Fe–Cr alloys. Using experimental data obtained by

    an atom probe field ion microscopy PoSAP analysis,

    they study whether the observed microstructures are

    topologically equivalent to the structures generated byseveral model equations.

    Two structures are topologically equivalent, if one

    can be deformed into the other without cutting or glu-

    ing. While it is difficult in general to determine when

    two structures are equivalent, one can easily obtain a

    measure for the degree of similarity by studying topo-

    logical invariants. These are objects (such as numbers,

    or other algebraic objects) assigned to each structure,which remain unchanged under deformations. In other

    words, if these invariants differ for two structures, the

    structures cannot be topologically equivalent. Yet, sim-

    ilarities between the invariants indicate similarities be-

    tween the considered structures.

    The topological invariant used in [18] is the number

    of handles in the microstructure occupied by one of

    the two material phases, which is introduced as a char-acteristic measure for the topological complexity of per-

    colated structures. Since moving a dislocation through a

    sponge-like interconnected microstructure generally re-

    quires cutting through the handles of the structure, the

    handle density can be viewed as an analogue of the par-

    ticle density for systems containing isolated particles.

    From a theoretical point of view, the number of handles

    in a microstructure is a special case of topological invar-iants called Betti numbers. In this paper, we demon-

    strate how the information contained in the Betti

    numbers can be used to further quantify the geometry

    of complex microstructures. We rely on recent progress

    in computational homology which makes efficient com-

    putation of the Betti numbers possible. Specifically, we

    use the recently released public-domain software pack-

    age CHomP [20,21].Our results are illustrated for the morphologies gen-

    erated by the deterministic Cahn–Hilliard model (3)

    and the stochastic model (4). We show that knowledge

    of the Betti numbers can detect significant differences

    in the dynamic behavior of these two models. By com-

    bining the information provided by different Betti num-

    bers, we are able to quantify boundary effects in finite

    size systems. In addition, our method can be used to de-tect fundamental changes in the microstructure topology

    due to parameter variation, such as the transition from

    percolated structures to isolated particles. Many struc-

    ture–property relationships in materials science are

    based on fundamental assumptions concerning the

    underlying microstructure, and our topological analysis

    can aid in unveiling the underlying topology. To sim-

    plify the presentation, we only consider two-dimensionalmicrostructures generated by spinodal decomposition.

    Nevertheless, the presented software works in arbitrary

    dimensions. Aside from the study of 3D microstruc-

    tures, this also allows for the analysis of spatio-temporal

    structure changes, by including time as a fourth dimen-

    sion [12].

    2. Homology and Betti numbers

    As is indicated in the introduction we claim that

    homology provides a useful technique for identifying

    and distinguishing the evolving microstructures of (3)

    and (4). While this paper is not the appropriate forum

    for an in-depth description of algebraic topology and

    homology, four points do need to be discussed: (i) Whatstructures do we want to understand the geometry of,

    (ii) what geometric features does homology measure,

    (iii) how is the homology being computed, and (iv) what

    additional information does homology provide when

    compared to existing methods of microstructure

    analysis?

    Since we are interested in phase separation and u(t,x)

    represents the relative concentration difference betweenthe two materials at time t and location x, the simplest

    decomposition of the domain X consists of the two sets

    XþðtÞ :¼ x 2 Xjuðt; xÞ > mf g andX�ðtÞ :¼ x 2 Xjuðt; xÞ < mf g; ð5Þ

    where m denotes a suitable threshold value, such as thetotal mass. In this situation, the sets X±(t) represent the

    regions in the material where one or the other element

    dominates. Of course, u is the solution of a (stochastic)

    nonlinear partial differential equation and hence we can-

    not expect to have an explicit representation of the sets

    X±. For the simulations in this paper, we use a finite dif-

    ference scheme to numerically solve (3) and (4) and thus

    given the domain X = (0,1) · (0,1) with a spatial gridsize N · N we obtain values Uðtk; x1‘ ; x2nÞ where xj =1/(2N) + (j � 1)/N. We give a geometric interpretationto this numerical grid via the squares Q‘, n :¼ [(‘ � 1)/N, ‘/N] · [(n � 1)/N,n/N]. In particular we define a cubi-cal approximation to X±(t) by

    UþðtkÞ :¼ Q‘; njUðtk; x1‘ ; x2nÞ > m� �

    and

    UþðtkÞ :¼ Q‘; njUðtk; x1‘ ; x2nÞ < m� �

    :

    Fig. 1 indicates sets U±(tk) at various times steps for a

    particular solution to (3) and (4). It is easy to observe

    that patterns produced by U±(tk) are complicated, time

  • 696 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    dependent, and appear at intermediate time steps to

    differ qualitatively for the deterministic and stochastic

    models. Before moving on to describe how homology

    can be used to quantify these observations, notice that

    if the computations had been performed in a three-

    dimensional domain, then the same approach could beused except that the two-dimensional squares Q‘, nwould be replaced by three-dimensional cubes.

