Research article Topical Sections

Dynamic boundary conditions in the interface modeling of binary alloys

  • We study the initial boundary value problem with dynamic boundary conditions to the Penrose-Fife equations with a 'memory e ect' for the order parameter and temperature time evolutions. The dynamic boundary conditions describe the process of production and degradation of surface crystallite near the walls, which confine the disordered binary alloy at a nearly melt temperature during the fast cooling process. The solid-liquid periodic distributions, which were obtained in 1D case, represent asymptotically periodic piecewise constant spatial-temporal impulses in a long time dynamics. It is confirmed that, depending on parameter values, the total number of discontinuity points of such periodic impulses can be finite or infinite. We refer to such wave solution types as relaxation or pre-turbulent, respectively. These results are compared with experimental data.

    Citation: Igor B. Krasnyuk, Roman M. Taranets, Marina Chugunova. Dynamic boundary conditions in the interface modeling of binary alloys[J]. AIMS Mathematics, 2018, 3(3): 409-425. doi: 10.3934/Math.2018.3.409

    Related Papers:

    [1] Luigi C. Berselli, Davide Catania . A note on the Euler–Voigt system in a 3D bounded domain: Propagation of singularities and absence of the boundary layer. AIMS Mathematics, 2019, 4(1): 1-11. doi: 10.3934/Math.2019.1.1
    [2] Hyun Geun Lee, Youngjin Hwang, Yunjae Nam, Sangkwon Kim, Junseok Kim . Benchmark problems for physics-informed neural networks: The Allen–Cahn equation. AIMS Mathematics, 2025, 10(3): 7319-7338. doi: 10.3934/math.2025335
    [3] Koichi Takahashi . Incorporating a tensor in the effective viscosity model of turbulence and the Reynolds stress. AIMS Mathematics, 2018, 3(4): 554-564. doi: 10.3934/Math.2018.4.554
    [4] Dieunel DOR . On the modified of the one-dimensional Cahn-Hilliard equation with a source term. AIMS Mathematics, 2022, 7(8): 14672-14695. doi: 10.3934/math.2022807
    [5] Álvaro Abucide, Koldo Portal, Unai Fernandez-Gamiz, Ekaitz Zulueta, Iker Azurmendi . Unsteady-state turbulent flow field predictions with a convolutional autoencoder architecture. AIMS Mathematics, 2023, 8(12): 29734-29758. doi: 10.3934/math.20231522
    [6] Narcisse Batangouna . A robust family of exponential attractors for a time semi-discretization of the Ginzburg-Landau equation. AIMS Mathematics, 2022, 7(1): 1399-1415. doi: 10.3934/math.2022082
    [7] Feng Zhou, Hongfang Li, Kaixuan Zhu, Xin Li . Dynamics of a damped quintic wave equation with time-dependent coefficients. AIMS Mathematics, 2024, 9(9): 24677-24698. doi: 10.3934/math.20241202
    [8] Grace Noveli Belvy Louvila, Armel Judice Ntsokongo, Franck Davhys Reval Langa, Benjamin Mampassi . A conserved Caginalp phase-field system with two temperatures and a nonlinear coupling term based on heat conduction. AIMS Mathematics, 2023, 8(6): 14485-14507. doi: 10.3934/math.2023740
    [9] Wafaa B. Rabie, Hamdy M. Ahmed, Taher A. Nofal, E. M. Mohamed . Novel analytical superposed nonlinear wave structures for the eighth-order (3+1)-dimensional Kac-Wakimoto equation using improved modified extended tanh function method. AIMS Mathematics, 2024, 9(12): 33386-33400. doi: 10.3934/math.20241593
    [10] Quan Zhou, Yang Liu, Dong Yang . Global attractors for a class of viscoelastic plate equations with past history. AIMS Mathematics, 2024, 9(9): 24887-24907. doi: 10.3934/math.20241212
  • We study the initial boundary value problem with dynamic boundary conditions to the Penrose-Fife equations with a 'memory e ect' for the order parameter and temperature time evolutions. The dynamic boundary conditions describe the process of production and degradation of surface crystallite near the walls, which confine the disordered binary alloy at a nearly melt temperature during the fast cooling process. The solid-liquid periodic distributions, which were obtained in 1D case, represent asymptotically periodic piecewise constant spatial-temporal impulses in a long time dynamics. It is confirmed that, depending on parameter values, the total number of discontinuity points of such periodic impulses can be finite or infinite. We refer to such wave solution types as relaxation or pre-turbulent, respectively. These results are compared with experimental data.


    1. Introduction

    This paper is a continuation of the previously published results [12], where the authors studied a boundary value problem for a system of Allen-Cahn, Cahn-Hilliard and heat transfer (with account of latent heat) equations with no-flux boundary conditions. This system models distributions of a conserved order parameter (u), non-conserved one (v) and temperature (θ). It was shown that there exists a solution of the problem which consists of long-time oscillating functions. Such solutions describe experimentally observed fluctuations of the order parameter, concentration and temperature in confined binary alloys or binary polymer mixtures. In the present paper, we consider the following system:

    τuutt+ut=[ω1Q(u,v)(Fu(u,v,θ)ε2uxx)x]x, (1)
    τvvtt+vtστvθt=ω2Q(u,v)(ε2vxxFv(u,v,θ))+σθ, (2)
    τθθtt+λτθvtt+θt+λvt=Dθxx (3)

    with the dynamic boundary conditions:

    ut(k,t)=Nk[u(k,t)], vt(k,t)=Gk[v(k,t)], θt(k,t)=Υk[θ(k,t)], k=0,l. (4)

    All parameters for the problem (1)–(4) will be given later in Section 2. We will refer to (1) as modified Cahn-Hilliard equation and to (2), (3) as modified Penrose-Fife equations.

