Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Heterogeneous Stefan problem and permafrost models with P0-P0 finite elements and fully implicit monolithic solver


  • We consider heat conduction models with phase change in heterogeneous materials. We are motivated by important applications including heat conduction in permafrost, phase change materials (PCM), and human tissue. We focus on the mathematical and computational challenges associated with the nonlinear and discontinuous character of constitutive relationships related to the presence of free boundaries and material interfaces. We propose a monolithic discretization framework based on lowest order mixed finite elements on rectangular grids well known for its conservative properties. We implement this scheme which we call P0-P0 as cell centered finite differences, and combine with a fully implicit time stepping scheme. We show that our algorithm is robust and compares well to piecewise linear approaches. While various basic theoretical properties of the algorithms are well known, we prove several results for the new heterogeneous framework, and point out challenges and open questions; these include the approximability of fluxes by piecewise continuous linears, while the true flux features a jump. We simulate a variety of scenarios of interest.

    Citation: Lisa Bigler, Malgorzata Peszynska, Naren Vohra. Heterogeneous Stefan problem and permafrost models with P0-P0 finite elements and fully implicit monolithic solver[J]. Electronic Research Archive, 2022, 30(4): 1477-1531. doi: 10.3934/era.2022078

    Related Papers:

    [1] Yujia Chang, Yi Jiang, Rongliang Chen . A parallel domain decomposition algorithm for fluid-structure interaction simulations of the left ventricle with patient-specific shape. Electronic Research Archive, 2022, 30(9): 3377-3396. doi: 10.3934/era.2022172
    [2] Ying Liu, Yanping Chen, Yunqing Huang, Yang Wang . Two-grid method for semiconductor device problem by mixed finite element method and characteristics finite element method. Electronic Research Archive, 2021, 29(1): 1859-1880. doi: 10.3934/era.2020095
    [3] Jun Pan, Yuelong Tang . Two-grid $ H^1 $-Galerkin mixed finite elements combined with $ L1 $ scheme for nonlinear time fractional parabolic equations. Electronic Research Archive, 2023, 31(12): 7207-7223. doi: 10.3934/era.2023365
    [4] Yiyuan Qian, Haiming Song, Xiaoshen Wang, Kai Zhang . Primal-dual active-set method for solving the unilateral pricing problem of American better-of options on two assets. Electronic Research Archive, 2022, 30(1): 90-115. doi: 10.3934/era.2022005
    [5] Noelia Bazarra, José R. Fernández, Ramón Quintanilla . Numerical analysis of a problem in micropolar thermoviscoelasticity. Electronic Research Archive, 2022, 30(2): 683-700. doi: 10.3934/era.2022036
    [6] Yijun Chen, Yaning Xie . A kernel-free boundary integral method for reaction-diffusion equations. Electronic Research Archive, 2025, 33(2): 556-581. doi: 10.3934/era.2025026
    [7] Hongze Zhu, Chenguang Zhou, Nana Sun . A weak Galerkin method for nonlinear stochastic parabolic partial differential equations with additive noise. Electronic Research Archive, 2022, 30(6): 2321-2334. doi: 10.3934/era.2022118
    [8] Wenya Qi, Padmanabhan Seshaiyer, Junping Wang . A four-field mixed finite element method for Biot's consolidation problems. Electronic Research Archive, 2021, 29(3): 2517-2532. doi: 10.3934/era.2020127
    [9] Qian He, Wenxin Du, Feng Shi, Jiaping Yu . A fast method for solving time-dependent nonlinear convection diffusion problems. Electronic Research Archive, 2022, 30(6): 2165-2182. doi: 10.3934/era.2022109
    [10] Changling Xu, Huilai Li . Two-grid methods of finite element approximation for parabolic integro-differential optimal control problems. Electronic Research Archive, 2023, 31(8): 4818-4842. doi: 10.3934/era.2023247
  • We consider heat conduction models with phase change in heterogeneous materials. We are motivated by important applications including heat conduction in permafrost, phase change materials (PCM), and human tissue. We focus on the mathematical and computational challenges associated with the nonlinear and discontinuous character of constitutive relationships related to the presence of free boundaries and material interfaces. We propose a monolithic discretization framework based on lowest order mixed finite elements on rectangular grids well known for its conservative properties. We implement this scheme which we call P0-P0 as cell centered finite differences, and combine with a fully implicit time stepping scheme. We show that our algorithm is robust and compares well to piecewise linear approaches. While various basic theoretical properties of the algorithms are well known, we prove several results for the new heterogeneous framework, and point out challenges and open questions; these include the approximability of fluxes by piecewise continuous linears, while the true flux features a jump. We simulate a variety of scenarios of interest.



    In this paper we are interested in numerical approximation of heat conduction with phase change in heterogeneous and composite materials. We are motivated by important applications in modeling permafrost, human tissue, and phase change materials (PCM) for thermal energy storage, smart textiles and buildings. These applications involve materials with drastically different thermal properties separated by an interface, thus require conservative algorithms. While we focus on mathematical and computational challenges rather than on practical engineering or geophysics scenarios, we aim to develop robust and accurate yet simple algorithms adequate for low regularity of solutions and easily extendable to future multiphysics simulators. Challenges similar to those discussed here occur in simulation of multiphase flow in fractures or rocks of different type [1,2], as well as in other applications which we give in what follows.

    Overview. We focus on two phases: solid and liquid, separated by a free boundary, or by a region in which these phases coexist. In a single material we consider the following nonlinear parabolic equation

    tw+q=f;q=(k(θ)θ); (1.1)

    to be supplemented by appropriate boundary and initial conditions. The model is solved for the temperature θ, the internal energy (enthalpy) density w and the heat flux q. Here f represents heat sources, k is the heat conductivity. The definition of w closes the model. For this definition, we consider one of the following equilibrium relationships

    (ST):wαST(θ),or (1.2a)
    (ST)ϵ:w=αSTϵ(θ),or (1.2b)
    (P):w=αP(θ). (1.2c)

    In particular, in the Stefan problem denoted by (ST), wαST(θ)=cθ+Lχ. Here c is the heat capacity, L is the latent heat, and the water fraction χH(θ) (where H is the Heaviside graph) "translates" the well known Stefan condition prescribing the velocity of the free boundary S between the liquid and solid regions to (1.1) defined over both solid and liquid regions. Its approximation is (ST)ϵ, in which αSTϵ(θ) is some ϵ-dependent single-valued piecewise smooth approximation to αST with Lipschitz constant LαSTϵϵ1; see Section 2 for details. In turn, in permafrost models (P) w=αP(θ) is given by a monotone piecewise smooth Lipschitz function αP(), with some large Lipschitz constant LαP; we give details in Section 5. See also a summary of notation in Table 1.

    Table 1.  Variables and symbols used in this paper. The units are chosen for consistency with the literature; [26] e.g.
    Variable Description Units
    θ Temperature [C]
    w Internal energy (enthalpy) density [J/cm3]
    q Heat flux [J/cm2 s]
    χ Water fraction []
    c Volumetric heat capacity [J/cm3 C]
    k Thermal conductivity [J/cm s C]
    L Latent heat of fusion/melting [J/cm3]
    u Kirchhoff temperature [J/cm s]
    η Porosity []
    χw Unfrozen water content []
    Domains Description Units
    xΩRd Spatial variable in the domain of heat conduction [cm]
    t(0,T) Time variable and interval [s]
    (x,t)Q Space-time cylinder [(cm,s)]
    NMAT Number of different materials []
    Ω(j) Subdomain with material j=1,NMAT []
    Relationship Description
    θ=β(w) Temperature-enthalpy relationship
    wα(θ) Enthalpy-temperature relationship
    u=βK(w) Kirchhoff temperature-enthalpy relationship
    wαK(u) Enthalpy-Kirchhoff temperature relationship
    Superscript Choice of constitutive relationships
    ST Stefan problem
    STϵ Regularized Stefan problem
    P Permafrost model

     | Show Table
    DownLoad: CSV

    Known work. Even in a single material the models (1.1) and (1.2) does not have classical solutions and features challenges to the analysis and approximation due to the nonlinear multi-valued or only piecewise smooth character of α(). Typically, (1.1) is posed in the sense of distributions, and its solutions (θ,w) have low regularity.

    The majority of numerical analysis for (1.1), under homogeneous Dirichlet boundary conditions, is carried out for (ST)ϵ after so-called Kirchhoff transformation which renders (ST) or (ST)ϵ in the equivalent form, and with a new variable u instead of θ

    twΔu=f,u=βK(w).

    Since u (as well as θ) are expected to be continuous, it appears natural to seek their continuous piecewise linear (P1) finite element approximations. In turn, wh is either also piecewise linear (P1) [3] or piecewise constant (P0) [4,5]. We denote these approaches, respectively, as P1-P1 or P1-P0. For the permafrost (P) problem, the literature reports on the node-centered finite difference or piecewise linear finite element implementations, but without theoretical analysis or studies of convergence; see, e.g., in [6,7,8,9,10,11]. These also fall in the category of P1-based approximations. In turn, it is common to use cell-centered finite differences (CCFD) or finite volumes (FV) for a variety of fluid flow problems including those in subsurface [12].

    Heterogeneous materials and conservative approximations. It is well-known that the P1-based approximations due to their poor approximation of fluxes are not well suited for computational modeling of problems with large heterogeneity of coefficients such as Darcy flow in porous media coupled to transport; see, e.g., discussions in [12,13]. For computational models of (1.1) and (1.2) alone in single material, this would not be an issue.

    However, our goals in this paper are to study (1.1) and (1.2) in heterogeneous domains; additionally, we seek robust algorithms suitable for coupled multiphysics system such as THM [14,15,16,17], and, simultaneously, for multiphase multicomponent systems at complicated geometries such as at pore-scale based on voxel grids [18,19,20,21]. Therefore, we draw from literature on conservative approximations using mixed finite element methods for multiphase flow in porous media and specifically on problems with nonlinearities similar to (1.2) [22,23,24]. These feature conservative fluxes and piecewise constant (P0) approximations to scalar unknowns. In this paper we approximate both θ and w with piecewise constants (P0), thus we refer to these algorithms as P0-P0. Finally, when fluxes are approximated in the space RT[0] (discussed below), the algorithms can be conveniently implemented as cell-centered finite differences (CCFD); this gives a robust easily extendable structure towards more complex physics across many scales. Although these algorithms feature lower order approximation error, this is not an issue given low regularity of solutions to (1.1). Unfortunately, the approximation of fluxes features a technical gap since the normal fluxes for (ST) feature a jump; we defer the associated theoretical issues to future work.

    Our results. (i) We first evaluate the feasibility of P0-P0 approximations with fully implicit in time approximations for (1.1) with (1.2) for a single material such as bulk water. We show our approach compares well to P1-based schemes, and that the solvers are robust, even if there are some theoretical concerns. (ii) Second, we develop P0-P0 algorithms for the heterogeneous case k=k(x;θ) and α=α(x;θ) in multiple materials, when the data features the realistic but most challenging case of piecewise constant k(x;) and α(x;). Another challenge is that θ may not be continuous across material interfaces with high thermal resistivity. We prove various results and show robustness of our P0-P0 and monolithic CCFD algorithm which is conservative. Finally, we apply P0-P0 scheme to permafrost modeling including in heterogeneous materials, tie the theoretical framework for αP to that developed for other αST and αSTϵ, and confirm convergence.

    Overall, we find that P0-P0 algorithm is robust for heterogeneous extensions of (1.1) with any of (1.2) we considered. The order of convergence in the scalar variables (θ,w) is, for the most part, about O(h) and, O(h) whenever time step τ=O(h) is used. The nonlinear solver performs also well, even though it might need improvements for strongly heterogeneous nonlinearities.

    Outline: We give details on the phase change problems modeled by (1.1) with (1.2a) and (1.2b) in Section 2. In Section 2.4 we overview, compare, and illustrate traditional piecewise linear approximations to (ST) and (ST)ϵ which we call P1-P1 or P1-P0. In Section 3 we define our P0-P0 approximations to (ST) and (ST)ϵ, prove some auxiliary results, and compare convergence and robustness of the scheme to P1-based approaches. In Section 4 we address heterogeneous (ST) such as in (1.3), and prove properties of the P0-P0 scheme for the solutions that feature a jump across the material interfaces. In Section 5 we provide details on the permafrost model (P) (1.2c), along with convergence results for P0-P0 scheme, and illustrate with simulations for both homogeneous and heterogeneous problem. We close in Section 6, with auxiliary results given in Appendix 7.

    The problem (1.1) is posed in Q=Ω×(0,T) where ΩRd is an open bounded spatial domain with a smooth boundary Ω. In heterogeneous case, we partition ¯Ω=NMATj=1¯Ω(j) into subdomains Ω(j), corresponding to j=1,NMAT materials, with the material interfaces Γ=ijΓ(ij),Γ(ij)=Ω(i)Ω(j). Thermal properties of a region depend on the material in that region. We assume that in material j the heat conduction and phase transition parameters are known fixed constants. For example, at a point xΩ(j), the freezing temperature is fixed for this material so that θfr(x)=θ(j)fr. In turn, the thermal conductivity k(x;θ)=k(j)(θ) depends on the temperature in a fixed way specific to material j. We denote this dependence of thermal properties as follows

    c(x;θ)|Ω(j)=c(j)(θ);k(x;θ)|Ω(j)=k(j)(θ);θfr(x)|Ω(j)=θ(j)fr;L(x)|Ω(j)=L(j). (1.3)

    The notation Q(j)p indicate the portion of Q in domain (j) and phase p.

    We assume that the subdomains and interfaces are smooth enough so that standard notation and results using the spaces Ck(Ω), and Lebesque and Sobolev spaces such as Lp(Ω) and HkΩ) make sense. For a Lipschitz continuous function f we denote its Lipschitz constant by Lf. We denote by V=H10(Ω) for the primal variational formulation, and by X=Hdiv(Ω) and M=L2(Ω) the spaces for fluxes and scalars, respectively, needed in the mixed setting. By (u,v)=Ωuv we denote the inner product of scalar functions on L2 as well as that for vector valued functions on (L2)d. The norms ||f||G of functions in space G are as indicated with a subscript G which is dropped if G=L2(Ω) or G can be inferred from context. For time-dependent problems when t(0,T) and, e.g., uL2(0,T;V) we use the shorthand notation L2(V). For discrete formulations, we recall the spaces VhV of piecewise linear finite elements based on a triangular grid. We will also use XhX some approximations to the vector valued functions such as heat flux, and MhM the space of piecewise constant approximations.

    By 1U(x) we denote the characteristic function of set URdx. For two sets Ω and Ω+ separated by an interface Ω+Ω, by [r(x)]=(r|Ω+r|Ω)|x we denote the jump of quantity r at some interface point xΩ+Ω. These two domains correspond to two materials, or to two phases of one material. When the sign of the jump is important, we make it precise on which side of the interface the limits are taken.

    Next, consider real functions u,w defined on Q, and a function β:RR. The statement u=β(w) means that u(x,t)=β(w(x,t)) a.e. on Q. Similarly, consider a graph of a relation αR×R for which α1=β is a non-injective function, but α() is multi-valued. In this case υα(r) is equivalent to (r,υ)α, and, for functions u,w on Q, the statement wα(u) means that w(x,t)α(u(x,t)) a.e. (x,t)Q.

    In time-discrete formulations we use uniform time stepping with time step τ>0 and 0=t0<t1, and tn=nτ. For an unknown u(,tn), we denote by un() its time approximation, and by unh() its fully discrete approximation, which may be identified by the set of values such as Uni. For a known function f(,t), we denote fn to be the value f(,tn) or its integral average 1τtntn1f(,υ)dυ, made precise in the context.

    Finally, when using physical data and units, we use SI units or state otherwise; see Table 1.

    In addition, we make some assumptions on the data. Assumption 1.1 is a standard assumption for data of heat conduction problems.

