interfacial dynamics for thermodynamically consistent phase-field models with nonconserved order...

Upload: imm-k-rod-phi

Post on 02-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    1/49

    Electronic Journal of Differential Equations, Vol. 1995(1995), No. 16, pp. 149.ISSN: 1072-6691. URL: http://ejde.math.swt.edu (147.26.103.110)telnet (login: ejde), ftp, and gopher access: ejde.math.swt.edu or ejde.math.unt.edu

    Interfacial Dynamics for Thermodynamically

    Consistent PhaseField Models with

    Nonconserved Order Parameter

    Paul C. Fife

    Oliver Penrose

    Abstract

    We study certain approximate solutions of a system of equations for-mulated in an earlier paper (Physica D 43 4462 (1990)) which in dimen-sionless form are

    ut+ w()t = 2u,

    2t = 2

    2 + F(,u),

    where u is (dimensionless) temperature, is an order parameter, w() isthe temperatureindependent part of the energy density, and F involvesthederivative of the free-energy density. The constants and are oforder 1 or smaller, whereas could be as small as 108. Assuming that asolution has two singlephase regions separated by a moving phase bound-ary (t), we obtain the differential equations and boundary conditions

    satisfied by the outer solution valid in the sense of formal asymptoticsaway from and the inner solution valid close to . Both first and sec-ond order transitions are treated. In the former case, the outer solutionobeys a free boundary problem for the heat equations with a Stefanlikecondition expressing conservation of energy at the interface and anothercondition relating the velocity of the interface to its curvature, the surfacetension and the local temperature. There are O() effects not present inthe standard phasefield model, e.g. a correction to the Stefan conditiondue to stretching of the interface. For secondorder transitions, the mainnew effect is a term proportional to the temperature gradient in the equa-tion for the interfacial velocity. This effect is related to the dependenceof surface tension on temperature.

    We also consider some cases in which the temperature u is very small,and possibly or as well; these lead to further free boundary problems,

    1991 Mathematics Subject Classifications: 35K55, 80A22, 35C20.Key words and phrases: phase transitions, phase field equations, order parameter,free boundary problems, interior layers.c1995 Southwest Texas State University and University of North Texas.

    Submitted: September 26, 1995. Published November 27, 1995.

    1

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    2/49

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    3/49

    EJDE1995/16 P. C. Fife & O. Penrose 3

    The conditions leading to a free boundary problem of MullinsSekerka type

    are elucidated and contrasted with those usually postulated within the frame-work of the traditional phase field model. In particular, the time evolution fromStefantype motion into MullinsSekerka motion is discussed (Sec. 14).

    The relation between the interface thickness, the surface tension, and theGibbsThompson law, is discussed, and our viewpoint corroborated by knownphysical data.

    Finally, the asymptotic procedure here used is developed and discussed withgreat care, and certain first order terms in the interface conditions are derivedhere for the first time.

    The phase field models developed in [PF1] were based on certain postulatedforms for the internal energy, free energy, and entropy of the system. Theseare made precise in Sec. 3 of this paper. Other assumptions, of a mathemat-ical nature, are made in the paper, particularly in Sec. 5. These latter areassumptions about the nature of the layered solutions we are investigating, andare made in order to carry out a matched asymptotic expansion. Assumptionsof this type are in fact nearly always made in formal asymptotic treatments ofapplied problems, but are rarely made explicit. We strive to spell them outcompletely.

    There has been good progress in rigorous justification of the type of formalasymptotics used here, which means proving the existence of solutions for whichour assumptions hold. See [CC], [St], [St2] for such a justification in the case ofthe traditional phase field model. (Such progress has been even more impressivein the case of the Allen-Cahn and Cahn-Hilliard models.)

    Other thermodynamically consistent models have been developed in recentyears; see [T], [UR], [AP1], [AP2], [WS], and the references given there (note

    also [K], described from a thermodynamically consistent point of view in [WS]).In many cases they are more complicated than ours, due to the inclusion ofeffects such as variable density.

    2 The main ideas and results

    As in [PF1] we start from a Helmholtz free energy function of the form

    f(, T) = w() Ts0() cTlog T,

    where is the order parameter, T the absolute temperature, w() and s0 arethe temperature-independent parts of the energy density and entropy density,

    andc is the heat capacity at constant , which for the time being we take to beconstant. The internal energy is then

    e(, T) =(f /T)

    (1/T) = w() + cT (1)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    4/49

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    5/49

    EJDE1995/16 P. C. Fife & O. Penrose 5

    -

    6

    h+(T)

    h(T)

    h0(T)

    T

    (a)

    -

    6

    f

    -1 1-

    6

    f

    1

    -1

    6

    -

    T

    (b)

    h+(T)

    h(T)

    Figure 1: Two possible functionsf(, T) plotted for T=T0 (thick)and T > T0 (thin), and their null sets (with the functions hi(T)) :(a)f(, T) = (2 1)(c(T T0) )(b)f(, T) = (1 2) T+ T0

    Our purpose is to study these layered solutions in detail. Our focus is on allsolutions of this type, rather than on solutions satisfying specific boundary orinitial conditions.

    We are mainly concerned with first-order phase transitions. A matchedasymptotic analysis for this case is given in Sections 5 12. Only the two-dimensional case is considered, but the method is easily extended to three di-mensions. Our analysis relies on the smallness of , a parameter (actually adimensionless surface tension) related to 1 which will be given later. We as-sume that the dimensionless width of the layers is O()) and that their internalstructure scales with in a way to be defined more carefully in Sec. 5. Underthese assumptions, the analysis allows one to deduce further information of adetailed nature about the layered solutions. For example, it provides approxi-mate information about how the interphase regions move. It is this property ofengendering further information which lends the assumption its credibility.

    The result of the analysis is that the layered solutions can be formally ap-

    proximated at the macroscopic level by the solution of a free boundary problem,the interphase layer being approximated by a sharp interface. The free bound-ary problem, set out in Section 11, consists of heat equations in each of thetwo single-phase domains, coupled through their common domain boundary(the interface) by means of two specific relations. One of them is a Stefanlike

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    6/49

    6 Interfacial Dynamics EJDE1995/16

    condition, and the other is a condition relating the temperature there to the

    velocity, curvature, and surface tension . These approximations, valid awayfrom the layer, are supplemented by fine structure approximations, valid in thevicinity of the layer, which give information about the phase and temperatureprofiles within the interphase region.

    The analysis reveals some new effects: (a) to order , the temperature may bediscontinuous at the interface; (b) the effect of interface stretching is accountedfor by an extra term in the Stefan condition; and ( c) there is in general a smallextra normal derivative term as well as a curvature term in the other interfacecondition.

    In Section 13, we consider the analogous question of phase boundary motionin the case of second order phase transitions. The previous development iseasily adapted to this case, but the results are strikingly different. The modelconsidered by Allen and Cahn [AC] is a particular case.

    The basic free boundary problem obtained in Section 11 has many particularlimiting cases when certain order of magnitude assumptions are made on theparameters of the problem; a few of these possibilities are explored in Sections14 and 15. As opposed to previous derivations of similar limiting cases, weshow that the various alternative free boundary problems obtained in [CF] and[C2] as formal approximations for small when certain parameters are taken todepend on , appear here as corollaries of our basic results. The same is truefor the classic motionbycurvature problem. Thus one general analysis does itall. In the case of curvaturedriven free boundary problems of various kinds,we elucidate in Section 15 the physical conditions under which they are validapproximations. These conditions are distinctly different from those which havebeen suggested in the past, and are motivated by thermodynamic considerations.

    In Section 16, the implications of allowing the coefficients to depend on andT are explored. Section 17 is devoted to the interesting case, not consideredbefore, when the thermal diffusivity is enhanced within the interphase zone.Again, the analysis in Sections 5 12 can be adapted. The most significant newfeature is the appearance, in the Stefan interface condition, of an extra termrepresenting diffusion within the zone. This term involves the second tangentialderivative (or, in three dimensions, the surface Laplacian) of the curvature of theinterface. An analogous result has been derived by Cahn, Elliott, and NovickCohen [CEN], in the case of CahnHilliard type equations. They show thatenhanced mobility within the interfacial zone results in a limiting free boundaryproblem in which the motion is driven by the Laplacian of the curvature. See[CT] for a materials scientific theory of a class of surface motions depending onthe Laplacian of curvature.

    In Section 18 the generalization, important in some applications, is madeto multi-component order parameters. As we shall see, this provides a possiblebasis for treating the motion of anisotropic interfaces. Free boundary problemsof the same general type as before are obtained, but the coefficients of thecurvature and velocity in the free boundary conditions are more complicated.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    7/49

    EJDE1995/16 P. C. Fife & O. Penrose 7

    A systematic matched asymptotic analysis of moving layer problems of this

    sort was first carried out in [CF]; see also [C2] and [F2]. Similar problems weretreated using related techniques in [P] and in [RSK]. To an extent, our conclu-sions are analogous to those in [CF] and [C2], but there are many importantdifferences, as was mentioned above.

    3 The basic model and hypotheses for first order

    phase transitions; nondimensionalization.

    The models considered here consist of field equations (2), (3) for a temperaturefunctionT(x, t) and an order parameter function (x, t). In the main part ofthe paper we shall assume that c, 0, andD are positive constants, and that isa scalar function. Some assumptions about the functionfwill also be needed.These depend somewhat on whether the phase transition is of first or secondorder. We begin with the case of a first order transition (sections 512), forwhich the assumptions are set out below. Second order transitions are treatedin Sec. 13.

    A1. f(, T) is twice continuously differentiable in both variables. Consideredas a function of at fixed T for any T in some interval T < T < T+, f(, T)has two local minima = h(T) and = h+(T), which we order so that

    h(T)< h+(T).

    It also has a single intermediate local maximum at = h0(T) (h(T), h+(T)).Thus ftypically has one of the forms shown in Figure 1. The two numbersh(T) and h+(T) are the values of for which uniform phases can exist at

    temperature T. In general these two minima correspond to different values ofthe free energy. If that is the case, one of the phases is stable and the other ismetastable, so that they cannot coexist at equilibrium. Only if they correspondto the same free energy density,

    f(h(T), T) = f(h+(T), T), (5)

    can the two phases coexist at equilibrium.

    Our second assumption (which holds only for first order transitions) will bethat phase equilibrium is possible only at a single temperatureT0 (the meltingtemperature, in the case of solidliquid transitions):

    A2. Equation (5) is satisfied if and only ifT =T0 (T, T+).Our third assumption strengthens the local minimum condition on f(, T)

    at = h(T) in A1, to

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    8/49

    8 Interfacial Dynamics EJDE1995/16

    A3.

    2f2

    (h(T), T)> 0. (6)

    Our fourth assumption concerns the latent heat. To write it simply, wedenote = h(T0); c =h0(T0). Since we are considering the case of a firstorder phase transition, the energies of the two phases are different : w(+)=w(). The more ordered phase (with the order parameter near+) will havethe lower energy, and so the latent heat is

    = w() w(+), (7)

    satisfying

    A4.

    > 0. (8)

    In view of (1) and (7), A4 is equivalent to the condition

    T

    +

    f(, T)/

    T d > 0 at T=T0.

    Using A2, we see that this is equivalent to

    d

    dT [f(h(T), T) f(h+(T), T)] |T=T0 >0, (9)

    so that if (5) holds for T=T0, it cannot hold for T=T0, and hence A4 impliesthe only if part of A2.

    It will be convenient to recast our equations (2), (3) in dimensionless form.Recall that x and t are physical variables. We define dimensionless space andtime by

    x= x/L; t=tD/cL2. (10)

    where L is a characteristic macrolength for our system. For example, we maychoose it to be he diameter of the spatial domain of definition of our functions and T or the minimum radius of curvature of the initial interface, defined tobe the curve {= c}. Each term of Equation (2) has the dimensions of energydensity per unit time, and the terms in (3) have dimensions of energy densityper unit temperature. We divide (2) by DT0

    L2 and (3) byc, to make each term

    dimensionless.To simplify the notation further, we use a new temperature variable u =

    TT0

    1, whereT0 is given in A2 above, and define

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    9/49

    EJDE1995/16 P. C. Fife & O. Penrose 9

    w() w()cT0 , cT0 ,

    F(, u) = 1cT(u)

    f(, T(u)),

    (11)

    where is a dimensionless parameter chosen so that

    F

    (c, 0) = 1. (12)

    (We are assuming that+is of order 1.) This is our method of normalizingthe seat functionF. But we have also incorporated into the definitions of thedimensionless w and above; this is natural since f, w and are related by(1), (4), and (7) and constitute an important point of departure from previous

    phase-field models (see Sec. 15). The use ofallows us to obtain approximatebut simpler forms of the laws of interfacial motion whenis small (Sec. 14).Clearly, (7) continues to hold with the overbars removed. Since w and w

    are only defined up to an arbitrary additive constant, we are free to choose thatconstant so that

    w w() = 2

    . (13)

    With these representations, (2) and (3) become

    ut+ w()t = 2u, (14)

    2

    t= 2

    2

    + F(, u), (15)

    where now denotes differentiation with respect to x and we have set 2 =1/L2c and = 0D/1c. We expect to be O(1) but, as we shall see inSec. 4 ,2 is typically very small. Equations (14) and (15) form the basis of theremainder of the paper.

    Our assumptions A1 A4 can now be reexpressed in terms of the newnotation:

    Equivalent of A1: For each small enoughu, the function F(, u) is bistablein; that is, it has the form of a seat function of, as exemplified by the graphsin Fig. 1. (Again, we denote the outer zeros ofF byh(u).)

    Equivalent of A2:

    h+(u)h(u)

    F(, u)d= 0 if and only if u= 0. (16)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    10/49

    10 Interfacial Dynamics EJDE1995/16

    Equivalent of A3:

    F

    (h(u), u)< 0. (17)

    It follows from (1), (5), (7), and (11) that

    = +

    Fu(, 0)d. (18)

    We therefore have:

    Equivalent of A4:

    d

    du h+(u)

    h(u)

    F(, u)d < 0 when u= 0. (19)

    Again, note the relation between this and the only if part of (16).As a first consequence of these assumptions, we note the fact, which is guar-

    anteed (see [F1] and its references, for instance) by (16) and (17), that theboundary value problem

    + F(, 0) = 0, z (, ); () = , (0) =c. (20)has a unique solution (z). Changing the integration variable in (19) from toz by the relation = (z), we see that (19), and hence A4, is in turn equivalentto

    F

    u((z), 0)(z)dz

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    11/49

    EJDE1995/16 P. C. Fife & O. Penrose 11

    = w(1) w(1) = 4af0.Another measurable quantity giving information about the parameters of

    the model is the surface tension . It is equal to the excess free energy per unitarea in a plane interface, which for T =T0 is ([CA])

    =

    f((r/), T0) +

    1

    21T0

    d

    dr

    2dr. (23)

    According to (22), the functionF(, 0) is f0cT0

    (2 1), so that the definition(12) ofgives F(0, 0) =

    f0cT0

    = 1 and

    f0 = cT0.

    Since nowF(, 0) = (2 1), we have from (20)

    (z) = tanh z

    2.

    Using the relations z = r/ = r/L,(r) = (z) = tanh(z/

    2), and f(, T0) =14

    f0(2 1)2 = 14f0sech4(z/

    2), we can simplify (23) to

    =L

    14

    f0sech4 z2

    + 1T0

    42L2sech4

    z2

    dz

    =LcT022

    3 since f0 = cT0 and 2 =

    1cL2

    .

    (24)

    Therefore we can think of the productas a measure for the magnitude of the

    surface tension.The surface tension at a solidliquid interface can be deduced from the value

    of the GibbsThompson coefficient

    G=T0

    =

    T04af0

    .

    From (24) and this we obtain for the width of the interface

    L= 3

    2

    2cT0=

    3

    2aG

    T0.

    The value ofa can be estimated as follows: first, if the entropy is to be aconcave function of, then, as shown in [PF3], we must havea > 1

    2; secondly,

    if the liquid can be supercooled to a temperature T then fmust have a localminimum at = 1 whenT =T, which with (22) implies T/T0 > a/(1 + a)i.e. a < T/(T0 T). For example in the case of the ice-water transition wemight take T0 = 273K, T = 233K, giving a < 5.8. We shall take the valuea= 1 to be typical.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    12/49

    12 Interfacial Dynamics EJDE1995/16

    Typical values ofG and T0 are 105 cm-deg and 300K, respectively. Using

    them, we obtain

    L 1.5 107 cm,which is of the order of a few lattice spacings of an ice crystal, a not unreasonableinterface thickness. IfL equals, say, 10 cm., then is less than 107 and theapproximations to be developed in this paper should be very accurate.

    5 The approximation scheme.

    Our procedure is based on assumptions which have been used implicitly invarious earlier studies of similar problems ([CF], [RSK], [P], etc.) and havebeen rigorously justified, under certain conditions, in the analogous cases of theCahnAllen equations [MSc], [Chen1] and the CahnHilliard equation [ABC].We spell them out completely. Their plausibility rests in large part on thefact that they lead to a succession of reasonable formal approximations. Forsimplicity we consider only the two dimensional case.

    The core assumption is that there exist families of solutions (u(x, t; ),(x, t; ))of (14), (15), defined for all small > 0, all x in a domainD R2, and all tin an interval [0, t1], with internal layers. This concept is defined precisely inthe form of assumptions (a) (e) below, as follows.

    For such a family, we assume that, for all small 0, the domainDcan ateach timet be divided into two open regions D+(t; ) and D(t; ), with a curve(t; ) separating them. This curve does not intersectD. It is smooth, anddepends smoothly on t and . In particular, its curvature and its velocity are

    bounded independently of. These regions are related to the family of solutionsas follows.(a) Let be any open set of points (x, t) inD [0, t1] such that

    dist(x, (t; 0)) is bounded away from 0. Then for some 0 > 0, u can, weassume, be extended to be a smooth (say, three times differentiable) function ofthe three variables x, t, and uniformly for 0 < 0, (x, t) in . The sameis assumed true of(x, t; ).

    It follows in particular that the functionsuk(x, t) 1k!k u|=0, k= 0, 1, 2, 3,are defined in all ofD \ (t; 0). A similar statement holds for k |=0.

    It also follows from (15) that for (x, t) in any region as described above,F(, u) = O(2). This implies, by the definition ofh, that is close either toh+(u) or h(u). (The third possibility would be near h0(u); but in view ofthe instability of this constant solution of (15) (for fixed u), we assume there

    are no extended regions where is close to this value.)(b) For inD(t; 0) [0, t1], we assume that is close to h(u). Our

    interpretation is that the material where is close toh(u) is in phase I (theless ordered phase, since < +) and that where is near h+ is in phase II,the more ordered phase.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    13/49

    EJDE1995/16 P. C. Fife & O. Penrose 13

    Much of our analysis will refer to a local orthogonal spatial coordinate system

    (r, s) depending parametrically on t and , defined in a neighborhood of (t; ),which we define precisely as the set where = c. We definer(x, t; ) to be thesigned distance from x to (t; ), positive on theD+ side of (t; ). Then forsmall enough, in a neighborhood

    N(t; ) = {x: r(x, t; )< } ,we can define an orthogonal curvilinear coordinate system (r, s) inN, wheres(x, t; ) is defined so that when x (t; ), s(x, t; ) is the arc length along(t; ) tox from some point x1(t; ) (t; ) (which always moves normal to as t varies).

    Transforming u and to such a coordinate system, we obtain the functions

    u(r,s,t; ) = u(x, t; ), (r,s,t; ) = (x, t; ).

    Let uk(r,s,t), k(r,s,t) be defined in the same way as uk and k above, interms of derivatives at = 0. They exist, by virtue of (a) above.

    (c) For eacht [0, t1],we assume that the-derivativesuk, k 3, restrictedto the open domain D{r >0}, can be extended to be smooth functions on theclosure D{r 0} (on , they no longer signify the derivatives indicated above).Similarly, we assume that the restrictions to D {r 0, z0 > 0, the Taylor series approximations

    u(x, t; ) = u0(x, t) + u1(x, t) + o(), (25a)

    u(r,s,t; ) = u0(x, t) + u1(x, t) + o(), (25b)

    U(z, s, t; ) = U0(z, s, t) + U1(z, s, t) + o(), (25c)

    together with their differentiated versions, are valid for all sufficiently small :in the case of (25a,b), uniformly for dist(x, (t;0))> r0 >0 and in the case of

    (25c), for dist(x, (t; ))< z0. Similar statements hold for , , . Truncatedseries as in (25a) and (25b) will constitute our outer approximation; ones likethose in (25c) will constitute the inner approximation.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    14/49

    14 Interfacial Dynamics EJDE1995/16

    (e) The approximations in (25b,c) above are assumed to hold simultaneously

    in a suitable region: for some 0< ,

    and (25c) holds for

    dist(x, (t; )< 2.

    Differential equations for the functions uk, k, etc. can be obtained bysubstituting (25a) and its analog into (14) and (15) and equating coefficientsof powers of. For this purpose, the only conclusion from (15) which will beneeded is the relation

    F(, u) = 0(2), (26)

    which provides algebraic equations relating uk andk, k 1. For example,

    0 = h(u0) in D. (27)In the same way, we get from (14) that

    te0 = 2u0 in D,

    e0 = u0+ w(0).(28)

    Our object will be to find free boundary problems satisfied by the outerfunctionsuk,k. For this, we need not only differential equations and algebraicrelations such as (27) and (28) holding inD, but also extra conditions on .These extra conditions will be obtained by finding the inner functions Uk, kand using assumption (e) to obtain matching conditions relating them to theouter functionsu, . And to find these inner functions, we shall in turn need torelate the surface (t; ) to the family precisely and to specify a curvilinearcoordinate system near . This will be done in the next section.

    6 The r, s and z, s coordinate systems; matching

    relations.

    Recall that our definition of will be the level surface

    (t; ) = {x: (x, t; ) = c} , (29)

    and the (r, s) coordinate system is attached to .To go from Cartesian to (r, s) coordinates we transform derivatives as follows:

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    15/49

    EJDE1995/16 P. C. Fife & O. Penrose 15

    t is replaced by t+ rtr+ sts;

    2 is replaced by rr+ |s|2ss+ 2rr+ 2ss.(30)

    Here, we have used the fact that|r| 1.The derivatives of r and s in these expressions can be written in terms

    of kinematic and geometric properties of the interface . The details of thecalculation are given in the Appendix; we quote only the results here. Letv(s, t; ) denote the normal velocity of in the direction ofD+ at the point s,and let denote its curvature, defined by (s, t; ) =2r(x, t; )|. Then thetime derivatives ofr ands can be written

    rt(x, t; ) = v(s, t; ), st= rvs1 + r

    , (31a)

    where, here and below, the arguments ofv and are (s(x, t; ), t; ), and sub-scripts onv and denote differentiation. The corresponding expression for thespace derivatives ofr ands are

    2r(x, t; ) = 1 + r

    , (31b)

    2s(x, t; ) = rs(1 + r)3

    , |s|2 = 1(1 + r)2

    . (31c)

    To obtain the equations for the inner approximation we first define, in ac-cordance with (1) and (11), the nondimensional internal energy

    e= u + w(). (32)

    As in (25b), we shall denote by ethe same quantity e expressed as a function ofr,s,t,. In view of (30) and (31), our basic equations (14) and (15) then become

    te vre rvs1+rse=

    urr+ 1+r

    ur+ 1(1+r)2

    uss+ rs(1r)3us,

    2(t vr rvs1+r s) =

    (33)

    2rr+

    1 + rr+

    1

    (1 + r)2ss+

    rs(1 + r)3

    s+ F(, u). (34)To obtain the inner expansion we follow the procedure set out in Sec. 5(d),

    defining inN(t; ) the stretched normal coordinate

    z = r/ (35)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    16/49

    16 Interfacial Dynamics EJDE1995/16

    and the functions

    U(z, s, t; ) = u(r,s,t; ), (z, s, t; ) = (r,s,t; ), E(z, s, t; ) = e(r,s,t; ).

    To obtain differential equations for Uand , we substitute (35) into (33), (34),obtaining :

    Uzz+ Uz+ vEz 22Uz+ 2Uss 2Et= O(3), (36)zz+ F(, U) + z+ vz = O(

    2), (37)

    E= U+ w(). (38)

    The functions U, , Ecan be expanded in powers of as in (25c) :

    U(z, s, t; ) = U0(z, s, t) + U1(z, s, t) + o() (

    0),

    = 0+ + o() (39)

    and so on. By the regularity assumptions in Sec. 4, v(s, t; ) and (s, t; )can also be expanded in series like (39). The differential equations satisfiedby the functions U0, 0, etc. are obtained by substituting (39) into (36) and(37). This will be done in the next section, but first we formulate the matchingconditions (e.g. [F2]) obtained by requiring that the inner and outer expansionsrepresent the same function in their common domain of validity (which existsby Assumption (e) of the previous section). They are the following, where wehave omitted the carets from the symbols u and .

    limr0

    u0(r,s,t) = limz

    U0(z, s, t); (40)

    limr0

    ru0(r,s,t) = limz

    zU1(z, s, t). (41)

    IfU1(z, s, t) = A(s, t) + B(s, t)z+ o(1) as z , then

    A(s, t) = u1(0, s , t); B = ru0(0, s , t), (42)and so on. Similar relations apply, connecting 0, 1 to 0, 1, etc. Finally ifzU2 = A

    (s, t) + B

    (s, t)z+ o(1), then

    A(s, t) = ru1(0, s , t). (43)

    7 The zero-order inner approximation.

    We substitute (35) and (39) into (36) and (37) to obtain a series expansion inpowers of for each side of the latter. By equating the coefficients of each powerof we obtain a sequence of equations for the various terms Ui and i . Thefirst couple of them are analyzed as follows :

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    17/49

    EJDE1995/16 P. C. Fife & O. Penrose 17

    O(1) in (36):

    U0zz = 0.

    We want U0 to be bounded as z , because of (40); so U0 is independentofz :

    U0 = U0(s, t).

    O(1) in (37):

    0zz+ F(0, U0) = 0. (44)

    By (29), we have 0(0, s , t) =c. By the equation for analogous to (40),we seek a solution 0 which approaches distinct finite limits as z

    , and it

    is clear from (44) that these limits must be roots ofF(, U0) = 0. Moreover,it can be seen by multiplying (44) by 0z and integrating from to + thatthe integral in (16) withu replaced byU0 must vanish. By A2 (16), this implies

    U0 0. (45)Therefore 0 must satisfy the differential equation in (20), and by the definitionofr it satisfies the other conditions in (20) as well, so that it must actually bethe function defined in (20):

    0(z, s, t) (z). (46)satisfying the condition

    (

    ) = h

    (0)

    . (47)

    (Notice that 0 does not depend on s or t.)The matching condition (40) now gives, by (45), (46) and (47), the following

    boundary condition on the lowest-order outer solution :

    u0 |r=0= 0; 0|r=0 = . (48)

    At this point, we are in a position to define and evaluate, to lowest order,interfacial free energy and entropy densities at T =T0. The total free energy inthe system is ([PF1, eqs. 3.9 and 3.12])

    F[, T(u)] =

    (f(, T(u)) +1

    21T(u)||2)dx,

    T(u) = (u+ 1)T0. Its dimensionless form follows from our previous nondimen-sionalization procedure:

    F[, u] = F[, T]/cT0L2 =

    f(, T)

    cT0+

    T(u)

    2T02||2

    dx.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    18/49

    18 Interfacial Dynamics EJDE1995/16

    The (dimensionless) surface tension is the interfacial free energy per unit

    length of interface. It can be calculated, at T =T0 (u= 0) by subtracting thefree energy density of a uniform phase, which is f(, T0), from the function fin the integrand and then integrating with respect to z =r/ from to.To lowest order in we may use the approximation = (r/), T =T0 in thisintegral, obtaining

    =

    f((z)) +

    1

    2((z))2

    dz, (49)

    where we have set

    f() =f(, T0) f(, T0)

    cT0, (50)

    the dimensionless bulk free energy density. But since from (44), (45), and (46)

    = F(, 0) = dd

    f(),

    it follows that ()2 = 2f(), so that the contributions of the two terms in theintegral for are equal. We therefore have

    =

    ((z))2dz= 11

    2[f() f()]d 1. (51)

    This is a standard formula (see e.g. [AC]). We shall call 1 the scaleddimensionless surface tension.

    8 The jump condition at the interface.

    In this section, we show how the energy balance equation (33), applied to theinner approximation, leads to a jump condition for the outer solution at theinterface, from which the velocity of the interface can be determined once theouter solution is known. We first calculate this to lowest order, and then toorder.

    In view of (39), (45), and (46), we may set

    U=U , =+ ,

    where U , = O(1). We then have by (38)

    E= U+ w(+ ).

    Since does not depend on t, it follows that Et = O(), and hence from (36)that

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    19/49

    EJDE1995/16 P. C. Fife & O. Penrose 19

    Uzz+ vEz+ Uz = O(2). (52)

    Integrating (52), we get

    Uz+ vE+ U+ C1(s,t,) = O(2) (53)

    for some integration constant C1 = C11+ C12.Since by (39a) U=U1 + U2+ O(

    2), the lowest order approximation in (53)yields

    U1z= (vE)0 C11 = v0w((z)) C11, (54)where in the second equation we have used the fact that the expansions ofU, v,and induce an expansion ofvEin powers of, with (vE)0 = v0w() being

    the lowest order term. Similarly induced expansions will be used below.To obtain the lowest order jump condition for the outer approximation, we

    apply (40) and (41) to the left and right sides of (54). We thus obtain

    ru0 |r=0= v0e0|r=0 C11= v0/2 C11, (55)using (48) and (13). By subtraction we get the jump relation

    [ru0] = v0[e0] = v0, (56)where the square brackets indicate the limit from the right (r= 0+) minus thelimit from the left (r= 0).

    We now derive the jump condition to order analogous to (56). Considerthe terms of order in (53). Since U1 = U2, these terms are

    U2z = (vE)1 0U1 C12. (57)To evaluate U1, we integrate (54):

    U1= v0p(z) C11z+ C2, (58)where

    p(z) =

    zz0

    w((s))ds, (59)

    z0 will be chosen later, and C2 is another integration constant, unknown at thisstage, but depending onz0. Hence from (57), we have

    U2z = (vE)1+ 0v0p(z) C12+ 0C11z 0C2.Applying the matching relations (43) and (40), we obtain:

    ru1|r=0 = (ve)1|r=0+ 0v0P C12 0C2, (60)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    20/49

    20 Interfacial Dynamics EJDE1995/16

    where thePs are defined by the relation

    p(z) = w()z+ P+ o(1) (z ), (61)i.e.

    P =z0

    (w((z)) w()) dz.

    For the sake of symmetry in notation, we choose the lower limit z0 so thatP+ = P P. Thus we obtain (60) with P = P. Subtracting, we obtain

    [ru1] = [(ve)1] + 20v0P. (62)Combining (56) with (62), we get, to order ,

    [r

    u] =

    [ve] + 2vP+ O(2). (63)

    This relation can be used to determine v to order once the outer solution isknown to this order.

    Physically, eqn (63) expresses the conservation of energy at the interface.The left side represents the net flux of energy into per unit length; the firstterm on the right represents the portion of that energy which is taken up withphase change. The second term on the right of (63), which is a higher orderterm not usually displayed, represents the effect of stretching or contractingas it evolves; its presence is necessary to ensure conservation of energy to thisorder.

    9 The zero-order outer approximation.

    The zero-order outer approximation to our layered family of solutions consistsof a curve 0(t) dividing Dinto two subregions D+(t) and D(t), and functionsu0, 0, continuous in each ofD. These can now be determined: we obtain u0and 0 by solving a Stefan problemS0, defined below, and then we obtain 0from (27) by taking 0 = h(u0).

    (a) InD, u0 is to satisfy (27), (28)

    te0 = 2u0, (64)

    wheree0 =u0+ w(h(u0)). (65)

    Note that in the case where h(u) are independent ofu, (64) is the usual linear

    heat equation for u0. This was the case for the liquid phase in the densityfunctional model in [PF1], and for both phases in [WS].

    (b) On 0(t) we have the interface condition (48)

    u0 = 0 (66)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    21/49

    EJDE1995/16 P. C. Fife & O. Penrose 21

    and the Stefan condition (56)

    [ru0] = v0. (67)

    (c) Our focus has been on the properties of layered solutions in general,without reference to initial or boundary conditions. But we are now led to afree boundary problem for which it is natural to specify these extra conditions.In fact, there may be boundary conditions for the temperature u, and hence foru0, atD (exclusive of 0), and initial conditions u0(x, 0), 0(0).

    Ifu0(x, 0) is nonpositive in D+(the solid) and nonnegative in D(liquid), S0is the classical Stefan problem, and for smooth initial conditions has a uniqueclassical solution for a small time interval. Our basic assumptions about thefamilies (u, ) in Sec. 5 imply that in fact this is true for all t [0, t1], theinterval mentioned in Sec. 5.

    Ifu0(x, 0) has signs opposite from those, however, then it is generally be-lieved thatS0 is an ill-posed problem, in which case our assumptions in Sec.5 will hold only in very special circumstances, such as when the domain andall data have radial symmetry. This ill-posed problem would correspond to amodel for crystal growth into a supersaturated liquid with no account takenof curvature or surface tension effects. We shall see in Sec. 14 that ifu0(x, 0)is very small in magnitude, then the assumptions in Sec. 5 become reasonableagain; in fact the lowest-order free boundary problem then contains regularizingcurvature terms.

    10 The first-order interface condition.

    To obtain a more accurate outer solution we must calculate u1, and for this weneed expressions for u1 on the interface .

    First, we apply the matching condition (42) to (58), taking into account (61)and the fact that P = P, to obtain

    u1|= v0P+ C2. (68)Here the subscript means the limit as is approached from the + side orthe side. To determine these limits, we must now find the constantC2. Itturns out that this can be done by examining the O() terms in (37). By using(46), one can put those terms into the form

    L1=

    Fu((z), 0)U1

    0

    (z)

    v0(z), (69)

    whereL is the operator defined by L zz+ F((z), 0).We know that the operator L has a nullfunction (z) which decays expo-

    nentially as z , obtained by differentiating (44) with respect to z andsetting 0 = . We are seeking a solution of (69) which grows at most as fast

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    22/49

    22 Interfacial Dynamics EJDE1995/16

    as a polynomial at

    . Multiplying (69) by and integrating, we see that the

    equation (69) has such a solution only if the right side is orthogonal to :

    Fu((z), 0)U1(z, s, t)(z)dz+ 0(s, t)1+ v0(s, t)1 = 0,

    where (recall (51))

    1 =

    ((z))2dz.

    Substituting from (58), we obtain

    Fu((z), 0)(v0p(z) C11z+ C2)(z)dz+ (0+ v0)1 = 0. (70)

    From (18) and (20), the coefficient of C2 here is seen to be. On thebasis of Assumption A4, we may therefore solve (70) for C2 as

    C2(s, t) = (+ p) v0(s, t) + 0(s, t) + qC11, (71)

    where

    p=

    p(z)(z)dz

    , q=

    z (z)dz

    , =

    1

    , (72a)

    with

    (z) = Fu((z), 0)(z). (72b)

    Substituting (71) into (68), we find

    u1|= (+ p P) v0+ 0+ qC11. (73)Finally, the constantC11 may be found from (55):

    C11(s, t) = ru0|r=0 12

    v0(s, t). (74)

    In the classical case when the Stefan problem (64) - (67) is well posed, it canbe used to determine the quantities on the right of (73) from the initial andboundary conditions for u0 and 0. In (74), either sign may be chosen; the

    right side is independent of the choice. We shall use the upper sign in D+ andthe lower one inD.Set (ru0) ru0|r=0. Then substituting (74) into (73), we have

    u1|+ q(ru0) = mv0+ 0, (75)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    23/49

    EJDE1995/16 P. C. Fife & O. Penrose 23

    where

    m= + (p P) 12

    q. (76)

    Ifm+=m, it is clear from (75) that the outer temperature distribution u willundergo a discontinuity of the order across the interface.

    Example: Consider the particular case when Fu(, 0) is an even function of, and (z) is even. Then from the above, we have q= 0, so that (75) becomes

    u1|= (+ (p P)) v0+ 0.If, in addition, (as in [PF1])

    w() =A B2 + const,then P vanishes whenever B does. In this case, then, the possibility thatu1is discontinuous across is associated with the presence of quadratic terms inw. The case when they are absent is the one treated, in the context of thetraditional phase field model, in [CF] and [F2].

    11 The first-order outer solution.

    We can now formulate the procedure for determining the first order approxima-tion to the outer solution. This approximation can be determined by solving thefollowing modified Stefan problem S, which generalizes the problem S0 definedin section 9:(a) InD, u is to satisfy (14)

    te= 2u, (77)where

    e= u + w(),

    = h(u) in D by (26). (78)(b) On we have, from (66), (75), and (63), the conditions

    (u + q(ru))| = mv+ ; (79)

    [ru] = v[e] + vP, (80)where the coefficients are given by (72a), (76).

    (c) In addition, boundary conditions, to hold on D, and initial data are tobe prescribed.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    24/49

    24 Interfacial Dynamics EJDE1995/16

    The coefficients in (79), (80) are the same as in (75) and (62). As men-

    tioned before, when h are constants, (77) is the heat equation with constantcoefficients. Even whenh are not constants, the heat equation is a reasonableapproximation in typical cases (see Sec. 14 ).

    The term in ru in (79) appears to introduce a singular perturbation intothe problem, but this is not likely to be true. We consider a model problemconsisting of (77) and (79) on the half line{r > 0} with the right side of (79)taken to be a known constant. The potential effect of such a singular perturba-tion can be ascertained from the inner equations associated with stretching thevariable r. In the model problem it is readily seen to be a regular perturbation.

    What we have shown so far is that under the assumptions in Sec. 5 , theexact solution family (u(x, t; ), (x, t; )) satisfies (77) - (80) except for errorterms of the order2. Let us now suppose thatS is a well-posed problem, andlet (u(x, t; ),(x, t; ), (t; )) denote its solution when the conditions in (c) arethe same as those of the exact family. Thus (u,, ) satisfy the same equationsand initial conditions as (u,, ), except that theO(2) terms are discarded. Itis natural to expect the following assertion, which is basic to the paper, to hold:

    Expectation: |u(x, t; )u(x, t; )|,|(...)(...)|,|(t; )(t; )| =O(2) (0),uniformly inD [0, t1].

    12 Discussion; surface tension.

    As in previous phase field models, a GibbsThompson term and a kineticundercooling termmvappear on the right of (79). In addition, there appearsan O() normal derivative term on the left, which can be important for secondorder transitions, as we shall discover.

    The last term in (79) may be compared with the thermodynamic formulafor the GibbsThompson effect, which can be written

    (T T0) | = T0

    where is the surface tension, the latent heat and the curvature in physicalunits. The corresponding formula in our dimensionless units is

    u|=

    (81)

    where = /cT0L. From (51) and (72a),

    =

    (z)2dz= 1 = .

    Thus (81) simplifies to u|= , which indeed agrees with the relevant termsin (79).

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    25/49

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    26/49

    26 Interfacial Dynamics EJDE1995/16

    In dealing with second order transitions, our formal assumptions will simply

    be A1 and (84).Following the asymptotic development in sections 411, we see that the

    following changes are necessary.

    The conclusion (45), hence also the left part of (48), no longer hold. In fact,the value ofu0on the interface is no longer determined a priori. Therefore in thelowest order outer problem (64)(67), (66) is to be replaced by [u0] = 0, and theright side of (67) is replaced by zero. Thus u0 and its derivative are continuousacross 0. In view of (82), we see thatu0 is determined as the solution of theheat equation (64) in all ofD, with no reference to (appropriate boundaryand initial conditions may be prescribed). The interfaces location is foundindependently, as we now describe.

    We pass to (70), which by virtue of (84) and (74) with = 0 becomes

    m1v0+ 0 = m2ru0, (85)

    where

    m1 = 1

    p(z)(z)dz, m2 = 11

    z(z)dz,

    the functions p and being defined in (59) and (72b).

    To find the interface (t), then, u0 is first determined by (64), and then is found from the forced motionbycurvature problem (85), with knownforcing term m2ru0 dependent on position and on time.

    In Sec. 16 , we shall examine the more realistic case when the thermal

    diffusivity D is different in the two phases; then it can be checked that theproblem for u0 can no longer be decoupled from that for .

    The interface condition (85) is similar to the motion-by-curvature law givenby the CahnAllen theory of isothermal phase transitions ([AC], [MSc]), butthere is now an extra term proportional to the temperature gradient (whichis continuous across the interface). For a physical interpretation of this term,suppose that the surface tension (excess free energy of the interface) decreaseswith temperature. Then the interface will tend to move so as to increase itstemperature. This tendency is borne out by (85) in the typical case thatm1and m2 are positive. There is an analogous forcing term in the correspondingequation (79) for firstorder transitions, but in that context its effect is relativelysmall.

    When the temperature deviationuis small (

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    27/49

    EJDE1995/16 P. C. Fife & O. Penrose 27

    14 The transition from Stefan to Mullins-Sekerka

    evolution. Other free boundary problems.

    There are several special circumstances in which our basic free boundary prob-lem (77) - (80) can be approximated formally by simpler free boundary problems.

    For example, the nonlinear diffusion equation (77) can typically be approx-imated by a linear one. In fact, the left side can be written as

    c(u)ut, where c(u) 1 + w(h(u))h(u),

    and the specific heat functions c(u) can be approximated by constants whenthe functionsh(u) are constants (as in [WS] and for one case in [PF1]), nearlyconstant, and/or when u is small enough. In such cases, they may by replacedby c

    (0). In the following, we shall assume that this approximation is valid.

    It should be noted that the assumption u 1 is entirely reasonable in manycases. It simply says

    |T T0| T0. (86)In the case of water, for example, it means that the temperature range (inCentigrade degrees) in the phenomenon under consideration is much smallerthan 273.

    Anticipating that the simpler problems to be examined may involve dy-namics on a longer time scale and hence slower speed, we proceed formally byrescalingu and t. We set

    u= u, t = t, v= v,

    where the parameters and are 1 and may be small. We make thesesubstitutions in (77) - (80) with the error terms O(2) appended, and divide byto obtain

    cut = 2u + O(2/) inD, (87a)

    (u+ q(ru))| =

    mv+

    + O(2/) on , (87b)

    [ru] =

    v +

    vP+ O(2/) on . (87c)

    To show how one kind of evolution may develop into a different kind at large

    times, we consider first the normal case when = 1. Then if we set= 1 aswell (so no rescaling actually occurs) and disregard terms of orders and higherorder in (87), we get the classical Stefan problem

    cut= 2u inD, (88a)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    28/49

    28 Interfacial Dynamics EJDE1995/16

    u= 0 on , (88b)

    [ru] = v on . (88c)

    LetD represent a bounded vessel containing the material under consid-eration, and suppose it is thermally insulated, so that the normal derivativeru = u = 0 on D. Suppose thatu > 0 in the liquid phase, u < 0 in thesolid. Consider layered solutions with interface (t) evolving according to (88).Then it can be checked that

    d

    dt

    c+

    D+

    u2dx + c

    D

    u2dx

    = 2

    D

    |u|2dx

    which in view of (88b) is strictly negative as long asudoes not vanish identically.Suppose that (t), the solution of (88), exists for all t and does not intersectD. Then it is natural to conjecture that as t, u(x, t)0 uniformly inDand that (t) approaches a limiting configuration .

    Eventually, then, u = O() and the quantity on the right of (79) and(87b) can no longer be neglected on relative to u. (This term indicates,in fact, that the temperature u can in general never achieve smaller orders ofmagnitude than .) Whenu achieves this order of smallness, we set = in(87) and observe that if we select = as well, and drop higher order terms,we obtain a reasonable problem of Mullins-Sekerka type [MS]:

    2u= 0 in D, (89a)

    u= on , (89b)

    [ru] = v on . (89c)

    An existence theory for the solution (u, ) of (89) has recently been given byChen [Ch2]; for the Hele-Shaw problem, which bears some similarity, see [CP].

    In short, when the temperature becomes small enough, the evolution accord-ing to (88) is conceptually replaced by the much slower evolution according to(89). The evolution (89) is well known to decrease the length of (t) and topreserve the area inside it, so it is expected that under the slow process, willtypically evolve from (or something near it) into a circle with the same area.

    Other free boundary problems can be obtained by assuming and/or aresmall, i.e. that the latent heat is small or the relaxation process for is quick.For example, if 2 = , we may set = 1 to obtain the classicalmotion-by-curvature law

    v= on , (90)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    29/49

    EJDE1995/16 P. C. Fife & O. Penrose 29

    coupled to a heat equation with prescribed jump condition (88c) on (known

    from solving (90)).A number of other possibilities can occur; we leave them for the reader to

    discover.

    15 Comparison with other models and limiting

    arguments.

    The free boundary problem (89) and its companion with (89a) replaced by

    ut= 2u,have been derived by asymptotic methods from other phase field models, most

    notably in [C2] and [WS]. Here we compare both the model in [WS] and itsasymptotic development with ours; similar considerations hold for the Langermodel in [C2].

    The model in [WS] has the property that our functions introduced in A1 areconstant:h(T) 0, h+(T) 1. Thus the order parameter in the purely liquidor solid phase does not depend on T. On the other hand, the latent heat (T)may depend on T. To compare the model in [WS] to ours, we must replaceby 1 (distinguishing order from disorder parameters). In our notation,their evolution problem (their eqns. 40, 41) corresponds to our (14), (15) with

    F(, u(T)) = 1

    c

    1

    T0 1

    T

    w() 4( 1)( 1

    2) (91)

    and w() = p()|=1, where p() is a function with p(0) = 0, p(1) = 1,whose first and second derivatives vanish at = 0 and 1.

    Note that this model provides no theoretical limit to the extent of super-cooling of the liquid or superheating of a solid, unlike the equations depictedin our Fig. 1 and the example in Sec. 4 . In fact, at each value ofT the freeenergyFhas local minima (in ) at = 0 and 1, representing stable liquid andsolid phases. If the temperatures under consideration are kept fairly near to T0,this should not be an important deficiency.

    The temperature-independent part of the correponding dimensionless bulkentropy density given by (4), s0=

    1c

    s0, satisfies

    s0() = 4( 1)( 1

    2) +

    w()cT0

    ,

    and since w(12

    ) = 0, we haves0(12

    ) = 1, showing that the entropy in this modelis not a concave function of.

    Further notational comparisons are the following: the parameters , m, anda in [WS] correspond to our parameters /2, 1, and 1/4c, respectively.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    30/49

    30 Interfacial Dynamics EJDE1995/16

    Their asymptotics is based on the assumption that the first term in (91) is

    O(). We rewrite the assumption as:

    cT0

    1 T0

    T

    p(1 ) =O().

    Let us assume that the dimensionless quantity cT0

    =O(1); this is reasonable intypical scenarios. Since p() is also O(1), this implies that

    1

    1 T0T =O().

    This relation can be guaranteed, for example, by requiring either(a) = O(1), or(b) TT0

    T0=O().

    Physical meaning can be given to both of these cases. In case (a), theimplication is that

    TF

    T

    cT0p(1 ) = O(1) = O(),

    whereasF

    =O(1).

    The meaning is that the free energy depends much more (by a factor 1)sensitively on than on the temperature T.

    In case (b), we have, in our notation, u = O(), which is exactly the as-sumption (= ) under which we have derived (89). This says simply that the

    temperature is close, as measured by , to the melting temperature T0.

    16 Variable c, D, and .

    All of the preceding can be extended in a straightforward way to the case whenc, D, and are given functions of and T, and w depends on T as well ason . This allows these first three physical parameters to differ in the differentphases. It was observed by Chen [Chen3] that ifc differs in the two phases, then must depend on T. It is often the case that the temperature variation in theproblem under consideration is small enough that it does not by itself induce asignificant variation in the values of these physical constants. For simplicity, weassume this is the case, i.e. that c and D depend on but not on D.

    The results under this generalization are quite analogous to those obtainedbefore, and so are not surprising. We record them here for the sake of complete-ness.

    The first observation is that the expression (4) for a term in the right sideof (3) must be supplemented by the extra termc()log T.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    31/49

    EJDE1995/16 P. C. Fife & O. Penrose 31

    By way of notation, we set

    D() =D1d(), c() = c1k(),

    whereD1 andc1 are defined to be the minimal values ofD and c, respectively.We assume that the dimensionless functions d and k areO(1).

    In the definitions of dimensionless variables given in (10) and (11), we nowreplace the symbols D and c by D1 and c1. The basic equations (14) and (15)now become

    t(k()u+ w()) = d()u, (92)

    ()2t= 22 + F(, u). (93)

    All of the analysis in the previous sections has its analog in the presentmore general context. The resulting free boundary problem (77)(80) takes thefollowing form:InD,

    te(u) = d(u)u, (77) = (94)

    e(u) = k(h(u))u + w(h(u)), d(u) = d(h(u)),

    = h(u) in D.(78) = (95)

    On

    ,

    u + n(dru)|= mv+ ; (79) = (96)

    [d(u)ru] = [ve] + vP. (80) = (97)Here the constants m andn have to be defined as follows. Let

    d0(z) = d((z)), p(z) =

    (w((z))/d0(z))dz, q (z) =

    (1/d0(z))dz, (98)

    and let the constants P and Q be such that

    p(z) = wd()z P+ o(1), z ,

    q(z) = 1d()

    z Q+ o(1).In the following, , are the same as before. Let

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    32/49

    32 Interfacial Dynamics EJDE1995/16

    m1 =

    ((z), 0)((z))2dz >0,

    p=

    p(z)(z)dz

    , q=

    q(z)(z)dz

    .

    Then

    m= m1+ (p P) 12

    (q Q),

    n= q Q.For future reference, we give here the modified versions of relations (33) and(52):

    te vre rvs1+rse=

    (dur)r+ d 1+r

    ur+ 1(1+r)2

    (dus)s+ d rs(1+r)3

    us,(33) = (99)

    (dUz)z+ vEz+ dUz= O(2). (52) = (100)

    Referring to Sec. 14 , we obtain the following generalizations of the examplesin (88)(90).

    The normal case (88):

    ctu= d(u)u (88a) = (101a)u= 0 on , (88b) = (101b)

    [dru] = v on , (88c) = (101c)

    where a different but obvious definition is given for c.

    Motion by curvature (90) (= , = 1).

    m1v= 1 on . (90) = (102)

    The Mullins-Sekerka case (89) (= = ).

    du= 0 in D, (89a) = (103a)

    u= on , (89b) = (103b)

    [dru] = v on . (89c) = (103c)

    Similar considerations hold for second order transitions. The coupling of thegeneralization of (85) with that of (64) is particularly interesting.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    33/49

    EJDE1995/16 P. C. Fife & O. Penrose 33

    17 Enhanced diffusion in the interface.

    In some materials science connections ([CT], and references therein), it is im-portant to take into account increased material diffusivity in surfaces. In ourmodels it is heat rather than material which is diffusing. Nevertheless, there isclearly an analogy with material diffusion, and so it may be interesting to adaptour results to the case when diffusivity is much greater in the interfacial regionthan it is elsewhere. Specifically, we treat the case when this ratio is of theorder1. We show, among other things, that an effect of this is the presence,in (89c) or (103c), of an extra term on the right taking the form of the secondderivative of the curvature with respect to arc length along the interface. Asimilar result was obtained by a different route for a Cahn-Hilliard model in[CFN].

    Our basic assumption is that the function D can be written as the sum of

    two terms as follows. Let D1 = Min[D(+), D()]. Then

    D= D(; ) = D1

    1D() + D()

    D1d(; ), (104)

    where D and D are O(1) functions, and D() vanishes when = . In fact,we assume that for some positive number ,

    D()> 0 for + < < + ; D() = 0 otherwise. (105)

    The function d(; ) is analogous to the function d() used in Sec. 16 , so inparticularD1 is used again in the definitions of the nondimensional variables.

    We shall direct attention to the modifications in the previous treatment

    occasioned by (104). They begin in Sec. 6 . Concentrating on the first termin (36), we see that it must be replaced by

    (1D+ D)(U0+ U1+ ...)z

    z

    .

    Therefore the terms of orders O(1) and O(1) in the revised version of (36)(analog of (100)) are:

    z

    (1D(0(z)) + D(0(z)) + D(0(z))1)U0z+ D(0(z))U1z

    = 0.

    Integrating this with use of the fact that U0z(z) = D(0(z)) = 0 at z =,we obtain

    (1

    D(0(z)) + D(0(z)) + D(0(z))1)U0z+ DU1z = 0. (106)

    TheO(1) term in this equation tells us that

    D(0(z))U0z = 0. (107)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    34/49

    34 Interfacial Dynamics EJDE1995/16

    Now letI=

    {z : 0(z)

    [

    + , +

    ]

    }. It follows from (105) and (107)

    that

    U0z= 0, z I.Moreover since D(0(z)) = 0 for z I, we see that the third term in (106)vanishes. Hence

    D(0(z))U0z+ D(0(z))U1z = 0.

    This tells us (a) that for z I, U0z= 0, so that in fact for all z ,

    U0z 0;and therefore (b) that

    U1z= 0, z I. (108)In Sec. 7 , (44)(48) remain unchanged. In particular, 0(z) = (z) as

    before.Recall (99), which shows the way that our d enters into the energy balance

    equation. In view of that and (104), equation (52) = (100) must be correctedto

    z

    dUz

    + vEz+ dUz 2DzUz+ s

    DUs

    = O(2), (52) = (109)

    and the new version of (53) is

    dUz+ vE+

    dUzdz 2

    DzUzdz+

    s

    DUs

    dz

    +C1(s,t,) = O(2).

    (53) = (110)

    Here C1 = C11 + C12. As before (54), (vE)0 = v0w((z)). Therefore theO(1) and O(1) terms in (110) are

    1D((z)) + D((z))1+ D((z))

    U1z+D((z))U2z = v0w((z))C11.

    But we know from (108) that DU1z= DU1z = 0, so in fact our alternate version

    of (54) is:

    D((z))U2z+ D((z))U1z = v0w((z)) C11. (54) = (111)Let(z) be the characteristic function ofI, i.e. (z) = 0 forz I, and = 1

    otherwise. In view of (108), multiplying (111) by (z) does nothing to the terminU1z, but annihilates the term in D. We so obtain

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    35/49

    EJDE1995/16 P. C. Fife & O. Penrose 35

    U1z= v0 (z)w((z))D((z))

    C11 D

    , (112)

    U1 = v0

    (w/D)dz C11

    (/D)dz+ C2, (113)

    C2 being an integration constant and the arguments ofw and D being (z).From (69), which continues to hold as written, and (113), we have the other

    condition

    0 = Fu((z), 0)

    (z)U1(z) + (0+ v0)1=

    Fu

    (z)

    v0

    w

    Ddz C11

    Ddz+ C2

    + (0+ v0)1,

    hence solving for C2, we get

    C2 = (+p)v0+ 0+ qC11,

    where now

    p(z) =zz0

    w

    Ddz = w

    Dz P/2 + o(1) (z ),

    q(z) =zz1

    Ddz= 1

    Dz Q/2 + o(1) (z ),

    p = pdz

    ,

    q = qdz

    ,

    and , , are the same as before (see (72a), (72b), (18)).From (113), (42), we have

    u1| = v0P/2 C11Q/2 + C2= v0P/2 C11Q/2 + (+ p)v0+ 0+ qC11= 0(+ p

    P/2) + 0+ (q Q/2)C11.

    Also from (112), (41), we have

    Dru0| = v02

    C11,

    C11 = v

    0

    2 Dru0|.Hence

    u1|+ n(Dru0)| = mv0+ 0,

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    36/49

    36 Interfacial Dynamics EJDE1995/16

    m = + (p P/2) 12 (q Q/2), n = q Q/2.

    This equation provides the interface condition on analogous to (75). Hencethe analog of (79) is

    u + n(Dru) = mv+ .

    We now construct the analog of (80). For this purpose, we write down theO() terms in (110) (recall DU1z = 0):

    (dUz)2 + (vE)1 + 0

    DU2z+ DU1z

    dz +

    s

    DU1s

    dz + C12= 0. (114)

    We apply (43) to obtain an equation for

    Dru1|, which is given in terms

    of the asymptotic behavior of the first term in (114) according to (43). Forthis, we have to write each of the other 4 terms in (114) in the form A(s, t) +B(s, t)z + o(1) (z ). Only the terms A will be relevant (see (43)).The contribution of (vE)1, by (42), is (ve)1|, and the last integral in (114)is bounded by virtue of the compact support ofD. By (111), the first integralcan be expressed by means of (59) and (61) as

    v0

    w((z))dz C11z= v0 (wz P/2) C11z+ o(1) (z ).

    Using all of these facts, we obtain

    Dru

    1|+ (ve)1 0v0P/2 +

    z0

    sD((z))sU1(z, s, t)dz+ C12 = 0.

    We take the difference between the upper and the lower signs and recall thatD((z)) does not depend on s, to obtain

    (Dru)1

    + [(ve)1] 0v0P+

    D((z))2sU1(z, s, t)dz= 0. (115)

    From (113) and the definitions ofp, q, we may write

    U1 = v0M1(z; , , ) + 0+ M3(z)(Dru0)|,

    M1(z) = (p(z) p) 12

    (q(z) q) + ,

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    37/49

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    38/49

    38 Interfacial Dynamics EJDE1995/16

    It is easily shown that our present revision of this problem with (118) re-

    placing (103c) has the same properties (119). In fact

    0 =

    D

    u Du= D

    D |u|2

    u

    Dru

    u

    v + bss

    =

    v + bss

    =

    v+ b

    ss= dL

    dt b

    (s)2

    dLdt

    ,

    hencedL

    dt 0.

    The other relation in (119) is derived in the same manner.

    18 Multicomponent order parameter.

    It is commonplace to describe the local microscopic or macroscopic state of acrystalline material by means of a multicomponent order parameter (see, e.g.,[IS] and recent papers [Lai], [BB]).

    Going further, we remark that in some densityfunctional theories (e.g.[Ha]), the microscopic probability distribution of atoms at any location is de-scribed by a number density function. This function provides detailed informa-tion about the degree to which, and sense in which, the material is ordered atthat location. It is therefore an infinitedimensional generalization of a scalarorder parameter. One way to extract a scalar order parameter from such atheory was described in [PF1], namely artificially to restrict the allowed densityfunctions to a onedimensional subspace of function space. Then the scalarserves to designate locations on that 1D subspace, which can be pictured asa straight line. In the resulting phasefield model, the transition from liquidto solid across an interface corresponds to the density function changing whilerestricted to that line, whereas in the unrestricted model, it may change alongsome other curve in function space. In principle, this discrepancy could be par-tially remedied by restricting to a higher dimensional subspace, in which case

    we would be dealing with several order parameters.We sketch now how our analysis of interfacial motion can be extended to thecase when the order parameter has m components,

    = (1, m). (120)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    39/49

    EJDE1995/16 P. C. Fife & O. Penrose 39

    To avoid compounding complications, we treat D, c, and as constants, and

    continue to operate in 2-dimensional physical space. We consider only first ordertransitions.

    As mentioned in Sec. 1 , the system (2), (3) was derived in [PF1] as agradient system with respect to the entropy functional displayed following (3).

    Thus, the right side of (3) is S

    .

    Note that 1Tf(,T)

    = s(,e)

    .

    The gradient term 12

    1||2 in that functional was chosen as the simplestnegative definite quadratic function of (/x1,/x2). In the casewhen has several components, the analogous term must still be a negativedefinite function of, and we still choose it to be a quadratic form. Wedenote it by1

    2Q(). Thus the more general entropy functional is

    S[, e] = s((x), e(x))

    1

    2

    Q(

    ) dx. (121)

    We have considerable latitude in choosing Q, including the possibility ofmaking it anisotropic. Even with one order parameter, anisotropy can alterna-tively be modelled [MW] by making 1 depend on the direction of, but ifthis dependence is not such that 1()||2 continues to be a quadratic form,the Laplacian in (3) is replaced by a quasi-linear second order partial differen-tial operator whose coefficients depend discontinuously on at places where= 0.

    The most general form Q, in the case ofm components, is

    Q() =1ijkaijkikj,wherei = /xi and the as form a positive definite array and are symmetric

    in the pair (i, j) (which run from 1 to 2) as well as the pair (k, ) (which runfrom 1 to m). The normalizing parameter 1 is chosen so that

    MinXaijkki

    j = 1,

    whereX=

    ki :

    k,i |ki |2 = 1

    .

    Our new assumptions A1 A4 are analogous to those given before in section2:

    A1. As before, f(, T) has two and only two local minima with respect to, which are now m-tuples denoted as before by h(T). We can no longerorder them or speak of a single intermediate maximum. (The following can bepartially extended to the case when there are more than two minima; we do notpursue this.)

    A2. Again, for a firstorder transition, equation (5) holds if and only ifT=T0 (T, T+).

    Set = h(T0).The inner layer equation generalizing (20) and (44) is the system of equa-

    tions

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    40/49

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    41/49

    EJDE1995/16 P. C. Fife & O. Penrose 41

    (B())k =ij

    aijkij,

    (C())k =ij

    aijk(ij+ ji) .

    Notice that

    dA()/d= C(), dC()/d= 2(B() A()) (125)From (124), we see that the inner profile is governed by the following system

    of differential equations to replace (44):

    A()0zz + F(0, 0) = 0, 0(

    ) =

    . (126)

    This is (122), (123) with A depending on ; we again denote the solution by(z, ).

    Also, in place of (69), we obtain the following system from the O() termsin (124); here we use the fact that s = , and let primes denote /z:

    L1 = Fu((z, ), 0)U1 0B()(z, ) 0C()(z, ) v0(z, ),(127)

    L is the self-adjoint ordinary differential operator defined by

    L A() G(z, ),and the symmetric matrix functionG(z, ) is the Hessian of the function

    1

    cT0f(, T

    0)

    with respect to , evaluated at = (z, ).Differentiating (126) with respect to z, we find that

    L = 0; (128)

    taking the (L2)m scalar product of (127) with , we find a necessary conditionfor solvability:

    Fu((z, ), 0)U1, + () + v, = 0, (129)where

    () = B , C, . (130)

    From this point on the analysis proceeds as before. The basic first orderouter approximation satisfies a free boundary problem like (77) - (80) withdifferent coefficients of and v in theO() terms. These coefficients depend on

    . For example, in place of in (79), we have ()

    , and in place of, we have, /1(), where now

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    42/49

    42 Interfacial Dynamics EJDE1995/16

    1 = 1() = A(), . (131)The various limiting problems are obtained as before. Of special interest

    is the Mullins-Sekerka problem (89), in which the coefficient of in (89b) isreplaced by the-dependent coefficient given above. A similar statement is trueof the motion-by-curvature problem (90), in which the present analog will havea coefficient of proportional to . Very probably the sign of this coefficient,which is governed by the sign of, determines whether these problems are wellposed. For this reason, it is of some interest to obtain a simpler expression for.

    In their treatment of a different phase-field model with scalar order parame-ter but interfacial energy a given arbitrary function of , McFadden et al [MW]obtained an expression, which in our notation would be

    () =1() + 1 (). (132)

    This also holds in our framework, of course, when m = 1, as can be verified by acalculation based on our notation. However, this calculation depends very muchon the matrices A, B, Cbeing able to commute with one another. It appearsdoubtful that the expression (132) will continue to hold in the multi-componentcase.

    Appendix. Derivation of (31)

    Let us represent points on the curve (t) by the position vectorR(, t), whereis a parameter on defined as follows. When t = 0 it is arc length from somechosen point. When t increases, the point R(, t) (for fixed ) has trajectory

    normal to (t) at eacht. We will denote the unit tangent to (t) byT(, t) andthe unit normal in the direction of increasing r byN(, t).The normal velocity v(, t) is defined by

    Rt= v(, t)N(, t). (133)

    Let be arc length on , and

    (, t) =

    (, t). (134)

    Att = 0, we have chosen= , so that

    (, 0) = 1. (135)

    In all of the following, we suppose (t) is regular; and in particular that itsradius at curvature is bounded away from zero.

    Consider the point x(r,,t) represented by

    x(r,,t) = R(, t) + rN(, t). (136)

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    43/49

    EJDE1995/16 P. C. Fife & O. Penrose 43

    For fixedr and t the curvex(r,,t) will be parallel to (t) at a distancer. On

    this new curve, we can find the relation between arc length (which we continueto denote by) and the parameter. First, taking the differential of (136) witht and r fixed, we obtain:

    dx= (R+ rN)d.

    But by definition we also have dx = T d. Therefore

    T d= (R+ rN)d. (137)

    We also have from (134)R =T , (138)

    and sinceT N= 0,

    T N+ T N = 0. (139)The curvature(, t) of (t) is given by

    N= T = T

    (, t) 1T. (140)Hence from (139), (140) and the fact that N is in the direction ofT(note that

    |N|2 =N N), we have

    N = T. (141)

    From (137), (138), and (141) we now obtain

    =(2 + r). (142)

    SinceRt and R are in the directions N and Trespectively, we have

    Rt(, t) R(, t) = 0.

    Differentiating this equation with respect to t, we obtain

    Rtt R+ Rt Rt = 0. (143)

    But from (133)Rtt= vtN+ vNt, (144)

    Rt = vN+ vN; (145)

    So from (143)(145), (133), (138), we have

    Nt= 1vT. (146)If we setdx = 0 in (136) and use (138), (133), (141), (146), we find:

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    44/49

    44 Interfacial Dynamics EJDE1995/16

    rt

    = v, (147)and

    t =

    rv2(2 + r)

    . (148)

    Now recall the coordinate system r(x, t), s(x, t) defined in section 4 (wesuppressdependence). This r(x, t) is that in (136). At t = 0, this s coincideswith the previous . So at t = 0, we may replace with s in (142), set = 1,and obtain

    |s(x, 0)| = s

    = (2 + r(s, 0))1.

    Howerver, if t = t0 is any other value of t, we can reset the clock in theoriginal coordinate system (r,,t) so that time starts att0, and make the sameconclusion. Therefore in general we will have

    |xs(x, t)| = (2 + r(s, t))1. (149)In the same way, we find from (147), (148)

    r

    t(x, t) = v(s(x, t), t), (150)

    s

    t(x, t) =

    rvs(s, t)

    1 + r(s, t). (151)

    In these expressions, we have taken = 1 because of (135) and the change oftime setting.

    Along with (149), we also have the obvious relation

    |xr(x, t)| = 1. (152)We have established (31a) and part of (31c). Let us now calculate2r and

    2s. We have, for any domain ,

    2xr(x, t)dx=

    nr(x, t)d, (153)

    wheren is the normal derivative andd is arc length on .Let be the curvilinear rectangle (in the r, s coordinate system) shown in

    the diagram, bounded by sides L1 L4.Asnr = 0 on L2 andL4, we get:

    L1L3nrd= |L3| |L1|.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    45/49

    EJDE1995/16 P. C. Fife & O. Penrose 45

    !

    !

    "

    "

    (r0, s0)

    (r0+ dr, s0+ ds)

    (t)

    L1

    L4L3

    Figure 2: The domain

    OnL1 andL3 we have (149), so that

    |L1| (2 + r0(s0, t))ds,

    |L3| (2 + (r0+ dr)(s0, t))ds,

    so that

    2xr(x, t)dx (s0, t)drds.

    But the left side of this equation, to lowest order is 2r(s0, t) times the area of, and this area is approximately

    || |L1|dr= (2 + r0(s0, t))drds.

    Thus

    2r(x, t) = (s, t)1 + r(s, t)

    . (154)

    In a similar way, we have

    2xs(x, t)dx=L4

    (2 + r(s0+ ds,t))1dr

    L2

    (2 + r(s0, t))1dr.

    Now writing(s0 + ds,t) = (s0, t) + dand integrating betweenr0 and r0 + dr,we obtain to lowest order,

    2s(x, t) = rs(2 + r)3

    .

    This establishes (31b,c).

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    46/49

    46 Interfacial Dynamics EJDE1995/16

    Asknowledgements. Supported by National Science Foundation under Grants

    DMS 8901771 and 9201714, and by the Science and Engineering Research Coun-cil. Part of this research was done during a workshop sponsored by the Inter-national Centre for Mathematical Sciences at HeriotWatt University. The au-thors are grateful to John Cahn for informative discussions, particularly aboutthe possibility of terms involving the Laplacian of the curvature, as in Section17, and to Cecile Schreiber for carefully checking the details of an earlier version.

    References

    [ABC] N. Alikakos, P. Bates, and X. Chen, Convergence of the CahnHilliardequation to the HeleShaw model, Arch. Rat. Mech. Anal.128, 165-205(1994).

    [AC] S.M. Allen and J.W. Cahn, A microscopic theory for antiphase bound-ary motion and its application to antiphase domain coarsening, ActaMetallurgica27, 10851095 (1970).

    [AP1] H.W. Alt and I. Pawlow, A mathematical model of dynamics of nonisothermal phase separation, Physica D, in press.

    [AP2] H. W. Alt and I. Pawlow, Models of nonisothermal phase transitionsin multicomponent systems, Part I: Theory. preprint.

    [BB] B. Boettinger, R. Braun, J. Cahn, J. McFadden, and A. Wheeler,Phase-field model of order-disorder transitions in an FCC crystal,

    preprint.

    [Cahn] Personal communication.

    [C1] G. Caginalp, An analysis of a phase field model of a free boundary,Arch. Rat. Mech. Anal. 92, 205245 (1986).

    [C2] G. Caginalp, Stefan and HeleShaw type models as asymptotic limitsof the phase field equations, Phys. Rev. A 39, 58875896 (1989).

    [C3] G. Caginalp, PenroseFife modification of solidification equations hasno freezing or melting, Appl. Math. letters5, 9396 (1992).

    [CC] G. Caginalp and X. Chen, Phase field equations in the singular limit

    of sharp interface problems, in On the Evolution of Phase Boundaries,IMA Vol. of Mathematics and its Applications, Springer-Verlag, 1992.

    [CF] G. Caginalp and P.C. Fife, Dynamics of layered interfaces arising fromphase boundaries, SIAM J. Appl. Math.48, 506518 (1988).

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    47/49

    EJDE1995/16 P. C. Fife & O. Penrose 47

    [CA] J.W. Cahn and S.M. Allen, A microscopic theory for antiphase bound-

    ary motion and its experimental verification in Fe-Al alloy domaingrowth kinetics, Jour. de Physique, Colloque C7, 54 (1977).

    [CEN] J.W. Cahn, C.N. Elliott, and A. NovickCohen, The Cahn-Hilliardequation with a concentration dependent mobility: motion by minusthe Laplacian of the mean curvature, preprint.

    [CT] J.W. Cahn and J.E. Taylor, Surface motion by surface diffusion, ActaMetall, in press.

    [Chen1] Xinfu Chen, Generation and propagation of interface in reactiondiffusion equations, J. Diff. Equations 96, 116141 (1992).

    [Chen2] Xinfu Chen, Hele-Shaw problem and area preserving curve shortening

    motion, Arch. Rat. Mech. Anal. 123, 117-151 (1993).

    [Chen3] Xinfu Chen, Personal communication.

    [Chm] A. Chmaj, On the existence of a heteroclinic solution of a phase tran-sition Hamiltonian system: a variational approach, preprint.

    [CP] P. Constantin and M. Pugh, Global solutions for small data to theHele-Shaw problem, preprint.

    [F1] P. C. Fife, Mathematical Aspects of Reacting and Diffusing Systems,Lecture Notes in Biomathematics 28, SpringerVerlag, New York, 1978.

    [F2] P. C. Fife, Dynamics of Internal Layers and Diffusive Interfaces,

    CBMSNSF Regional Conference Series in Applied Mathematics #53,Soc. Ind. Appl. Math, Philadelphia (1988).

    [F3] P.C. Fife, Pattern dynamics for parabolic P.D.E.s, preprint.

    [F4] P. C. Fife, Models for phase separation and their mathematics, inNonlinear Partial Differential Equations with Applications to Patterns,Waves, and Interfaces, M. Mimura and T. Nishida, eds., pp. 183-212,KTK Sci. Pubs., Tokyo (1992).

    [Fix] G. Fix, Phase field methods for free boundary problems, in: FreeBoundary Problems, A. Fasano and M. Primicerio, eds., 589589, Pit-man, London (1983).

    [Ha] A.D.J. Haymet, Theory of the equilibrium liquidsolid transition, Ann.Rev. Phys. Chem 38, 89108 (1987).

    [He] C. Herring, in Structure and Properties of Solid Surfaces, R. Gomerand C. S. Smith, eds., Univ. of Chicago Press, Chicago (1952).

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    48/49

    48 Interfacial Dynamics EJDE1995/16

    [IS] Yu. A. Izyumov and V. N. Syromyatnikov, Phase Transitions and Crys-

    tal Symmetry, Kluwer Acad. Pubs., Dordrecht, 1990.

    [K] R. Kobayashi, Bull. Japan Soc. Ind. Appl. Math. 1, 22 (1991).

    [L] J. S. Langer, Models of pattern formation in firstorder phase transi-tions, pp. 164186 in Directions in Condensed Matter Physics, WorldScience Publishers (1986).

    [Lai] Theory of ordering dynamics forC u3Au, Phys. Rev. B 41, 9239-9256(1990).

    [MW] G. B. McFadden, A. A. Wheeler, R. J. Braun, S. R. Coriell, and R.F. Sekerka, Phase-field models for anisotropic interfaces, Phys. Rev. E,48, 2016-2024 (1993).

    [MS] W.W. Mullins and R. F. Sekerka, Morphological stability of a particlegrowing by diffusion or heat flow, J. Appl. Phys. 34, 323329 (1963).

    [MSc] P. de Mottoni, and M. Schatzmann, Geometrical evolution of developedinterfaces, Trans. Amer. Math. Soc., in press.

    [P] R.L. Pego, Front migration in the nonlinear CahnHilliard euqation,Proc. Roy. Soc. London A 422, 261278 (1989).

    [PF1] O. Penrose and P.C. Fife, Thermodynamically consistent models ofphasefield type for the kinetics of phase transitions, Physica D 43,4462 (1990).

    [PF2] O. Penrose and P.C. Fife, Comment on Caginalps paper PenroseFifemodification of the solidification equations has no freezing or melting,App. Math. Letter 5, 99 (1992).

    [PF3] O. Penrose and P.C. Fife, On the relation between the standard phasefield model and a thermodynamically consistent phasefield model,Physica D 69, 107113 (1993).

    [RSK] J. Rubinstein, P. Sternberg, and J.B. Keller, Fast reaction, slow diffu-sion and curve shortening, SIAM J. Appl. Math. 49, 116133 (1989).

    [S] P. Sternberg, The effect of a singular perturbation on nonconvex vari-ational problems, Arch. Rat. Mech. Ana. 101, 209-260 (1988).

    [St] B. Stoth, A model with sharp interface as limit of phase-field equationsin one space dimension, preprint.

    [St2] B. Stoth, The Stefan problem coupled with the Gibbs-Thompson law asa singular limit of the phase-field equations in the radial case, preprint.

  • 8/11/2019 INTERFACIAL DYNAMICS for THERMODYNAMICALLY CONSISTENT PHASE-FIELD MODELS with NONCONSERVED ORDE

    49/49

    EJDE1995/16 P. C. Fife & O. Penrose 49

    [T] L. Truskinovsky, Kinks versus shocks, in Shock Induced Phase Transi-

    tions and Phase Structures in General Media (R. Fosdick et al, eds.)SpringerVerlag, Berlin (1992).

    [UR] A.P. Umantsev and A.L. Roitburd, Nonisothermal relaxation in non-local media, Sov. Phys. Solid State 30, 651 (1988).

    [WS] SL. Wang, R.F. Sekerka, A.A.Wheeler, B.T. Murray, S.R. Coriell, R.J.Braun, and G.B. McFadden, Thermodynamicallyconsistent phasefield models for solidification, Physica D 69, 189-200 (1993).

    Paul C. Fife

    Mathematics Department

    University of Utah

    Salt Lake City, UT 84112

    U.S.A.

    E-mail: [email protected]

    Oliver Penrose

    Mathematics Department

    Heriot-Watt University

    Riccarton, Edinburgh, EH14 4AS

    U.K.

    E-mail: [email protected]