    Unlike the conditions assumed for the model studied in [12], in the present paper we take into account 'memory effects' for fluxes and we also implement dynamic boundary conditions to capture a feedback effect. As noted in [21], "Feedback processes are fundamental in all exact sciences. In fact, they were first introduced by Sir Isaac Newton and Gottfried W. Leibniz some 300 years ago in the form of dynamic laws". It turns out that a general melting process exhibits 'surface' order-disorder transition, so that wave perturbations with feedback become dominant in the organization of bulk structures of waves. In this sense, one can talk about surface-induced structures of relaxation, pre-turbulent, or turbulent types in real physical systems with the surface feedbacks. The feedback machine consists of the following components: input, output, control unit, processing unit, and one main processor, which are all connected by four transmission lines (see [21,Figure 1.3]). One-step machine algorithms can be characterized by the iterations u(t+1)=f[u(t)], where f:II is a non-linear function and I is some bounded interval. For example, we can consider a well-known logistic map f:uau(1u), where a[0,4] is the bifurcation parameter, and I:=[0,1]. In our situation, we propose to control behaviour of solution near the boundary by the dynamic boundary conditions, i. e. by using a special choice of inputs and outputs. From experimental point of view, the dynamic boundary conditions describe a process in an alloy near the surface where bubbles propagate into a bulk. These surface bubbles penetrate into the bulk along trajectories which are close to characteristics of correspondent linear hyperbolic equations. Using dynamic boundary conditions, we can reduce our original PDE problem to a system of difference equations. In some particular cases, asymptotic properties of solutions for this system can be deduced from iterations of a one-dimensional map (for example, quadratic or logistic).

    The evolution process of the conserved order parameter u will be described by the modified Cahn-Hilliard (mCH) equation that was originally introduced by Galenko et al. [16]. This equation models the non-Fickian diffusion of the binary alloys in so-called tau-approximation, which describes the 'memory' of the alloy. Then the mCH equation follows from the relation

    τuJt(,t)+J(,t)=M(u,θ)(Fu(u,ux)ε2uuxx)x (5)

    with the non-Fickian diffusion flux J, and θ=1T/Tm, where Tm is the melt temperature of crystallization on front of liquid phase at a neighbourhood of the disordered state, u is one component of a binary alloy. Here, F:=F(u,ux) is the free energy, ε2u=2F0ξ2b, F0 is the dimensionless energy of interaction between atoms of A and B types, ξb is the characteristic length, M is the mobility of atoms, and τu is the relaxation time. The mCH equation has simple physical meaning. Namely, if the front velocity for crystalline phase is small enough then the evolution of wetting phase is described by the classical Cahn-Hilliard (CH) equation. If this velocity is large enough then the crystalline phase is described by the mCH equation because the non-Fickian flow corresponds to a change of the front velocity due to a time delay. In the last case, we have to replace the CH equation with the mCH equation. In [16], it is shown that the decomposition in binary systems can be described by the mCH equation in the following cases: (a) local in time dynamics; (b) large characteristic velocities; (c) large gradients of the concentration; (d) deep super-cooling.

    Note that unlike the modified Cahn-Hilliard equation the Penrose-Fife type equations are very difficult to solve by using Onsager thermodynamic formalism, because one have to deal with the non-conserved order parameter and temperature which are coupled by time derivatives, and as a result, the problem does not allow a variational formulation. According to [7,Eq. (73),(74),p.113]), the system of Penrose-Fife equations is written as:

    γϵ2ϕt=ϵ2(1+τ(TTm))ϕxxFϕαβ2ϵ2Tϕx(1/T)x, (6)
    et=(KTx)x+β2ϵ2[ϕtϕxxα(ϕxϕt)x]. (7)

    where ϕ is order parameter, e is internal energy, T is temperature, Tm is a melting temperature, K is the thermal conductivity, F=F(T,ϕ) is the local density of the Helmholtz free energy, β2=1τTm, γ is the basic order parameter kinetic coefficient, τ and ϵ are some dimensionless parameters. These equations are reduced to the well-known model A if α=0, and to the model B if α=1. The A, B models are both of Ginsburg-Landau type and hence, are examples of phenomenological models. If we assume that K is constant, e=T+λφ, β=0 and TTm, then equations (6), (7) can be reduced to

    γϵ2ϕt=ϵ2ϕxxfϕ(T,ϕ), (8)
    Tt+λϕt=KTxx. (9)

    Therefore, the system (8)–(9) corresponds to the modified Penrose-Fife equations (2)–(3) with v=ϕ, θ=T, τv=τθ=0, with Q and F linearised about (u,v,θ)=(1/2,0,0). Also, the modified Cahn-Hilliard equation and the modified Penrose-Fife equations describe the non-Fickian processes (see Section 2).

    In [12,15] we have studied a boundary value problem for a system of Allen-Cahn, Cahn-Hilliard and heat transfer equations with Neumann boundary conditions. It was shown that there exists a family of solutions with long-time oscillations. Stationary oscillations for the concentration and the order parameter were known in binary alloy theory but stationary oscillations for the temperature was a new result. What happens with long-time oscillations if instead of the Neumann boundary conditions, one implement non-linear dynamic boundary conditions? In this case we obtain spatial-temporal limit distributions of concentration, order parameter and temperature, depending on the travelling wave variable s=tx/V, where V is the velocity of propagation. It should be mentioned that this result is still holds true if instead of the classical Penrose-Fife equations (see [6,22]) we consider the modified hyperbolic Penrose-Fife equations.

    The main idea of deriving modified equations is the following one: consider a general continuity equation for some quantity u:

    ut=Jx+f, (10)

    where J is a flux, f is the mass force. According to the Fourier's law (or Fick's law), we know that

    J=Dux, (11)

    where D is the diffusion coefficient. If D is a constant then by (10), (11) we arrive to the classical heat equation:

    ut=Duxx+f. (12)

    It is well-known that (12) has infinite speed of support propagation for perturbations but this definitely contradicts experimental results. There are many ways how to approach this paradox. One of possibilities is to take into account inertia effects of the flux. This method leads to the following non-Fickian law:

    J(t+τ)=Dux,

    where τ is some relaxation time. For example, for the most of different types of metals the thermal relaxation time is of the order of picoseconds. Approximating J(t+τ)J(t)+τJt(t), we get the Maxwell-Cattaneo law:

    J(t)+τJt(t)=Dux. (13)

    Differentiating (10) with respect to t and (13) with respect to x, we obtain

    utt=Jxt+ft,  Jx+τJxt=Duxx,

    whence

    τutt=Duxx+Jx+τft.