    Assumption 1.1. We assume that the latent heat coefficients L(j)0, and that heat capacity and conductivity coefficients are cmaxc(j)pcmin>0;kmaxk(j)pkmin>0 for all phases p=s,l, and domains j=1,NMAT.

    For simplicity of notation when reviewing literature as well as in the proofs of some results we assume the following. (More realistic cases are given in examples.)

    Assumption 1.2. Let θ satisfy homogeneous Dirichlet b.c. θ|Ω=0. We also assume that some initial value winitL2(Ω) is given.

    Last but not least, our focus in this paper is on the nonlinear relationship α given as one of the three specific choices α(ST), α(ST)ϵ, α(P) defined in (1.2) which come from phase change problems. However, our results apply to other possible α() with properties summarized in Assumption 1.3. These properties are well known for α(ST), α(ST)ϵ; see, e.g., [25]. For α(P) in the permafrost problem, these properties follow from the algebraic formulas discussed in Section 5.

    Assumption 1.3. We assume that α satisfies one of the following properties (a)–(c), similar to (ST), (ST)ϵ, (P), respectively: (a) α is a maximal monotone graph with a non-injective Lipschitz inverse β=α1, similarly to αST, (b) α is a monotone strictly increasing function which is piecewise smooth and globally Lipschitz, with an injective Lipschitz inverse, similarly to αSTϵ, (c) α is a smooth strictly increasing globally Lipschitz function which has one point of non-differentiability (similar to αP).

    In this section we provide details on (ST) and (ST)ϵ. We suppress these superscripts now. A description of the variables used in this paper is given in Table 1.

    We follow closely [27] as well as the applications literature [28] for thermodynamics of multiple phases. We consider the temperature field θ(x,t),xΩ,t>0 and define Ωl={xΩ:θ>0}, Ωs={xΩ:θ<0}, with Ql,Qs defined analogously. We also consider the phase interface S=QlQs, and St its instance at time t. The motion of St is described by its velocity v, with the normal component vν. In turn, the normal n to S is oriented towards Ql, with components (nx,nt) with nx parallel to ν, so that we have nt=vnx.

    We recall first the so-called strong (classical) formulation of Stefan problem: we seek the temperature θ in each region Ωp,p=s,l which satisfies the heat equation

    t(cθ)+q=f;q=kθ,xΩp. (2.1a)

    The first part of (2.1a) is the energy conservation involving heat flux q, external source f, and the internal energy cθ dependent on the temperature. The second part is the Fourier heat conduction law.

    The coefficients of volumetric heat capacity c and heat conductivity k depend on the phase of a material; see Table 2 for typical values. In this paper we consider the simplest realistic case in which c,k are piecewise constant in each Ωs,Ωl, respectively.

    c(θ)={cs,θ<0cl,θ>0,k(θ)={ks,θ<0kl,θ>0. (2.1b)
    Table 2.  Thermal properties of some materials including water, components of phase change materials (PCM), human tissue, minerals, rock grains and insulators. Given are the latent heat of fusion/melting L, volumetric heat capacity c, thermal conductivity k, and melting point. For each material, the available thermal properties of its solid (s) and liquid (l) phase are provided. () Depends on proportion of silicone. () Depends on water content. The properties of (W) component will be used frequently below.
    Material units L [J/cm3] cs/cl [J/cm3 C] ks/kl [J/cm s C] Melting [C]Point Ref.
    (W) Water/ice 306.00 1.90/4.19 0.0230/0.0058 0 [26]
    Paraffin solid/liquid 183.18 1.58/1.84 0.0040/0.0040 60 [35]
    Rubitherm RT 55 149.60 1.76/1.54 0.0020/0.0020 55 [36]
    Octadecanol (silicone) 168 to 213 0.002 to 0.002 55 to 60 [37]
    Human skin (epidermis) 0.8 0.0019 [38,39]
    Human muscle 0.0017 to 0.011 [38]
    Silica 0.7 0.014 1713 [39,40]
    Styrofoam 0.00029 [41]

     | Show Table
    DownLoad: CSV

    One can extend (2.1b) to θ=0 using arithmetic averages [27](Ⅳ.4.1); see also Section 3.2.3.

    Finding the free boundary S=QlQs, is part of the problem, with S(x,t):θ(x,t)=0. Additionally, the Stefan condition governs the velocity of St as follows

    [(qlqs)ν]=Lvν;or[(klνθlksνθs]=Lvν; (2.1c)
    or[(klθlksθs]nx=Lnt (2.1d)

    where ν is the unit normal to St=SΩ×{t}.

    The strong (classical) form of the Stefan problem seeks {S,θ}, with θ is expected to have continuous derivatives to all the relevant order in QS. As is well known, such a solution may not exist [27] (v)Section Ⅳ.1. Hence, there is need to define weak solutions of (2.1).

    The weak form of (2.1) is derived upon integration by parts using test functions from D(Q) and assuming that θC(¯Q) but that its derivatives need only be in L1(QS). The energy conservation (2.1a) is written in all of Q, and the weak form, in the sense of distributions, reads

    tw+q=f;q=kθ,inD(Q). (2.2a)

    The definition of enthalpy w follows by integration of parts of (2.2a) written in the sense of distributions, the use of (2.1a)–(2.1c) along with the assumption that S is smooth and θ continuous across S; see, e.g., [27](Ⅳ.1p.101) [25](A1 p244). This definition "translates" the Stefan condition (2.1c) as follows, in one of the many variants

    w=cθ+L2ϕ;orw=cθ+Lχ. (2.2b)

    Here ϕ is the order (phase) parameter and χ=ϕ+12 is the water fraction. In equilibrium, these are set to ϕ=1,χ=1 in Ωl and ϕ=1,χ=0 in Ωs. On S where θ=0, ϕ and χ are independent variables determined uniquely by the heat content w; in particular, 0χ=wL1 when θ=0. However, when written in terms of temperature, the relationships (θ,χ) or (θ,ϕ) appear multi-valued

    ϕsgn(θ)={1,θ<0[1,1],θ=01,θ>0,χH(θ)={0,θ<0[0,1],θ=01,θ>0. (2.3)

    This only appears when the fact that χ (and ϕ) are independent variables is ignored.

    The definition (2.2b) can be also explained from thermodynamic principles: here we have dw=c(θ,χ)dθ+Ldχ which expresses a general dependence of c on θ in (2.1b) as a weighted fraction of the heat capacities cl in the liquid and cs in solid phases

    c=clχ+cs(1χ), (2.4)

    with a possibly variable cp=cp(θ). Then one integrates dw=cdθ+Ldχ to yield

    w(θ,χ)=θθfr[cl(υ)χ(υ)+cs(υ)(1χ(υ))]dυ+Lχ. (2.5)

    This formula implies (2.1b) and has a nice nontrivial counterpart in permafrost; see Section 5.

    In relaxation or in phase field models, i.e., away from equilibrium, ϕ is not given by (2.3), but is governed by its own dynamics, with its range possibly outside [1,1]. We describe and illustrate such models in Section 2.4.1.

    The analysis of (2.2) proceeds after we change variables and write

    twΔu=f,wαK(u)u=βK(w);ortαK(u)Δuf. (2.6)

    The variable u is called the "Kirchhoff temperature" and is distinguished from the true temperature θ. To transform (2.2) to (2.6) we replace kθ=k(θ)θ=u with u(θ)=θ0k(r)dr, which, with (2.1b), gives

    u=u(θ)={klθ,θ0,ksθ,θ0,orθ=θ(u)={ukl,u0,uks,u0. (2.7)

    Next, from (2.2b) with (2.1b) we have the relationships for (θ,w)

    θ=β(w)={wcs,w<00,w[0,L]wLcl,w>Lwα(θ)={csθ,θ<0[0,L],θ=0L+clθ,θ>0. (2.8)

    Finally, we combine (2.8) with (2.7) and see that u=u(θ)=u(β(w))=βK(w) with

    u=βK(w)={kswcs,w<00,w[0,L]kl(wL)cl,w>LwαK(u)={csuks,u<0[0,L],u=0L+clukl,u>0. (2.9)

    We see that β and βK are non-injective functions, while α and αK are maximal monotone graphs, i.e., the range of identity plus the graph is R, and they are monotone. The functions βK,β are Lipschitz, with Lipschitz constants

    Lβ=max{c1s,c1l}; LβK=max{ksc1s,klc1l}. (2.10)

    The functions βK,β along with the graphs αK,α are affine bounded, i.e., there are constants Cβ=max{c1s,c1l,Lc1l},CβK=max{ksc1s,klc1l,Lklc1l} and Cα=max{cs,cl,L},CαK=max{csk1s,clk1l,L} with which

    |β(w)|Cβ(1+|w|);|βK(w)|CβK(1+|w|), wR, (2.11a)
    |w|Cα(1+|θ|),θR,wα(θ); (2.11b)
    |αK(u)|CαK(1+|u|),uR,wαK(u). (2.11c)

    There are many ways to approximate a multi-valued graph α (or αK, sgn or H) by a single-valued Lipschitz function so that its inverse is injective. One very specific (one-sided) approximation is the Yosida approximation

    αλ=1λ(I(I+λα)1). (2.12)

    We see that Lαλ=O(λ1) which blows up as λ0. For this and other similar approximations of αST by αSTϵ there is no jump of qn or of w across the free boundary S. Rather, these variables vary sharply in the region

    Ωϵl={x:0<θ(x)<ϵ}. (2.13)

    See, e.g., illustration in Figure 1(c).

    Figure 1.  Illustration of domains of interest in this paper. (a) Liquid and solid domains in (ST) with a sharp liquid-solid interface S:θ=0 (dashed curve), which is advancing during melting; shown is its normal velocity given by Stefan condition and its components. (b) Domains in the regularized (ST)ϵ problem including the light colored region ΩϵlΩl given by (2.13): 0<θ(x,t)<ϵ;xΩϵl; (c-d) Domains with with piecewise constant thermal properties. (c) Pore-scale domain with two materials: grains in black and void space filled with ice and water. (d) Phase change material at mesoscale.

    The models (1.1) with (1.2) and in particular (2.2) and (2.6) are nonlinear monotone evolution equations with maximal monotone graphs α. Theory for such equations in Banach spaces is given in [25]; the specific case of Stefan problem is discussed in [27] in an abstract Hilbert space setting; see also [3].

    We recall only the case of Dirichlet homogeneous boundary conditions, in the Hilbert space case with V=H10(Ω). We seek wL(L2)H1(H1), uL2(V) such that u=βK(w), w(,0)=winitL2(Ω) and (see, e.g., [4])

    (wt,ψ)+(u,ψ)=0, ψV. (2.14)

    Assuming some smooth enough winit is given, the maximal monotonicity of αK and the fact that both αK and α1K are affine bounded are sufficient for [27](1.20, p34; Thm.Ⅱ.5.1, p59). In particular, with fL1(L2)L2(V), and winitL2(Ω), one obtains that there exists a unique solution T0uL(V)H1(L2) (note that this implies uL2(L2)). Further refinement of the theory [27](Prop.Ⅱ.1.3 and Thm.Ⅱ.1.4))[27] gives u,wL(L2), with uniqueness in [27](Thm 5.2). Further regularity [27](Thm 2.5) indicates uH1(L2)L(V). Since θ can be found from (2.7) which is continuous, one generally obtains similar qualitative properties of θ as those of u.

    These results indicate quite low regularity of w in contrast to the case when αϵ is single valued and Lipschitz continuous, since then wϵ has the same regularity as uϵ [27](Thm 2.6). Further results hold for in-homogeneous Dirichlet data as long as it is smooth enough; these results elucidate the connection between weak and strong formulations [27](Ⅳ.6). Under nonlinear Neumann boundary conditions [29] for NMAT=1 one obtains θL2(H1),wL2(L2).

    In this section we review several approaches towards the numerical approximation of (2.2) and the Kirchhoff transformed version (2.6) which combine time-stepping with some spatial discretization such as finite elements or finite differences. We focus on the traditional mesh-based PDE discretization; other approaches include explicit tracking of the free boundary, but without solving for (θ,w) on a mesh; see, e.g., [30] for review. These latter approaches do not apply very well to the simulation at pore-scale, permafrost, or heterogeneous materials.

    In Section 2.4.1 we recall and illustrate time stepping schemes. Next in Section 2.4.2 we review spatial discretization approaches focusing on piecewise linear approximations uhVhV; the methods differ by how wh is approximated and how (ST)ϵ is selected.

    Consider some f=f(t), and A>0 in

    w+Aθ=f(t), θ(t)=β(w(t)), t>0;w(0)=winit. (2.15)

    We illustrate several time discrete schemes which approximate the relationship of (θ,w) either by regularization, Chernoff formula, or by relaxation.

    The backward Euler scheme is

    wn+τAθn=τfn+wn1,wnα(θn), n1,w0=winit. (2.16)

    and the solution for every n1 is guaranteed from Lemma 7.1; in practice, we use Newton's iteration substituting θn=β(wn).

    The solution to (2.15) can be also computed using so-called Chernoff formula [4]

    θn+τμAθn=τμfn+β(wn1), (2.17a)
    wn=wn1+μ(θnβ(wn1)), n1,w0=winit, (2.17b)

    where μ is the relaxation parameter, a constant approximation to 1β, which satisfies 0<μL1β, with Lβ from (2.10). The Chernoff formula offers a way to linearize the non-linear relationship β. Chernoff formula does not require any nonlinear iteration, since it solves a linear problem for θ followed by an update of wn. However, it produces consistency errors growing with μ1.

    We also consider the phase-relaxation approach: we replace the equilibrium relationship (2.8) w=cθ+Lχ=cθ+L2(ϕ+1) in which ϕsgn(θ) by a coupled problem which allows ϕ=ϕ(t) to evolve towards this equilibrium written as sgn1(ϕ)θ. [27](Ⅴ.1). This relaxation is another form of approximation to the equilibrium problem; see [31] for the PDE (2.2). In a more general phase-field approach, one can consider an additional dissipative term similar to Aϕ, plus a coarsening term proportional to ϕ which acts to counteract Aϕ; these together moderate the evolution of ϕ coupled to the PDE (2.2).

    Let ε>0 and γ>0. The phase-relaxation approach to (2.15) is given as

    w+Aθ=f,w=c(θ)θ+L2(ϕ+1),w(0)=winit, (2.18a)
    ϕ+1εg(ϕ)=γθ, ϕ(0)=sgn(winit). (2.18b)

    Here g is one of two possible choices which approximate sgn1() and contain the linear destabilisation term. We consider g(ϕ)=ϕ3ϕ similar to that in Allen-Cahn equation, [32] which we call "smooth". The other is g(ϕ)=g(ϕ)=sgn1(ϕ)ϕ which we call "non-smooth" [27](Section Ⅵ.5) [33,34]. We discretize (2.18) fully implicitly and solve by Newton's iteration

    wnwn1+τAθn=τfn,wn=c(θn)θn+L2(ϕn+1),w0=winit, (2.19a)
    ϕnϕn1+τεg(ϕn)=τγθn,ϕ0=sgn(winit). (2.19b)

    Example 2.1. We solve (2.15) with α=αST given by (2.8) with parameters in Table 2 (W). We choose A=102,winit=1, and a forcing term f(t)=sin(πt3600) for t[0,3600]. We derive the exact solution given in Section 7.1.

    In Figure 2 we plot numerical solutions using τ=40 and one of the three schemes: fully implicit, Chernoff formula, and phase relaxation with (2.18). For Chernoff formula, we choose μ=1.5 so that μL1β=cs=1.90. For phase relaxation, the key challenge here is a choice of the parameters ε=1/10, γ=1, so that the time-relaxed dynamics resembles that of (2.15).

    Figure 2.  Numerical solution of Example 2.1 using Backward Euler, Chernoff formula, Smooth (S) and Non-smooth (NS) phase relaxation approximation. The thin black line shows the exact solution (7.1). From left to right: plots of θ(t), w(t) and (θ,w).