    This leads us to the question of what geometric prop-

    erties homology can measure. We begin with the fact

    that for any topological space X there exist homology

    groups Hi(X), i = 0,1,2, . . . (see [20] and the referencestherein). For a general space Hi(X) takes the form of

    an arbitrary abelian group. However, in the context ofour investigations there are two simplifying conditions.

    The first is that from the point of view of materials sci-

    ence we are only interested in structures that can occur

    in three-dimensional Euclidean space. The second is that

    the spaces we compute the homology of are represented

    in terms of a finite number of cubes (squares if we re-

    strict our attention to a two-dimensional model as is

    being done in this paper). In this setting the homologygroups are much simpler; first, Hi(X) = 0 for i P 3 andsecond, for i = 0,1,2,

    HiðX Þ ffi Zbi ;where Z denotes the integers and bi is a non-negativeinteger called the ith Betti number. Thus, for our pur-

    poses the homology groups are characterized by theirBetti numbers.

    Each of the homology groups measures a different

    geometric property. H0(X) counts the number of con-

    nected components (pieces) of the space X. More pre-

    cisely, if b0 = k, then X has exactly k components. Aspecific example of this is shown in the left diagram in

    Fig. 2, where b0 = 26 corresponding to the 26 differentconnected components. Observe that the size or the

    Fig. 2. Betti numbers for the darker region of the r = 0 and t = 0.0036 snadifferent colors, the right diagram indicates the location of the b1 = 4 loops.reader is referred to the web version of this article.)

    shape of the components does not play a role in the va-

    lue of b0. Depending on the application this can be ta-ken as a strength or weakness of this approach. For

    the problems being discussed in this paper we see it as

    a strength, since we have no a priori knowledge of the

    specific geometry of the material microstructures.H1(X), or more precisely b1, provides a measure of

    the number of tunnels in the structure, though the cor-

    respondence is slightly more complicated. In a two-di-

    mensional domain, such as that indicated in Fig. 2,

    tunnels are reduced to loops. In particular, as is shown

    in the right diagram of Fig. 2, the green structure en-

    closes exactly four regions indicated by the black curves

    and hence b1 = 4. Notice that each of the remainingwhite regions hits the boundary. As before, size or shape

    plays no role in the definition of a loop.

    For a three-dimensional structure loops become tun-

    nels, where again the length or width of the tunnel is irrel-

    evant. For example a washer has one wide but very short

    tunnel while a garden hose has a narrow but long tunnel.

    In either case b1 = 1. As mentioned earlier we are onlyconsidering a two-dimensional domain in which caseH2(X) = 0. However, for a general three-dimensional

    domain b2 equals the number of enclosed volumes orcavities within the material.

    From a computational point of view, determining the

    Betti numbers of a complex 3D structure is far from triv-

    ial. In fact, it was pointed out explicitly in [18] that ‘‘de-

    signing a computer program to directly count the handle

    density in a complex structure would be extremely diffi-cult [18, p. 3419]’’. The authors therefore compute the

    number of handles indirectly, using techniques from dig-

    ital topology [22]. These techniques relate the number of

    handles to the Euler characteristic of the structure,

    which can be computed easily in three space dimensions.

    Our study is made possible by recent progress in compu-

    tational homology, which does allow for the direct com-

    pshot from Fig. 1. The left diagram shows the b0 = 26 components in(For interpretation of the references to color in this figure legend, the

  • M. Gameiro et al. / Acta Materialia 53 (2005) 693–704 697

    putation of the Betti numbers. As we already indicated

    in the introduction, we use the software package

    CHomP [21]. An elementary introduction to homology

    and the underlying algorithms can be found in [20]. Typ-

    ical computation times of the software in our situation

    are described in the following section.To close this section, we would like to place our

    methodology in the proper context and comment on

    its merits. Characterizing material properties is an old

    subject, and numerous tools have been devised over

    the course of time. Many of these tools use spatial aver-

    aging, such as the structure factor or point correlation

    functions. While these quantities do provide informa-

    tion on predominant wavelengths and length scales,the averaging process used in their definition removes

    much of the local connectivity information of the micro-

    structure. This fact is illustrated in [36], where several

    different correlation functions are used to reconstruct

    microstructures.

    In order to obtain quantitative connectivity informa-

    tion on microstructures, several authors have therefore

    employed topological methods. See for example[1,9,15,18,19,29], as well as the references therein.

    Due to the computational issues mentioned above,

    the easily computable Euler characteristic takes a pre-

    dominant role in these studies. Yet, the Euler charac-

    teristic, which is the alternating sum of the Betti

    numbers, provides considerably less topological infor-

    mation than the complete set of Betti numbers. This

    is true even in the two-dimensional case considered inthis paper. In Section 5 it will be shown that the Euler

    characteristic is closely related to boundary effects,

    whereas the Betti numbers also contain information

    on the bulk structure.

    3. Effects of noise on the pattern morphology

    In this section we use computational homology to

    study the effects of the stochastic forcing term in (4)

    on the temporal evolution of the pattern complexity.

    We consider the case of total mass m = 0 and W as in(2), and to begin with restrict ourselves to c1/2 =0.005. The simulations are performed on the unit

    square X = (0,1) · (0,1). The deterministic part of theevolution equation (4) is approximated using a finitedifference scheme with linearly implicit time-stepping,

    similar to schemes described in [14,31], for the stochas-

    tic noise term we follow [35]. This numerical scheme is

    implemented efficiently using the fast Fourier trans-

    form. For the simulations in this paper we used an

    implementation in C in combination with the fast Fou-

    rier transform package FFTW [11]. As for the noise

    term, we consider the case of cut-off noise which guar-antees a spatial correlation function q which closely

    approximates d.

    For varying values of the noise intensity r, we numer-ically integrate the Cahn–Hilliard–Cook equation (4)

    starting from a random perturbation of the homogene-

    ous state m = 0 with amplitude 0.0001 up to time

    tend ¼160c

    W00ðmÞ2; where W00ðmÞ ¼ 3m2 � 1; ð6Þ

    which in our situation reduces to tend = 0.004. This time

    frame covers the complete spinodal decompositionphase, as well as early coarsening. The domain is discre-

    tized by a 512 · 512-grid, the time interval [0, tend] is cov-ered by 10,000 integration steps. Every 50 time steps, the

    sets X±(t) introduced in the last section are determined,

    and finally their Betti numbers are computed using [21].

    On a 2 GHz Dual Xeon Linux PC it took about 15 min

    to create the 402 sets X±(t) for t/tend = 0, 0.005,

    0.010, . . ., 0.995, 1, as well as a total of 46 min to com-pute their Betti Numbers.

    Fig. 3 contains the results of these computations for

    two different values of the noise intensity. The solid

    red curves show the Betti number evolution for the

    deterministic case r = 0, the dashed blue curves are fornoise intensity r = 0.01; the solution snapshots in Fig. 1are taken from these two simulations. At first glance,

    the graphs in Fig. 3 are not surprising. For the initialtime t = 0 both b0 and b1 are large, since the initial per-turbation was chosen randomly on the computational

    grid. In fact, the actual values lie well outside the dis-

    played range. The smoothing effect of (4) leads to a ra-

    pid decrease of the Betti numbers for times t close to 0,

    and coarsening behavior in the Cahn–Hilliard–Cook

    model is responsible for the decrease observed towards

    the end of the time window. All shown evolutions exhi-bit small-scale fluctuations, which are most likely arti-

    facts caused by the cubical approximation of the sets

    X±(t). Even though these fluctuations are not necessarily

    desirable, they are a more or less automatic consequence

    of the numerical approximation of the partial differen-

    tial equation (4).

    Despite these similarities, there are some obvious dif-

    ferences between the deterministic and the stochasticevolution. In the deterministic situation, the initial com-

    plexity decay occurs sooner than in the stochastic case.

    In addition, it appears that the Betti numbers of X+(t)

    for r = 0.01 decay more or less monotonically, providedwe ignore the above-mentioned small-scale fluctuations.

    In contrast, in the deterministic case the initial decrease

    seems to be followed by a period of stagnation or even

    growth of the Betti numbers, see for example b0 forX+(t) or b1 for X

    �(t).

    The observations made in the last paragraph could

    indicate fundamental differences between the determinis-

    tic and the stochastic Cahn–Hilliard model. However,

    these observations are based on a single randomly cho-

    sen initial condition, and it is therefore far from clear

  • Fig. 3. Evolution of the Betti numbers for the solutions of Fig. 1. In each diagram, the solid red curve corresponds to r = 0, the dashed blue curve isfor r = 0.01. In both cases we used c1/2 = 0.005 and m = 0. The diagrams in the left column show the results for X+(t), the right column is for X�(t).The top row contains the evolutions of b0, the bottom row shows b1. (For interpretation of the references to color in this figure legend, the reader isreferred to the web version of this article.)

    698 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    whether they represent behavior typical for either of the

    models. For this we have to observe ensembles of solu-

    tions for each of the models and study the statistics oftheir complexity evolutions. For the purposes of this pa-

    per, we concentrate on the averaged complexity evolu-

    tion. Fig. 4 shows the averaged Betti number evolution

    Fig. 4. Evolution of Betti number averages for a sample size of 100, with ccorrespond to the values r = 0.1, 0.03, 0.01, 0.003, 0.001, and 0, from left to

    for six values of the noise intensity r ranging from 0to 0.1, in each case based on solution ensembles of size

    100. The qualitative form of the complexity evolutioncurves differs substantially. For large noise the complex-

    ity decays monotonically, while it shows a surprising in-

    crease in the deterministic situation. The change

    1/2 = 0.005 and m = 0. In the top row at dimension level 60 the curves

    right; analogously in the bottom row at level 25.

  • M. Gameiro et al. / Acta Materialia 53 (2005) 693–704 699

    between the two behaviors occurs gradually, and there

    seems to be a specific threshold rcr for the noise intensitybeyond which monotone decay is observed. Notice also

    that despite the small ensemble size, the evolution curves

    of X+(t) and X�(t) are in good agreement, which for our

    choice of m = 0 and W as in (2) has to be expected. Inthis sense, the ensemble behavior is a reflection of typical

    solution dynamics, reinforcing our above observations.

    The simulations discussed so far consider the specific

    value c1/2 = 0.005. We performed analogous simulationsfor various values of c, for various grid sizes, and a vari-ety of time steps. In each case, the time window ex-

    tended from 0 to tend as defined in (6). These results

    show that the behavior shown in Fig. 4 is typical. Inall cases, the time window covers the complete spinodal

    decomposition phase, as well as early coarsening; the

    resulting evolution curves for the deterministic case

    show the characteristic non-monotone behavior; for suf-

    ficiently large noise intensity we observe monotone de-

    cay. The only parameters changing with c appear to bethe absolute height of the evolution curves and (possi-

    bly) the critical noise intensity rcr. While these scalingswill be addressed in more detail in Section 5, we close

    this section with a brief discussion of the dependence

    of rcr on c.To get a more accurate picture of the monotonicity

    properties of the evolution curves we compute the aver-

    aged Betti number evolution from larger solution

    ensembles and for times up to only tend/4. As before, tendis given by (6), and we consider the case m = 0. The noiseintensity r is chosen equal to c1/2, and the ensemble sizeis taken as 1000. The results of these simulations for

    c1/2 = 0.005, 0.006, 0.007, and 0.01 are shown in Fig. 5.Notice that each of these curves exhibits monotone de-

    cay of the Betti numbers with a pronounced plateau,

    so one would expect that these curves are for values

    r � rcr. Yet, it appears that the length of the plateau de-creases with increasing c. This latter observation could

    0 0.05 0.1 0.15 0.2 0.250

    10

    20

    30

    40

    50

    60

    70

    80

    t/tend

    1

    1

    2

    2

    3

    3

    4

    4

    b0

    Fig. 5. Betti number averages for X+(t) with varying c1/2 = r and m = 0. Fromand 0.01, respectively. The sample size for each of these simulations is 1000

    indicate that for the larger c-values the noise intensitiesused in Fig. 5 are too large. In fact, our scaling analysis

    in Section 5 will show that this is the case.

    The results of this section demonstrate that Betti

    numbers can be used to quantitatively distinguish be-

    tween microstructure morphologies generated by differ-ent models. Ultimately we hope that this quantitative

    information can be used to match models to actual

    experimental data. In fact, the limited experimental data

    in [18] seems to indicate that the handle density of the

    microstructures generated through spinodal decomposi-

    tion decreases monotonically. Thus, these experimental

    results favor the stochastic Cahn–Hilliard–Cook model

    with sufficiently large noise intensity.

    4. Morphology changes due to mass variation

    So far we restricted our study to the case of equal

    mass, i.e., we assumed m = 0. Yet, spinodal decomposi-

    tion can be observed as long asW00(m) < 0, which for ourchoice of W is equivalent to |m| < 3�1/2 � 0.577. Totalmass values outside this range lead to nucleation and

    growth behavior which produces microstructures con-

    sisting of isolated droplets – in contrast to the micro-

    structures shown in Fig. 1. In this section we use

    computational homology to quantify the pattern mor-

    phology changes during spinodal decomposition in the

    deterministic Cahn–Hilliard model (3) as the total mass

    is increased from m = 0 towards m � 3�1/2.Before presenting our numerical results, we have to

    address one technical issue. Even though the homogene-

    ous state m is unstable as long asW00(m) < 0, the strengthof the instability changes with m. One can easily show

    that the growth rate of the most unstable perturbation of

    the homogeneous state is close to W00(m)2/(4c) [27]. Thus,as the total mass approaches the boundary of the spi-

    nodal region, the time frame for spinodal decomposition

    0 0.05 0.1 0.15 0.2 0.250

    5

    0

    5

    0

    5

    0

    5

    0

    5

    t/tend

    b1

    top to bottom the curves correspond to c1/2 = r = 0.005, 0.006, 0.007,.