    From here, due to (10), we obtain the following telegraph equation

    τutt+ut=Duxx+g, (14)

    where g=f+τft. It is well-known that (14) has finite speed of support propagation for perturbations. Hence, the paradox is finally resolved. Obviously, (14) coincides with (12) if τ=0. Following this way in Section 2, we derive a system (1)–(3).

    In the present paper, we show that the process of crystallisation of a melt can be described by the coupled modified Cahn-Hilliard (CH) and Penrose-Fife (PF) equations which describe the alloy's dynamic in a neighbourhood of the equilibrium point a=(u,v,θ)=(1/2,0,1T0/Tm), where Tm is a melting temperature. At the point a this system splits into a linear CH equation and two coupled linear PF equations. In this case the original problem is getting reduced to the study of PF equations only, that describe a process of ordering of atoms' types A and B which are initially in the disordered state which normalized by A+B=1. Then, after cooling of the alloy on the flat walls, the front of crystallization arises at the boundary of a pattern and propagates into the melt. Next, we assume that this process can be formally described by the dynamic boundary conditions (4).

    We are interested in existence of travelling wave type solutions for the linearised system (1)–(3). We use the method of reduction of the boundary value problem to the initial value problem for the system of difference equations with continuous time. These difference equations form hyperbolic dynamical system for which we can to apply the method developed by Sharkovsky (see [29]). By this reduction we show that there exist locally (in a neighborhood of an equilibrium point) oscillating spatia-temporal asymptotic solutions with a finite or an infinite number of discontinuity points on their periods. For example, if one of these boundary functions has at least one internal extremum then wave oscillations of order parameter and temperature arising in the bulk belong to the pre-turbulent type. If one of these functions is monotone on the interval (0,l) with two attractive and one repelling fixed points, then the PF system is bistable. In this case, we obtain asymptotic periodic impulse function with constant amplitudes and the one point of discontinuities on a period (see, Figure 2). Note that these periodic functions are smooth for smooth initial data, excluding the point t=+ (see [29]). Note that discontinuities of smooth solutions across the interface arise when t+. Otherwise, the solutions are smooth at each finite time t>0 under an assumption that initial data of the problem are also smooth. An attractor of the system represents a set of generalized functions (see [29]) which are piecewise constant periodic functions. Solutions can have different sets of discontinuities: finite, infinite, countable or uncountable, and we refer to such solutions as relaxation, pre-turbulent or turbulent type, respectively. We reference a reader to [29,Definition 1.5,p. 154] for details about solutions of relaxation, pre-turbulent and turbulent types. The type of the long time limit of a periodic solution depends on the topological structure of the mapping Φ:R2R2 which is controlled by the dynamic boundary conditions of the problem. The mapping Φ leads to a system of two difference equations with continuous time. This mapping is hyperbolic, and structurally stable with only a finite number of attractive and repelling points in R2.

    Figure 1. The upper pictures illustrate convergence of v(t) to zero and divergence of θ(t) for μ=0.28>1/4, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1. The middle pictures illustrate convergence of v(t) to zero and convergence from above of θ(t) to θ10.15 (dashed line) for μ=1/8<1/4, initial data v(t)=0.02+0.05exp(2t), θ(t)=0.05+0.01exp(t), b=1/2, and l/V=1. The lower pictures illustrate convergence of v(t) to zero and oscillating convergence of θ(t) to θ10.37 (dashed line) for μ=1/2>3/4, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1.
    Figure 2. The upper pictures illustrate convergence of v(t) to zero and convergence of θ(t) to piecewise constant function θ(t)=θ3=1,θ(t)=θ4=0 for 5/4<μ=1<3/4, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1. The lower pictures illustrate convergence of v(t) to zero and convergence of θ(t) to pre-turbulent type solution for 1.401μ<μ=1.3<=5/4, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1.

    Also, we show that this linearised equations (1)–(3) with non-linear boundary conditions (4) accurately describe the surface-induced spatial-temporal structures of wave type which enter into the bulk as 'solitons'. Limit distributions of the order parameter and temperature are periodic piecewise constant functions with a finite or an infinite number of discontinuity points on a period. The topological form of these functions is determined by the topological form of the boundary conditions. Moreover, these functions are elements of attractors of the corresponding dynamical system. The structure of such an attractor also depends on the initial data. Elements of the attractor can be deterministic or random functions (see [14,26,27,28,29]). As noted in [8,p.70], "Despite decades of research, the growth of lamellar eutectic in bulk sample is still not well understood". We consider a mathematical model which can be applied to the study of formation and evolution of spatial-temporal lamellar eutectic structures of relaxation, pre-turbulent and turbulent type (see Figure 3). These solutions describe 'one-dimensional' micro-structures with finite, countable or uncountable boundaries, which arise due to the solidification process.

    Figure 3. The upper pictures illustrate convergence of v(t) to zero and convergence of θ(t) to 3l/V=3 periodic function for 2<μ=1.5<μ, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1. The lower pictures illustrate convergence of v(t) to zero and convergence of θ(t) to turbulent type solution for μ=2, initial data v(t)=0.002+0.005exp(2t), θ(t)=0.005+0.001exp(t), b=1/2, and l/V=1.