    In the end we see that the fully implicit solution (2.16) trails the exact solutions θ(t) and w(t) fairly well. The Chernoff formula and phase relaxation approaches seem to have less close agreement, especially in w(t). While they offer other advantages, this experiment informs our subsequent choice to focus on the fully implicit time stepping.

    Early approaches to numerical solution to (ST) include the nodal finite difference approach in [42] for which convergence (but no specific order) is proven, independently of ϵ in the regularization (ST)ϵ of (ST) problem. In [26] some time error analysis is provided.

    The majority of rigorous work is on nodal piecewise linear approximations VhuhuV in (2.14); we call these P1-based. The approaches differ in time stepping (as in Section 2.4.1), in how the original problem (ST) is approximated by some (ST)ϵ, and in how wnh is defined. In particular, [43] prove L2(L2) order of convergence close to O(h) for fully implicit approximations for (ST)ϵ, when ϵ=O(h2),τ=O(h2) but require twϵ to be bounded, and refer actually to simulations with the (P) problem discussed in [7] instead of (ST)ϵ. In turn, [4,5,44] approximate solutions for some (ST)ϵ rather than (ST); [4,5] make use of the Chernoff formula similar to (2.17), while [44] and [31] use phase relaxation; these are closely related as shown in Figure 2.

    The main difference between the individual approaches is in the treatment of spatial integrals Ωwψ and Ωuψ with  ψVh, and in adjustments to how wh is found. In some schemes the numerical integration, or one of projection operators such as P0h,P1h,Πh are used. In some, piecewise constants and wnhMh are used, this is similar to our P0-P0 schemes to be defined in Section 3.

    The theoretically estimated convergence error depends, as usual, on h,τ and ϵ. Generally, θ is predicted to be approximated well, qualitatively and quantitatively, by all the P1-based schemes, with the order about O(h), in the weaker norms. However, the errors wwh are, as expected, higher, and wh appears "smeared" near the free boundary St.

    We set up detailed experiments and provide illustrations as part of our subsequent studies of P0-P0 schemes for (ST) problem in comparison to P1-based schemes. Our tests given in detail in Section 7.3 along with fine details on the schemes show somewhat better rates than those predicted and tested in [26,31,44]. Of the schemes studied, [5] produces the best approximation and convergence rates. We acknowledge the limitations of our study only in d=1, but believe these provide good starting point for subsequent comparisons with P0-P0 schemes.

    In this section we propose P0-P0 spatial discretization for (1.1) combined with any one of (1.2) combined with fully implicit in time scheme. We start with a mixed finite element formulation for (1.1), using M=L2(Ω) for scalar unknowns such as θ and w, and qX=Hdiv(Ω) which features continuous normal components across any smooth surface. Next we choose P0-P0 approximations θh,whMhM. We also seek fluxes qhXh=RT[0] on a rectangular grid as in [22,45]. The pair (Xh,Mh) is a stable pair for the Darcy problem satisfying the Banach-Neĉas-Babuŝka conditions [46]. For linear problems they approximate the fluxes and scalar unknowns to the same order O(h), with superconvergence for smooth solutions and some norms [45,46,47,48].

    Remark 3.1. For nonlinear relationship represented by (1.2) we encounter here the major challenge. The choice X=Hdiv works for the (ST)ϵ and (P) problems when α() in (1.2b) and (1.2c) is single-valued. However, in the Stefan problem (ST) with (1.2a), the fluxes qHdiv(Ω) since their normal takes a jump across St, which ties to the multivalued character of α(). This raises concerns on the approximability of qX by qhXh. We acknowledge this difficulty and formally develop the theory only for the single-valued α such as for (ST)ϵ and (P) problems, but we extend the algorithms in Xh×Mh to the (ST) problem. We defer further study to future considerations.

    The P0-P0 algorithm has several attractive features. 1) First, the normal fluxes of qhXh are continuous, which leads to conservative schemes across element and material interfaces; to support this, we work in the (θ,w) formulation instead with Kirchhoff variables (u,w). 2) Second, if θh is piecewise constant, it is natural to define wh=α(θh)Mh in a consistent fashion so that (2.8) is enforced at every degree of freedom. 3) Third, the P0-P0 equivalent to the mixed framework features approximation properties known from the literature; in particular, the results in [22] are most relevant for the present nonlinear case of (ST)ϵ and (P) problems. In fact, we demonstrate that the convergence in (θ,w) is not inferior to that for P1-based schemes from Section 2.4 even for (ST) problem; this is done in Section 3.4. 4) Last, the approximations, up to quadrature, are equivalent to a CCFD scheme for θhMh from which the fluxes qhXh follow post-processing; these features, recalled in Section 3.2, make implementation easy and allow its extensions to more complex nonlinear problems and multiple materials.

    Below we first set-up the notation and recall main results on the mixed finite element discretization leading to P0-P0 algorithm. For simplicity of notation we consider ΩRd,d=2 and assume that it is well covered by a rectangular grid Th=(ωij)ij so that ¯Ω=ijωij, with maxij|ωij|=maxijhx,ihy,j=h2. Each cell ωijTh has a center at some (xij,yij) and edges γi1/2,j,γi,j1/2,γi+1/2,j,γi,j+1/2, when listed clockwise from the left edge. Throughout this section we use Assumption 1.2.

    We recall the mixed formulation for the linear case of (1.1) and (1.2b) with L=0 but allowing k=k(x),c=c(x), with some given initial condition w(x,0)=winit(x). We have then k1q=θ;tw+q=f;w=α(θ)=cθ. Its weak formulation follows after we multiply each of the equations in (1.1) by test functions ψX, ηM, integrate by parts, respectively, and apply the boundary conditions. We seek (q,θ)X×M which satisfy

    (k1q,ψ)(θ,ψ)=0,ψX, (3.1a)
    (q,η)+(tw,η)=(f,η);ηM;w=α(θ)=c(x)θ. (3.1b)

    (In these equations the symbols (a,b) mean inner product Ωab in L2(Ω) as given in Section 1.1). The approximations qnh,θnh with wh=α(θh) are formulated after (3.1) is discretized in time, and when at each time step tn we seek (qnh,θnh)Xh×Mh which satisfy a system similar to (3.1). We make these precise now.

    We consider the well known spaces (Xh×Mh) built on Th with Xh=RT[0], the lowest order Raviart-Thomas space on rectangles [22,45,48]. The space Mh contains piecewise constants on Th; the basis functions spanning Mh are simply 1ωij, and θh|ωij=Θij associated with the cell centers of each ωij. The vector valued functions in Xh are tensor products of piecewise linears in one coordinate with piecewise constants in the other. In particular, (qh)1 is identified by their edge values at the left and right edges (i±1/2,j) so we have, e.g., (qh)1|γi+1/2,j=qi+1/2,j; analogously (qh)2 is identified by values at the bottom and top edges i,j±1/2, respectively, (qh)2|γi,j1/2=qi,j1/2. The basis functions for the vector valued functions in Xh are ψi±1/2,j for (qh)1 and ψi,j±1/2 for (qh)2. Let (Qn,Θn) denote the degrees of freedom for qnh,θnh in their bases.

    We define the fully implicit approximations (qnh,θnh)Xh×Mh to (3.1) as those that satisfy

    (k1qnh,ψ)h(θnh,ψ)=0,ψXh, (3.2a)
    (qnh,η)+(wnhwn1hτ,η)=(fn,η);ηMh;wh=α(θnh). (3.2b)

    We applied here numerical integration to the first integral in (3.2a) and replaced (k1qnh,ψ) by its approximation (k1qnh,ψ)h. Specifically, the following numerical integration is used: on every ωij=(xi1/2,xi+1/2)×(xj1/2,xj+1/2), the trapezoidal (T) scheme over (xi1/2,xi+1/2) and (M) midpoint scheme on (xj1/2,xj+1/2) for the first component of qhψh, and M×T scheme for the second component. This leads to some useful simplifications, very well known and described in [49]. In particular, (k1qnh,ψ)h gives KQ with a diagonal matrix K of positive edge factors Ti±1/2,j and Ti,j±1/2 involving k1; see Section 7.5 for more details. We also get (qnh,η)=BQn while (θnh,ψ) becomes BTΘn. The linear system reads

    KQn+BTΘn=0, (3.3a)
    BQn+1τWn=Gn,Wn=CΘn. (3.3b)

    with Gn=Fn+1τWn1. Note that C is the diagonal matrix of positive coefficients c|ωij.

    Remark 3.2. After we multiply the second equation in (3.3) by (1), the system (3.3) has a saddle-point structure. Thus it is well-posed, i.e., the operator N=[KBTBC] is an isomorphism [45](Prop.3.3.1 and Thm 3.6.2), because K is coercive (positive definite) on the kernel of B, C is positive definite. and B is surjective from XhMh.

    Remark 3.3. The system (3.3) can be easily modified to account for non-homogeneous Dirichlet boundary conditions; see Section 7.5 for details.

    Connection to CCFD. Since K is diagonal, every degree of freedom of Qn has an easy discrete interpretation, thus one can eliminate Qn, and (3.3) is equivalent to

    (τBK1BT+C)Θn=τGn. (3.4)

    This system is known as the cell-centered finite difference (CCFD) formulation. Now BK1BT is symmetric and at least nonnegative definite for Neumann boundary conditions, and positive definite for Dirichlet conditions. Since C is positive definite, we have a unique solution Θ from which Q follows.

    For linear problems such as (3.1), mixed FE solution (qh,θh) features optimal first order convergence of the errors ||qqh||X and ||θθh||M for the choice of RT[0]×Mh [46]. For Darcy and potential flow problems the quadrature error is lower order, and the mixed approach provides formal interpretation of the CCFD algorithm [49]. For parabolic problems under Neumann boundary conditions and strong assumptions on the smoothness of θ and q, [48] shows that

    ||θhθ||,2=O(τ+h2). (3.5)

    This order, is, in general, not featured for nonlinear problems such as (1.1).

    Now we consider (1.1) or Kirchhoff-transformed problem (2.6) with (1.2b) or (1.2c) in the mixed formulation. We seek (q,θ) or (q,u)X×M and replace w=cθ by w=α(θ) or w=αK(u) in (3.1b) to get the weak mixed and discrete mixed formulations similar to that for (3.1).

    The challenge, well described in [22] is that one must respect the regularity of the unknowns (q,θ,w) which is usually inferior to that for the linear case when twL2(Ω). In particular, when βK() has derivative vanishing pointwise, it may happen that twL2(Ω). This difficulty is partially overcome with a Kirchhoff transformation and upon integration in time, and/or discretization in time; we refer to e.g., [22,46,50] for thorough discussion. The approach of taking finite differences in time is also known as the Rothe or Crandall-Ligget or Hille-Yosida framework [25,29].

    For nonlinear problems which are Kirchhoff-transformed the mixed framework is set-up for the solutions (wnh,unh,qnh) in [22,23,51]. However, the error estimates depend on the smoothness of w and u. Disregarding the error in the fluxes, these results, when applied to (1.1) state that, as in e.g., [22]

    ||Wnwn||H1(Ω)C(u,q,βK;h),nk=1(Wkwk,Ukuk)τ¯C(u,q,βK;h),

    with C,¯C0 as h0 depending on the smoothness of u, but with the order not specified directly. The work [22] does not directly address existence and uniqueness of the solutions. Overall, the results in [22,23,51] are well-suited for problems such as Richards equation, with α which features somewhat different challenges than those for (ST), (ST)ϵ, (P) problems listed in Assumption 1.3.

    We extend now (3.3) to the nonlinear case when k=k(θ) and w=α(θ), working with (θ,w) as scalar unknowns. We define some approximations ˜cnc(θn) and ˜knk(θn), and take some ˜αα; we make these precise in Section 3.2.3. The fully discrete problem reads

    ˜KQn+BTΘn=0, (3.6a)
    BQn+1τWn=Gn,Wn=˜α(Θn). (3.6b)

    This nonlinear system uses ˜K based on ˜k. We prove first that the system has a unique solution; this is needed since Remark 3.2 does not apply to this nonlinear case. We complete details on the algorithm in Section 3.2.3.

    Lemma 3.1. Let Assumption 1.1 hold on ˜k, and let ˜K and B be computed as in Section 3.1; in particular, let ˜K be positive definite. Then there exists a unique solution (Q,Θ,W) to (3.6) and its generalization

    ˜KQ+BTΘ=0, (3.7a)
    BQ+1τW=g,W˜α(Θ). (3.7b)

    Proof. We discuss only (3.7) since the result for (3.6) follows as its special single-valued case. We proceed in one of two alternative ways which are each worthwhile discussing. One is that we rewrite (3.7) eliminating Q as in (3.8)

    B˜K1BTΘ+1τW=G,W˜α(Θ). (3.8)

    We set A=B˜K1BT which is at least nonnegative definite as discussed earlier. The system is as in Lemma 7.1, thus the existence of a unique solution (Θ,W) follows, with Q found by postprocessing from (3.6a).

    Yet another proof follows ideas in [50](Thm 3.2), and is worthwhile mentioning because it extends to the abstract weak formulation of (1.1) in (X,M) for the case when qX. We note that the system (3.7) can be written as M([Q,Θ]T)=[0,G]T; the nonlinear operator M a sum of diagonal matrix of maximal monotone operators [˜K00Φα] (we define Φα as in Lemma 7.1), and the accretive linear thus Lipschitz operator [0BTB0]. This means that M is maximal monotone, and that there exists a unique solution to (3.7).

    Now we need to define ˜k,˜c and ˜α. These are needed since we extend (3.3) to the nonlinear case when k=k(θ) and w=α(θ). Typically, we assign these cell-wise based on the Θ=(Θij)ijθh

    ˜kij={kl,ωijΩhlks,ωijΩhsk,ωijΩh0;˜cij={cl,ωijΩhlcs,ωijΩhsc,ωijΩh0. (3.9)

    (The value k,c can be one of many including kl+ks2 and cl+cs2, respectively [27](Ⅳ.4.1)). Also,

    given(Θij)ij,defineΩhl=Θij>0ωij;Ωhs=Θij<0ωij;Ωh0=Θij=0ωij. (3.10)

    (In practice, the set Ωh0 is empty). From these the formula for ˜α follows by (2.8).

    Lastly, we need to make precise which Θ we use in (3.10). When entering a new time step tn, we have the previous time step value Θn1. When iterating on (3.6), in iteration m, we have Θn,(m1) to denote the iteration-lagged value, while we seek the new Θn,(m). We must therefore make precise in (3.10) whether it depends on the old Θn1, or on iteration-lagged value Θn,(m1). Adopting the notation for the sets in (3.10) from that for Θ we get, e.g., Ωh,n1l or Ωh,n,(m1)l. In other words, we can calculate ˜kij from (3.9) as either kn1ij or kn,(m1)ij. These choices give the matrix ˜K=Kn1 or ˜K=Kn,(m1), respectively. Thanks to Assumption 1.1 these have positive entries; see also Section 7.5. Therefore Lemma 3.1 applies to (3.6).

    Remark 3.4. The sets in (3.10) are not the same as

    givent>0,and(Ωl(t),Ωs(t)),define¯Ωhl(t)=ωijΩlωij;¯Ωhs(t)=ωijΩsωij. (3.11)

    If neither of Ωl(t),Ωs(t) is empty, Ω(¯Ωhl(tn)¯Ωhs(tn)), as illustrated in Figure 3.

    Figure 3.  Illustrations of a grid over Ω, or over Q, for (ST) problem and (ST)ϵ. Left: the free boundary S is not aligned with grid cell interfaces, and qHdiv(Ω). Middle: the approximation of q by qh should be reasonable in the (shaded) region (¯Ωhl(tn)¯Ωhs(tn)) defined in (3.11), even if its complement in Ω is not empty. Right: in (ST)ϵ, qHdiv(Ω) but q features sharp gradient in the region Ωϵ defined in (2.13).

    Next we discuss the nonlinear solver for (3.7). The solution to (3.7) exists and is unique according to Lemma 3.1, but the proof makes a reference to some (iterative) optimisation algorithm to find the desired minimizer Θn.

    In practice, the use of such a minimisation algorithm may be tedious and is unnecessary. Instead, we solve the problem using Newton's method: we rewrite (3.8) from the proof of Lemma 3.1 in residual form using the single-valued inverse β of α.

    In each iteration m=1,2,, given Wn,(m1), Θn,(m1) we solve

    F1(Θn,(m),Wn,(m))=τB(˜Kn,(m1))1BTΘn,(m)+Wn,(m)τGn=0, (3.12a)
    F2(Θn,(m),Wn,(m)=Θn,(m)β(Wn,(m))=0. (3.12b)

    This scheme means we are solving (3.12) simultaneously for two variables Θn,Wn. However, the second part of (3.12) is diagonal, thus we can, instead, eliminate Θn,(m) and solve the problem in terms of Wn,(m).

    It is known that Newton's iteration is not guaranteed to converge for an arbitrary initial guess even for smooth F=(F1,F2). Now the nonlinear function β in (3.12) is only piecewise differentiable, but such case is covered by the theory and practice of semi-smooth Newton methods in [52]. In our tests the Newton solver is robust and essentially grid-independent as long as the time step τ is not too large. We discuss performance of this iteration in Section 3.4.

    As we showed in Section 3.2, there is no difficulty formulating P0-P0 algorithm and solving the fully discrete case of (1.1) with multi-valued α (1.2a).

    In fact, given Θn we get Qn from (3.6a); this gives qnhXh. However, the true qHdiv. Thus any attempts to quantify the approximation error for qhq must take into account this important discrepancy expressed already in Remark 3.1. Thus there is a question whether the use of qhXh is appropriate for the (ST) problem. This challenge is somewhat more complex than the pointwise degeneracy with β=0 pointwise handled, e.g., in [22], which still keeps qX.

    At this time we see various avenues to address this challenge. One is to solve (ST)ϵ formally, and create a sequence qϵhXh for a collection of ϵ>0 adjusted to h, and defining ˜qnh as their limit. The fluxes qϵh would be reasonably accurate approximations to qϵX, and their limit ˜qnh to qX. One other is to find some approximation ~qnh to q by postprocessing qnh. One can also consider projecting q to Xh. We defer further analyses of possible improvements to qqnh to future work.

    Now we present examples and study the convergence of P0-P0 approximation for (ST) problem. We focus on the scalar unknowns (θ,w), and defer the study of the error qqh to future work. We study the approximation error θerr=θθh and werr=wwh. We choose only those norms that are easy to use when analytical solution is not available, and are easy to compare to the theoretical results on P1-based schemes. For completeness, the definitions of these norms are given in Section 7.4.1.

    In convergence studies we use uniform spatial and temporal discretization. We also note that when using fine grid, the error rates, especially those for w, are sensitive to interpolation and machine precision; thus, some care in grid refinement is needed to obtain consistent rates.

    The goal now is to compare the performance of our P0-P0 scheme with the P1-based schemes from Section 2.4. These results provide confidence in our work on multiple materials, as well as guide the choice of time step τ depending on h. We use two test cases which we call (RBC) from [26] and another called (VV) from [44]. These examples feature an assumed free boundary S moving with some given prescribed velocity, winit and time-dependent Dirichlet boundary conditions found from the exact solution. For illustration, the plots of the solutions and their approximations are given in Figure 4 and Figure 5.

    Figure 4.  Solution to (VV) Example 3.1 with M=10,τ=102 at three different times t=0.01, t=0.1, and t=0.2.
    Figure 5.  Solution to (RBC) Example 3.2 with M=20,τ=500, at three different time steps t=20000,t=100000, and t=200000. See also more details in Figure 13 in Section 7.4.2.

    Example 3.1. (VV) This example from [44] is not connected to any particular physical scenario, but results in very simple mathematical calculations. Let Ω=(0,0.4) and T=0.2, f=0 and L=c=k=1. Note that since the data is not physical, no particular units are used, even if our code assumes units such as those in Table 1. We have the free boundary for (2.2) S:ψ(x,t)=0, with ψ(x,t)=x+t+0.1, and

    (VV){w(x,t)=2(eψ(x,t)1)+1,θ=w1ψ(x,t)0,w(x,t)=eψ(x,t)1,θ=wψ(x,t)<0. (3.13)

    In experiments we stop the simulation at T=0.2, a time at which the free boundary position is still inside Ω. This choice helps to analyse how well the free boundary is approximated by the different numerical schemes.

    Example 3.2. (RBC) example uses realistic physically meaningful data from [26]. We have Ω=(0,20)[cm], L=306[J cm3], cs=1.90 and cl=4.19 [J cm3 C 1], ks=0.023 and kl=0.0058 [J sec1 cm1 C 1]. The free boundary s(t)=s0+vst, and

    (RBC){w=B+(B+L)eαw(vstx+s0),θ=(wL)/clxs(t),w=B+Beαs(vstx+s0),θ=w/csx>s(t), (3.14)

    where αl=vscl/kl and αs=vscs/ks, and data as in Table 2. We also use the parameters in Table 3.

    Table 3.  Parameters in Example 3.2.
    s0 B vs
    15 [cm] -594 [J/cm3] 5×105 [cm/sec]

     | Show Table
    DownLoad: CSV

    Convergence results for P0-P0 algorithm and (VV) example are given in Table 4, with τ=h/10. We also provide the fine grid-study with hfine=6.4×105 in Table 5; these results are very similar to those in Table 4, thus they validate our process for using θhfine as a proxy for θ.

    Table 4.  Temperature error calculated with the exact solution in Example 3.1 (VV) as described in Section 3.4.
    (VV) example convergence rates for error with exact solution
    Mx hx τ ||θerr||,1 Order ||θerr||,2 Order ||θerr||2,2 Order
    10 4.0×102 1.0×102 5.6635×103 - 1.1472×102 - 2.4479×103 -
    50 8.0×103 2.0×102 8.6400×104 1.1682 1.8488×103 1.1342 3.9386×104 1.1352
    250 1.6×103 4.0×104 1.5112×104 1.0833 3.0694×104 1.1157 6.6943×105 1.1011
    1250 3.2×104 8.0×105 2.8084×105 1.0456 5.5618×105 1.0613 1.2197×105 1.0579
    6250 6.4×105 1.6×105 5.4205×106 1.0221 1.0502×105 1.0358 2.3245×106 1.0300
    Mx hx τ ||werr||,1 Order ||werr||,2 Order ||werr||2,2 Order
    10 4.0×102 1.0×102 2.3537×102 - 1.1428×101 - 2.5765×102 -
    50 8.0×103 2.0×103 5.6481×103 0.8868 5.7288×102 0.4291 1.1383×102 0.5076
    250 1.6×103 4.0×102 1.4820×103 0.8313 3.4627×102 0.3128 6.4965×103 0.3485
    1250 3.2×104 8.0×105 2.6131×104 1.0783 1.3404×102 0.5897 3.0606×103 0.4677
    6250 6.4×105 1.6×105 5.7018×105 0.9459 6.5021×103 0.4495 1.2067×103 0.5783
    Mx hx τ ||θerr||,L1 Order ||θerr||,L2 Order
    10 4.0×102 1.0×102 7.8606×103 - 1.5139×102 -
    50 8.0×103 2.0×103 1.5822×103 0.9960 3.0487×103 0.9957
    250 1.6×103 4.0×104 3.1399×104 1.0048 6.0509×104 1.0048
    1250 3.2×104 8.0×105 6.2432×105 1.0036 1.2031×104 1.0036
    Mx hx τ ||werr||,L1 Order ||werr||,L2 Order
    10 4.0×102 1.0×102 2.7565×102 - 1.0452×101 -
    50 8.0×103 2.0×103 5.7124×103 0.9779 5.7020×102 0.3765
    250 1.6×103 4.0×104 1.1932×103 0.9730 2.5900×102 0.4903
    1250 3.2×104 8.0×105 2.4169×104 0.9921 1.1867×102 0.4849

     | Show Table
    DownLoad: CSV
    Table 5.  Convergence error for Example 3.1 (VV) as described in Section 3.4, error calculated with fine grid where Mx,fine=6250 in both ||||,1, ||||,2 and ||||,L1, ||||,L2, in comparison with errors calculated using the exact solution as in Table 4.
    (VV) example convergence rates for error with fine grid solution
    Mx hx τ ||θerr||,1 Order ||θerr||,2 Order ||θerr||2,2 Order
    10 4.0×102 1.0×102 5.6586×103 - 1.1462×102 - 2.4464×103 -
    50 8.0×103 2.0×103 8.5860×104 1.1716 1.8394×103 1.1368 3.9202×104 1.1377
    250 1.6×103 4.0×104 1.4570×104 1.1021 2.9652×104 1.1340 6.4815×105 1.1183
    1250 3.2×104 8.0×105 2.2663×105 1.1562 4.5146×105 1.1695 9.9217×106 1.1661
    Mx hx τ ||werr||,1 Order ||werr||,2 Order ||werr||2,2 Order
    10 4.0×102 1.0×102 2.0291×102 - 8.8338×102 - 1.5892×102 -
    50 8.0×103 2.0×103 5.6438×103 0.7951 5.7286×102 0.2691 8.7731×103 0.3691
    250 1.6×103 4.0×104 1.1927×103 0.9658 2.8148×102 0.4415 4.2063×103 0.4568
    1250 3.2×104 8.0×105 2.3874×104 0.9995 1.3043×102 0.4779 1.9348×103 0.4825
    Mx hx τ ||θerr||,L1 Order ||θerr||,L2 Order
    10 4.0×102 1.0×102 7.8603×103 - 1.5140×102 -
    50 8.0×103 2.0×103 1.5819×103 0.9961 3.0482×103 0.9959
    250 1.6×103 4.0×104 3.1365×104 1.0054 6.0466×104 1.0051
    1250 3.2×104 8.0×105 6.2148×105 1.0058 1.1999×104 1.0049
    Mx hx τ ||werr||,L1 Order ||werr||,L2 Order
    10 4.0×102 1.0×102 2.4335×102 - 1.0102×101 -
    50 8.0×103 2.0×103 5.7104×103 0.9007 5.3364×102 0.3965
    250 1.6×103 4.0×104 1.1920×103 0.9734 2.5740×102 0.4530
    1250 3.2×104 8.0×105 1.9459×104 1.1261 1.1151×102 0.5198

     | Show Table
    DownLoad: CSV

    The convergence results for (RBC) using the fine grid solution and exact solution are shown in Tables 6 and 7 respectively. We also tabulate the comparison to P1-based methods in Section 7.4.2 in Table 13.

    Table 6.  Convergence error with exact solution from (RBC) example described in Example 3.2 from Section 3.4.
    (RBC) example convergence rates for error with exact solution
    Mx hx τ ||θerr||,1 Order ||θerr||,2 Order ||θerr||2,2 Order
    20 1.0 5.0×103 7.4093×100 - 2.5085×100 - 4.5435×102 -
    200 1.0×101 5.0×102 5.9623×101 1.0944 2.4650×101 1.0076 4.3065×101 1.0233
    2000 1.0×102 5.0×101 2.9571×102 1.3045 7.8455×103 1.4972 2.2161×100 1.2885
    20000 1.0×103 5.0 3.1198×103 0.9767 8.2743×104 0.9769 2.4358×101 0.9589
    Mx hx τ ||werr||,1 Order ||werr||,2 Order ||werr||2,2 Order
    20 1.0 5.0×103 2.1277×102 - 2.0829×102 - 3.7168×104 -
    200 1.0×101 5.0×102 1.3797×101 1.1881 3.9387×101 0.7233 9.3584×103 0.5990
    2000 1.0×102 5.0×101 1.3045×100 1.0243 1.2142×101 0.5111 2.8891×103 0.5104
    20000 1.0×103 5.0 1.3097×101 0.9983 3.8480×100 0.4990 9.1760×102 0.4981
    Mx hx τ ||θerr||,L1 Order ||θerr||,L2 Order
    20 1.0 5.0×103 1.2507×101 - 3.3887×100 -
    200 1.0×101 5.0×102 1.1890×100 1.0220 3.3678×101 1.0027
    2000 1.0×102 5.0×101 9.0292×102 1.1195 2.4144×102 1.1445
    Mx hx τ ||werr||,L1 Order ||werr||,L2 Order
    20 1.0 5.0×103 1.9895×102 - 2.0449×102 -
    200 1.0×101 5.0×102 1.9028×101 1.0193 6.2413×101 0.5154
    2000 1.0×103 5.0×101 1.5034×100 1.1023 1.2142×101 0.7110

     | Show Table
    DownLoad: CSV
    Table 7.  Convergence error calculated with fine grid for Example 3.2 (RBC) from Section 3.4.
    (RBC) example convergence rates for error with fine grid solution
    Mx hx τ ||θerr||,1 Order ||θerr||,2 Order ||θerr||2,2 Order
    20 1.0 5.0×103 7.4070×100 - 2.5080×100 - 4.5425×102 -
    200 1.0×101 5.0×102 5.9330×101 1.0964 2.4588×101 1.0086 4.2914×101 1.0247
    2000 1.0×102 5.0×101 2.6568×102 1.3489 7.0525×103 1.5424 1.9892×100 1.3339
    Mx hx τ ||werr||,1 Order ||werr||,2 Order ||werr||2,2 Order
    20 1.0 5.0×103 2.1277×102 - 2.0829×102 - 3.4425×104 -
    200 1.0×101 5.0×102 1.2880×101 1.2180 3.8145×101 0.7372 7.4205×103 0.6664
    2000 1.0×102 5.0×101 1.2954×100 0.9975 1.2142×101 0.4972 2.8891×103 0.4097
    Mx hx τ ||θerr||,L1 Order ||θerr||,L2 Order
    20 1.0 5.0×103 1.2505×101 - 3.3883×100 -
    200 1.0×101 5.0×102 1.1877×100 1.0224 3.3632×101 1.0032
    2000 1.0×102 5.0×101 8.9758×102 1.1216 2.3903×102 1.1483
    Mx hx τ ||werr||,L1 Order ||werr||,L2 Order
    20 1.0 5.0×103 1.9401×102 - 2.0449×102 -
    200 1.0×101 5.0×102 1.9025×101 1.0085 6.2413×101 0.5154
    2000 1.0×102 5.0×101 1.3806×100 1.1393 1.1519×101 0.7339

     | Show Table
    DownLoad: CSV

    Summary of convergence of P0-P0 schemes: Generally, we see that our P0-P0 schemes converge roughly with first order in θ and half order in w. These rates are similar to those for P1-P0 schemes from Section 2.4. However, our P0-P0 schemes seem to improve on P1-based schemes qualitatively and quantitatively; in particular, we see improvement in the quality of approximations to the enthalpy, which seems due to the lack of consistency errors such as those for Chernoff formulas or phase relaxation.

    We acknowledge that this might be due to only testing in d = 1; nevertheless, these results are promising.

    Last but not least we discuss performance of the nonlinear solver for (3.7) since it works without regularization for (ST) problem, and we have not found a discussion for (ST) problem in the literature.

    The maximum number of iterations required for each case to converge is listed in Table 8. The solver uses for stopping criterium a combination of absolute tolerance of 1012 and a relative tolerance of 106 and τ=h/10, relative to the first iteration.