  • 700 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    grows. In order to compare the pattern morphology for

    different values of the total mass we therefore scale the

    considered time window. As in the last section, we com-

    pute the solutions up to time tend defined in (6), whose

    scaling is motivated by largest growth rate mentioned

    above.How do changes in the total mass m affect the micro-

    structures generated through spinodal decomposition?

    Fig. 6 contains typical patterns for m = 0, 0.1, . . ., 0.5.In each case, the pattern was generated by solving (3)

    up to time t = 0.4 Æ tend, with tend as in (6). Notice thateven though all of these microstructures are a conse-

    quence of phase separation through spinodal decompo-

    sition, the last two microstructures resemble the onesgenerated by nucleation and growth. In other words,

    Fig. 6 indicates a gradual change from the highly inter-

    connected structures observed in the equal mass case to

    the disconnected structures observed in nucleation.

    In order to further quantify this gradual change, we

    use computational homology as in the last section. For

    c1/2 = 0.005 and various values of m between 0 and0.55 we computed the averaged Betti number evolutionof the sets X±(t) in (5) from t = 0 up to time t = tend. The

    resulting two-dimensional surfaces are shown in Fig. 7.

    To avoid the large Betti numbers close to the random

    initial state, these graphs do not include the times t = 0

    and t = 0.005 Æ tend. Of particular interest is the topologyof the sets X+(t) which correspond to the dominant

    phase. The graphs in the left column of Fig. 7 show that

    in terms of the quantitative topological information gi-ven by the averaged Betti numbers, one observes a grad-

    ual and continuous change from the interconnected

    Fig. 6. Microstructures obtained from the Cahn–Hilliard model (3) for c1/2 =are shown for total mass m = 0, 0.1, . . ., 0.5 at time t = 0.4 Æ tend, i.e., shortlyshown in dark blue. (For interpretation of the references to color in this fig

    microstructures for m = 0 to the nucleation morphology.

    In addition, the graph of b0 for X+(t) indicates that for

    t/tend > 0.25, i.e., after the completion of spinodal decom-

    position, and for m > 0.2 the dominant phase forms a

    connected structure; that is, the complementary phase

    breaks up completely. Yet, the typical droplet structurereminiscent of nucleation and growth behavior can only

    be observed for values of the total mass larger than 0.3.

    In the surfaces of Fig. 7 this is reflected by the fact that

    for fixed time t, the Betti number b1 of the dominantphase X+(t) continues to increase with m until m � 0.3.As the total mass approaches 3�1/2, the first Betti num-

    ber then decreases again, corresponding to an increasing

    droplet size.In addition to the results shown in Fig. 7, we per-

    formed analogous simulations for c1/2 = 0.0025, 0.0075,and 0.0100. In each case, the qualitative shape of the

    computed surfaces matched the one for c1/2 = 0.005,only the absolute height of the surfaces changed. In fact,

    our computations indicate that the units on the vertical

    axes of Fig. 7 scale proportionally to c�1 for these valuesof c.

    The computations of this section indicate how

    homology can be used to quantify global morphological

    changes, such as connectivity, due to mass variation. On

    a more general level, given a model of a particular mate-

    rial it would be interesting to seek correlations between

    macroscopic properties and the type of geometric infor-

    mation expressed in Fig. 7. Such studies would be in the

    spirit of structure–property relationships which havebeen obtained for certain types of microstructures, such

    as for systems of isolated particles using particle distri-

    0.005 for varying m. From top left to bottom right solution snapshots

    after completion of the spinodal decomposition phase. The set X+(t) is

    ure legend, the reader is referred to the web version of this article.)

  • Fig. 7. Evolution of Betti number averages for a sample size of 100, with c1/2 = 0.005 and varying mass m. The left column corresponds to X+(t), theright one to X�(t); the top row contains the graphs for the 0th Betti number, the bottom row for the first.

    M. Gameiro et al. / Acta Materialia 53 (2005) 693–704 701

    butions, or for interconnected structures [24,25] using

    the notion of matricity. Comparisons of this nature

    would probably be of even greater importance in the

    case of three-dimensional simulations, since visualiza-

    tion of these phenomena would be extremely

    complicated.

    5. Boundary effects and scalings

    While our study so far concentrated on the Betti

    numbers b0 and b1, significant additional informationcan be obtained by combining the topological informa-

    tion for X+(t) with the information for the complemen-

    tary sets X�(t). To illustrate this, consider again the

    microstructure in the left diagram of Fig. 2 consisting

    of 26 components. Most of these components are intro-

    duced through boundary effects – only 6 components areinternal. Even though the Betti number b0 of X

    +(t) does

    not distinguish between these two types of components,

    only internal components of X+(t) introduce loops in

    X�(t). In fact, if we denote the number of internal com-

    ponents of X+(t) by bint, 0(X+(t)) and the number of com-

    ponents touching the boundary of X by bbdy, 0(X+(t)),

    then we have

    bint; 0ðXþðtÞÞ ¼ b1ðX�ðtÞÞ andbbdy; 0ðXþðtÞÞ ¼ b0ðXþðtÞÞ � b1ðX�ðtÞÞ:

    For c1/2 = 0.005 and m = r = 0 the averaged temporalevolution of these quantities is shown in the left diagram

    of Fig. 8, for a sample size of 100. While the number of

    internal components shows the typical non-monotone

    behavior described in the last sections, the number of

    boundary components remains basically constant, wellinto the coarsening regime. Notice also that due to

    m = 0 and our choice of W, the ensemble averages ofb1(X

    +(t)) and b1(X�(t)) have to be equal. Thus, the num-

    ber bbdy, 0(X+(t)) of boundary components equals in fact

    the Euler characteristic of the set X+(t). Similarly, the

    right diagram of Fig. 8 shows the evolutions of bint, 0and bbdy, 0 for the corresponding stochastic case withr = 0.01. While the number of internal componentsexhibits the monotone decay described in Section 3,

    the number of boundary components reaches the same

    level as in the deterministic situation, but starts its decay

    earlier.

    In view of Fig. 8, one can rightfully question the con-

    clusions obtained so far. During the spinodal decompo-

    sition regime, bint, 0 and bbdy, 0 are comparable, furtherinto the coarsening regime the number of boundarycomponents clearly dominates. In order to validate

    our results of the last two sections, it is therefore neces-

    sary to consider various system sizes. If instead of the

    unit square we consider the domain X = (0, ‘) · (0, ‘),then a combination of the spatial rescaling x ´ x/‘and the temporal rescaling t´ t‘2 transforms both (3)

  • 0 0.2 0.4 0.6 0.8 10

    10

    20

    30

    40

    50

    t/tend

    0 0.2 0.4 0.6 0.8 10

    10

    20

    30

    40

    50

    t/tend

    Fig. 8. Averaged temporal evolution of the number bint, 0 of internal components of the set X+(t), and of the number bbdy, 0 of components of X

    +(t)

    touching the boundary of the base domain X. In each diagram, the solid red curve corresponds to bint, 0, the dashed blue curve is for bbdy, 0. The leftdiagram is for the deterministic case r = 0, the right one for the stochastic case with noise intensity r = 0.01. In both cases we used c1/2 = 0.005 andm = 0. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

    702 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    and (4) into equations on the unit square, but with new

    interaction parameter c/‘2. In the stochastic case (4), therescaling leaves the noise intensity r unchanged. In otherwords, we can consider larger system sizes by consider-

    ing smaller values of c.The results of such a system size scaling analysis are

    shown in Fig. 9, where we consider both the determin-

    istic case r = 0 and the stochastic case with r = 0.01 for

    0 0.2 0.4 0.6 0.8 10

    100

    200

    300

    400

    500

    t/tend

    0 0.2 0.4 0.6 0.8 10

    100

    200

    300

    400

    500

    t/tend

    σ = 0

    σ = 0.01

    b int, 0

    Fig. 9. Evolution of internal and boundary component averages of X+(t) for

    case r = 0, the bottom row is for r = 0.01. The left column is for the numberbint, 0 Æ c/c*, with c

    1=2� ¼ 0:0015; from top to bottom in each of the diagrams

    column is for the number of components touching the boundary and shows th

    of c1/2.

    m = 0 and a variety of c-values. The resulting evolutioncurves for bint, 0 have been multiplied by c/c*, withc1=2� ¼ 0:0015, which is the smallest considered valueof the interaction parameter. Thus, the left column of

    Fig. 9 indicates a scaling of bint, 0 proportional toc�1, i.e., proportional to the area of the underlying do-main. For the number of boundary components shown

    in the right column of Fig. 9, we used the scaling factor

    0 0.2 0.4 0.6 0.8 10

    20

    40

    60

    80

    100

    120

    140

    160

    t/tend

    0 0.2 0.4 0.6 0.8 10

    20

    40

    60

    80

    100

    120

    140

    160

    t/tend

    bbdy, 0

    varying c and a sample size of 100, with m = 0. The top row shows theof internal components and shows the evolutions of the scaled quantity

    the curves are for c1/2 = 0.0015, 0.0025, 0.005, 0.0075, 0.01. The righte evolutions of the scaled quantity bbdy, 0 Æ (c/c*)