    The paper is organized as follows. In Section 2, we formulate the initial boundary value problem for the linear Penrose-Fife equation and the Cahn-Hilliard equation with dynamic boundary conditions and initial data satisfying the smooth fitting conditions at the endpoints x=0 and x=l, where l is the size of a sample. We use the observation that these equations at the equilibrium point can be decomposed as the CH equation, which is independent of order parameter and temperature, and as the coupled Penrose-Fife equations. By this reason, we will study the boundary value problem for the PF equations only. In Section 3, it will be shown that the initial boundary value problem can be reduced to the initial value problem for the system of difference equations with continuous time. Moreover, in Sections 3 and 4, we will consider an example when all functions in boundary conditions are linear except one that is in the boundary condition for temperature. In this case, the problem reduces to the quasi-one-dimensional difference equation in R2, which, respectively, can be analyzed by application of the quadratic map. Additionally, in Section 5 we consider applications to experiments and discuss the physical interpretation of the dynamic boundary conditions.


    2. The problem statement

    In this section, we linearise the modified Cahn-Hilliard equation, the Penrose-Fife equation and the heat transfer equation about the equilibrium (u,v,θ)=(1/2,0,θ0), where θ0 is a critical temperature, and we consider the long-time asymptotic behaviour of solutions. First of all, we derive these modified equations. Hence, let us consider the following system of equations (see [12]):

    ut=[ω1Q(u,v)(Fu(u,v,θ)ε2uxx)x]x,ω1R+, (15)
    vt=ω2Q(u,v)(ε2vxxFv(u,v,θ))+σθ,σR,ω2R+, (16)
    θt+λvt=Dθxx,λR, (17)

    where u is a conserved order parameter, v is a non-conserved order parameter, θ is temperature,

    Q(u,v)=u(1u)(1/4v2)0 (18)

    is mobility, λ is latent heat parameter, ω1,ω2 are positive constants (below, we take ω1=ω2=16). These equations represent the simplest form of the phase field model (see [6,22]). In this case, the Ginzburg-Landau functional can be written as:

    E=l0[F(u,v,θ)+ε22(u2x+v2x)]dx, (19)

    where

    F(u,v,θ)=θ2[G(u+v)+G(uv)]+χu(1u)βv2 (20)

    is a free energy, χ and βR,

    G(s)=slns+(1s)ln(1s) (21)

    is entropy. Equations (16)–(17) are Penrose-Fife equations. Next, if we assume that a characteristic relaxation time of the temperature field is much faster than a relaxation time of concentration and that the heat transfer of both phases is described by the parabolic equation. If a characteristic relaxation time for the order parameter is much smaller than the relaxation time for the temperature then we obtain the parabolic equation for the order parameter (the classic Allen-Cahn equation) and the hyperbolic equation for the temperature in the Penrose-Fife system. Note that if it is not true then both Penrose-Fife equations are of hyperbolic type.

    Next, assume that all fluxes satisfy the non-Fickian generalised law, i. e.

    τuJ1,t+J1+ω1Q(u,v)(Fu(u,v,θ)ε2uxx)x=0, (22)
    τvJ2,t+J2+ω2Q(u,v)(ε2vxxFv(u,v,θ))=0, (23)
    τθJ3,t+J3+Dθx=0, (24)

    where τu,τv,τθ are the corresponding relaxation times. Using the following relations

    ut=J1,x, vt=J2+σθ, θt+λvt=J3,x,

    by (22)–(24) we arrive at

    τuutt+ut=[ω1Q(u,v)(Fu(u,v,θ)ε2uxx)x]x, (25)
    τvvtt+vtστvθt=ω2Q(u,v)(ε2vxxFv(u,v,θ))+σθ, (26)
    τθθtt+λτθvtt+θt+λvt=Dθxx. (27)

    Note that the system (25)–(27) coincides with (15)–(17) when τu=τv=τθ=0. This approach was used in the study of the Cahn-Hilliard equation with delay argument for application to polymer blends with dynamic boundary conditions in [11,14]. These equations describe evolution of distributions with non-Fickian diffusion and represent 'tau-approximation' for 'numerical turbulence'. Moreover, this idea was also used by James Clerk Maxwell for heat transfer equation (see [3,19]). Next, linearising (25)–(27) about the equilibrium point (u,v,θ)=(1/2,0,0), we obtain the following system of equations

    τuutt+ut=ε2uxxxx2χuxx, (28)
    τvvtt+vtστvθt=ε2vxx+2βv+σθ, (29)
    τθθtt+λτθvt+θt+λvt=Dθxx. (30)

    As a result, the linearised Cahn-Hilliard and Penrose-Fife equations are uncoupled, and we can consider the Penrose-Fife equations separately by using the following dynamic boundary conditions (4), where Nk,Gk,Υk:II;k=0,1 are the given smooth functions, I:=[0,l] and θ=1T/Tm. Note that the boundary conditions describe 'probability' density of crystallite injection with feedback into the bulk after cooling below T<Tm. We conclude that, for special initial conditions of exponential type, the attractor of the Penrose-Fife problem contains piecewise constant periodic functions p1(s), p2(s) (see Figure 2), where s=tx/V.


    3. Travelling wave solutions to the initial boundary value problem

    Using that the Cahn-Hilliard equation (28) is uncoupled from the system (29), (30) we can study separately the linearised hyperbolic Penrose-Fife equations:

    τvvtt+vtστvθt=ε2vxx+2βv+σθ, (31)
    τθθtt+λτθvtt+θt+λvt=Dθxx (32)

    coupled with

    vt=G0[v]atx=0,vt=G1[v]atx=l, (33)
    θt=Υ0[θ]atx=0,θt=Υ1[θ]atx=l, (34)

    where τv,τθ are the corresponding relaxation times. We will look for a solution of these equations in the form of travelling waves, namely,

    v(x,t)=v(s),θ(x,t)=θ(s), where s=tx/V. (35)

    Substituting (35) into (31), (32), we get that

    (τvε2V2)v"+v2βv=στvθ+σθ, (36)
    (τθDV2)θ"+θ=λτθv"λv. (37)

    Here, β=χχc is the interaction energy between atoms of type A and B, χc is an energy of decomposition of a disordered phase on two ordered phases. By (36), (37) we deduce that

    α0v+α1v+α2v"+α3v=0 if τθτvDV2, (38)
    ˜α1v+˜α2v"+˜α3v=0 if τθτv=DV2, (39)

    where

    α0=(τvε2V2)(τθDV2), α1=τvε2V2+τθDV2+σλτvτθ,
    α2=1+σλ(τv+τθ)2β(τθDV2), α3=σλ2β,
    ˜α1=τvε2V2, ˜α2=1+σλτθ, ˜α3=σλ2β.

    Next, for simplicity, we will consider the case V2=Dτθ=ε2τv only.


    4. Asymptotic solutions for the Penrose-Fife equations


    4.1. Asymptotic solutions for the order parameter

    If V2=Dτθ=ε2τv then from (38) we obtain that

    σλτvτθv+(1+σλ(τv+τθ))v+(σλ2β)v=0. (40)

    By (40) we deduce that

    v+a0v+a1v=0, where a0=1+σλ(τv+τθ)σλτvτθ, a1=σλ2βσλτvτθ. (41)

    Integrating (41) from s=t to s=tl/V, we arrive at

    v(t)+a0v(t)+a1v(t)=v(tl/V)+a0v(tl/V)+a1v(tl/V), (42)

    whence, taking into account the boundary conditions (33), we get

    G0[v(t)]G0[v(t)]+a0G0[v(t)]+a1v(t)=G1[v(tl/V)]G1[v(tl/V)]+a0G1[v(tl/V)]+a1v(tl/V). (43)

    Let us denote by ˜G0:=G0G0+a0G0+a1Id, ˜G1:=G1G1+a0G1+a1Id, where Id is the identity map. Then from (43) we find the following difference equation

    v(t)=G[v(tl/V)], where G:=˜G10˜G1. (44)

    On the other hand, integrating (41) on s, we arrive at

    v(s)=k1+k2eλ1s+k3eλ2s if a04a10,v(s)=k1+k2ea02s+k3sea02s if a04a1=0 (45)

    kiR, where

    λ1=a0±a204a0a12.

    So, (45) forms admissible class of initial functions for difference equation (44), i. e.

    v0(t)=k1+k2eλ1t+k3eλ2t if a04a10,
    v0(t)=k1+k2ea02t+k3tea02t if a04a1=0

    for all t[l/V,0).

    For example, consider the boundary conditions (33) in the form

    vt=αv at x=0,  vt=f(v) at x=l, (46)

    where f(v) satisfies f(v)(f(v)+a0)=(α(α+a0)+a1)(v2+δ)a1v. Here α(α+a0)+a10 and δ are arbitrary parameters, then (44) is reduced to the logistic equation:

    v(t)=v2(tl/V)+δ. (47)

    It can be shown (see [18,29]) that solutions of this equation tend to 2Nl/V — periodic function p1(t)P+ with a finite or an infinite number of discontinuity points tΓ on a period as t+, where P+ is a set of attractive circles of a map G:II, N is a common multiple of the attractive circles.

    Now, we show what happens, for example, if we linearise boundary condition (33) at a disordered state v=0. By (33) we arrive at

    vt(0,t)=G0[0]+G0[0]v(0,t),  vt(l,t)=G1[0]+G1[0]v(l,t). (48)

    In this case, similar to (44) we get the following linear difference equation:

    v(tl/V)=m1v(t)+m2, (49)

    where

    m1=G0[0](G0[0]+a0)+a1G1[0](G1[0]+a0)+a1, m2=G0[0](G0[0]+a0)G1[0](G1[0]+a0)G1[0](G1[0]+a0)+a1.

    Equation (49) has a general solution

    v(t)=Θ(t)mVlt1+m21m1 if m11, and v(t)=Θ(t)Vlm2t if m1=1, (50)

    where Θ(t) is an arbitrary l/V-periodic function. So, the linearised boundary conditions give us very simple asymptotic behaviour.


    4.2. Asymptotic solutions for temperature

    If V2=Dτθ=ε2τv then by (36), (37) we deduce that

    θ+b1θ+b2v=0, where b1=1+σλτθσλτvτθ, b2=1+2βτθστvτθ. (51)

    Subtracting (51) at s=tl/V from (51) at s=t, we arrive at

    θ(t)θ(tl/V)+b1[θ(t)θ(tl/V)]+b2[v(t)v(tl/V)]=0. (52)

    Taking into account the boundary conditions (34) and equation (44), we get

    Υ0[θ(t)]Υ0[θ(t)]+b1Υ0[θ(t)]=Υ1[θ(tl/V)]Υ1[θ(tl/V)]+b1Υ1[θ(tl/V)]b2(G0G[v(tl/V)]G1[v(tl/V)]). (53)

    Let us denote by ˜Υ0:=Υ0Υ0+b1Υ0, ˜Υ1:=Υ1Υ1+b1Υ1, and ˜G:=G1G0G. Then from (53) we find the following difference equation

    θ(t)=˜Υ10(˜Υ1[θ(tl/V)]+b2˜G[v(tl/V)]). (54)

    On the other hand, integrating (51) on s, we get

    θ(s)=k4eb1sk1b2b1k2b2b1+λ1eλ1sk3b2b1+λ2eλ2s kiR, a04a10,λib1, (55)

    where k1,k2,k3 are from (45). Thus, (55) provides admissible class of initial functions for difference equation (54), i. e. θ0(t)=k4eb1tk1b2b1k2b2b1+λ1eλ1tk3b2b1+λ2eλ2t for all t[l/V,0).

    For example, if b2=0 and Υ:=˜Υ10˜Υ1:II is structurally stable hyperbolic map then again we obtain the same difference equation to (44). It can be shown (see [18,29]) that solutions of this equation tend to periodic piecewise constant function with a finite or an infinite number of discontinuity points on one period as t+. If b20 then we have to consider the coupled system of difference equations.


    4.3. Example

    In the general situation, we do not have any classical theory to apply. Therefore we will study one of simple examples to illustrate some possible scenarios of asymptotic behaviour of solutions. Consider the boundary conditions (33), (34) in the form

    vt=αv at x=0,  vt=βv at x=l, (56)
    θt=γθ at x=0,θt=g(θ) at x=l, (57)

    where g(θ) satisfies g(θ)(g(θ)+b1)=γ(γ+b1)(θ2+μ). Here γ{0,b1}, α(α+a0)+a10 and μ are arbitrary parameters. Then the system of difference equations (44), (45) are reduced to

    θ(t)=θ2(tl/V)+av(tl/V)+μ,  a=b2(αβ)(αβa1)γ(γ+b1)(α(α+a0)+a1), (58)
    v(t)=bv(tl/V),  b=β(β+a0)+a1α(α+a0)+a1. (59)

    Note that α=β then (58), (59) reduces to the following uncoupled system:

    θ(t)=θ2(tl/V)+μ,  v(t)=v(tl/V). (60)

    If |b|<1 then the map

    ˆfμ:(θ,v)(θ2+v+μ,bv) (61)

    describes all trajectories of the dynamical system attracted by a line v=0. Thus asymptotic behaviour of equation (58) is determined by properties of the one-dimensional logistic map

    fμ:θθ2+μ (62)

    of the line v=0 mapped into itself. For example, if μ>1/4 then θ(t)+ as t (see the upper picture on Figure 1) because there are no any fixed points. For μ=1/4, we have the saddle-node type fixed point (1/2,0). For μ<1/4, the map ˆfμ has two fixed points (θ1,0)=(114μ2,0), (θ2,0)=(1+14μ2,0). As a result, we have a saddle-node bifurcation at μ=1/4.

    If μ(3/4,1/4) then the point (θ1,0) is stable node type, and the point (θ2,0) is unstable saddle type (see Figure 1). The attractive region of the point (θ1,0) is an open unbounded region W at the plane (θ,v) with a boundary which contains the saddle point (θ2,0) of codimension one and separatrix of this point, and also countable set of curves. These curves are pre-image of the separatrix for iterations ˆfnμ,n=0,1,.... The limit solution fΔ(θ,v) is (1) (θ1,0) as (θ,v)W; (2) (θ2,0) as (θ,v)W; (3) (+,0) as (θ,v)¯W.

    If μ(5/4,3/4) then the points (θ1,2,0) are saddle type and the map has attractive circle of period 2 formed by points (θ3,0)=(134μ2,0), (θ4,0)=(1+34μ2,0). In this case, we have solutions of relaxation type (see, Figure 1).

    If μ[μ,5/4], where μ=1.401, then the map fμ has circles of periods 1,2,22,..,2n (ones of each periods!), where 2n=n(μ)+ as μμ. A circle of a period 2n is attractive but another circles are repelling. In this case, a solution tends to a piecewise constant 2Nl/V-periodic function as t+, excluding the solution θ(t)=θ1. The limit function has, at least, a countable set of points of discontinuities on a period. A number of oscillations of the limit solutions on each interval (t,t+l/V) tends to a power function as t+. We will call such solutions as solutions of pre-turbulent type (see Figure 2).

    If μ(7/4,μ) then a set of non-wandering points of the map consists from attractive circle of a period 3 and Cantor set which represents closure of a set of points for repelling circles (see [17]). As a result, a solution tends to a 3l/V-periodic function as t+ (see Figure 3).

    If μΛ:=(2,7/4) then bifurcations of solution accompanied by a change of periods with respect to the universal ordering (see [29,30]):

    122223...722522322725232...753. (63)

    The period doubling bifurcations arise with universal velocity ν=4.669 and ones characterised by the universal relations of amplitudes of arising oscillations ρ=2.502. From (63) it follows that for each m exists a map which has a circle of the period m and one has not of circles of periods m<m as mm. For the period doubling bifurcations from (63) we get that 2im for all i0 if m2i,i=0,1,2,... Next, define by μ[n] a least value of a parameter μ for which the map fμ has a circle of a period n. Then for fμ there is the ordering (see [30]):

    μ[1]μ[2]μ[4]...μ[52]μ[32]...μ[5]μ[3]. (64)

    Assume that

    Λnm={μΛ:fμ has a circle of a period (2m+1)2n},m,nZ+, (65)

    and μnm=infΛnm. Then from (64) we arrive at

    μ00μ10μ20μ13μ12μ11...μ03μ02μ01. (66)

    If μ=2 and initial function θ0(t):2θ0(t)2 t[l/V,0) then a solution tends to a function, which values are equal to [2,2] for any t, as t+. It means that a solution of θ(t)=θ2(tl/V)+μ for any t<t" has a number of oscillations with the amplitude [2,2] on interval (t+t,t"+t) as t+. The number of oscillations increases by 2 every time as time interval t increases by l/V and eventually goes to infinity. In this case, we will talk about limit solutions of turbulent type (see Figure 3). In addition, if μ<2 then all solutions become unbounded, except θ(t)=θ1,2.

    Note that the experiment which proves the existence of surface oscillating distributions of temperature has been done by Gao et al. (see, [9,Figure 2a]). The influence of latent heat on formation of surface heat structures in a pattern has been also explained in this article.


    5. Discussion


    5.1. Comparison with experiment

    As an example of possible application of our results, we present some data from an experimental study of ordering (segregation) at the CuAu(100) surface. In [1], it is shown that Au enrichment in the top surface layer persists up to temperature far beyond the bulk order–disorder transition temperature. The segregation, which happens below the bulk order-disorder transition temperature Tm, depends on compositional depth profile that gives information about the binding characteristics of such alloys. So, monotone profile indicates weak pair interaction between the two components but oscillating profile results from stronger pair interaction. Therefore it is useful to perform layer-selective composition analysis in the near-surface region. For example, for low-energy interaction there are no layers but for lager energy there arise three or more surface-layers (see [1,4,10,20,31,32]). The theoretical study by Tersoff [32] for CuAu (100) predicts Au segregation with oscillating segregation depth profile. So, Au rich layers can alternate with Au depleted layers. The average amplitude of these oscillations decreases while temperature increases. This average amplitude, according to Tersoff, decays exponentially into the bulk when value of Tm is above the bulk ordering transition temperature.

    Thus, investigation of surface segregation requires layer composition analysis as a function of temperature. Such information can be obtained by the low-energy ion scattering beyond first layer chemical composition. The Cu and Au concentrations in the top layer are measured with He+ ion scattering. Information about the first and second layer composition is obtained from Na+ scattering spectra (see [1,Figure 2]). The spectra taken on the two different azimuthal directions show that Au is the dominant species in the top layer where both Cu and Au are detected in the second layer. The experimental results are well reproduced by the solutions of relaxation type with a unique point of 'discontinuities' on a period (see, Figure 1). So, the quantitative evaluation is possible.

    The surface composition as a function of temperature range is obtained from quantitative evaluation of both, He+ and Na+ measurements. For this purpose, the Cu and Au concentrations in the surface layers were varied in the simulations until the best agreement between measured and calculated spectra were obtained.

    To make it the crystal must be heated up to temperature of data points. At temperature T<Tm, the rapid cooling with liquid nitrogen must be necessary. Then at temperature T<0C arise mixing in surface layers (or disordered state). The Au concentration at a neighbourhood T=Tm is 0.95 of common concentration Au+Cu=1 and decreases with higher temperature while the Cu concentration increases. As a result, there is the typical graphic form for oscillations of relaxation type with finite points of 'discontinuities' on a period. These oscillations describe the layers which are parallel to the (100) plane of the pattern.

    Note that the remaining small Au concentration in the second layer indicates slight deviation from 'ideal' bulk temperature that qualitatively corresponds to the mathematical results. That is the oscillations are spatial-temporal piecewise constant distributions. Indeed, as noted in [1]: 'The asymmetric development of the Au concentration in the two layers relative to the bulk value XAuVol is also assign of a damped oscillating concentration depth profile, which is similar to the case of Cu3Au'. Thus, these results indicate on existence of a continuous phase transition in the surface region. Next, the asymmetric development of the Au concentration in the two layers which are relative to the bulk value X of the concentration AuVol is a sign of a damped oscillating concentration depth profile that are similar to the case of Cu3Au [5,24].


    5.2. Physical sense of boundary conditions

    At higher temperatures gradual desegregation is observed, i. e. decreasing of the Au concentration in the first layer accompanied by increasing Au concentration in the second layer. The degree of desegregation can be used to estimate the segregation energy ΔH. The Langmuir-McLean relation should be applied [23] to describe temperature dependence of the concentration X of the segregating components as

    Xsurface1Xsurface=Xbulk1XbulkeΔHkBT, (67)

    where kB is the Stefan-Boltzmann constant. For example, if ΔHkBT1 then the solutions of the Penrose-Fife equations are determined in main by the boundary conditions, i. e. by the surface segregation energy. Thus, the cooling of a crystal with impurity leads to the formation of meta-stable states (or a number of clusters) which contain defects. These states are long lived (or meta-stable). Note that the crystallization in the bulk is difficult to identify but the crystallization at the surface can be determined by using of scanning tunneling microscope [5,24].

    Further, let us consider an example of simplest dynamic boundary conditions which are due to surface defects. If T>Tm, then positions of impurities are not correlated. But below Tm defects are correlated so that the rate of change of the order parameter v is proportional to v2. It leads to phenomenon that if T<Tm, then defects are structured in clusters, and one would expect an increase in the number of defect fluctuations. In 1D approximation, the radiuses of clusters and its density may be described by distributions of relaxation type. Such distributions are captured by the boundary condition

    vt=kv2+μ,k,μR (68)

    at the left wall that confines the binary alloy in the liquid state, and the same linear boundary conditions at the right wall. In a similar way, one can construct a special dynamic boundary conditions for the temperature, or even more complex boundary conditions, which are connected the order parameter and temperature in some nonlinear way. To conclude, we note that the dynamic boundary conditions were previously discussed in [13] for the binary alloys. At first, this type of boundary conditions was considered for polymer mixtures by Binder et al. [2,25]. The Cahn-Hilliard equation in 3D geometry with dynamic boundary conditions was studied in [13], where 3D-wave structures were obtained for the unit cube domain. These results can also be applied to the Penrose-Fife equations.


    6. Conclusion

    We consider the self-organization phenomenon in a binary alloy with memory which is in the disordered state and confined by the two flat walls. This problem postulated as an initial boundary value problem for the hyperbolic Penrose-Fife equations with dynamic boundary conditions. Such type model describes evolution of order parameter and temperature in a binary alloy. It is shown that the solutions of the problem can be represented in the form of travelling waves. This allows us to reduce the PDE problem to the initial value problem for two difference equations with continuous time delay. In particular case when these equations have special quadratic form, it is proved that asymptotic solutions satisfy to the Sharkovsky ordering. These solutions have the finite or infinite points of discontinuities on a period. As a result, we get the oscillating solutions for order parameter and temperature. Thus, behaviour of order parameter and temperature about the walls, due to the dynamic boundary conditions, leads to appearance of surface induced spatia-temporal oscillations into a bulk.


    Conflict of interest

    All authors declare no conflicts of interest in this paper.