    Table 8.  Newton iterations for Examples 2 (VV), 3 (RBC), and multiple material Example 4.2 as discussed in Section 3.5.
    (VV) (RBC) (4.2)
    Table 5 Table 7 Table 10
    Mx iter Mx iter Mx iter
    10 4 20 4 20 3
    50 5 200 5 100 5
    250 5 2000 5 500 5
    1250 5 - - 2500 5

     | Show Table
    DownLoad: CSV

    Overall, the solver performs reliably under a variety of circumstances, including with uniform coefficients used in Example 3.1 (VV), realistic thermal properties used in Example 3.2 (RBC) and in the presence of multiple materials in Example 4.2 to be discussed below. It is also important that the performance seems to be mesh independent.

    Though not illustrated in Table 8, we see in practice that the number of Newton iterations may increase with larger discrepancies between parameters and larger latent heat L values. We add the notes on the specifics of the heterogeneous permafrost case in Section 5.

    More work on the theoretical underpinnings of P0-P0 algorithm is needed for Stefan problem, but overall our P0-P0 (CCFD) algorithm seems well suited to all the choices of α in (1.2) including even the most challenging case of αST().

    We consider now the main challenge addressed in this paper, that of simulation of phase change problem in a heterogeneous domain, a generalization of (2.2) to the case when the region Ω is occupied by NMAT materials, each in Ω(j),j=1,NMAT, with interfaces denoted by Γij=Ω(i)Ω(j), and Γ=ijΓij. We also have Σ=Γ×(0,T). See Figures 1 and 6 for illustration. The data corresponding to material (j) are denoted by c(j)l,c(j)s,k(j)l,k(j)s,θ(j)fr,L(j). With these we formulate α(j) as in (2.8), and assume that the data satisfies Assumption 1.1, and that Γ is at least as smooth as Ω. We denote the appropriate functional spaces local to Ω(j) with superscript (j), e.g., in X(j), and so on.

    Figure 6.  Illustration of thermal conduction in domains where the fluxes are continuous as in (4.1b) but the temperature appears to take a jump modeled by (4.1c). Left: two materials separated by a layer of third material with very low conductivity; temperature in function of x is shown as a black curve. Middle: the same materials as on left when the width of the interface is very small, thus it is only practical to model this region as low dimensional interface, and the temperature features behavior with a jump. Right: the results of simulation of stationary and non-stationary heat conduction in Example 4.1 similar to the case illustrated in the middle.
    Figure 7.  Solutions to Example 4.1 in Section 4.3 obtained with monolithic CCFD, with comparison to the corresponding stationary solution θstat given in (7.15). The graphs correspond to the different ratios k(2)k(1) and ρR as indicated. The magnitude of the jump [θh]Γ scales with ρR, as expected, but is robust with respect to k(2)k(1). For small k(2)k(1) and small t, the jump [θh]Γ is sensitive to h.

    Let (2.2) hold in each Ω(j). Denote by w(j),θ(j) the restriction of w,θ to Ω(j), so we have

    tw(j)+q(j)=f(j);q(j)=kθ(j),w(j)α(j)(θ(j)). (4.1a)

    This problem requires some initial conditions and some boundary conditions on Ω.

    We also need some interface conditions on each Γij. For these, it is natural to assume (i) continuity of fluxes, and (ii) of the temperatures across each Γij. However, in some applications including PCM, the condition (ii) must be relaxed to model the heat conduction across a very thin region of very low heat conductivity. When the width of that layer is very small compared to the width of the regions surrounding this layer, and if the interface region is approximated by a lower dimensional interface Γij, the temperature appears to take a jump, since the limits of temperature from both sides of Γij are quite distinct, while the flux is preserved [37]. See Figure 6 for illustration.

    This jump condition is formulated, e.g., in [29,53,54]

    q(i)ν=q(j)ν,onΓij. (4.1b)
    θ(j)θ(i)=ρRq(i)ν,onΓij. (4.1c)

    where ρR0 is called "thermal resisitivity" and where the orientation of the unit normal ν to Γij is from Qi to Qj. (More generally, one can consider ρR to be specific to an interface).

    If ρR=0 or the fluxes q(j) vanish, we have continuity of temperatures θ across Γij (but not necessarily of u). If ρR>0, the condition (4.1c) is a Robin condition which slows down the process of reaching thermal equilibrium across the interface. This latter case is important for applications. However, its mathematical and computational challenges have not been well studied. We address some of the challenges that (4.1) brings from theoretical and algorithmic point of view when ρR>0.

    The mixed formulation is particularly convenient for problems involving interfaces and multiple materials. In particular, the condition (4.1b) prescribes the continuity of normal fluxes across Γ which is naturally preserved for the fluxes qX. The condition also prescribes a possible jump of the scalar unknowns, but this is not an issue when extending θM. Therefore, if the problems on each Ω(j) are well posed in X(j)×M(j) for each j, then the global problem can be well studied in X×M; see, e.g., [55,56]. Similarly, one can easily define the P0-P0 (CCFD) algorithms on each subdomain coupled to the discrete counterparts of (4.1b)–(4.1c).

    As concerns the overall solution algorithm based on CCFD for (4.1a), in principle, some domain decomposition approach iteration is required to satisfy (4.1b)–(4.1c). However, one of our contributions in this paper is that we are able to formulate a monolithic P0-P0 algorithm on Ω. We also formulate various theoretical results and estimates for qhXh. We note however that even though we do not state approximation properties in this paper, these may involve θ considered in some broken spaces such as H(Ω)=H1(Ω(1))×H1(Ω(2)) rather than H1(Ω) due to the discontinuity of θ across Γ.

    After some literature review in Section 4.1, we formulate and illustrate our algorithms in Sections 4.2 for linear and nonlinear problems when qX. We extend these to (ST) problem and apply the same scheme, even if some of the theoretical results do not apply when qX. We provide examples and study convergence in Section 4.3.2.

    The mathematical literature on (4.1) is not abundant, but we review what is available. The problem is a special case of more general heterogeneous nonlinear parabolic problem in which

    θfr=θfr(x);L=L(x);c=c(x;θ);k=k(x;θ). (4.2)

    When the data (4.2) vary smoothly as functions of its first argument, it is possible to apply Kirchhoff transformation on Ω but this introduces various lower order terms in the PDE. This approach considered, e.g., in [57,58], allows some well-posedness analysis.

    Another class of approaches in [59,60,61] study a problem similar to Stefan problem via minimization of a collection of normal convex integrands, and assumes smoothness of α(x;θ) in x, for an application with phase change of a different type where the multivalued graph α(x;) representing the solubility constraint depends smoothly on x. At the same time, the challenges discussed here are similar to those revealed by simulations in [61] when applied to piecewise constant data.

    The piecewise constant case (1.3) of interest in this paper does not allow Kirchhoff transformation but is also most realistic. The well-posedness of the case ρR>0 is studied in the primal setting in [29] under external Neumann boundary conditions on Ω and with some initial data. Existence of solutions is proven for ρR>0, with estimates which allow the limiting case ρR0. Uniqueness is also proven [29]. For ρR>0, the paper predicts wL(Q), θL2(H), and that the jump is

    (i)||[θ]||L2(Σ)=O(ρR),(ii)||tw||L2(H)C1+C2ρR. (4.3)

    We note that as as ρR0, the estimates (4.3) suggest continuity of the temperature across Γ but also predict the blowup of tw.

    As concerns numerical approximation, [54] describes the approximation in (V(j)h)j similar to the P1-P1 schemes described in Section 2.4. A-priori bounds for (θnh,wnh) including those similar to (4.3) are proved in [53] for ρR>0. Finally, a time discrete approximation to twh is measured in some seminorms which involve h2τ2, 1ρR; these suggest to use hτ bounded as h0,τ0, and ρR such that τ2ρR0. Based on these a-priori bounds, convergence to the solution corresponding to ρR=0 is established. However, no rate of convergence for the approximation errors θθh is given.

    Last but not least we mention other problems and applications which feature jumps of the primary unknowns and fluxes. These works fit in the framework of heterogeneous domain decomposition [56] and inform our results, but are not closely related. The closest is our prior work on the domain decomposition approach for semiconductor modeling of heterogeneous junctions in [62,63,64]. Further, there is closely related substantial work on multiphase flow and multiple rock types in, e.g., [1,2] as well as the abundant literature on Beavers-Joseph conditions on the Stokes-Darcy interfaces, e.g. in [65] and other heterogenous domain decomposition settings.

    We approximate (4.1) with P0-P0 discrete scheme from Section 3. We recall that even though the formal approximation with XqqhXh is appropriate only for the single-valued setting w=α(θ) (as in Remark 3.1), the P0-P0 algorithm produces discrete conservative fluxes qhXh also for the multivalued case wα(θ).

    We start with P0-P0 (or CCFD) algorithm on each material domain, a counterpart of (3.7), which we couple with discrete counterparts of (4.1b)–(4.1c). Then we show that this domain decomposition formulation can be handled by a monolithic solver. With this, the theoretical results including convergence analyses available for single domain formulations in X(j)h×M(j)h extend to that on Xh×Mh (as long as qX). Moreover, the problem does not require any iteration on the interface.

    We explain the details first for the linear heat equation with the notation from Section 3.2; the generalization to the nonlinear problem is straightforward. For simplicity, we focus on only two materials NMAT=2 which occupy the domains Ω(1),Ω(2) separated by an interface Γ=Γ12.

    We rewrite (4.1) in a time-discrete mixed form, with subdomain problems (4.4a)–(4.4b) to be solved by (θn,(j),qn,(j))j. We also make precise the initial, external, and coupling boundary conditions (4.4c) and coupling via (4.4d)

    (k(j))1q(j)=θ(j),xΩ(j),t>0 (4.4a)
    q(j)+1τw(j)=g,xΩ(j),w(j)=α(j)(θ(j)). (4.4b)
    w(j)(x,0)=winit(x)|Ω(j),θ|ΩΩ(j)=θD|ΩΩ(j), (4.4c)
    (q(2)q(1))νΓ=0,θ(2)θ(1)=ρRq(1),(x,t)Σ. (4.4d)

    We approximate (4.4) with P0-P0 algorithm for (4.4a)–(4.4c) on each Ω(j) as in Section 3, each implemented as CCFD, and corresponding to a block linear system similar to (3.3) is solved, with off-diagonal coupling terms expressing (4.4d).

    This coupling could be solved by iteration; we refer to [56] for a discussion of domain decomposition iterative algorithms. In contrast, we propose to satisfy (4.4d) without an iteration which simplifies the implementation substantially. We explain the ideas briefly for d=1 in Section 4.2.2.

    Let Ω=(a,b) and Ω(1)=(a,x) and Ω(2)=(x,b), with a grid (ωi)i covering Ω(1)Ω(2) so that Γ=x is one of the cell edges. With global numbering of the cells ωi in Ω, we have i=1,j,j+1,M, where Γ=x=xi+1/2=γi+1/2 is between some cells ωi and ωi+1. Now the problem is discretized as in Section 3.2 with q(1)h and q(2)h identified by the edge values Q(1)=(Q1/2,Q3/2,Qi+1/2) and Q(2)=(Q+i+1/2,Qi+3/2,QM+1/2), respectively. We note the double value of the flux Qi+1/2,Q+i+1/2 to be set equal due to (4.4d). The temperatures θ(1)hM(1)h and θ(2)hM(2)h, respectively, are identified by the cell values Θ(1)=(Θ1,Θ2,Θi), and Θ(2)=(Θi+1ΘM). Finally, the Dirichlet conditions at the external boundaries are with Θ1/2=θD(x1/2), and ΘM+1/2=θD(xM+1/2). These enter (3.3) as in Remark 3.3.

    The approximation to the first part of the interface condition (4.4d) equates the fluxes Qi+1/2=Q+i+1/2. We also require Dirichlet values Θ and Θ+ to be used by each subdomain problem at x so that the second part of (4.4d) holds. Finding these is part of the problem. In summary, the approximation to (4.4) is as follows.

    (4.5a)

    where we find and from the solutions on and defined as follows.

    (4.5b)
    (4.5c)

    The solution to (4.5) requires formulation of appropriate matrices , to solve (4.5b) and (4.5c), respectively. These come from the transmissivities on every and are given as in Section 7.5 by (7.11). We also have two boundary edge factors and given by (7.13) independently on and , respectively.

    The problem (4.5) could be solved by iteration. However, our idea is that (4.5) can be solved by a much simpler monolithic CCFD solver on for and for which . The system takes the form

    (4.6a)
    (4.6b)

    The definition of is a straightforward extension of , but involves as well as an appropriately chosen transmissibility to guarantee (4.4d).

    Lemma 4.1. Solution to (4.5) exists and is unique; it is equivalent to that of (4.6) provided

    (4.7)

    where we use the shorthand notation and .

    Proof. The proof is purely algebraic. Assume (4.5) is satisfied. Denote , , so (4.5a) holds for some , . We know also from (4.5b) and (4.5c).

    1. Recall and . Setting these equal from the first condition in (4.5a), and expressing from the second part of (4.5a), we work to eliminate and get a formula for in terms of and . After some lengthy calculations we get

    2. Next we recalculate . After some algebra we get

    3. Now the expression on the right hand side can be written as

    This provides the definition of (4.7), i.e., of for the monolithic formulation. Existence and uniqueness of the solutions to (4.6) follows directly from that for (3.3) with the special definition of which satisfies the same properties as all other .

    Remark 4.1. When , (4.7) reduces to give as the usual weighted harmonic average of and in (7.11), and provides the interface value of

    In other words, a monolithic CCFD approach on matching grids is equivalent to the domain decomposition approach. This observation is perhaps not new, but is nevertheless very useful.

    Corollary 4.1. The results in Lemma 4.1 easily extend to when is aligned with cell interfaces. In addition, these results extend to nonlinear problems such as (3.6) and (3.7) with

    Here in each time step and iteration one updates the transmissibilities based on the current guess of . This system uses special interface transmissivities involving resulting in the dependence of matrix on .

    Remark 4.2. Convergence of the solutions to for the linear problem is expected to be qualitatively similar to the case without jump, since, upon Lemma 4.1, it follows from those for CCFD on each domain, e.g., in [48]. At the same time, the solution is not globally smooth enough, thus broken norms must be used when referring to the approximation error's dependence on higher order norms.

    We provide an example with a numerical approximation in Section 4.3. We also prove theoretical estimates on the jump .

    We aim to derive results similar to (4.3) in the mixed setting. We write the fully discrete mixed form of (4.4), and work with the discrete counterparts of , with . As explained for (4.6), we use which can be , or or , formed, respectively, either in a fully implicit, time-lagging or iteration lagging fashion.

    Lemma 4.2. Let . Then the numerical solution of (4.1) satisfies, at every

    (4.8)

    Proof. We suppress and . The weak mixed formulation of (4.1a) after integrating and adding the equations over reads: find such that for every we have the following

    (4.9a)
    (4.9b)

    To derive the estimates we now choose and . Adding the two equations and cancelling the skew-symmetric terms we obtain, after rearranging

    The right hand side now involves the boundary and the right hand side . We have control over the boundary term since . The first term on the left hand side is nonnegative. The second term can be bounded from below for , as well as for and

    In fact, for (ST) problem we have . For (ST), we have , with a similar calculation for from Section 5. Then we apply Cauchy-Schwarz inequality to the integral and follow up with the inequality . With sufficiently large we can now move the term to the left hand side to balance it with . We conclude that thus, upon (4.4d) we finally obtain

    (4.10)

    from which (4.8) follows.

    This result is similar to that predicted in (4.3) [29] in the primal formulation. With more work, we can also cover the case .

    Corollary 4.2. The estimate in Lemma 4.2 extends to the solutions for (ST) or (P) problems with single-valued when .

    Now we illustrate the problem (4.1) with simulations using our P0-P0 solver.

    We consider first an example in for which we study the dependence of the jump of on .

    Example 4.1. Let with some . Let some be given, as well as , , and so that . We also impose boundary conditions , The numerical solutions are computed with , and . We consider different ratios , and .

    We plot the solutions to Example 4.1 found numerically with our monolithic P0-P0 solver. We also consider the corresponding stationary limit of (4.1); see the details in (7.15) and (7.14) in Section 7.6.

    The numerical solution to the stationary problem in this simple case essentially coincides with the exact solution; this is typical for P0-P0 solution. The non-stationary solutions evolve towards this stationary solution. The qualitative nature is consistent with the imposed interface jump condition.

    In the end we also test the scaling of the jump predicted by (4.8) given what we found for the stationary solution in Section 7.6; we check how it behaves over time. The details in Section 7.6 predict that the jump of scales linearly with as , unlike predicted for the evolution problem in (4.3). For large ratio and small we find that the jump behaves similarly to that for the stationary problem. For small , and before the solutions are close to the stationary limit, we see that the size of the jump depends significantly on the grid discretization . This behavior correlates with the large magnitude of when is small. In fact, it also correlates with the estimates of in (4.3) (see also subsequent discussion in Example 4.3). We defer further study of these features to the future.

    Now we consider (ST) problem, and start by checking convergence of our P0-P0 scheme for the case of heterogeneous materials. To this end, we modify Example 3.2.

    Example 4.2. We let with with parameters given in Table 9. Note that the material within has while has ; the freezing temperature for both.

    Table 9.  Heterogeneous materials from Example 4.2.
    0.5 0.15 0.5 1 0 0
    1 0.25 1 2 10 0

     | Show Table
    DownLoad: CSV

    The initial conditions are which corresponds to for all . Boundary conditions are and for . Figure 8 shows the temperature increase throughout the domain due to the heat transfer through the left boundary.

    Figure 8.  Simulation results in Example 4.2. Top row: solution at ; Bottom row: solution at . The interface separating and is shown by the vertical magenta line.

    We do not have an exact solution, thus we must use fine grid solution for convergence studies. The coarse grid solution for , is compared with that for the fine grid, at the first fine grid time step of and the time . In convergence study we disregard the initial time period in which the solution has very steep due to the discrepancy between the initial and boundary data. The convergence of the solution is of lower order during that time and pollutes the rest of the error analysis.

    Convergence results, calculated for are shown in Table 10. We see similar rates of convergence for this heterogeneous example as for the homogeneous Example 3.2.

    Table 10.  Convergence error for and in Example 4.2 calculated for .
    Order Order Order
    20 - - -
    100 1.1404 1.1801 1.2245
    500 0.9414 0.9729 0.9975
    2500 1.0842 1.0853 1.0851
    Order Order Order
    20 - - -
    100 1.1876 0.5580 0.6404
    500 0.9426 0.4977 0.4823
    2500 1.0809 0.6072 0.5975

     | Show Table
    DownLoad: CSV

    Next we illustrate the need for multiple materials by simulating heat flow at the pore-scale. We proceed to the simulation in at the pore-scale which is motivated by our interest in permafrost. We use our P0-P0 solver with , since there is no interface with high resistivity between the mineral and water components.

    Example 4.3. Let with , with water saturated pore space , and the rock grains ; see Figure 9. The material parameters are those of water for and inferred from those for silica for as given in Table 2. We start from thermal equilibrium, with constant temperature with which we calculate .

    Figure 9.  Simulation described in Example 4.3 within Section 4.3.2 of heat conduction in the pore space, with geometry within depicted on top. We simulate the thawing front moving from the left to the right starting from thermal equilibrium and with left boundary subject to increased temperature. Displayed is the first time step (left), two middle time steps (middle) and steady state (right) for temperature (top) and enthalpy (bottom).

    The heat in increases due to the boundary condition at where we set . We also assume insulated boundary conditions elsewhere for .

    In Figure 9 we show temperature profiles of four time steps. The melting front moves from left to right with a phase change in the void space , but the grains do not undergo phase change. The heat flux moves faster through the areas with more grains, since this change requires less energy required for phase change than the heat conduction in the water component.

    More work on testing the heterogeneous Stefan problem and even the linear heat equation is needed, but overall we believe our monolithic P0-P0 scheme performs well for multiple materials. While not reported here, we tested the simulation cases of different freezing temperatures, and drastically different parameter between domains, and see that the algorithms perform in a robust way across these scenarios, even if a decrease of the time step is needed in the most challenging cases.

    Permafrost is defined as the ground that remains frozen for two or more years [41,66,67]. Temperatures in permafrost respond to the changes in ambient temperatures at the soil-atmosphere interface, possibly including the effects of precipitation and vegetation. In the upper portion of permafrost called active layer the temperatures increase in the summer and decrease in other parts of the year. The bottom of the active layer is isothermal but the depth of this layer is changing due to the increase of ambient temperatures, and this causes the thawing of some portions of permafrost, with further environmental consequences. The importance and impact of coupled phenomena within permafrost models is of current interest. Modeling of the coupled phenomena in permafrost regions is complex, and requires careful conservative approximations across the scales [10,14,15,67,68,69].

    We recall now the (ST) problem discussed in Section 2 for conservation of energy in any volume occupied only by water component in one of two phases: liquid or solid. For any small sub-volume centered at some we can calculate the water fraction . In equilibrium, it equals 1 whenever in the entire . On the other hand, if on , then . In the case , we have , and can be considered as an independent thermodynamic variable at constant volume; see [27]. In other words, the water fraction can be considered a pointwise or a volumetric quantity, and in (ST) problem we assumed (2.3) and .

    The model for conservation of energy in permafrost must extend this formulation to account for the presence of rock grains which do not change phase. We assume here that the soil rock grains have a constant heat capacity and a constant thermal conductivity .

    The presence of rock grains has several consequences. The first consequence is that in the energy balance we have to account for the rock as a separate material; we discuss this in Section 5.1.2. Second, the presence of rocks affect the local energy landscape at the fluid-rock interfaces; in particular, it causes depression of freezing temperatures in small pores, as well as premelting, the presence of a thin film of water around rock grains [70,71]. Both phenomena have significant effects on the qualitative behavior of phase transitions in permafrost.

    In modeling, one must consider the scale at which we wish to consider the phenomena. At the pore-scale of to scale, the rock grains and water occupied domains should be treated as separate materials such as in Example 4.3. In practice, modeling large scale changes in permafrost in response to the temperature in the environment must occur at larger scale such as that of , i.e., at the so-called Darcy scale. Typically, models at Darcy scale take advantage of constitutive relationships measured experimentally in a laboratory. Thanks to modern computational science the pore-scale modeling can be connected to Darcy scale such as in our work and in particular [18,19,20,21], but a thorough discussion is outside our present scope.

    We introduce notation which helps to explain the difference and connections between Stefan problem and permafrost models.

    As mentioned earlier, porous medium is made of rock grains and non-rock "void space" occupied by fluids such as water and air. In this paper we consider only the water component. The rock grains occupy a fixed portion of , and the water component in occupies the remainder so that . The void (non-rock) proportion is called porosity, and . The water component can be in liquid or solid phases so that , and we have

    (5.1)

    The quantity is called "unfrozen water content" in permafrost and is determined experimentally for a particular soil type depending on the temperature.

    The presence of rock grains influences the energy balance. In permafrost the balance of energy must account for the energy content in and , thus the heat capacity is a weighted fraction of that in rock grains and in fluid space,

    (5.2)

    which can be also written, after rearranging, as

    (5.3)

    where is the (constant) heat capacity of the "unfrozen soil" and is the (constant) heat capacity of the "frozen soil". In the end, we get [8]

    (5.4)

    Remark 5.1. We can now compare the formulas for in the classical Stefan problem (2.5) and in the permafrost model (5.4). The former sees the change between "solid" and "liquid", but the latter (5.4) presents the phase change problem between the "frozen soil" and the "unfrozen soil", where the amount of each is controlled by the unfrozen water content .

    One of important features of permafrost is the presence of unfrozen water at low temperatures i.e. even when . This phenomenon is not fully explained, and is accompanied by the depression of the freezing temperatures in small pores. The permafrost models take advantage of the empirical data which fits one of the parametric relationships for such as (5.1), where is the unfrozen water content determined experimentally and dependent on soil type. We follow a formulation based on [14]: here is parametrized with some soil-type dependent parameters , as well as with which denotes the residual unfrozen water content at some really low reference temperature

    (5.5)

    Now is the threshold temperature called freezing point depression such that for , only liquid water is present. Unlike in bulk water, typically . Parameter depends on the soil type. For example, for coarse-grained soils such as sand, the value of is large, but for fine-grained soils such as clay is small. For a particular soil type, we assume to be a constant.

    Remark 5.2. A variety of algebraic models for other than (5.5) are given in applications literature; see Table 11 for expressions adapted so that they fit (5.1). The models are qualitatively similar to each other. For a particular expression for plugged in (5.4) they produce which is a monotone increasing injective Lipschitz function. In particular, with (5.5) we get, after some rearranging

    (5.6)
    Table 11.  Different models for liquid water fraction in Section 5.1.
    Model (reference) Unfrozen water content Parameters Typical values
    [L] [72] Silt: [73]
    [W] [6] Sand: , Silt: [6]
    [M] [14] Clay: , [15]
    [A] [41] Silt: , [41]
    Basalt: [41]

     | Show Table
    DownLoad: CSV

    For comparison, we plot corresponding to the the different models in Figure 10. Here we use the parameters [16] , , and along with the thermal properties of water as given in Table 2.

    Figure 10.  Enthalpy given by (5.4) for different experimental models for unfrozen water content tabulated in Table 11. The plots mimic the properties of clay [15]. Although the graphs appear to have a singular derivative near , the functions are all Lipschitz. For the particular curves, we use for [M] [14], for [L] [72], for [W] [6]. Also shown is the enthalpy curve (ST) calculated without considering the effects of unfrozen water content when at i.e.. when the enthalpy (5.4) is calculated assuming that .

    We see that each can be seen as an approximation to the monotone graph , and each is a strictly monotone smooth increasing function of , differentiable everywhere except at (as we stated in Assumption 1.3 for ).

    To complete the permafrost model, we need to define the thermal conductivity of the porous medium which is made of and , with the latter partitioned between and .

    It is well known that, unlike capacity given in (5.3), conductivity of composite materials depends not just on the values of , as well as on the fractions , but also on the geometry of how the phases are arranged. Still, some authors consider straightforward weighting with volumetric fractions as in (5.3) which gives, e.g., in [16]

    (5.7)

    where is the thermal conductivity of the "unfrozen soil" and is the thermal conductivity of the "frozen soil". While (5.7) is a good first order approximation, it does not take into account the geometry of the rock grains; more accurate expressions based on data are considered in [10,14,15] [41](.) Other general approaches can be considered and involve upscaling or homogenization; see, e.g., [18,74]. More work is needed on more accurate expression for .

    For the needs of this paper we use (5.7), combined with some selected model for . This completes the permafrost model, once boundary and initial conditions are specified. To construct these for realistic scenarios, care is needed. The former depend on environmental conditions, while the latter must be found carefully since thermal equilibrium might not be always realistic to assume.

    We recall now the P0-P0 (equivalent to CCFD) approximation to (1.1) with (1.2) given in Section 3. The function described in Section 5.1.3 features a piecewise smooth nonlinearity, thus fits the general class of problems that Section 3 applies to, and we expect that the scheme P0-P0 will work well.

    In fact, the fluxes permafrost problem feature continuous normal components since there is no jump prescribed. Therefore, we expect that is approximated to at least to first order accuracy, similarly to what we encountered for the (harder) (ST) problem.

    To put our P0-P0 approach in perspective, we first briefly review relevant literature; see Section 5.2.1. Then we follow up with the test of convergence of P0-P0 in Section 5.2.2 along with examples in Section 5.2.3.

    The work [6] uses a variational formulation, implicitly discretized in time, to generate a set of nonlinear algebraic equations which are solved using the Newton's method; they test their algorithm on examples; see also discussion in [11]. Next, [8] use a fixed grid finite element method further employing Picard's method to deal with the consequential nonlinearity in the mass and stiffness matrix. In turn, [9] use a node-centered finite difference scheme and solve the nonlinear equations using Newton's methods, and [15] implement their finite element model within the commercially available solver ABAQUS.

    We adopt the fully discrete formulation for P0-P0 as given in Section 3 and extended to heterogeneous media in Section 4.

    Example 5.1. For convergence studies, we choose an example based on [26]. We consider the scenario as in Example 3.2, but with instead of , and with the following initial and boundary conditions

    (5.8)

    The soil properties are chosen to be . We run the simulation over a time period of . Since the exact solution is unknown, the convergence rates are calculated using a fine grid solution computed on a mesh with cells and time step .

    The results are given in Table 12. We see a modest improvement in the present case with over the rates obtained for in the Stefan problem, with order robustly for and for . Further, we also observe that the order of convergence seems independent of small variations in the soil type, or in the parameter in (5.5). These results are for as in (ST) problem.

    Table 12.  Convergence orders using P0-P0 scheme to permafrost problem for Example 5.1, and a fine grid solution. Different soil types are used, as shown by parameter in Column 3.
    0.37 0.50 0.81 0.52 0.63 0.93
    0.34 0.55 0.80 0.43 0.61 0.86
    0.65 1.06 1.30 0.53 0.95 1.30
    0.54 0.99 1.46 0.54 1.01 1.44
    0.57 0.66 1.08 0.95 1.15 1.36
    0.49 0.61 0.95 0.69 0.85 1.18
    0.56 1.00 1.35 0.50 0.97 1.29
    0.54 1.03 1.48 0.53 1.03 1.42
    0.62 0.73 1.22 1.55 1.77 1.56
    0.60 0.63 1.02 0.50 0.72 1.28
    0.54 0.96 1.35 0.51 0.98 1.18
    0.53 1.03 1.47 0.54 1.03 1.32

     | Show Table
    DownLoad: CSV

    Next we attempt which is optimal in linear and mildly nonlinear parabolic problems. We attempt the latter since the graph is much smoother than any of , hence, we hope to have better than first order convergence. However, there is only mild improvement in the convergence rate when using compared to ; specifically, the order of convergence is roughly . In the end, this improvement may or may not justify the extra computational effort of using .

    Next we apply our P0-P0 algorithm to physical examples pertaining to permafrost thawing. The goal of these examples is to test the robustness of our scheme and to illustrate the modeling aspects of permafrost. Since most of the interesting dynamics in permafrost occurs due to the varying surface boundary conditions and along the depth, we confine ourselves to the case.

    We use the expression (5.5) for the unfrozen water content .

    Example 5.2. Homogeneous case. Consider a small column of soil with physical parameters as in Table 2 along with , . We use the following initial and boundary conditions

    and run the simulation over a time period of . The solution at different times is shown in Figure 11.

    Figure 11.  Solution at different time steps of the permafrost examples. Left: results for homogeneous soil in Example 5.3; Right: simulation for heterogeneous soil in Example 5.3. Also shown is the interface separating the fine and coarse soil. Bottom: the corresponding curves at for the homogeneous and heterogeneous example, respectively.

    We see the temperature profile trending slowly towards the stationary distribution. When the simulation is run over a time period of at least hours, steady state is achieved in which the temperature qualitatively has an almost linear profile.

    Example 5.3. Heterogeneous case. We extend Example 5.2 to a heterogeneous case. We consider be composed of two different soil types: and . Thermal properties of water are as in Table 2, and we use . The difference in the properties of coarse and fine soil is in parameter in (5.5). We choose and . We start with the initial and boundary conditions

    and run the simulation over a time period of . The solution at different times is shown in Figure 11.

    The difference between Examples 5.2 and 5.3 is apparent from Figure 11. At a prominent jump in enthalpy appears near ; this effect is due to the soil heterogeneity.

    Next we aim to simulate the effect of permafrost warming due to variable temperature at the top boundary, and a possible effect of climate warming.

    Example 5.4. Homogeneous case with time dependent boundary condition. Let represent a column of soil of porosity , with the top boundary subject to time-varying temperature boundary conditions representing typical environmental change; see, e.g., [41](Figures 35) and [75] where the effect of climate warming is considered. At the bottom we apply fixed Dirichlet condition representing fixed temperature below the active layer. For simulation we use thermal properties of water from Table 2, along with , . Our goal is to simulate the temperature and enthalpy profile during the time period of years following an earlier period of time , initiated with

    We use the simulated (shown in the top row of Figure 12) as the initial data for the new time period , with

    Figure 12.  Top: and , for Example 5.4. Middle left: temperature plot. Included are (top boundary), and at the depth and for a period of years. Middle right: enthalpy plot at the corresponding . Bottom: temperature and enthalpy at the end of simulation [years].

    The simulation results are shown in Figure 12. The temperature at a depth and as a function of time is plotted along with the temperature and enthalpy at the end of years. We use the simulation results to get a rough estimate of the active layer thickness to be .

    The P0-P0 approximation scheme and the solver behaves similarly for permafrost problem (P) as for (ST) problem. For single material problems, the model is robust and features essentially mesh independent number of iterations under reasonable time steps.

    However, simulation in heterogeneous domains pose difficulties. The performance of Newton's method can be rather rugged, and is a (well-known) challenge [76]. For permafrost, the solver performance deteriorates in the case of highly disparate data, and/or and large time steps. Specifically, the former is an issue if there is a big difference, e.g., in the soil parameters (or ) in Section 5.1.

    One way to overcome this challenge is to use adaptive time stepping through which the time step is reduced if Newton iteration struggles to converge. Generally, reducing the time step leads to smoother convergence of residuals to zero. For very difficult cases related to high degree of heterogeneity, we regularize the problem by introducing an artificial "intermediate" layer in the region close to the interface in which the soil parameters vary more smoothly. Such a strategy can be used to find an initial guess for Newton step, or be used as the solution for a few initial time steps.

    More work on the solver and schemes is needed, but overall we believe our P0-P0 scheme performs well for the permafrost problem (P). We see that its behavior is smoother and the challenges are somewhat milder than those for the (ST) problem, except in heterogeneous domains.

    In this paper we consider a collection of models motivated by the applications to the phase change problems dubbed (ST), their (ST) modeling approximation, and the models (P) in permafrost.

    For approximation, we consider the technique well known for its conservative properties, mixed finite element method on rectangular elements. Because the solutions are expected to have low regularity, we choose to use only lowest order finite elements, with piecewise constant approximations to the scalar unknowns which we call P0-P0, and which are similar to finite volume or cell-centered finite difference approaches. For time discretization we employ a first order fully implicit scheme. We show that our P0-P0 approach works well for (ST) as well as (ST) and (P) permafrost, and that they compare well to the P1-based approaches that were primarily formulated for (ST). We extend the P0-P0 approach to heterogeneous case of multiple materials, and showed that a monolithic P0-P0 discretization we construct is robust and easy to implement. We set up theoretical framework relating to known literature results, proved several results, and point out the challenges and open questions.

    More work is underway. In particular, due to the challenges we acknowledged above, rigorous convergence analysis of P0-P0 is inaccessible at this time, even if it can be inferred in the limit of that for (ST). The main challenge is that that the use of for the heat flux is not directly possible for (ST). At the same time, our algorithms are robust even if they do not address this formal difficulty, but require more work which we defer to the future.

    On modeling side, in permafrost, or more generally, in freezing soils, there are additional micro-scale physical effects which contribute to the complex physics of phase change. These include freezing temperature depression, presence of air which acts as an insulator, configuration (geometry) of the pore space, as well as the chemical composition of liquid and mineral phases. While these are not fully explained, we aim to connect the computational models of multi-material pore-scale to the macro-scale experimental models such as those in Section 5.

    For (P) models in permafrost, in the paper we neglect convection and other physical phenomena which can potentially contribute to the temperature changes such as flow and mechanical behavior; these, along with other physical and environmental effects including those of radiation, vegetation and snow cover, and more, will be studied in forthcoming paper.

    The exact solution to (2.15) with data as in Example 2.1 is given by

    where is when phase change begins, is when all the solid has changed into liquid, and is the initial enthalpy.

    We consider a useful ODE system on with a matrix and right hand side

    (7.1)

    We define the solutions as limits of the solutions to the time-discrete problem

    (7.2)

    The result below will be used many times in this paper.

    Lemma 7.1. Let be maximal monotone and symmetric nonnegative definite. Then there exists a unique solution to (7.2).

    Proof. We find as the minimizer of

    where we set , with the (strictly convex) primitive of (i.e., , and the subgradient ). The functional is strictly convex due to the convexity of and strict convexity of , and thus it has a unique minimizer , and we set . With known , we find from , which follows from (7.2).

    Now we provide a detailed comparison of the P1-based formulations recalled in Section 2.4. We do so, for simplicity, for and a grid covering with cells with centers so that and . Other results not reviewed here include grid adaptive schemes in [77].

    We recalling that is identified by its nodal values . In turn, in P1-P0 approaches are identified by .

    We denote by the projection onto constants , by the projection onto , and by the Ritz projection onto .

    P1-P0 approximations: In [4]: one seeks such that

    (7.3a)
    (7.3b)

    Here we require , since (7.3b) implements the Chernoff formula. Note that the advantage of (7.3) is that (7.3a) is linear in . However, a consistency error arises since and , and

    Next, [5] are able to improve the convergence rates by modifying the scheme given by (7.3) through the use of regularization and projection. Their fully discrete regularized scheme reads [5]: find such that

    (7.4a)
    (7.4b)

    where is given by (7.5), essentially a Yosida approximation to ,

    (7.5)

    for some .

    The work of [44] uses a semi-implicit scheme for phase relaxation coupled with (2.6) given by They seek as an independent variable and solve

    (7.6a)
    (7.6b)

    where , with being the linear interpolator. We see that (7.6b) is equivalent to the Chernoff formula (7.3b) with . This shows the subtle difference in the scheme by [5] and [44]: in [44] numerical integration is employed whereas in [5] the projection operator is used to obtain .

    Remark 7.1. (7.6b) is obtained using the discretization

    (7.7)

    To account for in (7.7) instead of , the stability condition must be enforced [31].

    P1-P1 fully implicit approach in [3] A scheme of a different flavor is explored in [3]. Fully implicit in time formulation of (2.6) is approximated with

    (7.8)

    This discrete systems is, in practice, similar to (7.2) except it applies to (ST) and pays considerable attention to the identification and the role of in this last equation. While no regularization of is needed, the results are not accompanied by numerical computations. Optimal convergence rates [3](Theorem 2.7) apply to all models (1.2)

    (7.9)

    where is a constant such that

    (7.10)

    In particular, for (ST) and (2.8) we have . This corresponds to an order of convergence when . However, [3] do not include numerical implementation details or experiments.

    Other P1-P1 schemes include those in relaxation models [31].

    We carry out convergence analyses when the true solution is known, or when only its proxy, a fine grid solution is available. For time dependent problems the use of renders convergence order verification in some norms impractical. Therefore in this paper we only use norms; we also confine ourselves only to the error in scalar unknowns, deferring the discussion of the error in the fluxes to future work.

    As stated in Introduction, the notation means . We denote the norms appropriately, say as ; all these have to be approximated. In particular, we distinguish the grid norms

    We also use a tighter approximation than the grid norm

    The theoretically estimated convergence error depends, as usual, on and . Of the literature we have reviewed, [26,31,44] provide numerical results supporting their theoretically derived convergence rates. We set up additional experiments and convergence study as a prequel to the study of our P0-P0 schemes for (ST) problem: we use analytical solutions given in [26,44] for . The theoretical convergence orders as well as those we tested are tabulated in Table 13. We compare the [P1-P0] schemes given in [5,44] in Examples 3.1 and 3.2, respectively; see the plots shown in Figure 13. Additionally, we mention the work in [4] which provides a semi-discrete scheme which we further discretize in space following [5]. For this reason, along with using a small regularization parameter in (7.4), the numerical solution following the formulation as in [4] is virtually indistinguishable from [5] in Example 2 and hence is not shown in Figure 13.

    Table 13.  Convergence orders obtained in Examples 3.1 (VV) and 3.2 (RBC) for P1-based schemes from literature and for our P0-P0 scheme. Also listed is the order of convergence for derived theoretically in the literature which is denoted by (Lit.).
    Scheme - Case (Lit.) Notes
    P1-P0 [4] (VV) 0.47 0.52 0.52 0.25 0.27 0.53 0.24
    (RBC) 0.42 0.44 0.42 0.26 0.5 0.25
    P1-P0 [5] (VV) 0.84 0.84 0.91 0.66 0.51 0.96 0.52
    (RBC) 0.70 0.72 0.72 0.45 0.87 0.45
    P1-P0 [44] (VV) 0.66 0.69 0.73 0.50 0.35 0.70 0.35
    (RBC) 0.14 0.13 0.26 0.28 0.47 0.24
    P0-P0 (VV) 1.38 1.43 1.27 - 0.50 1.00 0.47
    (RBC) 1.08 1.09 1.07 0.45 0.96 0.47

     | Show Table
    DownLoad: CSV
    Figure 13.  Top row: Solutions with P1-P0 to (VV) Example 2 with at . Bottom row: Solutions P1-P0 to (RBC) Example 3 with at time .

    All together, we observe that is approximated well, qualitatively and quantitatively, by all the schemes. However, appears to be "smeared" near the interface and features higher errors. Of all the schemes, that in [5] produces the best approximation and convergence rates. The examples we show are consistent with the rates proved in [44]; see Table 13.

    We recall now one especially useful and well known feature of numerical implementation of (3.6) when numerical integration is used to get the entries of matrix . As discussed in [24,49], these follow from the use of trapezoidal (T) and midpoint (M) rules which leads to an algebraic expression involving the so-called transmissivity or transmissibility edge factor , where

    is the weighted scaled harmonic average of and so that (see, e.g., [24](eq.(15)))

    (7.11)

    We emphasize that the use of harmonic averages leads to conservative fluxes; this is in contrast with primal formulations in which arithmetic averages are used, and fluxes are not conservative.

    Next, we integrate and obtain, for the total normal flux across and the first component of

    (7.12a)

    The right hand side explains how the entries of matrix arise.

    A similar expression is obtained for and using .

    (7.12b)

    The use of factors allows the interpretation of (7.12) as a finite difference analogue of . It also explains why the matrix is diagonal.

    Last but not least, (7.11) is done also for the second component of flux across .

    Handling Dirichlet boundary conditions. The expression (7.12) is valid for the interior edges away from external boundaries and interfaces. If the edge is on a boundary on which Dirichlet condition is prescribed with some , then the analogue of (7.12) reads

    (7.13)

    The portion of the expression associated with moves to the right hand side of (3.3), but the structure of the matrix does not change.

    We consider the special linear case with piecewise constant coefficients of (4.1) from Example 4

    (7.14a)
    (7.14b)
    (7.14c)
    (7.14d)

    We also consider the stationary solution to (7.14) which satisfy

    (7.15a)
    (7.15b)
    (7.15c)

    It is not difficult to find its analytical solution, with and

    In particular in the special case , and we get

    (7.16)

    From this we have the jump scaling linearly with as , unlike predicted for the evolution problem in (4.3).

    This work was partially supported by the National Science Foundation DMS-1912938 and DMS-1522734, and by the NSF IRD plan 2019-21 for M. Peszynska.

    This material is based upon work supported by and while serving at the National Science Foundation. Any opinion, findings, and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation.

    The authors would like to thank the anonymous reviewers whose remarks helped to improve the paper.

    The authors declare there is no conflicts of interest.



    [1] E. Ahmed, J. Jaffré, J. E. Roberts, A reduced fracture model for two-phase flow with different rock types, Math. Comput. Simul., 137 (2017), 49–70. https://doi.org/10.1016/j.matcom.2016.10.005 doi: 10.1016/j.matcom.2016.10.005
    [2] C. Alboin, J. Jaffré, J. E. Roberts, X. Wang, C. Serres, Domain decomposition for some transmission problems in flow in porous media, Numer. Treat. Multiphase Flows Porous Media, 552 (2000), 22–34.
    [3] J. Rulla, N. J. Walkington, Optimal rates of convergence for degenerate parabolic problems in two dimensions, SIAM J. Numer. Anal., 33 (1996), 56–67. https://doi.org/10.1137/0733004 doi: 10.1137/0733004
    [4] E. Magenes, R. H. Nochetto, C. Verdi, Energy error estimates for a linear scheme to approximate nonlinear parabolic problems, ESAIM: M2AN, 21 (1987), 655–678.
    [5] R. H. Nochetto, C. Verdi, The combined use of a nonlinear Chernoff formula with a regularization procedure for two-phase Stefan problems, Numer. Funct. Anal. Optim., 9 (1988), 1177–1192. https://doi.org/10.1080/01630568808816279 doi: 10.1080/01630568808816279
    [6] J. A. Wheeler, Simulation of heat transfer from a warm pipeline buried in permafrost, Am. Inst. Chem. Eng., (1973), 267–284.
    [7] J. A. Wheeler, Permafrost thermal design for the trans-Alaska pipeline, Moving Boundary Probl., (1978), 267–284.
    [8] D. Nicolsky, V. Romanovsky, G. Tipenko, Using in-situ temperature measurements to estimate saturated soil thermal properties by solving a sequence of optimization problems, The Cryosphere, 1 (2007), 41–58. https://doi.org/10.5194/tc-1-41-2007 doi: 10.5194/tc-1-41-2007
    [9] S. Marchenko, V. Romanovsky, G. Tipenko, Numerical modeling of spatial permafrost dynamics in alaska, in Proceedings of Ninth International Conference on Permafrost, Ninth International Conference on Permafrost, (2008), 1125–1130.
    [10] E. E. Jafarov, S. S. Marchenko, V. E. Romanovsky, Numerical modeling of permafrost dynamics in alaska using a high spatial resolution dataset, The Cryosphere, 6 (2012), 613–624. https://doi.org/10.5194/tc-6-613-2012 doi: 10.5194/tc-6-613-2012
    [11] T. Kelley, J. Rulla, Solution of the discretized Stefan problem by Newton's method, Nonlinear Anal., 14 (1990), 851–872. https://doi.org/10.1016/0362-546X(90)90025-C doi: 10.1016/0362-546X(90)90025-C
    [12] M. F. Wheeler, M. Peszynska, Computational engineering and science methodologies for modeling and simulation of subsurface applications, Adv. Water Resour., 25 (2002), 1147–1173.
    [13] C. Dawson, S. Sun, M. F. Wheeler, Compatible algorithms for coupled flow and transport, Comput. Methods Appl. Mech. Eng., 193 (2004), 2565–2580. https://doi.org/10.1016/j.cma.2003.12.059 doi: 10.1016/j.cma.2003.12.059
    [14] R. L. Michalowski, A constitutive model of saturated soils for frost heave simulations, Cold Reg. Sci. Technol., 22 (1993), 47–63. https://doi.org/10.1016/0165-232X(93)90045-A doi: 10.1016/0165-232X(93)90045-A
    [15] Y. Zhang, R. Michalowski, Thermal-hydro-mechanical analysis of frost heave and thaw settlement, J. Geotech. Geoenviron. Eng., 141 (2015).
    [16] H. Liu, P. Maghoul, A. Shalaby, A. Bahari, Thermo-hydro-mechanical modeling of frost heave using the theory of poroelasticity for frost-susceptible soils in double-barrel culvert sites, Trans. Geotech., 20 (2019). https://doi.org/10.1016/j.trgeo.2019.100251 doi: 10.1016/j.trgeo.2019.100251
    [17] F. Yu, P. Guo, Y. Lai, D. Stolle, Frost heave and thaw consolidation modelling. part 2: One-dimensional thermohydromechanical (THM) framework, Can. Geotech. J., 57 (2020), 1595–1610.
    [18] M. Peszynska, A. Trykozko, Pore-to-core simulations of flow with large velocities using continuum models and imaging data, Comput. Geosci., 17 (2013), 623–645. https://doi.org/10.1007/s10596-013-9344-4 doi: 10.1007/s10596-013-9344-4
    [19] M. Peszynska, A. Trykozko, G. Iltis, S. Schlueter, D. Wildenschild, Biofilm growth in porous media: Experiments, computational modeling at the porescale, and upscaling, Adv. Water Res., 95 (2016), 288–301. https://doi.org/10.1016/j.advwatres.2015.07.008 doi: 10.1016/j.advwatres.2015.07.008
    [20] C. Shin, A. Alhammali, L. Bigler, N. Vohra, M. Peszynska, Coupled flow and biomass-nutrient growth at pore-scale with permeable biofilm, adaptive singularity and multiple species, Math. Biosci. Eng., 18 (2021), 2097–2149. https://doi.org/10.3934/mbe.2021108 doi: 10.3934/mbe.2021108
    [21] M. Peszynska, J. Umhoefer, C. Shin, Reduced model for properties of multiscale porous media with changing geometry, Computation, 9 (2021), 1–44.
    [22] T. Arbogast, M. F. Wheeler, N. Y. Zhang, A nonlinear mixed finite element method for a degenerate parabolic equation arising in flow in porous media, SIAM J. Numer. Anal., 33 (1996), 1669–1687. https://doi.org/10.1137/S0036142994266728 doi: 10.1137/S0036142994266728
    [23] C. S. Woodward, C. N. Dawson, Analysis of expanded mixed finite element methods for a nonlinear parabolic equation modeling flow into variably saturated porous media, SIAM J. Numer. Anal., 37 (2000), 701–724.
    [24] M. Peszynska, E. Jenkins, M. F. Wheeler, Boundary conditions for fully implicit two-phase flow model, in Recent Advances in Numerical Methods for Partial Differential Equations and Applications (eds. X. Feng and T. P. Schulze), Contemporary Mathematics Series, American Mathematical Society, 306 (2002), 85–106.
    [25] R. E. Showalter, Monotone operators in Banach space and nonlinear partial differential equations, vol. 49 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 1997. https://doi.org/10.1090/surv/049
    [26] J. C. Rogers, A. E. Berger, M. Ciment, The alternating phase truncation method for numerical solution of a Stefan problem, SIAM J. Numer. Anal., 16 (1979), 563–587.
    [27] A. Visintin, Models of phase transitions, vol. 28 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser Boston, Inc., Boston, MA, 1996. https://doi.org/10.1007/978-1-4612-4078-5
    [28] L. W. Lake, Enhanced oil recovery, Prentice Hall, 1989.
    [29] T. Roubicek, The Stefan problem in heterogeneous media, Ann. l'Inst. Henri Poincaré Anal. Linéaire, 6 (1989), 481–501.
    [30] E. Javierre, C. Vuik, F. Vermolen, S. van der Zwaag, A comparison of numerical models for one-dimensional Stefan problems, J. Comput. Appl. Math., 192 (2006), 445–459. https://doi.org/10.1016/j.cam.2005.04.062 doi: 10.1016/j.cam.2005.04.062
    [31] X. Jiang, R. Nochetto, A P1–P1 finite element method for a phase relaxation model Ⅰ: Quasiuniform mesh, Siam J. Numer. Anal., 35 (1998), 1176–1190.
    [32] S. M. Allen, J. M. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening, Acta Metall., 27 (1979), 1085–1095. https://doi.org/10.1016/0001-6160(79)90196-2 doi: 10.1016/0001-6160(79)90196-2
    [33] Y. Oono, S. Puri, Study of phase-separation dynamics by use of cell dynamical systems, I. modeling, Phys. Rev. A, 38 (1988), 434–453. https://link.aps.org/doi/10.1103/PhysRevA.38.434 doi: 10.1103/PhysRevA.38.434
    [34] J. F. Blowey, C. M. Elliott, The Cahn–Hilliard gradient theory for phase separation with non-smooth free energy part Ⅰ: Mathematical analysis, Eur. J. Appl. Math., 2 (1991), 233–280. https://doi.org/10.1017/S095679250000053X doi: 10.1017/S095679250000053X
    [35] P. Reddy, C. Gunasekar, A. Mhaske, V. Krishna, Enhancement of thermal conductivity of pcm using filler graphite powder materials, IOP Conf. Ser.: Mater. Sci. Eng., 402 (2018). https://doi.org/10.1088/1757-899X/402/1/012173 doi: 10.1088/1757-899X/402/1/012173
    [36] Rubitherm ® Technologies GmbH, 2021. https://www.rubitherm.eu
    [37] D. Yu, Z. He, Shape-remodeled macrocapsule of phase change materials for thermal energy storage and thermal management, Appl. Energy, 247 (2019), 503–516. https://doi.org/10.1016/j.apenergy.2019.04.072 doi: 10.1016/j.apenergy.2019.04.072
    [38] M. L. Cohen, Measurement of the thermal properties of human skin: A review, J. Invest. Dermatol., 69 (1977), 333–338. https://doi.org/10.1111/1523-1747.ep12507965 doi: 10.1111/1523-1747.ep12507965
    [39] Engineering Toolbox, 2021. https://www.engineeringtoolbox.com
    [40] Wikipedia, 2021. https://en.wikipedia.org
    [41] O. B. Andersland, B. Ladanyi, Frozen Ground Engineering, 2nd edition, Wiley, ASCE, Hoboken, 2004.
    [42] G. H. Meyer, Multidimensional Stefan problems, SIAM J. Numer. Anal., 10 (1973), 522–538. https://doi.org/10.1137/0710047 doi: 10.1137/0710047
    [43] J. W. Jerome, M. E. Rose, Error estimates for the multidimensional two-phase Stefan problem, Math. Comput., 39 (1982), 377–414. https://doi.org/10.1090/S0025-5718-1982-0669635-2 doi: 10.1090/S0025-5718-1982-0669635-2
    [44] C. Verdi, A. Visintin, Error estimates for a semi-explicit numerical scheme for Stefan-type problems., Numer. Math., 52 (1987/88), 165–186. http://eudml.org/doc/133231
    [45] D. Boffi, M. Fortin, F. Brezzi, Mixed Finite Element Methods and Applications, Springer series in computational mathematics, 2013.
    [46] A. Ern, J. L. Guermond, Theory and practice of finite elements, vol. 159 of Applied Mathematical Sciences, Springer-Verlag, New York, 2004. https://doi.org/10.1007/978-1-4757-4355-5
    [47] F. Brezzi, M. Fortin, Mixed and hybrid finite element methods, vol. 15 of Springer Series in Computational Mathematics, Springer-Verlag, New York, 1991.
    [48] A. Weiser, M. F. Wheeler, On convergence of block-centered finite differences for elliptic problems, SIAM J. Numer. Anal., 25 (1988), 351–375. https://doi.org/10.1137/0725025 doi: 10.1137/0725025
    [49] T. F. Russell, M. F. Wheeler, Finite element and finite difference methods for continuous flows in porous media, Math. Reservoir Simul., (1983), 35–106.
    [50] R. E. Showalter, Nonlinear degenerate evolution equations in mixed formulation, SIAM J. Math. Anal., 42 (2010), 2114–2131. https://doi.org/10.1137/100789427 doi: 10.1137/100789427
    [51] E. Schneid, P. Knabner, F. Radu, A priori error estimates for a mixed finite element discretization of the Richards' equation, Numer. Math., 98 (2004), 353–370. https://doi.org/10.1007/s00211-003-0509-2 doi: 10.1007/s00211-003-0509-2
    [52] M. Ulbrich, Semismooth Newton methods for variational inequalities and constrained optimization problems in function spaces, vol. 11 of MOS-SIAM Series on Optimization, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2011.
    [53] T. Roubicek, Numerical solution of the nonlinear heat equation in heterogeneous media, Numer. Funct. Anal. Optim., 11 (1990), 793–810.
    [54] T. Roubicek, A finite-element approximation of Stefan problems in heterogeneous media, in Free Boundary Value Problems, (1990), 267–275.
    [55] R. Glowinski, M. F. Wheeler, Domain decomposition and mixed finite element methods for elliptic problems, in First International Symposium on Domain Decomposition Methods for Partial Differential Equations (eds. R. Glowinski, G. H. Golub, G. A. Meurant and J. Periaux), SIAM, Philadelphia, (1988), 144–172.
    [56] A. Quarteroni, A. Valli, Domain decomposition methods for partial differential equations, Numerical Mathematics and Scientific Computation, Oxford University Press, New York, 1999. https://doi.org/10.1007/978-94-011-5412-38
    [57] I. Pawlow, A variational inequality approach to generalized two-phase Stefan problem in several space variables, Ann. Mat. Pura Appl., 131 (1982), 333–373. https://doi.org/10.1007/BF01765160 doi: 10.1007/BF01765160
    [58] M. Niezgodka, I. Pawłow, A generalized Stefan problem in several space variables, Appl. Math. Optim., 9 (1982), 193–224. https://doi.org/10.1007/BF01460125 doi: 10.1007/BF01460125
    [59] N. L. Gibson, F. P. Medina, M. Peszynska, R. E. Showalter, Evolution of phase transitions in methane hydrate, J. Math. Anal. Appl., 409 (2014), 816–833. https://doi.org/10.1016/j.jmaa.2013.07.023 doi: 10.1016/j.jmaa.2013.07.023
    [60] M. Peszynska, R. Showalter, J. Webster, Advection of methane in the hydrate zone: Model, analysis and examples, Mathe. Methods Appl. Sci., 38 (2015), 4613–4629. https://doi.org/10.1002/mma.3401 doi: 10.1002/mma.3401
    [61] M. Peszynska, C. Shin, Stability of a numerical scheme for methane transport in hydrate zone under equilibrium and non-equilibrium conditions, Comput. Geosci., 5 (2021), 1855–1886. https://doi.org/10.1007/s10596-021-10053-2 doi: 10.1007/s10596-021-10053-2
    [62] D. Foster, T. Costa, M. Peszynska, G. Schneider, Multiscale modeling of solar cells with interface phenomena, J. Coupled Syst. Multiscale Dyn., 1 (2013), 179–204. https://doi.org/10.1166/jcsmd.2013.1013 doi: 10.1166/jcsmd.2013.1013
    [63] T. Costa, D. Foster, M. Peszynska, Domain decomposition for heterojunction problems in semiconductors, in VECPAR 2014, High Performance Computing for Computational Science - VECPAR 2014, 11th International Conference, (2014), 92–101. http://arXiv.org/abs/1412.7946.
    [64] T. Costa, D. H. Foster, M. Peszynska, Progress in modeling of semiconductor structures with heterojunctions, J. Coupled Syst. Multiscale Dyn., 3 (2015), 66–86. https://doi.org/10.1166/jcsmd.2015.1066 doi: 10.1166/jcsmd.2015.1066
    [65] M. Discacciati, E. Miglio, A. Quarteroni, Mathematical and numerical models for coupling surface and groundwater flows, Appl. Numer. Math., 43 (2002), 57–74. https://doi.org/10.1016/S0168-9274(02)00125-3 doi: 10.1016/S0168-9274(02)00125-3
    [66] M. Sandells, D. Flocco, Introduction to the Physics of the Cryosphere, Morgan and Claypool, 2014.
    [67] T. Osterkamp, C. Burn, Permafrost, in Encyclopedia of Atmospheric Sciences, (2003), 1717–1729. https://doi.org/10.1016/B0-12-227090-8/00311-0
    [68] X. Zhang, Y. Wu, E. Zhai, P. Ye, Coupling analysis of the heat-water dynamics and frozen depth in a seasonally frozen zone, J. Hydrol., 593 (2021). https://doi.org/10.1016/j.jhydrol.2020.125603 doi: 10.1016/j.jhydrol.2020.125603
    [69] X. Zhang, E. Zhai, Y. Wu, D. Sun, Y. Lu, Theoretical and numerical analyses on hydro–thermal–salt–mechanical interaction of unsaturated salinized soil subjected to typical unidirectional freezing process, Int. J. Geomech., 21 (2021), 04021104. https://doi.org/10.1061/(ASCE)GM.1943-5622.0002036 doi: 10.1061/(ASCE)GM.1943-5622.0002036
    [70] J. Wettlaufer, M. G. Worster, Premelting dynamics, Annu. Rev. Fluid Mech., 38 (2006), 427–452. https://doi.org/10.1146/annurev.fluid.37.061903.175758 doi: 10.1146/annurev.fluid.37.061903.175758
    [71] A. W. Rempel, J. S. Wettlaufer, M. G. Worster, Premelting dynamics in a continuum model of frost heave, J. Fluid Mech., 498 (2004), 227–244. https://doi.org/10.1146/annurev.fluid.37.061903.175758 doi: 10.1146/annurev.fluid.37.061903.175758
    [72] C. W. Lovell, Temperature effects on phase composition and strength of partially-frozen soil, Highw. Res. Board Bull., 1957.
    [73] V. Romanovsky, T. Osterkamp, Effects of unfrozen water on heat and mass transport in the active layer and permafrost, Permafrost Periglacial Processes, 11 (2000), 219–239. https://doi.org/10.1002/1099-1530(200007/09) doi: 10.1002/1099-1530(200007/09)
    [74] Ulrich Hornung, Homogenization and porous media, vol. 6 of Interdisciplinary Applied Mathematics, Springer-Verlag, New York, 1997.
    [75] H. Zhang, J. Zhang, Z. Zhang, J. Chen, Y. You, A consolidation model for estimating the settlement of warm permafrost, Comput. Geotech., 76 (2016), 43–50. https://doi.org/10.1016/j.compgeo.2016.02.013 doi: 10.1016/j.compgeo.2016.02.013
    [76] C. T. Kelley, Iterative methods for linear and nonlinear equations, SIAM, Philadelphia, 1995.
    [77] M. Paolini, G. Sacchi, C. Verdi, Finite element approximations of singular parabolic problems, Int. J. Numer. Methods Eng., 26 (1988), 1989–2007. https://doi.org/10.1002/nme.1620260907 doi: 10.1002/nme.1620260907
  • This article has been cited by:

    1. Malgorzata Peszynska, Naren Vohra, Lisa Bigler, Upscaling an Extended Heterogeneous Stefan Problem from the Pore-Scale to the Darcy Scale in Permafrost, 2024, 22, 1540-3459, 436, 10.1137/23M1552000
    2. M. Peszynska, Z. Hilliard, N. Vohra, Coupled flow and energy models with phase change in permafrost from pore- to Darcy scale: Modeling and approximation, 2024, 450, 03770427, 115964, 10.1016/j.cam.2024.115964
    3. Naren Vohra, Malgorzata Peszynska, Iteratively coupled mixed finite element solver for thermo-hydro-mechanical modeling of permafrost thaw, 2024, 22, 25900374, 100439, 10.1016/j.rinam.2024.100439
    4. Naren Vohra, Malgorzata Peszynska, Robust conservative scheme and nonlinear solver for phase transitions in heterogeneous permafrost, 2024, 442, 03770427, 115719, 10.1016/j.cam.2023.115719
    5. Julieta Bollati, María F. Natale, José A. Semitiel, Domingo A. Tarzia, Determination of one unknown coefficient in a two-phase free boundary problem in an angular domain with variable thermal conductivity and specific heat, 2024, 531, 0022247X, 127775, 10.1016/j.jmaa.2023.127775
    6. Zizheng Zhang, Wenlong Tian, Yuli Hu, Bo Li, Numerical study on phase change propulsion mechanism of thermal underwater glider based on spectral method, 2024, 256, 13594311, 124066, 10.1016/j.applthermaleng.2024.124066
  • Reader Comments
  • © 2022 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(3326) PDF downloads(164) Cited by(6)

Article outline

Figures and Tables

Figures(13)  /  Tables(13)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog