This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quasistatic evolution and accretive phase-change
in finite-strain viscoelastic solids

Andrea Chiesa University of Vienna, Faculty of Mathematics and Vienna School of Mathematics, Oskar-Morgenstern-Platz 1, A-1090 Vienna, Austria andrea.chiesa@univie.ac.at http://www.mat.univie.ac.at/\simachiesa  and  Ulisse Stefanelli University of Vienna, Faculty of Mathematics, Oskar-Morgenstern-Platz 1, A-1090 Vienna, Austria, University of Vienna, Vienna Research Platform on Accelerating Photoreaction Discovery, Währingerstraße 17, 1090 Wien, Austria, and Istituto di Matematica Applicata e Tecnologie Informatiche E. Magenes, via Ferrata 1, I-27100 Pavia, Italy ulisse.stefanelli@univie.ac.at http://www.mat.univie.ac.at/\simstefanelli
Abstract.

We investigate the quasistatic evolution problem for a two-phase viscoelastic material at finite strains. Phase evolution is assumed to be irreversible as one phase accretes in time in its normal direction, at the expense of the other. Mechanical response depends on the phase, and growth is influenced by the mechanical state at the boundary of the accreting phase, making the model to be fully coupled. This setting is inspired by the early stage development of solid tumors, as well as by the swelling of polymer gels. We formulate the evolution problem by coupling the quasistatic balance of momenta in weak form and the growth dynamics in the viscosity sense. Both a diffused- and a sharp-interface variant of the model are proved to admit solutions.

Key words and phrases:
Accretive growth, viscoelastic solid, finite-strain, quasistatic evolution, variational formulation, viscosity solution, existence.
2020 Mathematics Subject Classification:
74F99, 74G22, 49L25

1. Introduction

This paper is concerned with the quasistatic evolution of a viscoelastic compressible solid undergoing a phase transition. We assume that the material presents two phases, of which one grows at the expense of the other by accretion. In particular, the phase-transition font evolves in a normal direction to the accreting phase, with a growing rate depending on the deformation. This behavior is indeed common to different material systems. It may be observed in the early stage development of solid tumors [6, 28, 56], where the neoplastic tissue invades the healthy one. Swelling in polymer gels also follows a similar dynamics, with the swollen phase accreting in the dry one [31, 50] and causing a volume increase. Accretive phase-growth can be observed in some solidification processes [43, 53], as well.

The focus of the modelization is on describing the interplay between mechanical deformation and accretion, On the one hand, the two phases are assumed to have a different mechanical response, having an effect on the viscoelastic evolution of the medium. On the other hand, the time-dependent mechanical deformation is assumed to influence the growth process, as is indeed common in biomaterials [21], polymeric gels [57], and solidification [44]. The mechanical and phase evolutions are thus fully coupled.

The state of the system is described by the pair (y,θ):[0,T]×Ud×[0,)(y,\theta):[0,T]\times U\to\mathbb{R}^{d}\times[0,\infty), where T>0T>0 is some final time and UU is the reference configuration of the body. Here, yy is the deformation of the medium while θ\theta determines its phase. More precisely, for all t[0,T]t\in[0,T] the accreting (growing) phase is identified as the sublevel Ω(t):={xU|θ(x)<t}\Omega(t):=\{x\in U\ |\ \theta(x)<t\}, whereas the receding phase corresponds to UΩ(t)¯U\setminus\overline{\Omega(t)}. The value θ(x)\theta(x) formally corresponds to the time at which the point xUx\in U is added to the growing phase. As such, θ\theta is usually referred to as time-of-attachment function. An illustration of the mechanical setting is given in Figure 1.

Refer to caption
Figure 1. Setting of the model in the sharp-interface case.

As growth processes and mechanical equilibration typically occur on very different time scales, we assume that the body is in quasistatic equilibrium. This calls for specifying the stored energy density W(θ(x)t,y)W(\theta(x){-}t,\nabla y) and the viscosity R(θ(x)t,y,y˙)R(\theta(x){-}t,\nabla y,\nabla\dot{y}) of the medium, as well as the applied body forces f(θ(x)t,x)f(\theta(x){-}t,x). All these quantities are assumed to be dependent on the phase via the sign of θ(x)t\theta(x){-}t, which indeed distinguishes the two phases, in the spirit of the celebrated level-set method [40, 45]. In addition, we include a second-gradient regularization term in the energy of the form H(2y)H(\nabla^{2}y), which we take to be phase independent, for simplicity. All in all, the quasistatic equilibrium system takes the form

div(yW(θ(x)t,y)+y˙R(θ(x)t,y,y˙)divDH(2y))=f(θ(x)t,x).{}-\operatorname{div}\left(\partial_{\nabla y}W(\theta(x){-}t,\nabla y)+\partial_{\nabla\dot{y}}R(\theta(x){-}t,\nabla y,\nabla\dot{y})-\operatorname{div}{\rm D}H(\nabla^{2}y)\right)=f(\theta(x){-}t,x). (1.1)

This system is solved weakly, complemented by mixed boundary conditions on yy and a homogeneous natural condition on the hyperstress divDH(2y)-\operatorname{div}\,{\rm D}H(\nabla^{2}y), namely,

y=idon[0,T]×ΓD,\displaystyle y={\rm id}\ \ \text{on}\ \ [0,T]\times\Gamma_{D}, (1.2)
DH(2y):(νν)=0on[0,T]×U,\displaystyle{\rm D}H(\nabla^{2}y){:}(\nu\otimes\nu)=0\ \ \text{on}\ \ [0,T]\times\partial U, (1.3)
(yW(θ(x)t,y)+y˙Rε(θ(x)t,y,y˙))ν\displaystyle\left(\partial_{\nabla y}W(\theta(x){-}t,\nabla y){+}\partial_{\nabla\dot{y}}R_{\varepsilon}(\theta(x){-}t,\nabla y,\nabla\dot{y})\right)\nu
divS(DH(2y))=0on[0,T]×ΓN,\displaystyle\quad-{\rm div}_{S}\,({\rm D}H(\nabla^{2}y))=0\ \ \text{on}\ \ [0,T]\times\Gamma_{N}, (1.4)

where ν\nu is the outer unit normal to U\partial U, ΓD\Gamma_{D} and ΓN\Gamma_{N} are the Dirichlet and Neumann part of the boundary U\partial U, respectively, and divS{\rm div}_{S} denotes the surface divergence on U\partial U [27].

The quasistatic equilibrium system is coupled to the phase evolution by requiring that the time-of-attachment function θ\theta solves the generalized eikonal equation

γ(y(θ(x),x),y(θ(x),x))|(θ)(x)|=1\gamma\big{(}y(\theta(x),x),\nabla y(\theta(x),x)\big{)}|\nabla(-\theta)(x)|=1 (1.5)

for all xx in the complement of a given initial set Ω0U\Omega_{0}\subset\subset U where we set θ=0\theta=0. This corresponds to assuming that Ω(t)\Omega(t) accretes in its normal direction, with growth rate γ()>0\gamma(\cdot)>0. More precisely, the evolution of the generic point x(t)x(t) on the boundary Ω(t)\partial\Omega(t) follows the ODE flow

x˙(t)=γ(y(t,x(t)),y(t,x(t)))n(x(t))\dot{x}(t)=\gamma\big{(}y(t,x(t)),\nabla y(t,x(t))\big{)}\,n(x(t))

where n(x(t))n(x(t)) indicates the normal to Ω(t)\partial\Omega(t) at x(t)x(t). Accretive growth is paramount to a wealth of different biological models [49], including plants and trees [15, 17] and the formation of hard tissues like horns or shells [36, 46, 51]. The dependence of the growth rate γ\gamma on the actual position and strain is intended to model the possible influence of local features such as nutrient concentrations, as well as of the local mechanical state [21].

Note that accretive growth occurs in a variety of nonbiological systems, as well. These include crystallization [29, 55], sedimentation of rocks [18], glacier formation, accretion of celestial bodies [8], as well as technological applications, from epitaxial deposition [32], to coating, masonry, and 3D printing [20, 30], just to mention a few.

By assuming smoothness and differentiating the equation θ(x(t))=t\theta(x(t))=t in time one obtains θ(x(t))x˙(t)=1\nabla\theta(x(t)){\cdot}\dot{x}(t)=1. This, together with the above flow rule for x(t)x(t) and n(x(t))=θ(x(t))/|θ(x(t))|n(x(t))=\nabla\theta(x(t))/|\nabla\theta(x(t))|, originates the generalized eikonal equation (1.5). Note that the growth rate γ\gamma in (1.5) depends on the actual deformation y(θ(x),x)y(\theta(x),x) and strain y(θ(x),x)\nabla y(\theta(x),x) at the growing interface, making the system (1.1)–(1.5) fully coupled.

We specify the initial conditions for the system by setting

θ=0onΩ0,\displaystyle\theta=0\ \ \text{on}\ \Omega_{0}, (1.6)
y(0,)=y0onU,\displaystyle y(0,\cdot)=y_{0}\ \ \text{on}\ U, (1.7)

where the initial deformation y0y_{0} and the initial portion of the growing phase Ω0\Omega_{0} are given. Note that Ω0\Omega_{0} is not required to be connected, in order to possibly model the onset of the accreting phase at different sites. On the other hand, the evolution described by (1.5) does not preserve the topology and disconnected accreting regions may coalesce over time.

The aim of this paper is to present an existence theory to the initial and boundary value problem (1.1)–(1.7). We tackle both a sharp-interface and a diffused-interface version of the model, by tuning the assumptions on WW and RR, see Sections 2.32.4. More precisely, in the diffused-interface model we assume that energy and dissipation densities change smoothly as functions of the phase indicator θ(x)t\theta(x){-}t across a region of width ε>0\varepsilon>0, namely for ε/2<θ(x)t<ε/2-\varepsilon/2<\theta(x)-t<\varepsilon/2. On the contrary, in the sharp-interface case material potentials are assumed to be discontinuous across the phase-change surface {θ(x)=t}\{\theta(x)=t\}.

In both regimes, we prove that the fully coupled system (1.1)–(1.7) admits a weak/viscosity solution, see Definition 2.1 below. More precisely, the quasistatic viscous evolution (1.1)–(1.4) is solved weakly, whereas the growth dynamics equation (1.5) is solved in the viscosity sense, see Theorem 2.1. We moreover prove that solutions fulfill the energy equality, where the energetic contribution of the phase transition is characterized, see Proposition 2.1. As a by-product, solutions of the diffused-interface model for ε>0\varepsilon>0 are proved to converge up to subsequences to solutions of the sharp-interface model as the parameter ε\varepsilon converges to 0, see Corollary 2.1.

Before going on, let us mention that the engineering literature on growth mechanics is vast. Without any claim of completeness, we mention [47, 58] and [33, 37, 38, 48, 52] as examples of linearized and finite-strain theories, respectively. On the other hand, mathematical existence theories in growth mechanics are scant, and we refer to [3, 12, 19] for some recent results. To the best of our knowledge, no existence result for finite-strain accretive-growth mechanics is currently available. In the linearized case, an existence result for the model [58] has been obtained in [13].

The paper is structured as follows. Section 2 is devoted to the statement of the main existence result, Theorem 2.1. In section 3, we give the proof of the energy identity. The proof of Theorem 2.1 is then split in Sections 4 and 5, respectively focusing on the diffused-interface and the sharp-interface setting. In the diffused-interface case, the proof relies on an iterative construction, where the mechanical and the growth problems are solved in alternation. The existence proof for the sharp-interface model is obtained by passing to the limit as ε0\varepsilon\to 0.

2. Main results

In this section, we specify assumptions, introduce the weak/viscosity notion of solution, and state the main results for problem (1.1)–(1.7).

2.1. Notation

In what follows, we denote by d×d\mathbb{R}^{d\times d} the Euclidean space of d×dd{\times}d real matrices, d2d\geq 2, by symd×d\mathbb{R}^{d\times d}_{\rm sym} the subspace of symmetric matrices, and by II the identity matrix. Given Ad×dA\in\mathbb{R}^{d\times d}, we indicate its transpose by AA^{\top} and its Frobenius norm by |A|2:=A:A|A|^{2}:=A{:}A, where the contraction product between matrices A,Bd×dA,\,B\in\mathbb{R}^{d\times d} is defined as A:B:=AijBijA{:}B:=A_{ij}B_{ij} (we use the summation convention over repeated indices). Analogously, let d×d×d\mathbb{R}^{d\times d\times d} be the set of real 33-tensors, and define their contraction product as AB:=AijkBijkA\smash{\vdots}B:=A_{ijk}B_{ijk} for A,Bd×d×dA,B\in\mathbb{R}^{d\times d\times d}. Moreover, given a symmetric positive definite 4-tensor d×d×d×d\mathbb{C}\in\mathbb{R}^{d\times d\times d\times d} and the matrix Ad×dA\in\mathbb{R}^{d\times d} we indicate by :Ad×d\mathbb{C}{:}A\in\mathbb{R}^{d\times d} and A:d×dA{:}\mathbb{C}\in\mathbb{R}^{d\times d} the matrices given in components by (:A)ij=ijkAk(\mathbb{C}{:}A)_{ij}=\mathbb{C}_{ijk\ell}A_{k\ell} and (A:)ij=AkCkij(A{:}\mathbb{C})_{ij}=A_{k\ell}C_{k\ell ij}, respectively. We shall use the following matrix sets

SL(d):={Ad×d|detA=1},\displaystyle{\rm SL}(d):=\{A\in\mathbb{R}^{d\times d}\;|\;\det A=1\},
SO(d):={ASL(d)|AA=I},\displaystyle{\rm SO}(d):=\{A\in{\rm SL}(d)\;|\;AA^{\top}=I\},
GL+(d):={Ad×d|detA>0}.\displaystyle{\rm GL}_{+}(d):=\{A\in\mathbb{R}^{d\times d}\;|\;\det A>0\}.

The scalar product of two vectors a,bda,\,b\in\mathbb{R}^{d} is classically indicated by aba{\cdot}b. The symbol BRdB_{R}\subset\mathbb{R}^{d} denotes the open ball of radius R>0R>0 and center 0d0\in\mathbb{R}^{d}, |E||E| indicates the Lebesgue measure of the Lebesgue-measurable set EdE\subset\mathbb{R}^{d}, and 𝟙E\mathbbm{1}_{E} is the corresponding characteristic function, namely, 𝟙E(x)=1\mathbbm{1}_{E}(x)=1 for xEx\in E and 𝟙E(x)=0\mathbbm{1}_{E}(x)=0 otherwise. For EdE\subset\mathbb{R}^{d} nonempty and xdx\in\mathbb{R}^{d} we define dist(x,E):=infeE|xe|{\rm dist}(x,E):=\inf_{e\in E}|x{-}e|. We denote by d1\mathcal{H}^{d-1} the (d1)(d{-}1)-dimensional Hausdorff measure.

In the following, we indicate by cc a generic positive constant, possibly depending on data but independent of ε\varepsilon and of the discretization step τ\tau. Note that the value of cc may change from line to line.

2.2. Admissible deformations

Fix the final time T>0T>0 and let the reference configuration UdU\subset\mathbb{R}^{d} be nonempty, open, connected, and bounded. We assume that the boundary U\partial U is Lipschitz, with ΓD,ΓNU\Gamma_{D},\,\Gamma_{N}\subset\partial U disjoint and open in the topology of U\partial U, ΓD\Gamma_{D}\not=\emptyset and ΓD¯ΓN¯=U\overline{\Gamma_{D}}\cap\overline{\Gamma_{N}}=\partial U, where the closure is taken in the topology of U\partial U. In the following, we use the short-hand notation Q:=(0,T)×UQ:=(0,T)\times U and ΣD:=(0,T)×ΓD\Sigma_{D}:=(0,T)\times\Gamma_{D}.

Deformations are assumed to belong to the space

WΓD2,p(U;d):={yW2,q(U;d)|y=id on ΓD},\displaystyle W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}):=\left\{y\in W^{2,q}(U;\mathbb{R}^{d})\,|\,y={\rm id}\text{ on }\Gamma_{D}\right\},

for almost all times, for some given p>dp>d fixed. Moreover, we impose local invertibility and orientation preservation. All in all, admissible deformations are defined as

𝒜{yWΓD2,p(U;d)|yGL+(d)a.e. in U}.{\mathcal{A}}\coloneqq\left\{y\in W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d})\ \Big{|}\ \nabla y\in{\rm GL}_{+}(d)\ \text{a.e. in }U\right\}.

2.3. Elastic energy

Let ε0\varepsilon\geq 0 be given and (hε)ε>0C(;[0,1])(h_{\varepsilon})_{\varepsilon>0}\subset C^{\infty}(\mathbb{R};[0,1]) be a sequence of increasing functions such that

hε(σ)={0ifσε/2,1ifσε/2,σhεL()2ε.h_{\varepsilon}(\sigma)=\begin{cases}0\quad&\text{if}\ \ \sigma\leq-\varepsilon/2,\\ 1\quad&\text{if}\ \ \sigma\geq\varepsilon/2,\end{cases}\quad\|\partial_{\sigma}h_{\varepsilon}\|_{L^{\infty}(\mathbb{R})}\leq\frac{2}{\varepsilon}. (2.1)

Moreover, let h0h_{0} be the discontinuous Heaviside-like function defined as h0(σ)=0h_{0}(\sigma)=0 if σ<0\sigma<0 and h0(σ)=1h_{0}(\sigma)=1 if σ0\sigma\geq 0. Note in particular that hεh0h_{\varepsilon}\rightarrow h_{0} in {0}\mathbb{R}\setminus\{0\} as ε0\varepsilon\rightarrow 0.

We define the elastic energy density Wε:×GL+(d)[0,)W_{\varepsilon}:\mathbb{R}\times{\rm GL}_{+}(d)\to[0,\infty) as

Wε(σ,F)(1hε(σ))Wa(F)+hε(σ)Wr(F).W_{\varepsilon}(\sigma,F)\coloneqq(1-h_{\varepsilon}(\sigma))W^{a}(F)+h_{\varepsilon}(\sigma)W^{r}(F). (2.2)

Here, σ\sigma is a placeholder for θ(x)t\theta(x){-}t, whose 0-sublevel set {xUθ(x)<t}\{x\in U\mid\theta(x)<t\} represents the accreting phase at time tt. In particular, Wε(σ,)=WaW_{\varepsilon}(\sigma,\cdot)=W^{a} for σ<ε/2\sigma<-\varepsilon/2, so that WaW^{a} is the elastic energy density of the accreting phase. On the other hand, Wε(σ,)=WrW_{\varepsilon}(\sigma,\cdot)=W^{r} for σ>ε/2\sigma>\varepsilon/2 and WrW^{r} is the elastic energy density of the receding phase.

On the two elastic-energy densities of the pure phases we require

Wa,WrC1(GL+(d);[0,)),\displaystyle W^{a},\,W^{r}\in C^{1}({\rm GL}_{+}(d);[0,\infty)), (2.3)
cW>0,q>pdpd:Wa(F),Wr(F)cW|F|2+cW(detF)qFGL+(d).\displaystyle\exists c_{W}>0,\ \exists q>\frac{pd}{p-d}:\quad W^{a}(F),\,W^{r}(F)\geq c_{W}|F|^{2}+\frac{c_{W}}{(\det F)^{q}}\quad\forall F\in{\rm GL}_{+}(d). (2.4)

Moreover, although not strictly needed for the analysis we require the frame indifference

Wa(QF)=Wa(F),Wr(QF)=Wr(F)FGL+(d),QSO(d).W^{a}(QF)=W^{a}(F),\ W^{r}(QF)=W^{r}(F)\quad\forall F\in{\rm GL}_{+}(d),\ Q\in{\rm SO}(d). (2.5)

As regards the second-order potential HH we ask for

HC1(d×d×d;[0,))convex,\displaystyle H\in C^{1}(\mathbb{R}^{d\times d\times d};[0,\infty))\ \ \text{convex}, (2.6)
H(QG)=H(G)for allGd×d×d,QSO(d),\displaystyle H(QG)=H(G)\ \ \text{for all}\ \ G\in\mathbb{R}^{d\times d\times d},\ Q\in{\rm SO}(d), (2.7)
0<cHCH:\displaystyle\exists 0<c_{H}\leq C_{H}:
cH|G|pH(G)CH(1+|G|)p,|DH(G)|CH|G|p1Gd×d×d,\displaystyle\qquad c_{H}|G|^{p}\leq H(G)\leq C_{H}(1+|G|)^{p},\quad|{\rm D}H(G)|\leq C_{H}|G|^{p-1}\quad\forall G\in\mathbb{R}^{d\times d\times d}, (2.8)
cH|G1G2|p(DH(G1)DH(G2))(G1G2)G1,G2d×d×d.\displaystyle\qquad c_{H}|G_{1}-G_{2}|^{p}\leq({\rm D}H(G_{1})-{\rm D}H(G_{2})){\vdots}(G_{1}-G_{2})\quad\forall G_{1},\,G_{2}\in\mathbb{R}^{d\times d\times d}. (2.9)

Again, the frame-indifference requirement (2.7) is not strictly needed for the analysis.

By integrating over the reference configuration UU we define 𝒲ε:C(U)×𝒜[0,)\mathcal{W}_{\varepsilon}\colon C(U)\times{\mathcal{A}}\rightarrow[0,\infty) and :WΓD2,p(U;d)[0,)\mathcal{H}\colon W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d})\rightarrow[0,\infty) as

𝒲ε(σ,y)UWε(σ,y)dxand(y)UH(2y)dx.\mathcal{W}_{\varepsilon}(\sigma,y)\coloneqq\int_{U}W_{\varepsilon}(\sigma,\nabla y)\,{\rm d}x\quad\text{and}\quad\mathcal{H}(y)\coloneqq\int_{U}H(\nabla^{2}y)\,{\rm d}x.

2.4. Viscous dissipation

We define the instantaneous viscous dissipation density Rε:×GL+(d)×d×d[0,)R_{\varepsilon}:\mathbb{R}\times{\rm GL}_{+}(d)\times\mathbb{R}^{d\times d}\to[0,\infty) as

Rε(σ,F,F˙)(1hε(σ))Ra(F,F˙)+hε(σ)Rr(F,F˙)R_{\varepsilon}(\sigma,F,\dot{F})\coloneqq(1-h_{\varepsilon}(\sigma))R^{a}(F,\dot{F})+h_{\varepsilon}(\sigma)R^{r}(F,\dot{F}) (2.10)

Here, Ra,Rr:GL+(d)×d×d[0,)R^{a},\,R^{r}:{\rm GL}_{+}(d)\times\mathbb{R}^{d\times d}\rightarrow[0,\infty) are assumed to be quadratic in the rate of the Cauchy–Green tensor, namely

Ra(F,F˙)12C˙:𝔻a(C):C˙,Rr(F,F˙)12C˙:𝔻r(C):C˙FGL+(d),F˙d×dR^{a}(F,\dot{F})\coloneqq\frac{1}{2}\dot{C}{:}\mathbb{D}^{a}(C){:}\dot{C},\quad R^{r}(F,\dot{F})\coloneqq\frac{1}{2}\dot{C}{:}\mathbb{D}^{r}(C){:}\dot{C}\qquad\forall F\in{\rm GL}_{+}(d),\ \dot{F}\in\mathbb{R}^{d\times d}

with CFFC\coloneqq F^{\top}F and C˙F˙F+FF˙\dot{C}\coloneqq\dot{F}^{\top}F+F^{\top}\dot{F}. We assume that

𝔻a,𝔻rC(symd×d;d×d×d×d)with(𝔻i)jkm=(𝔻i)kjm=(𝔻i)mjk\displaystyle\mathbb{D}^{a},\,\mathbb{D}^{r}\in C(\mathbb{R}^{d\times d}_{\rm sym};\mathbb{R}^{d\times d\times d\times d})\ \ \text{with}\ \ (\mathbb{D}^{i})_{jk\ell m}=(\mathbb{D}^{i})_{kj\ell m}=(\mathbb{D}^{i})_{\ell mjk}
j,k,,m=1,,d,fori=a,r,\displaystyle\quad\forall j,\,k,\,\ell,\,m=1,\dots,d,\ \text{for}\ i=a,\,r, (2.11)
c𝔻>0:c𝔻|C˙|2C˙:𝔻i(C):C˙C,C˙symd×d,fori=a,r.\displaystyle\exists c_{\mathbb{D}}>0:\quad c_{\mathbb{D}}|\dot{C}|^{2}\leq\dot{C}{:}\mathbb{D}^{i}(C){:}\dot{C}\quad\forall C,\,\dot{C}\in\mathbb{R}^{d\times d}_{\rm sym},\ \text{for}\ i=a,\,r. (2.12)

Here, RaR^{a} and RrR^{r} are the instantaneous viscous dissipation densities of the accreting and of the receding phase, respectively. Notice that this specific choice of RεR_{\varepsilon} ensures that

F˙Rε(σ,F,F˙)\displaystyle\partial_{\dot{F}}R_{\varepsilon}(\sigma,F,\dot{F}) =2(1hε(σ))F𝔻a(C):C˙+2hε(σ)F𝔻r(C):C˙\displaystyle=2(1{-}h_{\varepsilon}(\sigma))F\mathbb{D}^{a}(C){:}\dot{C}+2h_{\varepsilon}(\sigma)F\mathbb{D}^{r}(C){:}\dot{C}
=2(1hε(σ))F𝔻a(FF):(F˙F+FF˙)+2hε(σ)F𝔻r(FF):(F˙F+FF˙),\displaystyle=2(1{-}h_{\varepsilon}(\sigma))F\mathbb{D}^{a}(F^{\top}F){:}(\dot{F}^{\top}F{+}F^{\top}\dot{F})+2h_{\varepsilon}(\sigma)F\mathbb{D}^{r}(F^{\top}F){:}(\dot{F}^{\top}F{+}F^{\top}\dot{F}),

which is of course linear in F˙\dot{F}. By integrating on the reference configuration UU we define ε:C(U)×WΓD2,p(U;d)×H1(U;d)[0,)\mathcal{R}_{\varepsilon}\colon C(U)\times W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d})\times H^{1}(U;\mathbb{R}^{d})\rightarrow[0,\infty) as

ε(σ,y,y˙)URε(σ,y,y˙)dx.\mathcal{R}_{\varepsilon}(\sigma,y,\dot{y})\coloneqq\int_{U}R_{\varepsilon}(\sigma,\nabla y,\nabla\dot{y})\,{\rm d}x.

2.5. Loading and initial data

We assume that the body force density f=f(σ,x)f=f(\sigma,x) is (constant in time and) suitably smooth with respect to σ\sigma, namely

fW1,(;L2(U;d)).f\in W^{1,\infty}(\mathbb{R};L^{2}(U;\mathbb{R}^{d})). (2.13)

The σ\sigma-dependence of the force density ff is intended to cover the case of gravitation f=ρgf=\rho g, where the density ρ\rho depends on the phase, while the acceleration field gg is given.

We moreover assume that the initial deformation satisfies

y0𝒜withUWa(y0)+Wr(y0)+H(2y0)dx<.\displaystyle y_{0}\in\mathcal{A}\ \ \text{with}\ \int_{U}W^{a}(\nabla y_{0})+W^{r}(\nabla y_{0})+H(\nabla^{2}y_{0})\,{\rm d}x<\infty. (2.14)

In the classical setting one would have Wa(I)=Wr(I)=0W^{a}(I)=W^{r}(I)=0 and H(0)=0H(0)=0, so that one can simply choose y0=idy_{0}={\rm id}.

2.6. Growth

Concerning the accretive-growth model we ask for

γC0,1(d×GL+(d);(0,))withcγγ()Cγfor some 0<cγCγ.\displaystyle\gamma\in C^{0,1}(\mathbb{R}^{d}\times{\rm GL}_{+}(d);(0,\infty))\ \ \text{with}\ \ c_{\gamma}\leq\gamma(\cdot)\leq C_{\gamma}\ \ \text{for some}\ \ 0<c_{\gamma}\leq C_{\gamma}. (2.15)

Let moreover the initial location of the accreting phase be given by

Ω0UwithΩ0open andΩ0+BCγTU.\emptyset\not=\Omega_{0}\subset\subset U\ \ \text{with}\ \ \Omega_{0}\ \ \text{open and}\ \ \Omega_{0}+B_{C_{\gamma}T}\subset\subset U. (2.16)

As it will be clarified later, this last requirement guarantees that the accreting phase does not touch the boundary U\partial U over the time interval [0,T][0,T], see (4.14) below.

2.7. Main results

Assumptions (2.3)–(2.16) will be tacitly assumed throughout the remainder of the paper. We are interested in solving (1.1)–(1.7) in the following weak/viscosity sense.

Definition 2.1 (Weak/viscosity solution).

We say that a pair

(y,θ)(L(0,T;W2,p(U;d))H1(0,T;H1(U;d)))×C0,1(U¯)(y,\theta)\in\left(L^{\infty}(0,T;W^{2,p}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d}))\right)\times C^{0,1}(\overline{U})

is a weak/viscosity solution to the initial-boundary-value problem (1.1)–(1.7) if y(t,)𝒜y(t,\cdot)\in\mathcal{A} for all t(0,T)t\in(0,T), y(0,)=y0y(0,\cdot)=y_{0}, and

0TU(FWε(θt,y)+F˙Rε(θt,y,y˙)):z+DH(2y)2zdxdt\displaystyle\int_{0}^{T}\int_{U}\big{(}\partial_{F}W_{\varepsilon}(\theta{-}t,\nabla y){+}\partial_{\dot{F}}R_{\varepsilon}(\theta{-}t,\nabla y,\nabla\dot{y})\big{)}{:}\nabla z+{\rm D}H\left(\nabla^{2}y\right){\vdots}\nabla^{2}z\,{\rm d}x\,{\rm d}t
=0TUf(θt)zdxdtzC([0,T]×U¯;d)withz=0onΣD,\displaystyle\qquad=\int_{0}^{T}\int_{U}f(\theta{-}t){\cdot}z\,{\rm d}x\,{\rm d}t\quad\forall z\in C^{\infty}([0,T]\times\overline{U};\mathbb{R}^{d})\ \text{with}\ z=0\ \text{on}\ \Sigma_{D}, (2.17)

and θ\theta is a viscosity solution to

γ(y(θ(x),x),y(θ(x),x))|(θ(x))|=1inUΩ0¯,\displaystyle\gamma(y(\theta(x),x),\nabla y(\theta(x),x))|\nabla(-\theta(x))|=1\ \ \text{in}\ \ U\setminus\overline{\Omega_{0}}, (2.18)
θ(x)=0inΩ0.\displaystyle\theta(x)=0\ \ \text{in}\ \ \Omega_{0}. (2.19)

Namely, θ\theta satisfies (2.19), and, for all x0UΩ0¯x_{0}\in U\setminus\overline{\Omega_{0}} and any smooth function φ:U\varphi:U\to\mathbb{R} with φ(x0)=θ(x0)\varphi(x_{0})=-\theta(x_{0}) and φθ\varphi\geq-\theta (φθ\varphi\leq-\theta, respectively) in a neighborhood of x0x_{0}, it holds that γ(y(θ(x0),x0),y(θ(x0),x0))|φ(x0))|1(1,respectively)\gamma(y(\theta(x_{0}),x_{0}),\nabla y(\theta(x_{0}),x_{0}))|\nabla\varphi(x_{0}))|\leq 1(\geq 1,\ \text{respectively}). In particular,

0<1Cγ|θ|1cγa.e. inU.0<\frac{1}{C_{\gamma}}\leq|\nabla\theta|\leq\frac{1}{c_{\gamma}}\ \ \text{a.e. in}\ \ U. (2.20)

Note that this weak notion of solution still entails the validity of an energy equality.

Proposition 2.1 (Energy equality).

Under the assumptions (2.3)–(2.16), in the diffused-interface case ε>0\varepsilon>0 a weak/viscosity solution (y,θ)(y,\theta) fulfills for all t[0,T]t\in[0,T] the energy equality

U(Wε(θt,y)+H(2y)f(θt)y)dxU(Wε(θ,y0)+H(2y0)f(θ)y0)dx\displaystyle\int_{U}\big{(}W_{\varepsilon}(\theta{-}t,\nabla y)+H(\nabla^{2}y)-f(\theta{-}t){\cdot}y\big{)}\,{\rm d}x-\int_{U}\left(W_{\varepsilon}(\theta,\nabla y_{0})+H(\nabla^{2}y_{0})-f(\theta){\cdot}y_{0}\right)\,{\rm d}x
=20tURε(θs,y,y˙)dxds0tUf˙(θs)ydxds\displaystyle\quad=-2\int_{0}^{t}\!\!\int_{U}R_{\varepsilon}\left(\theta{-}s,\nabla y,\nabla\dot{y}\right)\,{\rm d}x\,{\rm d}s-\int_{0}^{t}\int_{U}\dot{f}(\theta{-}s){\cdot}y\,{\rm d}x\,{\rm d}s
0tUσWε(θs,y)dxds.\displaystyle\qquad-\int_{0}^{t}\!\!\int_{U}\partial_{\sigma}W_{\varepsilon}(\theta{-}s,\nabla y)\,{\rm d}x\,{\rm d}s. (2.21)

In the sharp-interface case ε=0\varepsilon=0, for all t[0,T]t\in[0,T], one has instead

U(W0(θt,y)+H(2y)f(θt)y)dxU(W0(θ,y0)+H(2y0)f(θ)y0)dx\displaystyle\int_{U}\big{(}W_{0}(\theta{-}t,\nabla y)+H(\nabla^{2}y)-f(\theta{-}t){\cdot}y\big{)}\,{\rm d}x-\int_{U}\left(W_{0}(\theta,\nabla y_{0})+H(\nabla^{2}y_{0})-f(\theta){\cdot}y_{0}\right)\,{\rm d}x
=20tUR0(θs,y,y˙)dxds0tUf˙(θs)ydxds\displaystyle\quad=-2\int_{0}^{t}\!\!\int_{U}R_{0}\left(\theta{-}s,\nabla y,\nabla\dot{y}\right)\,{\rm d}x\,{\rm d}s-\int_{0}^{t}\int_{U}\dot{f}(\theta{-}s){\cdot}y\,{\rm d}x\,{\rm d}s
0t{θ=s}Wr(y)Wa(y)|θ|dd1ds.\displaystyle\qquad-\int_{0}^{t}\!\!\int_{\{\theta=s\}}\frac{W^{r}(\nabla y)-W^{a}(\nabla y)}{|\nabla\theta|}\,{\rm d}\mathcal{H}^{d-1}\,{\rm d}s. (2.22)

Relations (2.21)–(2.22) express the energy balance in the model. In particular, the difference between the actual and the initial complementary energies (left-hand side in (2.21)–(2.22)) equals the sum of the total viscous dissipation, the work of external forces, and the energy stored in the medium in connection with the phase-transition process (the three terms in the right-hand side of (2.21)–(2.22), up to sign). Proposition 2.1 is proved in Section 3.

Our main result reads as follows.

Theorem 2.1 (Existence).

Under assumptions (2.3)–(2.16), for all given ε0\varepsilon\geq 0 there exists a weak/viscosity solution (y,θ)(y,\theta) of problem (1.1)–(1.7).

A proof of Theorem 2.1 in the diffused-interface case of ε>0\varepsilon>0 is presented in Section 4. The argument is based on an iterative strategy: for given yiy^{i} one finds the unique viscosity solution θi\theta^{i} to (2.18)–(2.19) (with yy replaced by yiy^{i}). Then, given θi\theta^{i} one can find yi+1y^{i+1} satisfying (2.1) (with θ\theta replaced by θi\theta^{i}). Note that such yi+1y^{i+1} may be nonunique. We however do not proceed via a fixed-point argument for multivalued maps (see, e.g., [24]) as the set of solutions yy for given θ\theta is generally not convex. To bypass this issue, we resort in directly proving the convergence of the iterative procedure.

Eventually, the proof of Theorem 2.1 in the sharp-interface case ε=0\varepsilon=0 will be obtained in Section 5 by passing to the limit as ε0\varepsilon\to 0 along a subsequence of weak/viscosity solutions (yε,θε)(y_{\varepsilon},\theta_{\varepsilon}) for ε>0\varepsilon>0. As a by-product, we have the following.

Corollary 2.1 (Sharp-interface limit).

Under assumptions (2.3)–(2.16), let (yε,θε)(y_{\varepsilon},\theta_{\varepsilon}) be weak/viscosity solutions of the diffused-interface problem (1.1)–(1.7) for ε>0\varepsilon>0. Then, there exists a not relabeled subsequence such that (yε,θε)(y,θ)(y_{\varepsilon},\theta_{\varepsilon})\to(y,\theta) strongly in C(U¯;d×)C(\overline{U};\mathbb{R}^{d}\times\mathbb{R}), where (y,θ)(y,\theta) is a weak/viscosity solution to the sharp-interface problem ε=0\varepsilon=0.

Before moving on, let us mention that the assumptions on the energy and of the instantaneous viscous dissipation density could be generalized by not requiring the specific forms (2.2) and (2.10). In fact, one could directly assume to be given Wε=Wε(σ,F)W_{\varepsilon}=W_{\varepsilon}(\sigma,F) and Rε=Rε(σ,F,F˙)R_{\varepsilon}=R_{\varepsilon}(\sigma,F,\dot{F}) of the form

Rε(σ,F,F˙)=12C˙:𝔻(σ,C):C˙R_{\varepsilon}(\sigma,F,\dot{F})=\frac{1}{2}\dot{C}{:}{\mathbb{D}}(\sigma,C){:}\dot{C}

with 𝔻C(×symd×d;d×d×d×d){\mathbb{D}}\in C(\mathbb{R}\times\mathbb{R}_{\rm sym}^{d\times d};\mathbb{R}^{d\times d\times d\times d}) by suitably adapting the smoothness and coercivity assumptions. Although the existence analysis could be carried out in this more general situation with no difficulties, we prefer to stick to the concrete choice of (2.2) and (2.10) as it allows a more transparent distinction of the diffused- and sharp-interface cases.

Moreover, let us point out that the admissible deformation yεy_{\varepsilon} is presently required to be solely locally injective, by means of the constraint detyε>0\det\nabla y_{\varepsilon}>0. On the other hand, global injectivity may also be enforced, in the spirit of [25], see also [41] in the static and [9, 10] in the dynamic case. This however calls for keeping track of reaction forces due to a possible self contact at the boundary ΓN\Gamma_{N}. From the technical viewpoint, one would need to include an extra variable in the state in order to model such reaction. The existence theory of Theorem 2.1 can be extended to cover this case, at the price of some notational intricacies. We however prefer to avoid discussing global injectivity here, for the sake of exposition clarity.

3. Proof of Proposition 2.1: energy equalities

We firstly consider the diffused-interface setting. Let (y,θ)(y,\theta) be a weak/viscosity solution to (1.1)–(1.7) for ε>0\varepsilon>0. The weak formulation (2.1) implies that equation (1.1) is actually solved in the distributional sense. In particular, owing to the regularity of yy and the standing assumptions, by comparison in the distributional equation one can conclude that

divdivDH(2yε)L2(0,T;(H1(U;d×d)))\operatorname{div}\operatorname{div}{\rm D}H(\nabla^{2}y_{\varepsilon})\in L^{2}(0,T;(H^{1}(U;\mathbb{R}^{d\times d}))^{*})

where (H1(U;d×d))(H^{1}(U;\mathbb{R}^{d\times d}))^{*} is the dual of H1(U;d×d)H^{1}(U;\mathbb{R}^{d\times d}). Hence, by using the duality product ,\langle\cdot,\cdot\rangle between (H1(U;d×d))(H^{1}(U;\mathbb{R}^{d\times d}))^{*} and H1(U;d×d)H^{1}(U;\mathbb{R}^{d\times d}) one can actually test the equation by y˙L2(0,T;H1(U;d×d))\dot{y}\in L^{2}(0,T;H^{1}(U;\mathbb{R}^{d\times d})) and integrate on [0,t][0,t] getting

0tUFWε(θs,y):y˙dxds+0tUF˙Rε(θs,y,y˙):y˙dxds\displaystyle\int_{0}^{t}\!\!\int_{U}\partial_{F}W_{\varepsilon}(\theta{-}s,\nabla y){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s+\int_{0}^{t}\!\!\int_{U}\partial_{\dot{F}}R_{\varepsilon}(\theta{-}s,\nabla y,\nabla\dot{y}){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s
+0tdivdivDH(2y),y˙ds=0tUf(θs)y˙dxds.\displaystyle\quad+\int_{0}^{t}\langle\operatorname{div}\operatorname{div}{\rm D}H(\nabla^{2}y),\dot{y}\rangle\,{\rm d}s=\int_{0}^{t}\!\!\int_{U}f(\theta{-}s){\cdot}\dot{y}\,{\rm d}x\,{\rm d}s. (3.1)

We readily have that

UWε(θt,y)dxUWε(θ,y0)dx=0tddtUFWε(θs,y)dxds\displaystyle\int_{U}W_{\varepsilon}(\theta{-}t,\nabla y)\,{\rm d}x-\int_{U}W_{\varepsilon}(\theta,\nabla y_{0})\,{\rm d}x=\int_{0}^{t}\frac{{\rm d}}{{\rm d}t}\int_{U}\partial_{F}W_{\varepsilon}(\theta{-}s,\nabla y)\,{\rm d}x\,{\rm d}s
=0tUFWε(θs,y):y˙dxds0tUσWε(θs,y)dxds.\displaystyle\quad=\int_{0}^{t}\!\!\int_{U}\partial_{F}W_{\varepsilon}(\theta{-}s,\nabla y){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s-\int_{0}^{t}\!\!\int_{U}\partial_{\sigma}W_{\varepsilon}(\theta{-}s,\nabla y)\,{\rm d}x\,{\rm d}s. (3.2)

Moreover, it is a standard matter to check that F˙Rε(σ,F,F˙):F˙=2Rε(σ,F,F˙)\partial_{\dot{F}}R_{\varepsilon}(\sigma,F,\dot{F}){:}\dot{F}=2R_{\varepsilon}(\sigma,F,\dot{F}), so that

0tUF˙Rε(θs,y,y˙):y˙dxds=20tURε(θs,y,y˙)dxds.\int_{0}^{t}\!\!\int_{U}\partial_{\dot{F}}R_{\varepsilon}(\theta{-}s,\nabla y,\nabla\dot{y}){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s=2\int_{0}^{t}\!\!\int_{U}R_{\varepsilon}(\theta{-}s,\nabla y,\nabla\dot{y})\,{\rm d}x\,{\rm d}s. (3.3)

The energy equality (2.21) hence follows by applying the chain rule [35, Prop. 3.6] to the duality term, owing to the fact that HH is convex, see (2.6), namely,

0tdivdivDH(2y),y˙ds=0tddsUH(2y)dxds\displaystyle\int_{0}^{t}\langle\operatorname{div}\operatorname{div}{\rm D}H(\nabla^{2}y),\dot{y}\rangle\,{\rm d}s=\int_{0}^{t}\frac{{\rm d}}{{\rm d}s}\int_{U}H(\nabla^{2}y)\,{\rm d}x\,{\rm d}s
=UH(2y(t,x))dxUH(2y0)dx.\displaystyle\quad=\int_{U}H(\nabla^{2}y(t,x))\,{\rm d}x-\int_{U}H(\nabla^{2}y_{0})\,{\rm d}x. (3.4)

The proof of energy equality (2.22) for the sharp-interface case ε=0\varepsilon=0 follows the same strategy, as one can again establish (3.1) (for W0W_{0} in place of WεW_{\varepsilon}) and (3.4). A notable difference is however in (3.2), which now requires some extra care as h0h_{0} is discontinuous. In particular, the energy equality (2.22) follows as soon as we prove that

UW0(θt,y)dxUW0(θ,y0)dx\displaystyle\int_{U}W_{0}(\theta{-}t,\nabla y)\,{\rm d}x-\int_{U}W_{0}(\theta,\nabla y_{0})\,{\rm d}x
=0tUFW0(θs,y):y˙dxds0t{θ=s}Wr(y)Wa(y)|θ|dd1ds.\displaystyle\quad=\int_{0}^{t}\!\!\int_{U}\partial_{F}W_{0}(\theta{-}s,\nabla y){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s-\int_{0}^{t}\!\!\int_{\{\theta=s\}}\frac{W^{r}(\nabla y)-W^{a}(\nabla y)}{|\nabla\theta|}\,{\rm d}\mathcal{H}^{d-1}\,{\rm d}s. (3.5)

The remainder of this section is devoted to the proof of (3.5).

To start with, let a nonnegative and even function ρC()\rho\in C^{\infty}(\mathbb{R}) be given with support in [1,1][-1,1] and with ρ(s)ds=1\int_{\mathbb{R}}\rho(s)\,{\rm d}s=1. For ε>0\varepsilon>0 we define ρε(t):=ρ(t/ε)/ε\rho_{\varepsilon}(t):=\rho(t/\varepsilon)/\varepsilon and ηε(t):=1tρε(s)ds\eta_{\varepsilon}(t):=\int_{-1}^{t}\rho_{\varepsilon}(s)\,{\rm d}s for all tt\in\mathbb{R}. As ηεh0\eta_{\varepsilon}\to h_{0} in {0}\mathbb{R}\setminus\{0\}, by letting

Gε(t):=U(Wr(y(t,x))+ηε(θ(x)t)(Wr(y(t,x))Wa(y(t,x))))dxG_{\varepsilon}(t):=\int_{U}\Big{(}W^{r}(\nabla y(t,x))+\eta_{\varepsilon}(\theta(x){-}t)\big{(}W^{r}(\nabla y(t,x))-W^{a}(\nabla y(t,x))\big{)}\Big{)}\,{\rm d}x

we readily check that

limε0Gε(t):=UW0(θ(x)t,y(t,x))dx\lim_{\varepsilon\to 0}G_{\varepsilon}(t):=\int_{U}W_{0}(\theta(x){-}t,\nabla y(t,x))\,{\rm d}x (3.6)

for all t[0,T]t\in[0,T]. As GεH1(0,T)G_{\varepsilon}\in H^{1}(0,T) we may compute its time derivative at almost all times getting

G˙ε(t)\displaystyle\dot{G}_{\varepsilon}(t) =U(FWr(y):y˙+ηε(θt)(FWr(y)FWa(y)):y˙)dx\displaystyle=\int_{U}\Big{(}\partial_{F}W^{r}(\nabla y){:}\nabla\dot{y}+\eta_{\varepsilon}(\theta{-}t)\big{(}\partial_{F}W^{r}(\nabla y){-}\partial_{F}W^{a}(\nabla y)\big{)}{:}\nabla\dot{y}\Big{)}\,{\rm d}x
Uρε(θt)(Wr(y)Wa(y))dx.\displaystyle\quad{}-\int_{U}\rho_{\varepsilon}(\theta{-}t)\big{(}W^{r}(\nabla y){-}W^{a}(\nabla y)\big{)}\,{\rm d}x.

By integrating in time, taking the limit ε0\varepsilon\to 0, and using (3.6), one hence gets

UW0(θt,y)dxUW0(θ,y0)dx=limε0(Gε(t)Gε(0))=limε00tG˙ε(s)ds\displaystyle\int_{U}W_{0}(\theta{-}t,\nabla y)\,{\rm d}x-\int_{U}W_{0}(\theta,\nabla y_{0})\,{\rm d}x=\lim_{\varepsilon\to 0}\big{(}G_{\varepsilon}(t)-G_{\varepsilon}(0)\big{)}=\lim_{\varepsilon\to 0}\int_{0}^{t}\dot{G}_{\varepsilon}(s)\,{\rm d}s
=limε00tU(FWr(y):y˙+ηε(θs)(FWr(y)FWa(y)):y˙)dxds\displaystyle\quad=\lim_{\varepsilon\to 0}\int_{0}^{t}\!\!\int_{U}\Big{(}\partial_{F}W^{r}(\nabla y){:}\nabla\dot{y}+\eta_{\varepsilon}(\theta{-}s)\big{(}\partial_{F}W^{r}(\nabla y){-}\partial_{F}W^{a}(\nabla y)\big{)}{:}\nabla\dot{y}\Big{)}\,{\rm d}x\,{\rm d}s
limε00tUρε(θs)(Wr(y)Wa(y))dxds\displaystyle\qquad{}-\lim_{\varepsilon\to 0}\int_{0}^{t}\!\!\int_{U}\rho_{\varepsilon}(\theta{-}s)\big{(}W^{r}(\nabla y){-}W^{a}(\nabla y)\big{)}\,{\rm d}x\,{\rm d}s
=0tUFW0(θs,y):y˙dxdslimε00tUρε(θs)(Wr(y)Wa(y))dxds.\displaystyle\quad=\int_{0}^{t}\!\!\int_{U}\partial_{F}W_{0}(\theta{-}s,\nabla y){:}\nabla\dot{y}\,{\rm d}x\,{\rm d}s-\lim_{\varepsilon\to 0}\int_{0}^{t}\!\!\int_{U}\rho_{\varepsilon}(\theta{-}s)\big{(}W^{r}(\nabla y){-}W^{a}(\nabla y)\big{)}\,{\rm d}x\,{\rm d}s.

In order to prove (3.5) it is hence sufficient to check that

limε00tUρε(θs)(Wr(y)Wa(y))dxds\displaystyle\lim_{\varepsilon\to 0}\int_{0}^{t}\!\!\int_{U}\rho_{\varepsilon}(\theta{-}s)\big{(}W^{r}(\nabla y){-}W^{a}(\nabla y)\big{)}\,{\rm d}x\,{\rm d}s
=0t{θ=s}Wr(y)Wa(y)|θ|dd1ds.\displaystyle\quad=\int_{0}^{t}\!\!\int_{\{\theta=s\}}\frac{W^{r}(\nabla y){-}W^{a}(\nabla y)}{|\nabla\theta|}\,{\rm d}\mathcal{H}^{d-1}\,{\rm d}s. (3.7)

By introducing the short-hand notation g=Wr(y)Wa(y)g=W^{r}(\nabla y)-W^{a}(\nabla y) and by using the coarea formula [16, Sec. 3.2.11] (recall that θ\theta is Lipschitz continuous and |θ|1/Cγ|\nabla\theta|\geq 1/C_{\gamma} almost everywhere) we can compute that

0tUρε(θs)(Wr(y)Wa(y))dxds=0t{θ=r}ρε(θs)g(s,x)|θ(x)|dd1(x)drds\displaystyle\int_{0}^{t}\!\!\int_{U}\rho_{\varepsilon}(\theta{-}s)\big{(}W^{r}(\nabla y){-}W^{a}(\nabla y)\big{)}\,{\rm d}x\,{\rm d}s=\int_{0}^{t}\!\!\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(\theta{-}s)\frac{g(s,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,{\rm d}r\,{\rm d}s
=0t{θ=r}ρε(rs)g(r,x)|θ(x)|dd1(x)drds\displaystyle\quad=\int_{0}^{t}\!\!\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(r{-}s)\frac{g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,{\rm d}r\,{\rm d}s
+0t{θ=r}ρε(rs)g(s,x)g(r,x)|θ(x)|dd1(x)drds.\displaystyle\qquad+\int_{0}^{t}\!\!\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(r{-}s)\frac{g(s,x){-}g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,\ {\rm d}r\,{\rm d}s. (3.8)

The coarea formula and the bound |θ|1/cγ|\nabla\theta|\leq 1/c_{\gamma} ensure that rh(r):=d1({θ=r})r\in\mathbb{R}\mapsto h(r):=\mathcal{H}^{d-1}(\{\theta=r\}) is integrable. Indeed,

hL1()=d1({θ=r})dr=U|θ|dx<.\|h\|_{L^{1}(\mathbb{R})}=\int_{\mathbb{R}}\mathcal{H}^{d-1}(\{\theta=r\})\,{\rm d}r=\int_{U}|\nabla\theta|\,{\rm d}x<\infty.

As g/|θ|g/|\nabla\theta| is bounded, setting

r(r):={θ=r}g(r,x)|θ(x)|dd1(x)r\in\mathbb{R}\mapsto\ell(r):=\int_{\{\theta=r\}}\frac{g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)

one has that L1()\ell\in L^{1}(\mathbb{R}), as well, and

{θ=r}ρε(rs)g(r,x)|θ(x)|dd1(x)dr\displaystyle\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(r{-}s)\frac{g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,{\rm d}r
=ρε(sr)({θ=r}g(r,x)|θ(x)|dd1(x))dr=(ρε)(s)\displaystyle\quad=\int_{\mathbb{R}}\rho_{\varepsilon}(s{-}r)\left(\int_{\{\theta=r\}}\frac{g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\right)\,{\rm d}r=(\rho_{\varepsilon}\ast\ell)(s)

where we used that ρε\rho_{\varepsilon} is even and where the symbol \ast stands for the convolution in \mathbb{R}. As ρε\rho_{\varepsilon}\ast\ell\to\ell strongly in L1()L^{1}(\mathbb{R}) for ε0\varepsilon\to 0, one can pass to the limit in the first term on the right-hand side of (3.8) and get

limε00t{θ=r}ρε(rs)g(r,x)|θ(x)|dd1(x)drds=limε00t(ρε)ds=0tds\displaystyle\lim_{\varepsilon\to 0}\int_{0}^{t}\!\!\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(r{-}s)\frac{g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,\ {\rm d}r\,{\rm d}s=\lim_{\varepsilon\to 0}\int_{0}^{t}(\rho_{\varepsilon}\ast\ell)\,{\rm d}s=\int_{0}^{t}\ell\,{\rm d}s
=0t{θ=s}g(s,x)|θ(x)|dd1(x)ds=0t{θ=s}Wr(y)Wa(y)|θ|dd1ds.\displaystyle\quad=\int_{0}^{t}\!\!\int_{\{\theta=s\}}\frac{g(s,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,{\rm d}s=\int_{0}^{t}\!\!\int_{\{\theta=s\}}\frac{W^{r}(\nabla y){-}W^{a}(\nabla y)}{|\nabla\theta|}{\rm d}\mathcal{H}^{d-1}\,{\rm d}s. (3.9)

As regards the second term in the right-hand side of (3.8), notice that ρε(rs)0\rho_{\varepsilon}(r{-}s)\neq 0 only if |rs|2ε|r-s|\leq 2\varepsilon. Hence, using the Hölder regularity of gg and the boundedness of 1/|θ|1/|\nabla\theta| and |θ||\nabla\theta|, we conclude that

limε0|0t{θ=r}ρε(rs)g(s,x)g(r,x)|θ(x)|dd1(x)drds|\displaystyle\lim_{\varepsilon\to 0}\left|\int_{0}^{t}\!\!\int_{\mathbb{R}}\int_{\{\theta=r\}}\rho_{\varepsilon}(r{-}s)\frac{g(s,x){-}g(r,x)}{|\nabla\theta(x)|}{\rm d}\mathcal{H}^{d-1}(x)\,{\rm d}r\,{\rm d}s\right|
limε0Cεα0tρε(rs)d1({θ=r})drds\displaystyle\quad\leq\lim_{\varepsilon\to 0}C\varepsilon^{\alpha}\int_{0}^{t}\!\!\int_{\mathbb{R}}\rho_{\varepsilon}(r{-}s)\,\mathcal{H}^{d-1}(\{\theta=r\})\,{\rm d}r\,{\rm d}s
=limε0CεαρεhL1()limε0CεαhL1()=0\displaystyle\quad=\lim_{\varepsilon\to 0}C\varepsilon^{\alpha}\|\rho_{\varepsilon}\ast h\|_{L^{1}(\mathbb{R})}\leq\lim_{\varepsilon\to 0}C\varepsilon^{\alpha}\|h\|_{L^{1}(\mathbb{R})}=0 (3.10)

for some α(0,1)\alpha\in(0,1). Relations (3.9)–(3) imply that the limit (3.7) holds true. This in turn proves (3.5) and the energy equality (2.22) follows.

4. Proof of Theorem 2.1: diffused-interface case

Let ε>0\varepsilon>0 be fixed. We prove existence of a weak/viscous solution (yε,θε)(y_{\varepsilon},\theta_{\varepsilon}) by an iterative construction. Let us start by checking that for all θC(U)\theta\in C(U) there exists an admissible yεL(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d))y_{\varepsilon}\in L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})) satisfying (2.1).

Proposition 4.1 (Existence of yεy_{\varepsilon} given θ\theta).

Let ε>0\varepsilon>0, hεh_{\varepsilon} defined as in (2.1), and θC(U)\theta\in C(U). Let (2.3)–(2.14) hold. Then, there exists yεL(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d))y_{\varepsilon}\in L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})) with yε(t,)𝒜y_{\varepsilon}(t,\cdot)\in\mathcal{A} for every t(0,T)t\in(0,T) satisfying (2.1).

Proof.

The assertion follows by adapting the arguments in [2] or [25]. We present the argument here, for completeness.

Notice at first that, thanks to the growth conditions (2.4) and (2.8), we have

cW(yL2(U;d×d)2+1detyLq(U)q)+cH2yLp(U;d×d×d)p𝒲ε(σ,y)+(y)c_{W}\left(\|\nabla y\|^{2}_{L^{2}(U;\mathbb{R}^{d\times d})}+\left\|\frac{1}{\det\nabla y}\right\|^{q}_{L^{q}(U)}\right)+c_{H}\|\nabla^{2}y\|^{p}_{L^{p}(U;\mathbb{R}^{d\times d\times d})}\leq\mathcal{W}_{\varepsilon}(\sigma,y)+\mathcal{H}(y)

for every y𝒜y\in\mathcal{A} and any σC(U)\sigma\in C(U). Following [35, Thm. 3.1], if 𝒲ε(σ,y)+(y)M\mathcal{W}_{\varepsilon}(\sigma,y)+\mathcal{H}(y)\leq M for some M>0M>0, it holds that

yW2,p(U;d)+yC1,1d/p(U;d)+(y)1C1d/p(U;d×d)c,minxU¯dety(x)1c,\|y\|_{W^{2,p}(U;\mathbb{R}^{d})}+\|y\|_{C^{1,1-d/p}(U;\mathbb{R}^{d})}+\|(\nabla y)^{-1}\|_{C^{1-d/p}(U;\mathbb{R}^{d\times d})}\leq c,\quad\min_{x\in\overline{U}}\det\nabla y(x)\geq\frac{1}{c}, (4.1)

where the constant c>0c>0 depends on MM but not on yy and σ\sigma.

The proof of Proposition 4.1 is based on a time-discretization scheme. Let the time step τT/Nτ\tau\coloneqq T/N_{\tau} with NτN_{\tau}\in\mathbb{N} be given and let tiiτt_{i}\coloneqq i\tau, for i=0,,Nτi=0,\dots,N_{\tau} be the corresponding uniform partition of the time interval [0,T][0,T]. For i=1,,Nτi=1,\dots,N_{\tau}, we define yτi𝒜y^{i}_{\tau}\in\mathcal{A} as a solution to the minimization problem

yτiargminy𝒜{𝒲ε(θti,y)+(y)+τε(θti,yτi1,yyτi1τ)Uf(θti)ydx}.\displaystyle y^{i}_{\tau}\in\operatorname*{arg\,min}_{y\in\mathcal{A}}\left\{\mathcal{W}_{\varepsilon}(\theta{-}t_{i},y)+\mathcal{H}(y)+\tau\mathcal{R}_{\varepsilon}\left(\theta{-}t_{i},y^{i-1}_{\tau},\frac{y{-}y^{i-1}_{\tau}}{\tau}\right)-\int_{U}f(\theta{-}t_{i}){\cdot}y\,{\rm d}x\right\}.

Under the growth conditions (2.4), (2.8), and (2.12), and the regularity assumptions (2.3), (2.6), (2.11), and (2.13), the existence of yτiy^{i}_{\tau} for every i=1,,Nτi=1,\dots,N_{\tau} follows by the Direct Method of the calculus of variations. Moreover, every minimizer yτiy_{\tau}^{i} satisfies the time-discrete Euler–Lagrange equation

U(FWε(θti,yτi)+F˙Rε(θti,yτi1,yτiyτi1τ)):zdx\displaystyle\int_{U}\left(\partial_{F}W_{\varepsilon}(\theta{-}t_{i},\nabla y^{i}_{\tau}){+}\partial_{\dot{F}}R_{\varepsilon}\left(\theta{-}t_{i},\nabla y^{i-1}_{\tau},\frac{\nabla y^{i}_{\tau}{-}\nabla y^{i-1}_{\tau}}{\tau}\right)\right){:}\nabla z\,{\rm d}x
+UDH(2yτi)2zdx=Uf(θti)zdx\displaystyle\quad+\int_{U}{\rm D}H\left(\nabla^{2}y^{i}_{\tau}\right){\vdots}\nabla^{2}z\,{\rm d}x=\int_{U}f(\theta{-}t_{i}){\cdot}z\,{\rm d}x (4.2)

for every zC(U¯;d)z\in C^{\infty}(\overline{U};\mathbb{R}^{d}) with z=0z=0 on ΓD\Gamma_{D} and for all i=1,,Nτi=1,\dots,N_{\tau}.

Let us introduce the following notation for the time interpolants on the partition: Given a vector (u0,,uNτ)(u_{0},...,u_{N_{\tau}}), we define its backward-constant interpolant u¯τ\overline{u}_{\tau}, its forward-constant interpolant u¯τ\underline{u}_{\tau}, and its piecewise-affine interpolant u^τ\hat{u}_{\tau} on the partition (ti)i=0Nτ(t_{i})_{i=0}^{N_{\tau}} as

u¯τ(0):=u0,u¯τ(t):=ui\displaystyle\overline{u}_{\tau}(0):=u_{0},\quad\quad\overline{u}_{\tau}(t):=u_{i} if t(ti1,ti] for i=1,,Nτ,\displaystyle\text{ if }t\in(t_{i-1},t_{i}]\quad\text{ for }i=1,\dots,N_{\tau},
u¯τ(T):=uNτ,u¯τ(t):=ui1\displaystyle\underline{u}_{\tau}(T):=u_{N_{\tau}},\quad\;\underline{u}_{\tau}(t):=u_{i-1} if t[ti1,ti) for i=1,,Nτ,\displaystyle\text{ if }t\in[t_{i-1},t_{i})\quad\text{ for }i=1,\dots,N_{\tau},
u^τ(0):=u0,u^τ(t):=uiui1titi1(tti1)+ui1\displaystyle\hat{u}_{\tau}(0):=u_{0},\quad\quad\hat{u}_{\tau}(t):=\frac{u_{i}-u_{i-1}}{t_{i}-t_{i-1}}(t-t_{i-1})+u_{i-1} if t(ti1,ti] for i=1,,Nτ.\displaystyle\text{ if }t\in(t_{i-1},t_{i}]\quad\text{ for }i=1,\dots,N_{\tau}.

Owing to this notation, we can take the sum in (4) for i=1,,Nτi=1,\dots,N_{\tau} and equivalently rewrite it in the compact form

0TU(FWε(θt¯τ,y¯τ)+F˙Rε(θt¯τ,y¯τ,y^˙τ)):zdxdt\displaystyle\int_{0}^{T}\!\!\int_{U}\left(\partial_{F}W_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\overline{y}_{\tau}){+}\partial_{\dot{F}}R_{\varepsilon}\left(\theta{-}\overline{t}_{\tau},\nabla\underline{y}_{\tau},\nabla\dot{\hat{y}}_{\tau}\right)\right){:}\nabla z\,{\rm d}x\,{\rm d}t
+0TUDH(2y¯τ)2zdxdt=0TUf(θt¯τ)zdxdt\displaystyle\quad+\int_{0}^{T}\!\!\int_{U}{\rm D}H\left(\nabla^{2}\overline{y}_{\tau}\right){\vdots}\nabla^{2}z\,{\rm d}x\,{\rm d}t=\int_{0}^{T}\!\!\int_{U}f(\theta{-}\overline{t}_{\tau}){\cdot}z\,{\rm d}x\,{\rm d}t
zC([0,T]×U¯;d)withz=0onΣD.\displaystyle\qquad\qquad\forall z\in C^{\infty}([0,T]\times\overline{U};\mathbb{R}^{d})\ \ \text{with}\ \ z=0\ \text{on}\ \Sigma_{D}. (4.3)

From the minimality of yτiy^{i}_{\tau} we get that

UWε(θti,yτi)+H(2yτi)+1τRε(θti,yτi1,yτiyτi1)dxUf(θti)yτidx\displaystyle\int_{U}W_{\varepsilon}(\theta{-}t_{i},\nabla y^{i}_{\tau})+H(\nabla^{2}y^{i}_{\tau})+\frac{1}{\tau}R_{\varepsilon}(\theta{-}t_{i},\nabla y^{i-1}_{\tau},\nabla y^{i}_{\tau}{-\nabla y^{i-1}_{\tau}})\,{\rm d}x-\int_{U}f(\theta{-}t_{i}){\cdot}y^{i}_{\tau}\,{\rm d}x
UWε(θti,yτi1)+H(2yτi1)dxUf(θti)yτi1dx.\displaystyle\leq\int_{U}W_{\varepsilon}(\theta{-}t_{i},\nabla y^{i-1}_{\tau})+H(\nabla^{2}y^{i-1}_{\tau})\,{\rm d}x-\int_{U}f(\theta{-}t_{i}){\cdot}y^{i-1}_{\tau}\,{\rm d}x.

Summing over i=1,,nNτi=1,\dots,n\leq N_{\tau}, we find

UWε(θtn,yτn)+i=1nti1tisWε(θr,yτi1)drdx\displaystyle\int_{U}W_{\varepsilon}(\theta{-}t_{n},\nabla y^{n}_{\tau})+\sum_{i=1}^{n}\int_{t_{i-1}}^{t_{i}}\partial_{s}W_{\varepsilon}(\theta{-}r,\nabla y^{i-1}_{\tau})\,{\rm d}r\,{\rm d}x
+UH(2yτn)+τi=1nRε(θti,yτi1,yτiyτi1τ)dxUf(θtn)yτndx\displaystyle+\int_{U}H(\nabla^{2}y^{n}_{\tau})+\tau\sum_{i=1}^{n}R_{\varepsilon}\left(\theta{-}t_{i},\nabla y^{i-1}_{\tau},\frac{\nabla y^{i}_{\tau}{-}\nabla y^{i-1}_{\tau}}{\tau}\right)\,{\rm d}x-\int_{U}f(\theta{-}t_{n}){\cdot}y^{n}_{\tau}\,{\rm d}x
UWε(θ,y0)+H(2y0)dxUf(θ)y0dxi=1nU(f(θti)f(θti1))yτi1dx,\displaystyle\quad\leq\int_{U}W_{\varepsilon}(\theta,\nabla y_{0})+H(\nabla^{2}y_{0})\,{\rm d}x-\!\int_{U}\!f(\theta){\cdot}y_{0}\,{\rm d}x-\sum_{i=1}^{n}\int_{U}(f(\theta{-}t_{i}){-}f(\theta{-}t_{i-1})){\cdot}y^{i-1}_{\tau}\,{\rm d}x,

which we can rewrite in terms of the time-interpolants, for every t[0,T]t\in[0,T], as

UWε(θt¯τ,y¯τ)+H(2y¯τ)dx+0t¯τUsWε(θr,y¯τ)dxdr\displaystyle\int_{U}W_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\overline{y}_{\tau})+H(\nabla^{2}\overline{y}_{\tau})\,{\rm d}x+\int_{0}^{\overline{t}_{\tau}}\int_{U}\partial_{s}W_{\varepsilon}(\theta{-}r,\nabla\underline{y}_{\tau})\,{\rm d}x\,{\rm d}r
+0t¯τURε(θt¯τ,y¯τ,y^˙τ)dxdrUf(θt¯τ)y¯τdx\displaystyle\qquad+\int_{0}^{\overline{t}_{\tau}}\int_{U}R_{\varepsilon}\left(\theta{-}\overline{t}_{\tau},\nabla\underline{y}_{\tau},\nabla\dot{\hat{y}}_{\tau}\right)\,{\rm d}x\,{\rm d}r-\int_{U}f(\theta{-}\overline{t}_{\tau}){\cdot}\overline{y}_{\tau}\,{\rm d}x
UWε(θ,y0)+H(2y0)dxUf(θ)y0dx0t¯τUf˙(θr)y¯τdxdr.\displaystyle\quad\leq\int_{U}W_{\varepsilon}(\theta,\nabla y_{0})+H(\nabla^{2}y_{0})\,{\rm d}x-\int_{U}f(\theta){{\cdot}}y_{0}\,{\rm d}x-\int_{0}^{\overline{t}_{\tau}}\int_{U}\dot{f}(\theta{-}r){\cdot}\underline{y}_{\tau}\,{\rm d}x\,{\rm d}r. (4.4)

By the growth conditions (2.4), (2.8), (2.12), and (2.13) we have

c\displaystyle c y¯τL2(U;d×d)2+c1dety¯τLq(U)q+c2y¯τLp(U;d×d×d)p\displaystyle\|\nabla\overline{y}_{\tau}\|^{2}_{L^{2}(U;\mathbb{R}^{d\times d})}+c\left\|\frac{1}{\det\nabla\overline{y}_{\tau}}\right\|^{q}_{L^{q}(U)}+c\|\nabla^{2}\overline{y}_{\tau}\|^{p}_{L^{p}(U;\mathbb{R}^{d\times d\times d})}
(2.1),(2.4),(2.8)cUWε(θt¯τ,y¯τ)+H(2y¯τ)dx\displaystyle\stackrel{{\scriptstyle\eqref{h_eps properties},\eqref{W3},\eqref{H3}}}{{\leq}}c\int_{U}W_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\overline{y}_{\tau})+H(\nabla^{2}\overline{y}_{\tau})\,{\rm d}x
(2.1),(4),(2.13)cεU0t¯τ(Wa(y¯τ)+Wr(y¯τ))𝟙{ε/2<θ(x)r<ε/2}drdx\displaystyle\stackrel{{\scriptstyle\eqref{h_eps properties},\eqref{energy ineq 1},\eqref{L1}}}{{\leq}}\frac{c}{\varepsilon}\int_{U}\int_{0}^{\overline{t}_{\tau}}\left(W^{a}(\nabla\underline{y}_{\tau})+W^{r}(\nabla\underline{y}_{\tau})\right)\mathbbm{1}_{\{-\varepsilon/2<\theta(x)-r<\varepsilon/2\}}\,{\rm d}r\,{\rm d}x
+c0t¯τy¯τL2(U;d)2dr+cy¯τL2(U;d)+c\displaystyle\qquad\qquad+c\int_{0}^{\overline{t}_{\tau}}\|\underline{y}_{\tau}\|^{2}_{L^{2}(U;\mathbb{R}^{d})}\,{\rm d}r+c\|\overline{y}_{\tau}\|_{L^{2}(U;\mathbb{R}^{d})}+c
(4.1),(2.3)cεUθ(x)ε/2θ(x)+ε/2drdx+c0t¯τy¯τL2(U;d)2dr+cy¯τL2(U;d)+c\displaystyle\stackrel{{\scriptstyle\eqref{eq a priori estimates},\eqref{W1}}}{{\leq}}\frac{c}{\varepsilon}\!\int_{U}\!\int_{\theta(x)-\varepsilon/2}^{\theta(x)+\varepsilon/2}\,{\rm d}r\,{\rm d}x+c\!\!\int_{0}^{\overline{t}_{\tau}}\!\!\|\underline{y}_{\tau}\|^{2}_{L^{2}(U;\mathbb{R}^{d})}\,{\rm d}r+c\|\overline{y}_{\tau}\|_{L^{2}(U;\mathbb{R}^{d})}+c
c0t¯τy¯τL2(U;d)2dr+cy¯τL2(U;d)+c,\displaystyle\quad\leq c\int_{0}^{\overline{t}_{\tau}}\|\underline{y}_{\tau}\|^{2}_{L^{2}(U;\mathbb{R}^{d})}\,{\rm d}r+c\|\overline{y}_{\tau}\|_{L^{2}(U;\mathbb{R}^{d})}+c, (4.5)

where cc is independent of θ\theta. Hence, by the Poincaré inequality and the Discrete Gronwall Lemma [27, (C.2.6), p. 534] we find

y¯τL(0,T;W2,p(U;d))c.\|\overline{y}_{\tau}\|_{L^{\infty}(0,T;W^{2,p}(U;\mathbb{R}^{d}))}\leq c. (4.6)

Moreover, by the classical result of [22, Thm. 3.1] we get

dety¯τc>0in[0,T]×U¯\det\nabla\overline{y}_{\tau}\geq c>0\ \ \text{in}\ \ [0,T]\times\overline{U} (4.7)

where cc may depend on U,T,y0,cW,cH,CH,d,,pU,\,T,\,y_{0},\,c_{W},\,c_{H},\,C_{H},\,d,,p, and qq, but is independent of θ\theta and ε\varepsilon. By the Poincaré inequality and the generalization of Korn’s first inequality by [39] and [42, Thm. 2.2], we have that

y^˙τL2(Q;d×d)cFy^˙τ+y^˙τFL2(Q;d×d)\|\nabla\dot{\hat{y}}_{\tau}\|_{L^{2}(Q;\mathbb{R}^{d\times d})}\leq c\|\nabla F\dot{\hat{y}}_{\tau}+\nabla\dot{\hat{y}}_{\tau}^{\top}F^{\top}\|_{L^{2}(Q;\mathbb{R}^{d\times d})}

for every FC(U¯;d×d)F\in C(\overline{U};\mathbb{R}^{d\times d}) with minxU¯detF(x)>0\min_{x\in\overline{U}}\det F(x)>0. Notice that, by (4.6)–(4.7) and by the Sobolev embedding of W2,p(U;d)W^{2,p}(U;\mathbb{R}^{d}) into C1d/p(U;d)C^{1-d/p}(U;\mathbb{R}^{d}), we can take F=y¯τ(t,)F=\nabla\overline{y}_{\tau}(t,\cdot) for every t[0,T]t\in[0,T]. Thus, by (2.12),

y^˙τL2(Q;d×d)c0TURε(θt¯τ,y¯τ,y^˙τ)dxdrc\|\nabla\dot{\hat{y}}_{\tau}\|_{L^{2}(Q;\mathbb{R}^{d\times d})}\leq c\int_{0}^{T}\int_{U}R_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\underline{y}_{\tau},\nabla\dot{\hat{y}}_{\tau})\,{\rm d}x\,{\rm d}r\leq c

and again by the Poincaré inequality, this time applied to y˙\dot{y}, we get that

y^τH1(0,T;H1(U;d))c.\|{\hat{y}}_{\tau}\|_{H^{1}(0,T;H^{1}(U;\mathbb{R}^{d}))}\leq c. (4.8)

By using these estimates, up to not relabeled subsequences we get

y¯τ,y¯τyεweakly- inL(0,T;WΓD2,p(U;d)),\displaystyle\overline{y}_{\tau},\,\underline{y}_{\tau}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}y_{\varepsilon}\quad\text{weakly-$*$ in}\ \ L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d})), (4.9)
y^˙τy˙εweakly inL2(Q;d),\displaystyle\nabla\dot{\hat{y}}_{\tau}\rightharpoonup\nabla\dot{y}_{\varepsilon}\quad\text{weakly in}\ \ L^{2}(Q;\mathbb{R}^{d}), (4.10)
y^τyεstrongly inC0,α(Q¯;d)\displaystyle\nabla\hat{y}_{\tau}\rightarrow\nabla y_{\varepsilon}\quad\text{strongly in}\ \ C^{0,\alpha}(\overline{Q};\mathbb{R}^{d}) (4.11)

for some α(0,1)\alpha\in(0,1). Moreover, from the convergences above we also get dety¯τdetyε\det\nabla\overline{y}_{\tau}\to\det\nabla y_{\varepsilon}. In combination with the lower bound (4.7), this implies that yεGL+(d)\nabla y_{\varepsilon}\in{\rm GL}_{+}(d) everywhere, hence yεy_{\varepsilon} is admissible, namely yε(t,)𝒜y_{\varepsilon}(t,\cdot)\in\mathcal{A} for every t(0,T)t\in(0,T).

We now pass to the limit in the time-discrete Euler–Lagrange equation (4.3). By (2.13) we have

0TUf(θt¯τ)zdxdt0TUf(θt)zdxdt.\int_{0}^{T}\int_{U}f(\theta{-}\overline{t}_{\tau}){\cdot}z\,{\rm d}x\,{\rm d}t\rightarrow\int_{0}^{T}\int_{U}f(\theta{-}t){\cdot}z\,{\rm d}x\,{\rm d}t.

The dissipation then goes to the limit as follows

0TUF˙Rε(θt¯τ,y¯τ,y^˙τ):zdxdt\displaystyle\int_{0}^{T}\int_{U}\partial_{\dot{F}}R_{\varepsilon}\left(\theta{-}\overline{t}_{\tau},\nabla\underline{y}_{\tau},\nabla\dot{\hat{y}}_{\tau}\right){:}\nabla z\,{\rm d}x\,{\rm d}t
=20TUhε(θt¯τ)y¯τ(𝔻r(y¯τy¯τ)(y^˙τy¯τ+y¯τy^˙τ)):zdxdt\displaystyle\quad=2\int_{0}^{T}\int_{U}h_{\varepsilon}(\theta{-}\overline{t}_{\tau})\nabla\underline{y}_{\tau}\left(\mathbb{D}^{r}(\nabla\underline{y}_{\tau}^{\top}\nabla\underline{y}_{\tau})(\nabla\dot{\hat{y}}_{\tau}^{\top}\nabla\underline{y}_{\tau}{+}\nabla\underline{y}_{\tau}^{\top}\nabla\dot{\hat{y}}_{\tau})\right){:}\nabla z\,{\rm d}x\,{\rm d}t
+20TU(1hε(θt¯τ))y¯τ(𝔻a(y¯τy¯τ)(y^˙τy¯τ+y¯τy^˙τ)):zdxdt\displaystyle\quad\quad+2\int_{0}^{T}\int_{U}(1{-}h_{\varepsilon}(\theta{-}\overline{t}_{\tau}))\nabla\underline{y}_{\tau}\left(\mathbb{D}^{a}(\nabla\underline{y}_{\tau}^{\top}\nabla\underline{y}_{\tau})(\nabla\dot{\hat{y}}_{\tau}^{\top}\nabla\underline{y}_{\tau}{+}\nabla\underline{y}_{\tau}^{\top}\nabla\dot{\hat{y}}_{\tau})\right){:}\nabla z\,{\rm d}x\,{\rm d}t
20TUhε(θt)yε(𝔻r(yεyε)(y˙εyε+yεy˙ε)):zdxdt\displaystyle\quad\rightarrow 2\int_{0}^{T}\int_{U}h_{\varepsilon}(\theta{-}t)\nabla y_{\varepsilon}\left(\mathbb{D}^{r}(\nabla y^{\top}_{\varepsilon}\nabla y_{\varepsilon})(\nabla\dot{y}^{\top}_{\varepsilon}\nabla y_{\varepsilon}{+}\nabla y^{\top}_{\varepsilon}\nabla\dot{y}_{\varepsilon})\right){:}\nabla z\,{\rm d}x\,{\rm d}t
+20TU(1hε(θt))yε(𝔻a(yεyε)(y˙εyε+yεy˙ε)):zdxdt\displaystyle\quad\quad+2\int_{0}^{T}\int_{U}(1{-}h_{\varepsilon}(\theta{-}t))\nabla y_{\varepsilon}\left(\mathbb{D}^{a}(\nabla y^{\top}_{\varepsilon}\nabla y_{\varepsilon})(\nabla\dot{y}^{\top}_{\varepsilon}\nabla y_{\varepsilon}{+}\nabla y^{\top}_{\varepsilon}\nabla\dot{y}_{\varepsilon})\right){:}\nabla z\,{\rm d}x\,{\rm d}t
=0TUF˙Rε(θt,yε,y˙ε):zdxdt\displaystyle\quad=\int_{0}^{T}\int_{U}\partial_{\dot{F}}R_{\varepsilon}\left(\theta{-}t,\nabla y_{\varepsilon},\nabla\dot{y}_{\varepsilon}\right){:}\nabla z\,{\rm d}x\,{\rm d}t

by the convergences (4.9)–(4.11), and (2.11). Moreover, we also have

0TUFWε(θt¯τ,y¯τ)dxdt0TUFWε(θt,yε)dxdt\int_{0}^{T}\int_{U}\partial_{F}W_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\overline{y}_{\tau})\,{\rm d}x\,{\rm d}t\rightarrow\int_{0}^{T}\int_{U}\partial_{F}W_{\varepsilon}(\theta{-}t,\nabla y_{\varepsilon})\,{\rm d}x\,{\rm d}t

by (4.9) and (2.3).

For the convergence of the second-gradient term \mathcal{H} we reproduce in this setting the argument from [25]. We consider test functions zτ(yεy¯τ)ϕz_{\tau}\coloneqq(y_{\varepsilon}-\overline{y}_{\tau})\phi with ϕ:U+\phi:U\to\mathbb{R}_{+} smooth in the time-discrete Euler–Lagrange equation (4.3) (note here that zτz_{\tau} is not smooth. Still, (4.3) holds by density for all zLp(0,T;W2,p(U;d))z\in L^{p}(0,T;W^{2,p}(U;\mathbb{R}^{d})) with z=0z=0 a.e. on ΣD\Sigma_{D}, as well). Notice that, by the convergences (4.9) and (4.10), zτ0z_{\tau}\rightarrow 0 strongly in L(0,T;H1(U;d))L^{\infty}(0,T;H^{1}(U;\mathbb{R}^{d})) and zτ0z_{\tau}\rightharpoonup 0 weakly in Lp(0,T;W2,p(U;d))L^{p}(0,T;W^{2,p}(U;\mathbb{R}^{d})). Thus, by the Euler–Lagrange equation (4.3), convergence (4.11), and the continuity assumptions on the densities (2.3), (2.11), and (2.13), we find that

lim supτ00TU(DH(2yε)DH(2y¯τ))2zτdxdt\displaystyle\limsup_{\tau\rightarrow 0}\int_{0}^{T}\int_{U}({\rm D}H(\nabla^{2}y_{\varepsilon})-{\rm D}H(\nabla^{2}\overline{y}_{\tau})){\vdots}\nabla^{2}z_{\tau}\,{\rm d}x\,{\rm d}t
=lim supτ0[0TUDH(2yε)2zτf(θt¯τ)zτdxdt\displaystyle\quad=\limsup_{\tau\rightarrow 0}\Bigg{[}\int_{0}^{T}\int_{U}{\rm D}H(\nabla^{2}y_{\varepsilon}){\vdots}\nabla^{2}z_{\tau}-f(\theta{-}\overline{t}_{\tau}){\cdot}z_{\tau}\,{\rm d}x\,{\rm d}t
+0TU(FWε(θt¯τ,y¯τ)+F˙Rε(θt¯τ,y¯τ,y^˙τ)):zτdxdt]=0\displaystyle\qquad+\int_{0}^{T}\int_{U}\left(\partial_{F}W_{\varepsilon}(\theta{-}\overline{t}_{\tau},\nabla\overline{y}_{\tau})+\partial_{\dot{F}}R_{\varepsilon}\left(\theta{-}\overline{t}_{\tau},\nabla\underline{y}_{\tau},\nabla\dot{\hat{y}}_{\tau}\right)\right){:}\nabla z_{\tau}\,{\rm d}x\,{\rm d}t\Bigg{]}=0

As ϕ\phi is arbitrary and we have the coercivity (2.8), this implies

2y¯τ2yεstrongly inLp(Q;d)\nabla^{2}\overline{y}_{\tau}\rightarrow\nabla^{2}y_{\varepsilon}\quad\text{strongly in}\ \ L^{p}(Q;\mathbb{R}^{d})

and thus

DH(2y¯τ)DH(2yε) strongly in Lp(Q;d).{\rm D}H(\nabla^{2}\overline{y}_{\tau})\rightarrow{\rm D}H(\nabla^{2}{y}_{\varepsilon})\quad\text{ strongly in }L^{p^{\prime}}(Q;\mathbb{R}^{d}).

Passing in the limit in (4.3) we then find (2.1) and conclude the proof. ∎

Let us now move to the proof of Theorem 2.1 in the diffused-interface case ε>0\varepsilon>0. As announced, the proof hinges on an iterative construction. To start with, let us remark that y0y_{0} from (2.14) is such that y0\nabla y_{0} is Hölder continuous in space and time. In particular, the mapping γ~:1+d(0,)\tilde{\gamma}:\mathbb{R}^{1+d}\to(0,\infty) defined by

γ~(θ,x):=γ(y0(θ,x),y0(θ,x))(θ,x)1+d\tilde{\gamma}(\theta,x):=\gamma(y_{0}(\theta,x),\nabla y_{0}(\theta,x))\quad\forall(\theta,x)\in\mathbb{R}^{1+d}

in Hölder continuous, as well. We let θ0C(U¯)\theta_{0}\in C(\overline{U}) be a viscosity solution to

γ(y0(θ0(x),x),y0(θ0(x),x))|(θ0(x))|=1inUΩ0¯,\displaystyle\gamma(y_{0}(\theta_{0}(x),x),\nabla y_{0}(\theta_{0}(x),x))|\nabla(-\theta_{0}(x))|=1\quad\text{in}\ U\setminus\overline{\Omega_{0}},
θ0(x)=0inΩ0.\displaystyle\theta_{0}(x)=0\quad\text{in}\ \Omega_{0}.

Indeed, the existence of viscosity solutions follows by the classical viscosity theory, see, e.g., [23, Prop. 1.3] as

(x,θ,p):U¯××dγ(y0(θ,x),y0(θ,x))|p|(x,\theta,p):\overline{U}\times\mathbb{R}\times\mathbb{R}^{d}\mapsto\gamma(y_{0}(\theta,x),\nabla y_{0}(\theta,x))\,|p|

is continuous and convex in pp, and xUdist(x,Ω0)/Cγx\in U\mapsto\operatorname{dist}(x,\Omega_{0})/{C_{\gamma}} and xUdist(x,Ω0)/cγx\in U\mapsto\operatorname{dist}(x,\Omega_{0})/{c_{\gamma}} are a viscosity subsolution and supersolution to the problem, respectively.

Let yε1L(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d))y^{1}_{\varepsilon}\in L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})) be defined via Proposition 4.1 for θ=θ0\theta=\theta_{0}. Then, for all i1i\geq 1, given yεiL(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d))y^{i}_{\varepsilon}\in L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})) we define θεiC(U¯)\theta^{i}_{\varepsilon}\in C(\overline{U}) to be a viscosity solution to

γ(yεi(θεi(x),x),yεi(θεi(x),x))|(θεi(x))|=1inUΩ0¯,\displaystyle\gamma(y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x),\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x))|\nabla(-\theta^{i}_{\varepsilon}(x))|=1\quad\text{in}\ U\setminus\overline{\Omega_{0}},
θεi(x)=0inΩ0.\displaystyle\theta^{i}_{\varepsilon}(x)=0\quad\text{in}\ \Omega_{0}.

The existence of such a viscosity solution follows again from the classical theory as

(x,θ,p):U¯××dγ(yεi(θεi(x),x),yεi(θεi(x),x))|p|(x,\theta,p):\overline{U}\times\mathbb{R}\times\mathbb{R}^{d}\mapsto\gamma(y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x),\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x))|p|

is continuous and convex in pp. In particular, we have that

dist(x,Ω0)Cγθεi(x)dist(x,Ω0)cγxU¯Ω0¯,\frac{\operatorname{dist}(x,\Omega_{0})}{C_{\gamma}}\leq\theta^{i}_{\varepsilon}(x)\leq\frac{\operatorname{dist}(x,\Omega_{0})}{c_{\gamma}}\quad\forall x\in\overline{U}\setminus\overline{\Omega_{0}}, (4.12)

as well as

1Cγ|θεi(x)|1cγfor a.e.xU.\frac{1}{C_{\gamma}}\leq|\nabla\theta^{i}_{\varepsilon}(x)|\leq\frac{1}{c_{\gamma}}\quad\text{for a.e.}\ x\in U. (4.13)

Note that (4.12) in particular implies that

Ωεi(t):={xU|θεi(x)<t}Ω0+BCγT.\Omega^{i}_{\varepsilon}(t):=\{x\in U\ |\ \theta^{i}_{\varepsilon}(x)<t\}\subset\Omega_{0}+B_{C_{\gamma}T}. (4.14)

As we have assumed that Ω0+BCγTU\Omega_{0}+B_{C_{\gamma}T}\subset\subset U in (2.16), inclusion (4.14) in particular guarantees that the accreting phase defined by θεi\theta^{i}_{\varepsilon} remains at positive distance from the boundary U\partial U, independently of ε\varepsilon and ii.

Given such θεi\theta^{i}_{\varepsilon}, we define yεi+1L(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d))y^{i+1}_{\varepsilon}\in L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})) by Proposition 4.1 applied for θ=θεi\theta=\theta^{i}_{\varepsilon}.

Notice that the sequence (yεi,θεi)i(y^{i}_{\varepsilon},\theta^{i}_{\varepsilon})_{i\in\mathbb{N}} defined by this iterative procedure is (possibly not unique but nonetheless) uniformly bounded in the space

(L(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d)))×C0,1(U¯),\left(L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d}))\right)\times C^{0,1}(\overline{U}),

thanks to (4.12) and the θ\theta-independent estimates (4.6) and (4.8). Moreover, (θεi)i(\theta^{i}_{\varepsilon})_{i\in\mathbb{N}} is uniformly Lipschitz continuous thanks to (4.13). Therefore, by the Ascoli–Arzelà and the Banach–Alaoglu Theorems there exists a pair (yε,θε)(y_{\varepsilon},\theta_{\varepsilon}) such that

yεiyεweakly- inL(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d)),\displaystyle y_{\varepsilon}^{i}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}y_{\varepsilon}\quad\text{weakly-$*$ in}\ L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})), (4.15)
yεiyεstrongly inC1,α(Q¯;d),\displaystyle y_{\varepsilon}^{i}\to y_{\varepsilon}\quad\text{strongly in}\ C^{1,\alpha}(\overline{Q};\mathbb{R}^{d}), (4.16)
θεiθεstrongly inC(U¯)\displaystyle\theta_{\varepsilon}^{i}\to\theta_{\varepsilon}\quad\text{strongly in}\ C(\overline{U}) (4.17)

for some α(0,1)\alpha\in(0,1) and θε\theta_{\varepsilon} fulfills (2.20). As (yεi)i(y^{i}_{\varepsilon})_{i\in\mathbb{N}} is uniformly Hölder continuous and γ\gamma is Lipschitz continuous, by (2.15) we have

|γ(yεi(θεi(x),x),yεi(θεi(x),x))γ(yεj(θεj(x),x),yεj(θεj(x),x)|\displaystyle|\gamma(y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x),\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x))-\gamma(y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x),\nabla y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x)|
c|yεi(θεi(x),x)yεj(θεj(x),x)|+c|yεi(θεi(x),x)yεj(θεj(x),x)|\displaystyle\quad\leq c|y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x)|+c|\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-\nabla y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x)|
c|yεi(θεi(x),x)yεj(θεi(x),x)|+c|yεi(θεi(x),x)yεj(θεi(x),x)|\displaystyle\quad\leq c|y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-y^{j}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)|+c|\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-\nabla y^{j}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)|
+c|yεj(θεi(x),x)yεj(θεj(x),x)|+c|yεj(θεi(x),x)yεj(θεj(x),x)|\displaystyle\qquad+c|y^{j}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x)|+c|\nabla y^{j}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)-\nabla y^{j}_{\varepsilon}(\theta^{j}_{\varepsilon}(x),x)|
cyεiyεjC1(Q¯)+cθεiθεjC(U¯)αxU¯.\displaystyle\quad\leq c\|y^{i}_{\varepsilon}-y^{j}_{\varepsilon}\|_{C^{1}(\overline{Q})}+c\|\theta^{i}_{\varepsilon}-\theta^{j}_{\varepsilon}\|^{\alpha}_{C(\overline{U})}\quad\forall\ x\in\overline{U}.

Together with convergences (4.16)–(4.17), this proves that xγ(yεi(θεi(x),x),yεi(θεi(x),x))x\mapsto\gamma(y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x),\nabla y^{i}_{\varepsilon}(\theta^{i}_{\varepsilon}(x),x)) converges to xγ(yε(θε(x),x),yε(θε(x),x))x\mapsto\gamma(y_{\varepsilon}(\theta_{\varepsilon}(x),x),\nabla y_{\varepsilon}(\theta_{\varepsilon}(x),x)) uniformly in U¯\overline{U}. By the stability of the eikonal equation with respect to the uniform convergence of the data, see, e.g., [23, Prop. 1.2], θε\theta_{\varepsilon} satisfies (2.18)–(2.19) along with the coefficient xγ(yε(θε(x),x),yε(θε(x),x))x\mapsto\gamma(y_{\varepsilon}(\theta_{\varepsilon}(x),x),\nabla y_{\varepsilon}(\theta_{\varepsilon}(x),x)). Moreover, since the bounds (4.6)–(4.8) are independent of θε\theta_{\varepsilon}, following the argument of the proof of Proposition 4.1, we can pass to the limit in the Euler–Lagrange equation (2.1) and conclude the proof of Theorem 2.1 in the case ε>0\varepsilon>0.

5. Proof of Theorem 2.1: sharp-interface case

The existence of weak/viscosity solutions in the sharp-interface case ε=0\varepsilon=0 is obtained by passing to the limit as ε0\varepsilon\to 0 in sequences of weak/viscosity solutions (yε,θε)(y_{\varepsilon},\theta_{\varepsilon}) of the diffused-interface problem.

Notice at first that the θε\theta_{\varepsilon} are uniformly Lipschitz continuous, see (4.12). Bounds (4.6)–(4.8) are independent of ε\varepsilon and imply that there exist not relabeled subsequences such that

yεyweakly- inL(0,T;WΓD2,p(U;d))H1(0,T;H1(U;d)),\displaystyle y_{\varepsilon}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}y\quad\text{weakly-$*$ in}\ L^{\infty}(0,T;W^{2,p}_{\Gamma_{D}}(U;\mathbb{R}^{d}))\cap H^{1}(0,T;H^{1}(U;\mathbb{R}^{d})), (5.1)
yεystrongly inC1,α(Q¯;d),\displaystyle y_{\varepsilon}\to y\quad\text{strongly in}\ C^{1,\alpha}(\overline{Q};\mathbb{R}^{d}), (5.2)
θεθstrongly inC(U¯)\displaystyle\theta_{\varepsilon}\to\theta\quad\text{strongly in}\ C(\overline{U}) (5.3)

for some α(0,1)\alpha\in(0,1). Hence, it is enough to prove that we can pass to the limit ε0\varepsilon\rightarrow 0 in equation (2.1). The convergence of the loading and second-gradient term can be obtained as in the proof of Proposition 4.1. Moreover, the level sets {θ(x)=t}\{\theta(x)=t\} have Lebesgue measure zero by (4.13). Hence, by the assumptions (2.1) on hεh_{\varepsilon} and the uniform convergence (5.3) of (θε)ε(\theta_{\varepsilon})_{\varepsilon}, we have

hεL()1,hε(θε(x)t)h0(θ(x)t) for a.e. (t,x)Q,\displaystyle\|h_{\varepsilon}\|_{L^{\infty}(\mathbb{R})}\leq 1,\quad h_{\varepsilon}(\theta_{\varepsilon}(x){-}t)\rightarrow h_{0}(\theta(x){-}t)\quad\text{ for a.e. }(t,x)\in Q,

and that (t,x)hε(θε(x)t)(t,x)\mapsto h_{\varepsilon}(\theta_{\varepsilon}(x){-}t) converges to (t,x)h0(θ(x)t)(t,x)\mapsto h_{0}(\theta(x){-}t) strongly in L2(Q)L^{2}(Q). On the other hand, by (2.3), the convergence (5.1), and the Aubin–Lions Lemma, for almost every (t,x)Q(t,x)\in Q and for i=0,1i=0,1, we have that

|FWi(yε)|c,FWi(yε)FWi(y).|\partial_{F}W_{i}(\nabla y_{\varepsilon})|\leq c,\quad\partial_{F}W_{i}(\nabla y_{\varepsilon})\rightarrow\partial_{F}W_{i}(\nabla y).

Thus, by Lebesgue’s Dominated Convergence Theorem

0TUFWε(θεt,yε)dxdt\displaystyle\int_{0}^{T}\int_{U}\partial_{F}W_{\varepsilon}(\theta_{\varepsilon}{-}t,\nabla y_{\varepsilon})\,{\rm d}x\,{\rm d}t\rightarrow 0TUh0(θt)FWr(y)dxdt\displaystyle\int_{0}^{T}\int_{U}h_{0}(\theta{-}t)\partial_{F}W^{r}(\nabla y)\,{\rm d}x\,{\rm d}t
+0TU(1h0(θt))FWa(y)dxdt.\displaystyle+\int_{0}^{T}\int_{U}(1{-}h_{0}(\theta{-}t))\partial_{F}W^{a}(\nabla y)\,{\rm d}x\,{\rm d}t.

Furthermore, by using convergence (5.1), we get

0TUF˙Rε(θεt,yε,y˙ε):zdxdt\displaystyle\int_{0}^{T}\int_{U}\partial_{\dot{F}}R_{\varepsilon}(\theta_{\varepsilon}{-}t,\nabla y_{\varepsilon},\nabla\dot{y}_{\varepsilon}){:}\nabla z\,{\rm d}x\,{\rm d}t\rightarrow 0TUh0(θt)F˙Rr(y,y˙):zdxdt\displaystyle\int_{0}^{T}\int_{U}h_{0}(\theta{-}t)\partial_{\dot{F}}R^{r}(\nabla y,\nabla\dot{y}){:}\nabla z\,{\rm d}x\,{\rm d}t
+0TU(1h0(θt))F˙Ra(y,y˙):zdxdt,\displaystyle+\int_{0}^{T}\int_{U}(1{-}h_{0}(\theta{-}t))\partial_{\dot{F}}R^{a}(\nabla y,\nabla\dot{y}){:}\nabla z\,{\rm d}x\,{\rm d}t,

which concludes the proof.

Acknowledgements

This research was funded in whole or in part by the Austrian Science Fund (FWF) projects 10.55776/F65, 10.55776/I5149, 10.55776/P32788, 10.55776/I4354, as well as by the OeAD-WTZ project CZ 09/2023. For open-access purposes, the authors have applied a CC BY public copyright license to any author-accepted manuscript version arising from this submission. Part of this research was conducted during a visit to the Mathematical Institute of Tohoku University, whose warm hospitality is gratefully acknowledged.

References

  • [1] G. Alberti, S. Bianchini, G. Crippa. Structure of level sets and Sard-type properties of Lipschitz maps. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 12 (2013), no. 4, 863–902.
  • [2] R. Badal, M. Friedrich, M. Kružík. Nonlinear and linearized models in thermoviscoelasticity. Arch. Ration. Mech. Anal. 247 (2023), no. 1, Paper No. 5, 73 pp.
  • [3] K. Bangert, G. Dolzmann. Stress-modulated growth in the presence of nutrients—existence and uniqueness in one spatial dimension. ZAMM Z. Angew. Math. Mech. 103 (2023), no. 10, Paper No. e202200558, 29 pp.
  • [4] M. Bardi, I. Capuzzo-Dolcetta. Optimal control and viscosity solutions of Hamilton-Jacobi-Bellman equations. Systems & Control Foundations & Applications Birkhäuser Boston, Inc., Boston, 1997.
  • [5] G. Barles. Solutions de viscosité des équations de Hamilton-Jacobi. Math. Appl., 17 Springer-Verlag, Paris, 1994.
  • [6] N. Bellomo, N. K. Li, P. K. Maini. On the foundations of cancer modelling: selected topics, speculations, and perspectives. Math. Models Methods Appl. Sci. 18 (2008), no.4, 593–646.
  • [7] A. Bressan, M. Lewicka. A model of controlled growth. Arch. Ration. Mech. Anal. 227 (2018), no. 3, 1223–1266.
  • [8] C. B. Brown, L. E. Goodman. Gravitational stresses in accreted bodies. Proc. R. Soc. Lond. A, 276 (1963), no. 1367, 571–576.
  • [9] A. Češík, G. Gravina, M. Kampschulte. Inertial evolution of non-linear viscoelastic solids in the face of (self-)collision. Calc. Var. Partial Differential Equations, 63 (2024), no. 2, 48 pp.
  • [10] A. Češík, G. Gravina, M. Kampschulte. Inertial (self-)collisions of viscoelastic solids with Lipschitz boundaries. arXiv:2312.00431, 2023.
  • [11] M. Crandall, H. Ishii, P. L. Lions. User’s guide to viscosity solutions of second order partial differential equations. Bull. Amer. Math. Soc. (N.S.), 27 (1992), no. 1, 1–67.
  • [12] E. Davoli, K. Nik, U. Stefanelli. Existence results for a morphoelastic model. ZAMM Z. Angew. Math. Mech. 103 (2023), no. 7, 25 pp.
  • [13] E. Davoli, K. Nik, U. Stefanelli, G. Tomassetti. An existence result for accretive growth in elastic solids. Math. Models Methods Appl. 34 (2024), no. 11, 2169–2190.
  • [14] Y. Dong-Hee, C. Pil-Ryung, K. Ji-Hee, G. Martin, Y. Jong-Kyu. A phase field model for phase transformation in an elastically stressed binary alloy. Modelling Simul. Mater. Sci. Eng., 13 (2005), 299.
  • [15] J. Dumais, D. Kwiatkowska. Analysis of surface growth in shoot apices, The Plant J. 31 (2001), no. 2, 229–241.
  • [16] H. Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York, Inc., New York, 1969.
  • [17] M. Fournier, H. Baillères, B. Chanson. Tree biomechanics: growth, cumulative prestresses, and reorientations. Biomimetics, 2 (1994), no. 3, 229–251.
  • [18] J.-F. Ganghoffer, I. Goda. A combined accretion and surface growth model in the framework of irreversible thermodynamics. Int. J. Engrg. Sci. 127 (2018), 53–79.
  • [19] J.-F. Ganghoffer, P. I. Plotnikov, J. Sokolowski. Nonconvex model of material growth: mathematical theory. Arch. Ration. Mech. Anal. 230 (2018), no. 3, 839–910.
  • [20] Q. Ge, A. Sakhaei, H. Lee, et al. Multimaterial 4D printing with tailorable shape memory polymers, Scientific Reports, 6 (2016).
  • [21] A. Goriely. The mathematics and mechanics of biological growth. Interdisciplinary Applied Mathematics, 45. Springer, New York, 2017.
  • [22] T. Healey, S. Krömer. Injective weak solutions in second-gradient nonlinear elasticity. ESAIM Control Optim. Calc. Var. 15 (2009), no. 4, 863–871.
  • [23] H. Ishii. A boundary value problem of the Dirichlet type for Hamilton-Jacobi equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 16 (1989), no. 1, 105–135.
  • [24] S. Kakutani. A generalization of Brouwer’s fixed point theorem. Duke Math. J. 8 (1941), 457–459.
  • [25] S. Krömer, T. Roubíček. Quasistatic viscoelasticity with self–contact at large strains. J. Elasticity, 142 (2020), 433–445.
  • [26] M. Kružík, D. Melching, U. Stefanelli. Quasistatic evolution for dislocation–free finite plasticity. ESAIM Control Optim. Calc. Var. 26 (2020), Paper No. 123.
  • [27] M. Kružík, T. Roubíček. Mathematical methods in continuum mechanics of solids. Interaction of Mechanics and Mathematics, Springer, Cham, 2019.
  • [28] Y. Kuang, J. D. Nagy, S. E. Eikenberry. Introduction to mathematical oncology. Chapman & Hall/CRC Mathematical and Computational Biology Series. CRC Press, Boca Raton, FL, 2016.
  • [29] J. S. Langer. Instabilities and pattern forestation in crystal growth. Rev. Mod. Phys. 52 (1980), no. 1.
  • [30] J. U. Lind et al. Instrumented cardiac microphysiological devices via multimaterial three-dimensional printing, Nature Mat. 16 (2017), 303–308.
  • [31] A. Lucantonio, P. Nardinocchi, L. Teresi. Transient analysis of swelling-induced large deformations in polymer gels. J. Mech. Phys. Solids, 61 (2013), no. 1, 205–218.
  • [32] M. Masi, S. Kommu. Chapter 6 Epitaxial growth modeling. D. Crippa, D. L. Rode, M. Masi (eds.), Semiconductors and Semimetals, Elsevier, vol. 72, 2001, 185–224.
  • [33] V. Metlov. On the accretion of inhomogeneous viscoelastic bodies under finite deformations. J. Appl. Math. Mech. 49 (1985), no. 4, 490–498.
  • [34] A. Mielke, T. Roubíček. Rate-independent elastoplasticity at finite strains and its numerical approximation. Math. Models Methods Appl. Sci. 26 (2016), no. 12, 2203–2236.
  • [35] A. Mielke, T. Roubíček. Thermoviscoelasticity in Kelvin-Voigt rheology at large strains. Arch. Ration. Mech. Anal. 238 (2020), no. 1, 1–45.
  • [36] D. E. Moulton, A. Goriely, R. Chirat. Mechanical growth and morphogenesis of seashells. J. Theor. Biol. 311 (2012), 69–79.
  • [37] S. K. Naghibzadeh, N. Walkington, K. Dayal. Surface growth in deformable solids using an Eulerian formulation. J. Mech. Phys. Solids, 154 (2021), no. 104499, 13 pp.
  • [38] S. K. Naghibzadeh, N. Walkington, K. Dayal. Accretion and ablation in deformable solids with an Eulerian description: examples using the method of characteristics. Math. Mech. Solids, 27 (2022), no. 6, 989–1010.
  • [39] P. Neff. On Korn’s first inequality with non-constant coefficients. Proc. Roy. Soc. Edinburgh Sect. A, 132 (2002), no. 1, 221–243.
  • [40] S. Osher, R. Fedkiw. Level set methods and dynamic implicit surfaces. Applied Mathematical Sciences, 153. Springer-Verlag, New York, 2003.
  • [41] A. Z. Palmer, T. J. Healey. Injectivity and self-contact in second-gradient nonlinear elasticity. Calc. Var. Partial Differential Equations, 56 (2017), no. 4, Paper No. 114, 11 pp.
  • [42] W. Pompe. Korn’s first inequality with variable coefficients and its generalization. Comment. Math. Univ. Carolin. 44 (2003), no. 1, 57–70.
  • [43] D. M. Rosa, J. E. Spinelli, I. L. Ferreira, et al. Cellular/dendritic transition and microstructure evolution during transient directional solidification of Pb-Sb alloys. Metall. Mater. Trans. A, 39 (2008), 2161–2174.
  • [44] K. Schwerdtfeger, M. Sato, K. H. Tacke. Stress formation in solidifying bodies. Metall. Mat. Trans. B, 29 (1998), 1057–1068.
  • [45] J. A. Sethian. Level set methods and fast marching methods. Evolving interfaces in computational geometry, fluid mechanics, computer vision, and materials science. Second edition. Cambridge Monographs on Applied and Computational Mathematics, 3. Cambridge University Press, Cambridge, 1999.
  • [46] R. Skalak, A. Hoger. Kinematics of surface growth. J. Math. Biol. 35 (1997), no. 8, 869–907.
  • [47] R. Southwell. Introduction to the theory of elasticity for engineers and physicists. Oxford University Press, Oxford, 1941.
  • [48] F. Sozio, A. Yavari. Nonlinear mechanics of accretion. J. Nonlinear Sci. 29 (2019), no. 4, 1813–1863.
  • [49] L. A. Taber. Biomechanics of growth, remodeling, and morphogenesis. Appl. Mech. Rev. 48 (1995), no. 8, 487–545.
  • [50] M. Shibayama, T. Tanaka. Volume phase transition and related phenomena of polymer gels. In: K. Dušek (ed.), Responsive Gels: Volume Transitions I. Advances in Polymer Science, vol 109. Springer, Berlin, Heidelberg, 1993.
  • [51] D. Thompson. On growth and forms: the complete revised edition. Dover, New York, 1992.
  • [52] L. Truskinovsky, G. Zurlo. Nonlinear elasticity of incompatible surface growth. Phys. Rev. E, 99 (2019), no. 5, 053001, 13 pp.
  • [53] Q. Wang, L. Wang, W. Zhang, K. Chou. Influence of cooling rate on solidification process Ce-High Mo austenite stainless steel: nucleation, growth, and microstructure evolution. Metals, 13 (2023), no. 2, 246.
  • [54] H. Wilbuer, P. Kurzeja, J. Mosler. Phase field modeling of hyperelastic material interfaces — theory, implementation and application to phase transformations. Comput. Methods Appl. Mech. Engrg. 426 (2024), 22 pp.
  • [55] S. Wildeman, S. Sterl, C. Sun, D. Loshe. Fast dynamics of water droplets freezing from the outside in, Phys. Rev. Lett. 118 (2017), 084101.
  • [56] D. Wodarz, N. L. Komarova. Dynamics of cancer. Mathematical foundations of oncology. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2014.
  • [57] T. Yamaue, D. Masao. The stress diffusion coupling in the swelling dynamics of cylindrical gels. J. Chem. Phys. 122 (2005), 084703.
  • [58] G. Zurlo, L. Truskinovsky. Inelastic surface growth. Mech. Res. Comm. 93 (2018), 174–179.