    1/2, for the same values

  • M. Gameiro et al. / Acta Materialia 53 (2005) 693–704 703

    (c/c*)1/2. The diagram therefore indicates that bbdy, 0

    scales proportional to c�1/2, i.e., proportional to thelength of the boundary of the domain. Due to the spe-

    cific form of the involved scaling factors, the absolute

    numbers given on the vertical axes in Fig. 9 are the

    correct component numbers for c1/2 = 0.0015. In thissituation, the number of internal components clearly

    dominates the number of boundary components

    throughout spinodal decomposition and early coarsen-

    ing. Notice also that the results for the stochastic case

    indicate that the critical noise intensity rcr discussed atthe end of Section 3 appears to be independent of the

    system size. This explains the findings of Fig. 5.

    While the scaling in the left column of Fig. 9 showsmore deviation than the one in the right column, these

    deviations are a consequence of the small system sizes

    for large values of c. Yet, we find it remarkable thatthe non-monotone evolution of bint, 0 can be observedeven for c1/2 = 0.01, despite the fact that in this casewe have bint, 0 � 0.5bbdy, 0 during spinodal decomposi-tion and early coarsening, and that the amplification

    factor used to scale the interior component curve is0.012/c* = 44.4. We believe that the accuracy of theabove scalings over the considered c-range indicatestheir validity also for smaller values of c.

    The results of this section demonstrate that the infor-

    mation provided by the collection of Betti numbers for

    X±(t) can be used to derive detailed quantitative state-

    ments about the relative sizes of boundary effects in fi-

    nite size systems, when compared to internal (or bulk)effects. This is in sharp contrast to what can be extracted

    from other (easier computable) topological invariants,

    such as the Euler characteristic. In the above situation,

    the Euler characteristic only describes the number of

    boundary components, but can provide no insight into

    the number of internal components.

    6. Conclusion

    In this paper we proposed the use of computational

    homology as an effective tool for quantifying and distin-

    guishing complicated microstructures. Rather than dis-

    cussing experimental data, we considered numerical

    simulations of the deterministic Cahn–Hilliard model,

    as well as its stochastic extension due to Cook. The ob-tained topological characterizations were used to (a) un-

    cover significant differences in the temporal evolution of

    the pattern complexity during spinodal decomposition

    between the deterministic and the stochastic model, as

    well as to (b) establish the existence of a gradual transi-

    tion of the spinodal decomposition microstructure mor-

    phology as the total mass approaches the boundary of

    the spinodal region. Furthermore, by combining differ-ent topological information, we managed to (c) obtain

    detailed information on boundary effects for various

    system sizes. In order to simplify our presentation, the

    results were presented in a two-dimensional setting.

    Nevertheless, the computational tools are available for

    and effective in arbitrary dimensions.

    We believe that the presented methods can serve as

    practical tools for assessing the quality of continuummodels for phase separation processes in materials. This

    is achieved by providing quantitative topological infor-

    mation which can readily uncover differences in models

    as in (a). In addition, this quantitative information can

    be used to compare the computed microstructure topol-

    ogy to experimental observations. For example, the

    study in [18] determined the handle density of spinodally

    decomposing iron–chromium alloys. This quantity cor-responds to the first Betti number, and the results of

    [18] indicate a monotone decay – which in combination

    with (a) favors the stochastic Cahn–Hilliard–Cook

    model over the deterministic one. Finally, combining

    topological information as in (c) makes it possible to

    quantify boundary effects, which are present in any finite

    size system.

    A second possible application is the identification ofmodel parameters based on the microstructure topol-

    ogy. For example, our considerations in (b) resulted in

    characteristic evolution curves as a function of the

    parameter m, as well as in scaling information with re-

    spect to c. Combined, these graphs could be used todetermine specific values of these parameters for certain

    experimental situations.

    It would be interesting to apply the methods pre-sented in this paper to more elaborate models which in-

    clude additional effects, such as for example anisotropic

    elastic forces [26] or polycrystalline structures [32].

    Moreover, we believe that combining our methods with

    the matricity concept introduced in [24,25] can further

    increase the classification potential of the method.

    Acknowledgements

    The work of M.G. was partially supported by

    CAPES, Brazil; the work of K.M. and T.W. was par-

    tially supported by NSF grants DMS-0107396 and

    DMS-0406231, respectively. We thank E. Sander for

    helpful discussions concerning the presentation of this

    material. We also thank the anonymous referee whosesuggestions led to significant improvements.

    References

    [1] Arns CH, Knackstedt MA, Pinczewski WV, Mecke KR. Phys

    Rev E 2001;63(3):031112.

    [2] Binder K. In: Arnold L, editor. Stochastic nonlinear sys-

    tems. Berlin: Springer; 1981. p. 62–71.

    [3] Blömker D, Maier-Paape S, Wanner T. Commun Math Phys

    2001;223(3):553–82.

  • 704 M. Gameiro et al. / Acta Materialia 53 (2005) 693–704

    [4] Cahn JW. J Chem Phys 1959;30:1121–4.

    [5] Cahn JW. J Chem Phys 1965;42:93–9.

    [6] Cahn JW. Trans Metall Soc AIME 1968;242:166–80.

    [7] Cahn JW, Hilliard JE. J Chem Phys 1958;28:258–67.

    [8] Cahn JW, Hilliard JE. J Chem Phys 1959;31:688–99.

    [9] Cerezo A, Hetherington MG, Hyde JM, Miller MK. Scripta

    Metall Mater 1991;25:1435–40.

    [10] Cook H. Acta Metall 1970;18:297–306.

    [11] Frigo M, Johnson SG. Proc ICASSP 1998 1998;3:1381–4.

    [12] Gameiro M, Kalies W, Mischaikow K. Phys Rev E 2004;

    70(3):035203 .

    [13] Garcke H, Niethammer B, Rumpf M, Weikard U. Acta Mater

    2003;51(10):2823–30.

    [14] Hairer E, Wanner G. Solving ordinary differential equations

    II. Berlin: Springer; 1991.

    [15] Hilfer R. Transport Porous Med 2002;46(2-3):373–90.

    [16] Hilliard JE. In: Aaronson HI, editor. Phase transforma-

    tions. Metals Park (OH): American Society for Metals; 1970.

    p. 497–560.

    [17] Hyde JM, Miller MK, Hetherington MG, Cerezo A, Smith

    GDW, Elliott CM. Acta Metall Mater 1995;43:3403–13.

    [18] Hyde JM, Miller MK, Hetherington MG, Cerezo A, Smith

    GDW, Elliott CM. Acta Metall Mater 1995;43:3415–26.

    [19] Ioannidis MA, Chatzis I. J Colloid Interface Sci

    2000;229(2):323–34.

    [20] Kaczynski T, Mischaikow K, Mrozek M. Computational homol-

    ogy. Applied mathematical sciences, vol. 157. New York: Sprin-

    ger; 2004.

    [21] Kalies W, Pilarczyk P. Computational homology program.

    Available from: http://www.math.gatech.edu/~chom/, 2003.

    [22] Kong TY, Rosenfeld A. Comput Vision Graph Image Process

    1989;48(3):357–93.

    [23] Langer JS. Ann Phys 1971;65:53–86.

    [24] Leßle P, Dong M, Schmauder S. Comput Mater Sci

    1999;15(4):455–65.

    [25] Leßle P, Dong M, Soppa E, Schmauder S. Scripta Mater

    1998;38(9):1327–32.

    [26] Löchte L, Gitt A, Gottstein G, Hurtado I. Acta Mater

    2000;48(11):2969–84.

    [27] Maier-Paape S, Wanner T. Commun Math Phys

    1998;195(2):435–64.

    [28] Maier-Paape S, Wanner T. Arch Ration Mech Anal

    2000;151(3):187–219.

    [29] Mendoza R, Alkemper J, Voorhees PW. In: Rappaz M, Beck-

    ermann C, Trivedi R, editors. Solidification processes and

    microstructures: a symposium in honor of Wilfried Kurz. TMS;

    2004. p. 123–29.

    [30] Miller MK, Hyde JM, Hetherington MG, Cerezo A, Smith

    GDW, Elliott CM. Acta Metall Mater 1995;43:3385–401.

    [31] Quarteroni A, Valli A. Numerical approximation of partial

    differential equations. Berlin: Springer; 1994.

    [32] Ramanarayan H, Abinandanan TA. Acta Mater

    2003;51(16):4761–72.

    [33] Sander E, Wanner T. J Stat Phys 1999;95(5-6):925–48.

    [34] Sander E, Wanner T. SIAM J Appl Math 2000;60(6):

    2182–2202.

    [35] Shardlow T. Numer Func Anal Opt 1999;20(1–2):121–45.

    [36] Torquato S. Annu Rev Mater Res 2002;32:77–111.

    [37] Ujihara T, Osamura K. Acta Mater 2000;48(7):1629–37.

    [38] Wanner T. Trans Am Math Soc 2004;356(6):2251–79.

    http://www.math.gatech.edu/~chom/

    Evolution of pattern complexity in the Cahn -- Hilliard theory of phase separationIntroductionHomology and Betti numbersEffects of noise on the pattern morphologyMorphology changes due to mass variationBoundary effects and scalingsConclusionAcknowledgementsReferences