    [1] R. Beikler, E. Taglauer, Surface segregation at the binary alloy Cu Au (100) studied by low-energy ion scattering, Surface Science, 643 (2016), 138-141.
    [2] K. Binder, S. Puri and H. L. Frisch, Surface-directed spinodal decomposition versus wetting phenomena: Computer simulations, Faraday Discuss, 112 (1999), 103-117.
    [3] A. Brandenburg, P. J. Käpylä and A. Mohammed, Non-Fickian diffusion and tau approximation for numerical turbulence, Phys. Fluids, 16 (2004), 1020-1028.
    [4] H. H. Brongersma, M. Draxler, M. de Ridder, et al. Surface composition analysis by low-energy ion scattering, Surface Science Reports, 62 (2007), 63-109.
    [5] T. M. Buck, G. H. Wheatley, L. Marchut, Order-disorder and segregation behavior at the Cu3Au(001) surface, Phys. Rev. Lett., 51 (1983), 43-46.
    [6] G. Caginalp, An analysis of a phase field model of a free boundary, Arch. Ration. Mech. An., 92 (1986), 205-245.
    [7] C. Charach, P. C. Fife, On thermodynamically consistent schemes for phase field equations, Open Syst. Inf. Dyn., 5 (1998), 99-123.
    [8] C. Ern, W. Donner, A. Rühm, et al. Surface density oscillations in disordered binary alloys: X-ray reflectivity study of Cu3Au(001), Appl. Phys. A, 64 (1997), 383-390.
    [9] J. Gao, V. Bojarevics, K. A. Pericleous, et al. Modeling of convection, temperature distribution and dendritic growth in glass-fluxed nickel melts, Journal Crystall Growth, 471 (2017), 66-72.
    [10] M. J. Harrison, D. P. Woodruff, J. Robinson, Surface alloys, surface rumpling and surface stress, Surface Science, 572 (2004), 309-317.
    [11] I. B. Krasnyuk, Spatial-temporal oscillations of order parameter in confined diblock copolymer mixtures, International Journal of Computational Materials, Science and Engineering, 2 (2013), 1350006.
    [12] I. B. Krasnyuk, R. M. Taranets, M. Chugunova, Long-time oscillating properties of confined disordered binary alloys, Journal of Advanced Research in Applied Mathematics, 7 (2015), 1-16.
    [13] I. B. Krasnyuk, Surface-directed multi-dimensional wave structures in confined binary mixture, International Journal of Computational Materials Science and Engineering, 4 (2015), 1550023.
    [14] I. B. Krasnyuk, Impulse Spatial-Temporal Domains in Semiconductor Laser with Feedback, Journal of Applied Mathematics and Physics, 4 (2016), 1714-1730.
    [15] I. B. Krasnyuk, R. M. Taranets, M. Chugunova, Stationary Solutions for the Cahn-Hilliard equation coupled with Neumann boundary conditiona, Bulletin of the South Ural State University. Ser., Mathematical Modelling, Programming Computer Software (Bulletin SUSU MMCS), 9 (2016), 60-74.
    [16] N. Lecoq, H. Zapolsky and P. Galenko, Evolution of the Structure Factor in a Hyperbolic Model of Spinodal Decomposition, European Physical Journal Special Topics, 179 (2009), 165-175.
    [17] T. Y. Li, J. A. Yorke, Period Three Implies Chaos, The American Mathematical Monthly, 82 (1975), 985-992.
    [18] Y. L. Maistrenko, E. Y. Romanenko, About qualitative behaviour of solutions of quasi-linear differential-difference equations, The research of differential-difference equations, Kiev, Institute Mathematics, 1980.
    [19] J. C. Maxwell, On the dynamical theory of gases, Philos. T. R. Soc. A, 157 (1867), 49-88.
    [20] H. Niehus, W. Heiland, E. Taglauer, Low-energy ion scattering at surfaces, Surface Science Reports, 17 (1993), 213-303.
    [21] H. O. Peitgen, H. Jürgens, D. Saupe, Chaos and Fractals, New Frontiers of Science, Springer-Verlag, New York, 2004.
    [22] O. Penrose, P. C. Fife, Thermodynamically consistent models of phase-field type for the kinetics of phase transitions, Physica D, 43 (1990), 44-62.
    [23] J. du Plessis, Surface Segregation, Series Solid State Phenomena, volume 11, Sci.-Tech. Pub., Vaduz, 1990.
    [24] M. Polak, R. Rabinovich, The interplay of surface segregation and atomic order in alloys, Surface Science Reports, 38 (2000), 127-194.
    [25] S. Puri, K. Binder, Surface effects on spinodal decomposition in binary mixtures and the interplay with wetting phenomena, Phys. Rev. E, 49 (1994), 5359-5377.
    [26] E. Y. Romanenko, A. N. Sharkovsky, Ideal turbulence: attractors of deterministic systems may lie in the space of random fields, Int. J. Bifurcat. Chaos, 2 (1992), 31-36.
    [27] E. Y. Romanenko, A. N. Sharkovsky, From one-dimensional to infinite-dimensional dynamical systems: Ideal turbulence, Ukrainian Math. J., 48 (1996), 1817-1842.
    [28] E. Yu. Romanenko, A. N. Sharkovsky, M. B. Vereikina, Self-stochasticity in deterministic boundary value problems, Nonlinear Boundary Value Problems, Institute of Applied Mathematiccs and Mechanics of the NAS of Ukraine, 9 (1999), 174-184.
    [29] A. N. Sharkovsky, Yu. L. Maistrenko, E. Yu. Romanenko, Difference equations and their applications, Ser. Mathematics and Its Applications, 250, Klüwer Academic, Dordrecht, The Netherlands, 1993.
    [30] A. Sharkovsky, A. Sivak, Universal Phenomena in Solution Bifurcations of Some Boundary Value Problems, J. Nonlinear Math. Phy., 1 (1994), 147-157.
    [31] E. Taglauer, Low-energy ion scattering and Rutherford backscattering, Surface Analysis - The Principal Techniques, 2nd ed., eds. by J.C. Vickerman and I.S. Gilmore, JohnWiley & Sons, Ltd., 269-331, 2009.
    [32] J. Tersoff, Oscillatory segregation at a metal alloy surface: Relation to ordered bulk phases, Phys. Rev. B, 42 (1990), 10965-10968.
  • Reader Comments
  • © 2018 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(4957) PDF downloads(941) Cited by(0)

Article outline

Figures and Tables

Figures(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog