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

A Cahn–Hilliard phase field model coupled to an Allen–Cahn model of viscoelasticity at large strains

A. Agosti111Corresponding author. E-mail address: abramo.agosti@unipv.it
Email addresses: abramo.agosti@unipv.it (A. Agosti), pierluigi.colli@unipv.it (P. Colli), harald.garcke@ur.de (H. Garcke), elisabetta.rocca@unipv.it (E. Rocca)
, P. Colli♯,‡, H. Garcke§, E. Rocca♯,‡
Abstract

We propose a new Cahn–Hilliard phase field model coupled to incompressible viscoelasticity at large strains, obtained from a diffuse interface mixture model and formulated in the Eulerian configuration. A new kind of diffusive regularization, of Allen–Cahn type, is introduced in the transport equation for the deformation gradient, together with a regularizing interface term depending on the gradient of the deformation gradient in the free energy density of the system. The designed regularization preserves the dissipative structure of the equations. We study the global existence of a weak solution for the model. While standard diffusive regularizations of the transport equation for the deformation gradient presented in literature for related phase field models coupled to viscoelasticity allows the study of existence of global weak solutions only for simplified cases, i.e. in two space dimensions and for convex elastic free energy densities of Neo–Hookean type which are independent from the phase field variable, the present diffusive regularization allows to study more general cases. In particular, we obtain the global existence of a weak solution in three space dimensions and for generic nonlinear elastic energy densities with polynomial growth, comprising the relevant cases of polyconvex Mooney–Rivlin and Ogden elastic energies. Also, our analysis considers elastic free energy densities which depend on the phase field variable and which can possibly degenerate for some values of the phase field variable. By means of an iterative argument based on elliptic regularity bootstrap steps, we find the maximum allowed polynomial growths of the Cahn–Hilliard potential and the elastic energy density which guarantee the existence of a solution in three space dimensions. We also propose two kinds of unconditionally energy stable finite element approximations of the model, based on convex splitting ideas and on the use of a scalar auxiliary variable respectively, proving the existence and stability of discrete solutions. We finally report numerical results for different test cases with shape memory alloy type free energy, showing the interplay between phase separation and finite elasticity in determining the topology of stationary states with pure phases characterized by different elastic properties.

Department of Mathematics, University of Pavia, 27100 Pavia, Italy.

Research Associate at the IMATI–C.N.R. Pavia, 27100 Pavia, Italy.

§ Fakultät für Mathematik, Universität Regensburg, 93040 Regensburg, Germany.

Keywords: Cahn–Hilliard \cdot Allen–Cahn \cdot Viscoelasticity \cdot Large elastic deformations \cdot Existence of weak solutions \cdot Gradient-stable finite element approximations

2020 Mathematics Subject Classification: 35A01 \cdot 35Q35 \cdot 35Q74 \cdot 65M60 \cdot 74B20 \cdot 74F10 \cdot 74H15 \cdot 74H20

1 Introduction

The study of Cahn–Hilliard phase field models coupled with finite viscoelasticity has gained increasing interest in the recent literature, cf., e.g., [6, 12, 2, 24]. These models describe the phase separation phenomena for multiphase materials in presence of elastic interactions between the materials constituents, and may find applications e.g. in soft matter dynamics [6], tumor growth dynamics [12], neurological and neuromuscular deseases (see the discussion in [2]).

In these works both the phase field and the viscoelasticity governing equations are formulated in the Eulerian reference configuration. Hence, the main state variable for elasticity is the velocity field, and the deformation gradient, entering in the elasticity constitutive assumptions, is determined solving a transport equation in terms of the velocity and the velocity gradient. A similar modeling approach for evolutionary models with finite viscoelasticity was implemented also in [3], which studies the Oldroyd-B model for a dilute polymeric fluid, in [7], which studies the motion of a class of incompressible heat-conducting viscoelastic rate-type fluids with stress-diffusion, in [4, 11], which study problems in magnetoelasticity, in [18], which studies an evolutionary non-isothermal viscoelastic problem, and in [20], which studies the diffusion of a solvent in a saturated hyperelastic porous solid of viscoelastic Kelvin-Voigt type at large strains.

A major challenge in studying the existence of weak solutions for phase field models coupled with finite viscoelasticity is to obtain results in three space dimensions and for generic elastic energy densities which are highly nonlinear in the deformation gradient and which may depend on the phase field variable. Indeed, in [4, 6, 11, 12], employing a second order diffusive regularization in the transport equation for the deformation gradient (i.e. adding a term proportional to the Laplacian of the deformation gradient), the existence of global weak solutions for the model is proved in two space dimensions and for elastic energy densities of Neo–Hookean type, i.e. quadratic in the deformation gradient, and independent of the phase field variable. The diffusive regularization of the transport equation, which is needed to increase the regularity of the deformation gradient and gain compactness of its approximations, breaks the kinematic relationship between the velocity and the deformation gradient variables. In [2], we designed a specific second order diffusive regularization of the transport equation for the deformation gradient, depending on both the phase field variable and on the deformation gradient, which allowed us to prove the existence of global weak solutions for elastic energy densities of Neo–Hookean type which depend on the phase field variable. The degeneracy of the elastic energy density depending on the phase field variable was not allowed in the framework of [2]. Also, existence results were obtained in three space dimensions by adding a further viscous regularization to the Cahn–Hilliard equation. By employing a different regularization approach of the model equations, obtained by adding a dissipative contribution to the Cauchy stress tensor which involves high order nonlinear terms in the small strain rate (i.e. the symmetric part of the velocity gradient), in [18] the author obtained existence and regularity results of a distributional solution for an evolutionary non-isothermal viscoelastic problem with nonlinear and compressible elastic energy. The latter regularization approach preserves the kinematic relationship between the velocity and the deformation gradient variables. Its main drawback is the introduction of high order nonlinear terms involving the velocity gradient in the definition of the Cauchy stress tensor, making numerical approximations of the model not straightforward.

In the present paper, we introduce a new kind of diffusive regularization, of Allen–Cahn type, in the transport equation for the deformation gradient, complemented by the introduction of a regularizing interface term depending on the gradient of the deformation gradient in the free energy density of the system. The designed regularization preserves the dissipative structure of the equations. The idea for the introduction of this kind of regularization comes from an observation by Roubicek [19, Remark 3], which suggested the possibility to consider a Cahn–Hilliard regularization of the transport equation for the deformation gradient in order to obtain stronger existence results with respect to those available in the literature and based on diffusive regularizations of the transport equation for the deformation gradient. The introduction of an interface contribution for the deformation gradient in the free energy enhances the space and time regularity of the deformation gradient, while increasing the degree of nonlinearity of the coupled system. In particular, the Cauchy stress tensor contains second order and quadratic first order terms in the deformation gradient. In our framework, the regularity of the latter terms is obtained by adding an Allen–Cahn type second order diffusive regularization in the transport equation for the deformation gradient. Following the Allen–Cahn literature, we consider a dual mixed Allen–Cahn formulation of the transport equation for the deformation gradient, introducing a dual variable of the deformation gradient which enters also in the expression for the Cauchy stress tensor. This theoretical framework allows us to obtain the existence of global weak solutions in three space dimensions and for generic nonlinear elastic energy densities with polynomial growth, which are coupled to the phase field variable and which possibly degenerate for some values of the phase field variable.

The resulting PDE system for the considered model is the following:

{νΔ𝐯+s=μϕ+div(𝐌𝐅T)+(𝐅)𝐌,div𝐯=0,𝐅t+(𝐯)𝐅(𝐯)𝐅+γ𝐌=0,𝐌=𝐅w(ϕ,𝐅)λΔ𝐅,ϕt+𝐯ϕdiv(b(ϕ)μ)=0,μ=ψ(ϕ)Δϕ+ϕw(ϕ,𝐅),\begin{cases}-\nu\Delta\mathbf{v}+\nabla s=\mu\nabla\phi+\operatorname{div}\left(\mathbf{M}\mathbf{F}^{T}\right)+\left(\nabla\mathbf{F}\right)\odot\mathbf{M},\\ \\ \operatorname{div}\mathbf{v}=0,\\ \\ \frac{\partial\mathbf{F}}{\partial t}+\left(\mathbf{v}\cdot\nabla\right)\mathbf{F}-(\nabla\mathbf{v})\mathbf{F}+\gamma\mathbf{M}=0,\\ \\ \mathbf{M}=\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F},\\ \\ \frac{\partial\phi}{\partial t}+\mathbf{v}\cdot\nabla\phi-\operatorname{div}(b(\phi)\nabla\mu)=0,\\ \\ \mu=\psi^{\prime}(\phi)-\Delta\phi+\partial_{\phi}w(\phi,\mathbf{F}),\end{cases} (1)

valid in ΩT\Omega_{T}, endowed with the boundary conditions

b(ϕ)μ𝐧=ϕ𝐧=0,[𝐅]𝐧=𝟎,𝐯=𝟎,b(\phi)\nabla\mu\cdot\mathbf{n}=\nabla\phi\cdot\mathbf{n}=0,\quad\left[\nabla\mathbf{F}\right]\mathbf{n}=\mathbf{0},\quad\mathbf{v}=\mathbf{0}, (2)

on Ω×[0,T]\partial\Omega\times[0,T], and with proper initial conditions. Here, 𝐯\mathbf{v} is the velocity field, ss is the pressure, 𝐅\mathbf{F} is the elastic deformation gradient, 𝐌\mathbf{M} is its dual variable, ϕ\phi is the phase field variable and μ\mu is the chemical potential. Moreover, ν\nu is a physical parameter, representing the viscosity of the material, while γ\gamma and λ\lambda are positive regularization parameters. The function b(ϕ)b(\phi) represents a positive phase-dependent mobility, ψ(ϕ)\psi(\phi) represents the bulk potential of the Cahn–Hilliard equation, and w(ϕ,𝐅)w(\phi,\mathbf{F}) is the elastic free energy density.

By means of an iterative argument based on elliptic regularity bootstrap steps applied to (1)4, we find the maximum allowed polynomial growths of the Cahn–Hilliard potential and the elastic energy density which guarantee existence of a solutions in three space dimensions.

We observe that the dual mixed formulation of System (1) is particularly suitable to design standard and efficient numerical approximations. Hence, we also propose two kinds of unconditionally energy stable finite element approximations of the model, based on convex splitting ideas and on the use of a scalar auxiliary variable, proving the existence and stability of discrete solutions. We finally show numerical tests for different test cases with shape memory alloy type free energy, proving the interplay between phase separation and finite elasticity in determining the topology of stationary states with pure phases characterized by different elastic properties.

Hence, the novelties of the present work with respect to previous studies in literature are the following:

  • Proof of existence of global weak solutions for a Cahn–Hilliard phase field model coupled to viscoelasticity at large strains in three space dimensions, based on a new Allen–Cahn type regularization of finite viscoelasticity, for generic finite elastic energy densities with polynomial growth, which are coupled to the phase field variable and which possibly degenerate for some values of the phase field variable;

  • Design and implementation of standard and efficient well posed and unconditionally energy stable finite element approximations of the model.

The paper is organized as follows. In Section 22 we introduce some notation regarding the employed tensor calculus and functional analysis. In Section 33 we develop the model derivation. In Section 44 we report the study of existence for a global weak solution of System (1). In Section 55 we report the design of finite element approximations of the model and the study of existence and stability of discrete solutions. In Section 66 we show results of numerical simulations. Finally, Section 77 contains concluding remarks and future perspectives.

2 Notation

Let Ω3\Omega\subset\mathbb{R}^{3} be an open bounded domain in 3\mathbb{R}^{3}. Let T>0T>0 denote some final time, and set ΩT:=Ω×(0,T)\Omega_{T}:=\Omega\times(0,T). We start by introducing the notation for vectorial and tensorial calculus. Given 𝐚,𝐛3\mathbf{a},\mathbf{b}\in\mathbb{R}^{3}, we denote by 𝐚𝐛\mathbf{a}\cdot\mathbf{b}\in\mathbb{R} their canonical scalar product in 3\mathbb{R}^{3}, with associated norm |𝐚|:=(𝐚𝐚)12|\mathbf{a}|:=(\mathbf{a}\cdot\mathbf{a})^{\frac{1}{2}}, and by 𝐚𝐛3×3\mathbf{a}\otimes\mathbf{b}\in\mathbb{R}^{3\times 3} their tensorial product. We indicate by {𝐞i}i=1,2,3\{\mathbf{e}_{i}\}_{i=1,2,3} the canonical basis of 3\mathbb{R}^{3}. Given two second order tensors 𝐀,𝐁3×3\mathbf{A},\mathbf{B}\in\mathbb{R}^{3\times 3}, we denote by 𝐀:𝐁\mathbf{A}\colon\mathbf{B}\in\mathbb{R} their Frobenius scalar product in 3×3\mathbb{R}^{3\times 3}, i.e. by components 𝐀:𝐁:=i,j=13AijBij\mathbf{A}\colon\mathbf{B}:=\sum_{i,j=1}^{3}A_{ij}B_{ij}, with associated norm |𝐀|:=(𝐀:𝐀)12|\mathbf{A}|:=(\mathbf{A}\colon\mathbf{A})^{\frac{1}{2}}. We indicate by {𝐞ij:=𝐞i𝐞j}i,j=1,2,3\{\mathbf{e}_{ij}:=\mathbf{e}_{i}\otimes\mathbf{e}_{j}\}_{i,j=1,2,3} the canonical basis of 3×3\mathbb{R}^{3\times 3}. Given a second order tensor 𝐀3×3\mathbf{A}\in\mathbb{R}^{3\times 3}, we indicate with the notation 𝐀[i]3\mathbf{A}^{[i]}\in\mathbb{R}^{3} the ii-th column vector of 𝐀\mathbf{A}. Given two third order tensors 𝐂,𝐃3×3×3\mathbf{C},\mathbf{D}\in\mathbb{R}^{3\times 3\times 3}, we denote by 𝐂\setstackgapS0.4ex\Shortstack𝐃\mathbf{C}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\mathbf{D}\in\mathbb{R} their scalar product in 3×3×3\mathbb{R}^{3\times 3\times 3}, i.e. by components

𝐂\setstackgapS0.4ex\Shortstack𝐃:=i,j,k=13CijkDijk.\mathbf{C}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\mathbf{D}:=\sum_{i,j,k=1}^{3}C_{ijk}D_{ijk}.

We also introduce the operation, defined by components,

(𝐂𝐁)k:=ij=13𝐂ijk𝐁ij,\displaystyle\left(\mathbf{C}\odot\mathbf{B}\right)_{k}:=\sum_{ij=1}^{3}\mathbf{C}_{ijk}\mathbf{B}_{ij},
(𝐂𝐃)kl:=ij=13𝐂ijk𝐃ijl,\displaystyle\left(\mathbf{C}\odot\mathbf{D}\right)_{kl}:=\sum_{ij=1}^{3}\mathbf{C}_{ijk}\mathbf{D}_{ijl},

which contracts respectively a third order tensor 𝐂3×3×3\mathbf{C}\in\mathbb{R}^{3\times 3\times 3} and a second order tensor 𝐁3×3\mathbf{B}\in\mathbb{R}^{3\times 3} to a vector 𝐂𝐁3\mathbf{C}\odot\mathbf{B}\in\mathbb{R}^{3} and two third order tensors 𝐂,𝐃3×3×3\mathbf{C},\mathbf{D}\in\mathbb{R}^{3\times 3\times 3} to a second order tensor 𝐂𝐃3×3\mathbf{C}\odot\mathbf{D}\in\mathbb{R}^{3\times 3}.

We denote by Lp(Ω;K)L^{p}(\Omega;K) and Wr,p(Ω;K)W^{r,p}(\Omega;K) the standard Lebesgue and Sobolev spaces of functions defined on Ω\Omega with values in a set KK, where KK may be \mathbb{R} or a multiple power of \mathbb{R}, and by Lp(0,t;V)L^{p}(0,t;V) the Bochner space of functions defined on (0,t)(0,t) with values in a Banach space VV, with 1p1\leq p\leq\infty. If KK\equiv\mathbb{R}, we simply write Lp(Ω)L^{p}(\Omega) and Wr,p(Ω)W^{r,p}(\Omega). Moreover, when stating general results which are valid for both functions with scalar or vectorial or tensorial values, we write fLpf\in L^{p}, fWr,pf\in W^{r,p}, without specifying if ff is a function with scalar, vectorial or tensorial values. For a normed space XX, the associated norm is denoted by ||||X||\cdot||_{X}. In the case p=2p=2, we use the notations H1:=W1,2H^{1}:=W^{1,2} and H2:=W2,2H^{2}:=W^{2,2}, and we denote by (,)(\cdot,\cdot) and ||||||\cdot|| the L2L^{2} scalar product and induced norm between functions with scalar, vectorial or tensorial values. The dual space of a Banach space YY is denoted by YY^{\prime}. The duality pairing between H1(Ω;K)H^{1}(\Omega;K) and (H1(Ω;K))\left(H^{1}(\Omega;K)\right)^{\prime} is denoted by <,><\cdot,\cdot>. Moreover, we denote by Ck(Ω;K),Cck(Ω;K)C^{k}(\Omega;K),C_{c}^{k}(\Omega;K) the spaces of continuously differentiable functions (respectively with compact support) up to order kk defined on Ω\Omega with values in a set KK, and with Ck([0,t];V)C^{k}([0,t];V), k0k\geq 0, the spaces of continuously differentiable functions up to order kk from [0,t][0,t] to the space VV. We finally introduce the spaces

Ldiv2(Ω,3):={𝐮Cc(Ω,3):div𝐮=0inΩ}¯||||L2(Ω;3),\displaystyle L_{\text{div}}^{2}(\Omega,\mathbb{R}^{3}):=\overline{\{\mathbf{u}\in C_{c}^{\infty}(\Omega,\mathbb{R}^{3}):\,\text{div}\mathbf{u}=0\;\text{in}\;\Omega\}}^{||\cdot||_{L^{2}(\Omega;\mathbb{R}^{3})}},
H0,div1(Ω,3):={𝐮Cc(Ω,3):div𝐮=0inΩ}¯||||H01(Ω;3).\displaystyle H_{0,\text{div}}^{1}(\Omega,\mathbb{R}^{3}):=\overline{\{\mathbf{u}\in C_{c}^{\infty}(\Omega,\mathbb{R}^{3}):\,\text{div}\mathbf{u}=0\;\text{in}\;\Omega\}}^{||\cdot||_{H_{0}^{1}(\Omega;\mathbb{R}^{3})}}.

In the following, CC denotes a generic positive constant independent of the unknown variables, the discretization and the regularization parameters, the value of which might change from line to line; C1,C2,C_{1},C_{2},\dots indicate generic positive constants whose particular value must be tracked through the calculations; C(a,b,)C(a,b,\dots) denotes a constant depending on the nonnegative parameters a,b,a,b,\dots.

3 Model derivation

The phase field model coupled with finite viscoelasticity which we analyze in this paper is a particular case of a class of phase field models derived in our previous contribution [2], obtained by considering a binary, saturated, closed and non reactive mixture of a solid elastic component and a liquid component, whose dynamics is driven by the microscopic interactions between its constituents and by their macroscopic visco-elastic behaviour. The mass and momentum balance of the mixture were derived using a generalized form of the principles of virtual powers, giving constitutive assumptions satisfying the first and second law of thermodynamics in isothermal situations.

In particular, we consider the following form for the free energy EE of the system in Eulerian coordinates, associated to an arbitrary subregion of the mixture R(t)ΩR(t)\subset\Omega moving with the mixture:

E(ϕ,ϕ,𝐅,𝐅)=R(t)e(ϕ,ϕ,𝐅,𝐅)=R(t)(12|ϕ|2+ψ(ϕ)+w(ϕ,𝐅)+λ2|𝐅|2),\displaystyle E(\phi,\nabla\phi,\mathbf{F},\nabla\mathbf{F})=\int_{R(t)}e\left(\phi,\nabla\phi,\mathbf{F},\nabla\mathbf{F}\right)=\int_{R(t)}\left(\frac{1}{2}|\nabla\phi|^{2}+\psi(\phi)+w(\phi,\mathbf{F})+\frac{\lambda}{2}|\nabla\mathbf{F}|^{2}\right), (3)

where ϕ\phi is the volume concentration of the elastic phase, 𝐅\mathbf{F} is the deformation gradient associated to the motion of the elastic phase, w(ϕ,𝐅)w(\phi,\mathbf{F}) is the hyperelastic free energy density and ψ(ϕ)\psi(\phi) represents a bulk energy due to the mechanical interactions of the micro–components. A term proportional to 12|ϕ|2+ψ(ϕ)\frac{1}{2}|\nabla\phi|^{2}+\psi(\phi) represents a surface energy contribution of the interface between the phases, expressed through a diffuse interface approach, while the term proportional to |𝐅|2|\nabla\mathbf{F}|^{2} is associated to elastic energy contributions from interfaces in the elastic material. Here, λ>0\lambda>0 is a regularization parameter, while the interface thickness parameter associated to the surface energy between the phases is taken to be equal to 11 for ease of notation. We also consider the incompressibility constraint for the partial volume of the elastic phase. The particular model is obtained from [2, System (37)], with the regularization parameters θ=δ=0\theta=\delta=0 and employing the following simplifying assumptions:

  • The momentum transfer in the mixture due to shear stresses in the liquid is negligible with respect to the momentum transfer between the solid and the liquid components;

  • The momentum exchange between the phases is induced by a Darcy-like flow of the liquid phase through the porous-permeable solid matrix associated to the solid phase, with constant permeability.

With these assumptions, the PDE system takes the form

{νΔ𝐯+s=div(𝐅w(ϕ,𝐅)𝐅T)div(ϕϕ+λ(Δ𝐅)𝐅T+λ𝐅𝐅),div𝐯=0,𝐅t+(𝐯)𝐅(𝐯)𝐅=0,ϕt+𝐯ϕdiv(b(ϕ)μ)=0,μ=ψ(ϕ)Δϕ+ϕw(ϕ,𝐅),\begin{cases}\displaystyle-\nu\Delta\mathbf{v}+\nabla s=\operatorname{div}\left(\partial_{\mathbf{F}}w(\phi,\mathbf{F})\mathbf{F}^{T}\right)-\operatorname{div}\left(\nabla\phi\otimes\nabla\phi+\lambda(\Delta\mathbf{F})\mathbf{F}^{T}+\lambda\nabla\mathbf{F}\odot\nabla\mathbf{F}\right),\\ \\ \displaystyle\operatorname{div}\mathbf{v}=0,\\ \\ \displaystyle\frac{\partial\mathbf{F}}{\partial t}+\left(\mathbf{v}\cdot\nabla\right)\mathbf{F}-(\nabla\mathbf{v})\mathbf{F}=0,\\ \\ \displaystyle\frac{\partial\phi}{\partial t}+\mathbf{v}\cdot\nabla\phi-\operatorname{div}(b(\phi)\nabla\mu)=0,\\ \\ \displaystyle\mu=\psi^{\prime}(\phi)-\Delta\phi+\partial_{\phi}w(\phi,\mathbf{F}),\end{cases} (4)

valid in ΩT\Omega_{T}, endowed with the boundary conditions

b(ϕ)μ𝐧=ϕ𝐧=0,[𝐅]𝐧=𝟎,𝐯=𝟎,b(\phi)\nabla\mu\cdot\mathbf{n}=\nabla\phi\cdot\mathbf{n}=0,\quad\left[\nabla\mathbf{F}\right]\mathbf{n}=\mathbf{0},\quad\mathbf{v}=\mathbf{0}, (5)

on Ω×[0,T]\partial\Omega\times[0,T], and with proper initial conditions. Here, ν>0\nu>0 is a viscosity parameter, ss is the solid pressure and b(ϕ)b(\phi) is a positive mobility function. System (4) is analogous to [2, System (39)], with (3) as the free energy of the system.

Using the relation

div(ϕϕ)λdiv(𝐅𝐅)\displaystyle-\operatorname{div}\left(\nabla\phi\otimes\nabla\phi\right)-\lambda\operatorname{div}\left(\nabla\mathbf{F}\odot\nabla\mathbf{F}\right) (6)
=(12|ϕ|2+ψ(ϕ))+μϕϕwϕλ(12|𝐅|2)λ(𝐅)Δ𝐅\displaystyle=-\nabla\left(\frac{1}{2}|\nabla\phi|^{2}+\psi(\phi)\right)+\mu\nabla\phi-\partial_{\phi}w\nabla\phi-\lambda\nabla\left(\frac{1}{2}|\nabla\mathbf{F}|^{2}\right)-\lambda\left(\nabla\mathbf{F}\right)\odot\Delta\mathbf{F} (7)
=(12|ϕ|2+ψ(ϕ)+w(ϕ,𝐅)+λ2|𝐅|2)e(ϕ,ϕ,𝐅,𝐅)+μϕ+(𝐅)(𝐅wλΔ𝐅),\displaystyle=-\nabla\underbrace{\left(\frac{1}{2}|\nabla\phi|^{2}+\psi(\phi)+w(\phi,\mathbf{F})+\frac{\lambda}{2}|\nabla\mathbf{F}|^{2}\right)}_{e\left(\phi,\nabla\phi,\mathbf{F},\nabla\mathbf{F}\right)}+\mu\nabla\phi+\left(\nabla\mathbf{F}\right)\odot\left(\partial_{\mathbf{F}}w-\lambda\Delta\mathbf{F}\right),

renaming ss+es\leftarrow s+e and adding a proper diffusive regularization, with regularization parameter γ>0\gamma>0, in the transport equation (4)3, we get

{νΔ𝐯+s=μϕ+div([𝐅w(ϕ,𝐅)λΔ𝐅]𝐅T)+(𝐅)(𝐅w(ϕ,𝐅)λΔ𝐅),div𝐯=0,𝐅t+(𝐯)𝐅(𝐯)𝐅+γ(𝐅w(ϕ,𝐅)λΔ𝐅)=0,ϕt+𝐯ϕdiv(b(ϕ)μ)=0,μ=ψ(ϕ)Δϕ+ϕw(ϕ,𝐅).\begin{cases}-\nu\Delta\mathbf{v}+\nabla s=\mu\nabla\phi+\operatorname{div}\left(\left[\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F}\right]\mathbf{F}^{T}\right)+\left(\nabla\mathbf{F}\right)\odot\left(\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F}\right),\\ \\ \operatorname{div}\mathbf{v}=0,\\ \\ \frac{\partial\mathbf{F}}{\partial t}+\left(\mathbf{v}\cdot\nabla\right)\mathbf{F}-(\nabla\mathbf{v})\mathbf{F}+\gamma\left(\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F}\right)=0,\\ \\ \frac{\partial\phi}{\partial t}+\mathbf{v}\cdot\nabla\phi-\operatorname{div}(b(\phi)\nabla\mu)=0,\\ \\ \mu=\psi^{\prime}(\phi)-\Delta\phi+\partial_{\phi}w(\phi,\mathbf{F}).\end{cases} (8)

We introduce the auxiliary variable

𝐌:=𝐅w(ϕ,𝐅)λΔ𝐅,\mathbf{M}:=\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F},

and rewrite system (8) as (cf. also System (1) in the Introduction)

{νΔ𝐯+s=μϕ+div(𝐌𝐅T)+𝐅𝐌,div𝐯=0,𝐅t+(𝐯)𝐅(𝐯)𝐅+γ𝐌=𝟎,𝐌=𝐅w(ϕ,𝐅)λΔ𝐅,ϕt+𝐯ϕdiv(b(ϕ)μ)=0,μ=ψ(ϕ)Δϕ+ϕw(ϕ,𝐅).\begin{cases}-\nu\Delta\mathbf{v}+\nabla s=\mu\nabla\phi+\operatorname{div}\left(\mathbf{M}\mathbf{F}^{T}\right)+\nabla\mathbf{F}\odot\mathbf{M},\\ \\ \operatorname{div}\mathbf{v}=0,\\ \\ \frac{\partial\mathbf{F}}{\partial t}+\left(\mathbf{v}\cdot\nabla\right)\mathbf{F}-(\nabla\mathbf{v})\mathbf{F}+\gamma\mathbf{M}=\mathbf{0},\\ \\ \mathbf{M}=\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F},\\ \\ \frac{\partial\phi}{\partial t}+\mathbf{v}\cdot\nabla\phi-\operatorname{div}(b(\phi)\nabla\mu)=0,\\ \\ \mu=\psi^{\prime}(\phi)-\Delta\phi+\partial_{\phi}w(\phi,\mathbf{F}).\end{cases} (9)

valid in ΩT\Omega_{T}, endowed with the boundary conditions

b(ϕ)μ𝐧=ϕ𝐧=0,[𝐅]𝐧=𝟎,𝐯=𝟎,b(\phi)\nabla\mu\cdot\mathbf{n}=\nabla\phi\cdot\mathbf{n}=0,\quad\left[\nabla\mathbf{F}\right]\mathbf{n}=\mathbf{0},\quad\mathbf{v}=\mathbf{0}, (10)

on Ω×[0,T]\partial\Omega\times[0,T], and with initial conditions 𝐅(𝐱,0)=𝐅0(𝐱)\mathbf{F}(\mathbf{x},0)=\mathbf{F}_{0}(\mathbf{x}), ϕ(𝐱,0)=ϕ0(𝐱)\phi(\mathbf{x},0)=\phi_{0}(\mathbf{x}) for 𝐱Ω\mathbf{x}\in\Omega.

In order to proceed, we make the following assumptions:

  • A0

    Ω3\Omega\subset\mathbb{R}^{3} is a bounded domain and the boundary Ω\partial\Omega is of class C2C^{2};

  • A1

    bC0()b\in C^{0}(\mathbb{R}) and there exist b0,b1>0b_{0},b_{1}>0 such that b0b(r)b1,b_{0}\leq b(r)\leq b_{1}, r\forall r\in\mathbb{R};

  • A2

    wC1(×3×3;)w\in C^{1}(\mathbb{R}\times\mathbb{R}^{3\times 3};\mathbb{R}), and there exist d10,d2>0d_{1}\geq 0,d_{2}>0 such that d1w(r,𝐓)d2(1+|𝐓|p)-d_{1}\leq w(r,\mathbf{T})\leq d_{2}\left(1+|\mathbf{T}|^{p}\right) for all r,𝐓3×3r\in\mathbb{R},\mathbf{T}\in\mathbb{R}^{3\times 3},with p[0,6)p\in[0,6). Moreover, there exist d3,d4>0d_{3},d_{4}>0 such that |𝐓w(r,𝐓)|d3(1+|𝐓|p1)\left|\partial_{\mathbf{T}}w(r,\mathbf{T})\right|\leq d_{3}\left(1+|\mathbf{T}|^{p-1}\right) and |rw(r,𝐓)|d4(1+|𝐓|p)\left|\partial_{r}w(r,\mathbf{T})\right|\leq d_{4}\left(1+|\mathbf{T}|^{p}\right) for all r,𝐓3×3r\in\mathbb{R},\mathbf{T}\in\mathbb{R}^{3\times 3};

  • A3

    ψC1()\psi\in C^{1}(\mathbb{R}) and there exist c1>0,c20c_{1}>0,c_{2}\geq 0 such that |ψ(r)|c1(|r|l+1)|\psi^{\prime}(r)|\leq c_{1}\left(|r|^{l}+1\right), ψ(r)c2\psi(r)\geq-c_{2}, r\forall r\in\mathbb{R}, with l[0,6+8max(2,2p6))l\in\left[0,6+\frac{8}{\max(2,2p-6)}\right). Moreover, there exists a convex decomposition of ψ=ψ++ψ\psi=\psi_{+}+\psi_{-}, where ψ+\psi_{+} is convex and ψ\psi_{-} is concave, such that |ψ′′(r)|c1(|r|q+1)|\psi_{-}^{\prime\prime}(r)|\leq c_{1}\left(|r|^{q}+1\right), r\forall r\in\mathbb{R}, with q[0,4)q\in[0,4);

  • A4

    The initial data have the regularity 𝐅0H1(Ω;3×3)\mathbf{F}_{0}\in H^{1}(\Omega;\mathbb{R}^{3\times 3}), ϕ0H1(Ω)\phi_{0}\in H^{1}(\Omega).

Remark 3.1

We observe that Assumption A2 includes the situation in which w(ϕ,𝐅)w(\phi,\mathbf{F}) may degenerate in the variable ϕ\phi, i.e. w(ϕ¯,𝐅)=0w(\bar{\phi},\mathbf{F})=0 for specific values ϕ=ϕ¯\phi=\bar{\phi} (e.g. ϕ¯=0\bar{\phi}=0) and for all 𝐅3×3\mathbf{F}\in\mathbb{R}^{3\times 3}. This situation could not be dealt with in the theoretical framework introduced in [2]. Moreover, we also observe that the growth law for ψ\psi in Assumption A3 depends on the growth law for ww in Assumption A2.

Remark 3.2

Assumption A2 concerning the form of the elastic energy density w(ϕ,𝐅)w(\phi,\mathbf{F}) is quite general, and includes as particular cases many realistic elastic energy densities of nonlinear elastic materials, e.g. the generalized Ogden and the Mooney-Rivlin energy densities [17]. These latter energy densities satisfy the polyconvexity and the coercivity properties, which are required by standard models in the theory of nonlinear elastic materials. We highlight that in our theoretical framework no polyconvexity and coercivity properties are generally required to obtain analytical results. Specifically, the Mooney–Rivlin energy density with phase-dependent elastic coefficients has the form

w(ϕ,𝐅)=f1(ϕ)2(𝐅:𝐅3)+f2(ϕ)2((𝐅:𝐅)2𝐅T𝐅:𝐅T𝐅6),w(\phi,\mathbf{F})=\frac{f_{1}(\phi)}{2}(\mathbf{F}\colon\mathbf{F}-3)+\frac{f_{2}(\phi)}{2}\left(\left(\mathbf{F}\colon\mathbf{F}\right)^{2}-\mathbf{F}^{T}\mathbf{F}\colon\mathbf{F}^{T}\mathbf{F}-6\right), (11)

which, if fiC1()f_{i}\in C^{1}(\mathbb{R}) with 0fi(r)k1,i0\leq f_{i}(r)\leq k_{1,i}, |fi(r)|k2,i|f_{i}^{\prime}(r)|\leq k_{2,i}, r\forall r\in\mathbb{R}, k1,i,k2,i>0k_{1,i},k_{2,i}>0 for i=1,2i=1,2, satisfies Assumption A2 with p=4p=4. The Ogden energy density with phase-dependent elastic coefficients has the form

w(ϕ,𝐅)=i=1Nfi(ϕ)(λ1pi+λ2pi+λ3pi3),w(\phi,\mathbf{F})=\sum_{i=1}^{N}f_{i}(\phi)\left(\lambda_{1}^{p_{i}}+\lambda_{2}^{p_{i}}+\lambda_{3}^{p_{i}}-3\right), (12)

where NN\in\mathbb{N} and p1,,pNp_{1},...,p_{N} are material specific parameters and λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are the principal stretches of deformation (i.e. λ12,λ22,λ32\lambda_{1}^{2},\lambda_{2}^{2},\lambda_{3}^{2} are the eigenvalues of 𝐅T𝐅\mathbf{F}^{T}\mathbf{F}). If fiC1()f_{i}\in C^{1}(\mathbb{R}) with 0fik1,i0\leq f_{i}\leq k_{1,i}, |fi(r)|k2,i|f_{i}^{\prime}(r)|\leq k_{2,i}, k1,i,k2,i>0k_{1,i},k_{2,i}>0 and 0<pi<60<p_{i}<6 for and i=1,,Ni=1,\dots,N, (12) satisfies Assumption A2.

We state here the main theorem of the present work concerning the existence of a global weak solution to (9) in three space dimensions, which will be proved in the forthcoming sections.

Theorem 3.1

Let the assumptions A0A4 be satisfied. Then, there exists a weak solution (𝐯,𝐅,𝐌,ϕ,μ)(\mathbf{v},\mathbf{F},\mathbf{M},\phi,\mu) of (9)-(10) with

𝐯L2(0,T;H0,div1(Ω;3)),\displaystyle\mathbf{v}\in L^{2}\left(0,T;H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right)\right),
𝐅L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3)),\displaystyle\mathbf{F}\in L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3})),
𝐌L2(0,T;L2(Ω;3×3)),\displaystyle\mathbf{M}\in L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})),
t𝐅L43s(0,T;L2(Ω;3×3)),s(0,13),\displaystyle\partial_{t}\mathbf{F}\in L^{\frac{4}{3}-s}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})\right),\;\;s\in\left(0,\frac{1}{3}\right),
ϕL(0,T;H1(Ω))L8max(2,2p6)(0,T;H2(Ω)),\displaystyle\phi\in L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{\max(2,2p-6)}}(0,T;H^{2}(\Omega)),
tϕL2(0,T;(H1(Ω))),\displaystyle\partial_{t}\phi\in L^{2}(0,T;\left(H^{1}(\Omega)\right)^{\prime}),
μL2(0,T;H1(Ω)),\displaystyle\mu\in L^{2}\left(0,T;H^{1}\left(\Omega\right)\right),

such that

{νΩ𝐯:𝐮=Ωμϕ𝐮Ω(𝐌𝐅T):𝐮+Ω(𝐅𝐌)𝐮,Ωt𝐅:𝚯+Ω(𝐯)𝐅:𝚯Ω(𝐯)𝐅:𝚯+γΩ𝐌:𝚯=0,Ω𝐌:𝚷=Ω𝐅w(ϕ,𝐅):𝚷+λΩ𝐅\setstackgapS0.4ex\Shortstack𝚷,<tϕ,q>+Ω(𝐯ϕ)q+Ωb(ϕ)μq=0,Ωμr=Ωϕr+Ωψ(ϕ)r+Ωϕw(ϕ,𝐅)r,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}\colon\nabla\mathbf{u}=\int_{\Omega}\mu\nabla\phi\cdot\mathbf{u}-\int_{\Omega}\left(\mathbf{M}\mathbf{F}^{T}\right)\colon\nabla\mathbf{u}+\int_{\Omega}\left(\nabla\mathbf{F}\odot\mathbf{M}\right)\cdot\mathbf{u},\\ \displaystyle\int_{\Omega}\partial_{t}\mathbf{F}\colon\boldsymbol{\Theta}+\int_{\Omega}\left({\mathbf{v}}\cdot\nabla\right)\mathbf{F}\colon\boldsymbol{\Theta}-\int_{\Omega}\left(\nabla{\mathbf{v}}\right)\mathbf{F}\colon\boldsymbol{\Theta}+\gamma\int_{\Omega}\mathbf{M}\colon\boldsymbol{\Theta}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}\colon\boldsymbol{\Pi}=\int_{\Omega}\partial_{\mathbf{F}}w(\phi,\mathbf{F})\colon\boldsymbol{\Pi}+\lambda\int_{\Omega}\nabla\mathbf{F}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi},\\ \displaystyle<\partial_{t}\phi,q>+\int_{\Omega}\left(\mathbf{v}\cdot\nabla\phi\right)q+\int_{\Omega}b\left(\phi\right)\nabla\mu\cdot\nabla q=0,\\ \displaystyle\int_{\Omega}\mu r=\int_{\Omega}\nabla\phi\cdot\nabla r+\int_{\Omega}\psi^{\prime}\left(\phi\right)r+\int_{\Omega}\partial_{\phi}w(\phi,\mathbf{F})r,\end{cases} (13)

for a.e. t(0,T)t\in(0,T) and for all 𝐮H0,div1(Ω;3)\mathbf{u}\in H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right), 𝚯L2(Ω;3×3)\boldsymbol{\Theta}\in L^{2}\left(\Omega;\mathbb{R}^{3\times 3}\right), 𝚷H1(Ω;3×3)\boldsymbol{\Pi}\in H^{1}\left(\Omega;\mathbb{R}^{3\times 3}\right), q,rH1(Ω)q,r\in H^{1}(\Omega), satisfying the initial conditions 𝐅(,0)=𝐅0\mathbf{F}(\cdot,0)=\mathbf{F}_{0} a.e. in Ω\Omega and ϕ(,0)=ϕ0\phi(\cdot,0)=\phi_{0} a.e. in Ω\Omega.

In the following, we will introduce a proper truncation of the growth behavior of w(ϕ,𝐅)w(\phi,\mathbf{F}), depending on a truncation parameter R>1R>1, and we will define a proper Faedo–Galerkin approximation of a truncated version of (9), proving the existence of a discrete solution and studying its convergence to a continuous weak solution, as the discretization parameter tends to zero, in three space dimensions. We will then study the limit problem, at the continuous level, as the truncation parameter RR\to\infty, thus removing the truncation operation and recovering an existence result for system (9) associated to the full growth laws assumed in Assumption A2.

Before proceeding, we state some preliminary results which will be used in the analysis.

3.1 Preliminary lemmas

We state here some Sobolev embedding and interpolation results which will be used in the following calculations. We start by recalling the Gagliardo-Nirenberg inequality (see e.g. [5]).

Lemma 3.1

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded domain with Lipschitz boundary and fWm,rLqf\in W^{m,r}\cap L^{q}, q1q\geq 1, rr\leq\infty, where ff can be a function with scalar, vectorial or tensorial values. For any integer jj with 0j<m0\leq j<m, suppose there is α\alpha\in\mathbb{R} such that

j3k=(m3r)α+(1α)(3q),jmα1.j-\frac{3}{k}=\left(m-\frac{3}{r}\right)\alpha+(1-\alpha)\left(-\frac{3}{q}\right),\quad\frac{j}{m}\leq\alpha\leq 1.

Then, there exists a positive constant CC depending on Ω\Omega, m, j, q, r, and α\alpha such that

DjfLkCfWm,rαfLq1α.||D^{j}f||_{L^{k}}\leq C||f||_{W^{m,r}}^{\alpha}||f||_{L^{q}}^{1-\alpha}. (14)

We also state the following interpolation results.

Lemma 3.2

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded domain with Lipschitz boundary, p[1,6)p\in[1,6) and fL(0,T;L2)L2(0,T;W1,6p1)f\in L^{\infty}(0,T;L^{2})\cap L^{2}(0,T;W^{1,\frac{6}{p-1}}), where f(𝐱,t)f(\mathbf{x},t), with t(0,T)t\in(0,T), 𝐱Ω\mathbf{x}\in\Omega, may be a scalar, a vector or a tensor. Then, there exists a positive constant CC depending on Ω\Omega such that

0TfL2+h2(2+h)(6p)3hC0TfL22[(6p)(2+h)3h]3hfW1,6p12,{h0if 1p3,h[0,2(6p)p3]if 3<p<6.\int_{0}^{T}||f||_{L^{2+h}}^{\frac{2(2+h)(6-p)}{3h}}\leq C\int_{0}^{T}||f||_{L^{2}}^{\frac{2[(6-p)(2+h)-3h]}{3h}}||f||_{W^{1,\frac{6}{p-1}}}^{2},\;\begin{cases}h\geq 0\;\text{if}\;1\leq p\leq 3,\\ h\in\left[0,\frac{2(6-p)}{p-3}\right]\;\text{if}\;3<p<6.\end{cases} (15)

Let moreover p[1,6)p\in[1,6) and fL(0,T;L6)L182p3(0,T;W1,182p3)f\in L^{\infty}(0,T;L^{6})\cap L^{\frac{18-2p}{3}}(0,T;W^{1,\frac{18-2p}{3}}), where f(𝐱,t)f(\mathbf{x},t), with t(0,T)t\in(0,T), 𝐱Ω\mathbf{x}\in\Omega, may be a scalar, a vector or a tensor. Then, there exists a positive constant CC depending on Ω\Omega such that

0TfL6+h2(6+h)(6p)hC0TfL62[(183p)(6+h)h(9p)]3hfW1,182p3182p3,{h0if 1p92,h[0,18(6p)2p9]if92<p<6.\int_{0}^{T}||f||_{L^{6+h}}^{\frac{2(6+h)(6-p)}{h}}\leq C\int_{0}^{T}||f||_{L^{6}}^{\frac{2[(18-3p)(6+h)-h(9-p)]}{3h}}||f||_{W^{1,\frac{18-2p}{3}}}^{\frac{18-2p}{3}},\;\begin{cases}h\geq 0\;\text{if}\;1\leq p\leq\frac{9}{2},\\ h\in\left[0,\frac{18(6-p)}{2p-9}\right]\;\text{if}\;\frac{9}{2}<p<6.\end{cases} (16)

Let finally fL(0,T;L6)Ls(0,T;W1,6)f\in L^{\infty}(0,T;L^{6})\cap L^{s}(0,T;W^{1,6}), with s1s\geq 1, where f(𝐱,t)f(\mathbf{x},t), with t(0,T)t\in(0,T), 𝐱Ω\mathbf{x}\in\Omega, may be a scalar, a vector or a tensor. Then, there exists a positive constant CC depending on Ω\Omega such that

0TfL6+h2s(6+h)hC0TfL6(12+h)shfW1,6s,h0.\int_{0}^{T}||f||_{L^{6+h}}^{\frac{2s(6+h)}{h}}\leq C\int_{0}^{T}||f||_{L^{6}}^{\frac{(12+h)s}{h}}||f||_{W^{1,6}}^{s},\;h\geq 0. (17)

We observe that (15), (16) and (17) are consequences of the Gagliardo–Nirenberg inequality (14) with j=0j=0, m=1m=1, k=2+hk=2+h, r=6p1r=\frac{6}{p-1}, q=2q=2 (for (15)), j=0j=0, m=1m=1, k=6+hk=6+h, r=182p3r=\frac{18-2p}{3}, q=6q=6 (for (16)) and j=0j=0, m=1m=1, k=6+hk=6+h, r=6r=6, q=6q=6 (for (17)). We moreover recall the following interpolation inequality.

Lemma 3.3

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded domain and fLqf\in L^{q}, q1q\geq 1, where ff can be a function with scalar, vectorial or tensorial values. Let also srqs\leq r\leq q. Then, there exists a positive constant CC depending on Ω\Omega such that

Ω|f|r(Ω|f|s)qrqs(Ω|f|q)rsqs.\int_{\Omega}|f|^{r}\leq\left(\int_{\Omega}|f|^{s}\right)^{\frac{q-r}{q-s}}\left(\int_{\Omega}|f|^{q}\right)^{\frac{r-s}{q-s}}. (18)

We finally state the Agmon type inequality in three space dimensions (see e.g. [1]) which will be used in the following calculations.

Lemma 3.4

Let Ω3\Omega\subset\mathbb{R}^{3}, be a bounded domain with Lipschitz boundary and fH2(Ω)f\in H^{2}(\Omega). Then, there exists a positive constant CC depending on Ω\Omega such that

fL(Ω)CfH1(Ω)12f|H2(Ω)12.||f||_{L^{\infty}(\Omega)}\leq C||f||_{H^{1}(\Omega)}^{\frac{1}{2}}||f|||_{H^{2}(\Omega)}^{\frac{1}{2}}. (19)

4 Existence result of a global weak solution

We first need to obtain an existence result for a truncated system.

Let us introduce the smooth step function gC1(+)g\in C^{1}(\mathbb{R}^{+}) with the properties

{0g()1;g(r)1forr1;g(r)0forr2;|g(r)|Cgr0.\displaystyle\begin{cases}0\leq g(\cdot)\leq 1;\\ g(r)\equiv 1\;\text{for}\;r\leq 1;\quad g(r)\equiv 0\;\text{for}\;r\geq 2;\\ |g^{\prime}(r)|\leq C_{g}\;\forall r\geq 0.\end{cases} (20)

Given R>1R>1, we define the smooth truncation function

gR(r):=g(rR)+(1g(rR))rmin(0,4p),g_{R}(r):=g\left(\frac{r}{R}\right)+\left(1-g\left(\frac{r}{R}\right)\right)r^{\min(0,4-p)}, (21)

where pp is defined in Assumption A2. We have that

gR(r)=1Rg(rR)(1rmin(0,4p))+min(0,4p)(1g(rR))rmin(1,3p).g^{\prime}_{R}(r)=\frac{1}{R}g^{\prime}\left(\frac{r}{R}\right)\left(1-r^{\min(0,4-p)}\right)+\min(0,4-p)\left(1-g\left(\frac{r}{R}\right)\right)r^{\min(-1,3-p)}. (22)

We observe that

0gR(r){1forrR;1+1Rmax(0,p4)forRr2R;1rmax(0,p4)1(2R)max(0,p4)forr2R,0\leq g_{R}(r)\leq\begin{cases}1\quad\text{for}\;r\leq R;\\ 1+\frac{1}{R^{\max(0,p-4)}}\quad\text{for}\;R\leq r\leq 2R;\\ \frac{1}{r^{\max(0,p-4)}}\leq\frac{1}{(2R)^{\max(0,p-4)}}\quad\text{for}\,r\geq 2R,\end{cases} (23)

and

|gR(r)|{0forrR,2Cg2R(11(2R)max(0,p4))+2max(1,p3)min(0,4p)(2R)max(1,p3)forRr2R,0forr2R.|g^{\prime}_{R}(r)|\leq\begin{cases}0\quad\text{for}\;r\leq R,\\ \frac{2C_{g}}{2R}\left(1-\frac{1}{(2R)^{\max(0,p-4)}}\right)+\frac{2^{\max(1,p-3)}\min(0,4-p)}{(2R)^{\max(1,p-3)}}\quad\text{for}\;R\leq r\leq 2R,\\ 0\quad\text{for}\,r\geq 2R.\end{cases} (24)

We introduce the truncated elastic energy density

wR(ϕ,𝐅):=gR(|𝐅|)w(ϕ,𝐅).w_{R}(\phi,\mathbf{F}):=g_{R}(|\mathbf{F}|)w(\phi,\mathbf{F}). (25)

Thanks to (23) we have that

d1wR(ϕ,𝐅){C(1+|2R|p)for|𝐅|2R;C|𝐅|min(0,4p)(1+|𝐅|p)C1(2R)max(0,p4)+C|𝐅|p+min(0,4p)C(1+|𝐅|p+min(0,4p))for|𝐅|2R.-d_{1}\leq w_{R}(\phi,\mathbf{F})\leq\begin{cases}C(1+|2R|^{p})\quad\text{for}\;|\mathbf{F}|\leq 2R;\\ \\ C|\mathbf{F}|^{\min(0,4-p)}\left(1+|\mathbf{F}|^{p}\right)\leq C\frac{1}{(2R)^{\max(0,p-4)}}+C|\mathbf{F}|^{p+\min(0,4-p)}\\ \leq C(1+|\mathbf{F}|^{p+\min(0,4-p)})\quad\text{for}\,|\mathbf{F}|\geq 2R.\end{cases} (26)

Given (24) and the relation

𝐅wR(ϕ,𝐅)=gR(|𝐅|)𝐅|𝐅|w(ϕ,𝐅)+gR(|𝐅|)𝐅w(ϕ,𝐅),\partial_{\mathbf{F}}w_{R}(\phi,\mathbf{F})=g^{\prime}_{R}(|\mathbf{F}|)\frac{\mathbf{F}}{|\mathbf{F}|}w(\phi,\mathbf{F})+g_{R}(|\mathbf{F}|)\partial_{\mathbf{F}}w(\phi,\mathbf{F}), (27)

we have that

|𝐅wR(ϕ,𝐅)|{C(1+|2R|p1)for|𝐅|2R;C|𝐅|min(0,4p)(1+|𝐅|p1)C1(2R)max(0,p4)+C|𝐅|p1+min(0,4p)C(1+|𝐅|p+min(1,3p))for|𝐅|2R.|\partial_{\mathbf{F}}w_{R}(\phi,\mathbf{F})|\leq\begin{cases}C(1+|2R|^{p-1})\quad\text{for}\;|\mathbf{F}|\leq 2R;\\ \\ C|\mathbf{F}|^{\min(0,4-p)}\left(1+|\mathbf{F}|^{p-1}\right)\leq C\frac{1}{(2R)^{\max(0,p-4)}}+C|\mathbf{F}|^{p-1+\min(0,4-p)}\\ \leq C(1+|\mathbf{F}|^{p+\min(-1,3-p)})\quad\text{for}\,|\mathbf{F}|\geq 2R.\end{cases} (28)

Thanks to (20), (26), (28) and the fact that ϕwR(ϕ,𝐅)=gR(|𝐅|)ϕw(ϕ,𝐅)\partial_{\phi}w_{R}(\phi,\mathbf{F})=g_{R}(|\mathbf{F}|)\partial_{\phi}w(\phi,\mathbf{F}), as a consequence of Assumption A2 we obtain the following property for wRw_{R}:

  • A2Bis

    wRC1(×3×3;)w_{R}\in C^{1}(\mathbb{R}\times\mathbb{R}^{3\times 3};\mathbb{R}), and there exist d10,d2>0d_{1}\geq 0,d_{2}>0 such that d1wR(r,𝐓)d2(1+|𝐓|p)-d_{1}\leq w_{R}(r,\mathbf{T})\leq d_{2}\left(1+|\mathbf{T}|^{p}\right) for all r,𝐓3×3r\in\mathbb{R},\mathbf{T}\in\mathbb{R}^{3\times 3}, with p[0,4]p\in[0,4]. Moreover, there exist d3,d4>0d_{3},d_{4}>0 such that |𝐓w(r,𝐓)|d3(1+|𝐓|p1)\left|\partial_{\mathbf{T}}w(r,\mathbf{T})\right|\leq d_{3}\left(1+|\mathbf{T}|^{p-1}\right) and |rw(r,𝐓)|d4(1+|𝐓|p)\left|\partial_{r}w(r,\mathbf{T})\right|\leq d_{4}\left(1+|\mathbf{T}|^{p}\right) for all r,𝐓3×3r\in\mathbb{R},\mathbf{T}\in\mathbb{R}^{3\times 3}.

Finally, we define the truncated system, depending on the finite parameter RR,

{νΔ𝐯R+qR=μRϕR+div(𝐌R𝐅RT)+𝐅R𝐌R,div𝐯R=0,𝐅Rt+(𝐯R)𝐅R(𝐯R)𝐅R+γ𝐌R=0,𝐌R=𝐅RwR(ϕR,𝐅R)λΔ𝐅R,ϕRt+𝐯RϕRdiv(b(ϕR)μR)=0,μR=ψ(ϕR)ΔϕR+ϕRwR(ϕR,𝐅R),\begin{cases}-\nu\Delta\mathbf{v}_{R}+\nabla q_{R}=\mu_{R}\nabla\phi_{R}+\operatorname{div}\left(\mathbf{M}_{R}\mathbf{F}_{R}^{T}\right)+\nabla\mathbf{F}_{R}\odot\mathbf{M}_{R},\\ \\ \operatorname{div}\mathbf{v}_{R}=0,\\ \\ \frac{\partial\mathbf{F}_{R}}{\partial t}+\left(\mathbf{v}_{R}\cdot\nabla\right)\mathbf{F}_{R}-(\nabla\mathbf{v}_{R})\mathbf{F}_{R}+\gamma\mathbf{M}_{R}=0,\\ \\ \mathbf{M}_{R}=\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})-\lambda\Delta\mathbf{F}_{R},\\ \\ \frac{\partial\phi_{R}}{\partial t}+\mathbf{v}_{R}\cdot\nabla\phi_{R}-\operatorname{div}(b(\phi_{R})\nabla\mu_{R})=0,\\ \\ \mu_{R}=\psi^{\prime}(\phi_{R})-\Delta\phi_{R}+\partial_{\phi_{R}}w_{R}(\phi_{R},\mathbf{F}_{R}),\end{cases} (29)

valid in ΩT\Omega_{T}, endowed with the same boundary conditions as (10). In the following, for ease of notation we will avoid to report the RR index for the solutions of system (29).

4.1 Faedo–Galerkin approximation scheme

We define the finite dimensional spaces which will be used to formulate the Galerkin ansatz to approximate the solutions of system (29). Let {𝜼𝒊}i\{\boldsymbol{\eta_{i}}\}_{i\in\mathbb{N}} be the eigenfunctions of the Stokes operator with homogeneous Dirichlet boundary conditions, i.e.

PL(Δ)𝜼𝒊=βi𝜼𝒊inΩ,𝜼𝒊=𝟎onΩ,P_{L}(-\Delta)\boldsymbol{\eta_{i}}=\beta_{i}\boldsymbol{\eta_{i}}\quad\text{in}\;\Omega,\quad\boldsymbol{\eta_{i}}=\mathbf{0}\quad\text{on}\;\partial\Omega,

where PL:L2(Ω;3)Ldiv2(Ω;3)P_{L}:L^{2}(\Omega;\mathbb{R}^{3})\to L_{\text{div}}^{2}(\Omega;\mathbb{R}^{3}) is the Leray projection operator, with 0<β0β1βm0<\beta_{0}\leq\beta_{1}\leq\dots\leq\beta_{m}\to\infty. The sequence {𝜼𝒊}i\{\boldsymbol{\eta_{i}}\}_{i\in\mathbb{N}} can be chosen as an orthonormal basis in Ldiv2(Ω;3)L_{\operatorname{div}}^{2}(\Omega;\mathbb{R}^{3}) and an orthogonal basis in H0,div1(Ω;3)H_{0,\operatorname{div}}^{1}(\Omega;\mathbb{R}^{3}), and, thanks to Assumption A0, we have that {𝜼𝒊}iH2(Ω;3)\{\boldsymbol{\eta_{i}}\}_{i\in\mathbb{N}}\subset H^{2}(\Omega;\mathbb{R}^{3}). We introduce the projection operator

PmS:H0,div1(Ω;3)span{𝜼𝟎,𝜼𝟏,,𝜼𝒎}.P_{m}^{S}:H_{0,\operatorname{div}}^{1}(\Omega;\mathbb{R}^{3})\to\text{span}\{\boldsymbol{\eta_{0}},\boldsymbol{\eta_{1}},\dots,\boldsymbol{\eta_{m}}\}.

Let {ξi}i\{\xi_{i}\}_{i\in\mathbb{N}} be the eigenfunctions of the Laplace operator with homogeneous Neumann boundary conditions, i.e.

Δξi=αiξiinΩ,ξi𝐧=𝟎onΩ,-\Delta\xi_{i}=\alpha_{i}\xi_{i}\quad\text{in}\;\Omega,\quad\nabla\xi_{i}\cdot\mathbf{n}=\mathbf{0}\quad\text{on}\;\partial\Omega,

with 0=α0<α1αm0=\alpha_{0}<\alpha_{1}\leq\dots\leq\alpha_{m}\to\infty. The sequence {ξi}i\{\xi_{i}\}_{i\in\mathbb{N}} can be chosen as an orthonormal basis in L2(Ω)L^{2}(\Omega) and an orthogonal basis in H1(Ω)H^{1}(\Omega), and, thanks to Assumption A0, {ξi}iH2(Ω)\{\xi_{i}\}_{i\in\mathbb{N}}\subset H^{2}(\Omega). Without loss of generality, we assume α0=0\alpha_{0}=0. We introduce the projection operator

PmL:H1(Ω)span{ξ0,ξ1,,ξm}.P_{m}^{L}:H^{1}(\Omega)\to\text{span}\{\xi_{0},\xi_{1},\dots,\xi_{m}\}.

Let moreover {𝚺i}i\{\boldsymbol{\Sigma}_{i}\}_{i\in\mathbb{N}} be the tensor eigenfunctions of the Laplace operator with homogeneous Neumann boundary conditions, i.e.

Δ𝚺i=γi𝚺iinΩ,𝚺i𝐧=𝟎onΩ,-\Delta\boldsymbol{\Sigma}_{i}=\gamma_{i}\boldsymbol{\Sigma}_{i}\quad\text{in}\;\Omega,\quad\nabla\boldsymbol{\Sigma}_{i}\mathbf{n}=\mathbf{0}\quad\text{on}\;\partial\Omega,

with 0=γ0=γ1==γ8<γ9γm0=\gamma_{0}=\gamma_{1}=\dots=\gamma_{8}<\gamma_{9}\leq\dots\leq\gamma_{m}\to\infty. Note that the eigenvalues γ0==γ8=0\gamma_{0}=\dots=\gamma_{8}=0 are associated to the eigentensors 𝚺0,,𝚺8\boldsymbol{\Sigma}_{0},\dots,\boldsymbol{\Sigma}_{8} which are proportional to the tensors 𝐞ij\mathbf{e}_{ij}, i,j=1,2,3i,j=1,2,3. The sequence {𝚺i}i\{\boldsymbol{\Sigma}_{i}\}_{i\in\mathbb{N}} can be chosen as an orthonormal basis in L2(Ω;3×3))L^{2}(\Omega;\mathbb{R}^{3\times 3})) and an orthogonal basis in H1(Ω;3×3))H^{1}(\Omega;\mathbb{R}^{3\times 3})), and, thanks to Assumption A0, {𝚺i}iH2(Ω;3×3)\{\boldsymbol{\Sigma}_{i}\}_{i\in\mathbb{N}}\subset H^{2}(\Omega;\mathbb{R}^{3\times 3}). We introduce the projection operator

PmL,Σ:H1(Ω;3×3)span{𝚺0,𝚺1,,𝚺m}.P_{m}^{L,\Sigma}:H^{1}(\Omega;\mathbb{R}^{3\times 3})\to\text{span}\{\boldsymbol{\Sigma}_{0},\boldsymbol{\Sigma}_{1},\dots,\boldsymbol{\Sigma}_{m}\}.

We make the Galerkin ansatz 𝐯m=idim(t)𝜼𝒊(𝐱)\mathbf{v}_{m}=\sum_{i}d_{i}^{m}(t)\boldsymbol{\eta_{i}}(\mathbf{x}), 𝐅m=ifim(t)𝚺𝒊(𝐱)\mathbf{F}_{m}=\sum_{i}f_{i}^{m}(t)\boldsymbol{\Sigma_{i}}(\mathbf{x}), 𝐌m=igim(t)𝚺𝒊(𝐱)\mathbf{M}_{m}=\sum_{i}g_{i}^{m}(t)\boldsymbol{\Sigma_{i}}(\mathbf{x}), ϕm=iaim(t)ξi(𝐱)\phi_{m}=\sum_{i}a_{i}^{m}(t)\xi_{i}(\mathbf{x}), μm=icim(t)ξi(𝐱)\mu_{m}=\sum_{i}c_{i}^{m}(t)\xi_{i}(\mathbf{x}) to approximate the solutions 𝐯,𝐅,𝐌,ϕ,μ\mathbf{v},\mathbf{F},\mathbf{M},\phi,\mu of system (29), and project the equation for 𝐯m\mathbf{v}_{m} onto span{𝜼0,𝜼1,,𝜼m}\text{span}\left\{\boldsymbol{\eta}_{0},\boldsymbol{\eta}_{1},\dots,\boldsymbol{\eta}_{m}\right\}, the equations for 𝐅m\mathbf{F}_{m} and 𝐌m\mathbf{M}_{m} onto span{𝚺𝟎,𝚺𝟏,,𝚺𝒎}\text{span}\{\boldsymbol{\Sigma_{0}},\boldsymbol{\Sigma_{1}},\dots,\boldsymbol{\Sigma_{m}}\} and the equations for ϕm\phi_{m} and μm\mu_{m} onto span{ξ0,ξ1,,ξm}\text{span}\left\{\xi_{0},\xi_{1},\dots,\xi_{m}\right\}, obtaining the following Galerkin approximation of system (29):

{νΩ𝐯m:𝜼i=Ωμmϕm𝜼iΩ(𝐌m𝐅mT):𝜼i+Ω(𝐅m𝐌m)𝜼i,Ωt𝐅m:𝚺i+Ω(𝐯m)𝐅m:𝚺iΩ(𝐯m)𝐅m:𝚺i+γΩ𝐌m:𝚺i=0,Ω𝐌m:𝚺i=Ω𝐅mwR(ϕm,𝐅m):𝚺i+λΩ𝐅m\setstackgapS0.4ex\Shortstack𝚺i,Ωtϕmξi+Ω(𝐯mϕm)ξi+Ωb(ϕm)μmξi=0,Ωμmξi=Ωϕmξi+Ωψ(ϕm)ξi+ΩϕmwR(ϕm,𝐅m)ξi,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{m}\colon\nabla\boldsymbol{\eta}_{i}=\int_{\Omega}\mu_{m}\nabla\phi_{m}\cdot\boldsymbol{\eta}_{i}-\int_{\Omega}\left(\mathbf{M}_{m}\mathbf{F}_{m}^{T}\right)\colon\nabla\boldsymbol{\eta}_{i}+\int_{\Omega}\left(\nabla\mathbf{F}_{m}\odot\mathbf{M}_{m}\right)\cdot\boldsymbol{\eta}_{i},\\ \displaystyle\int_{\Omega}\partial_{t}\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{i}+\int_{\Omega}\left({\mathbf{v}}_{m}\cdot\nabla\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{i}-\int_{\Omega}\left(\nabla{\mathbf{v}}_{m}\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{i}+\gamma\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Sigma}_{i}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Sigma}_{i}=\int_{\Omega}\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\colon\boldsymbol{\Sigma}_{i}+\lambda\int_{\Omega}\nabla\mathbf{F}_{m}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Sigma}_{i},\\ \displaystyle\int_{\Omega}\partial_{t}\phi_{m}\xi_{i}+\int_{\Omega}\left(\mathbf{v}_{m}\cdot\nabla\phi_{m}\right)\xi_{i}+\int_{\Omega}b\left(\phi_{m}\right)\nabla\mu_{m}\cdot\nabla\xi_{i}=0,\\ \displaystyle\int_{\Omega}\mu_{m}\xi_{i}=\int_{\Omega}\nabla\phi_{m}\cdot\nabla\xi_{i}+\int_{\Omega}\psi^{\prime}\left(\phi_{m}\right)\xi_{i}+\int_{\Omega}\partial_{\phi_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\xi_{i},\end{cases} (30)

in [0,t][0,t], with 0<tT0<t\leq T, for i=0,,mi=0,\dots,m and with initial conditions

ϕm(𝐱,0)=PmL(ϕ0),𝐅m(𝐱,0)=PmL,Σ(𝐅0).\phi_{m}(\mathbf{x},0)=P_{m}^{L}(\phi_{0}),\quad\mathbf{F}_{m}(\mathbf{x},0)=P_{m}^{L,\Sigma}(\mathbf{F}_{0}). (31)

System (30) defines a collection of initial value problems for a system of coupled ODEs

{νβidim=l,s[Ωξlξs𝜼i]blmasml,s[Ω(𝚺l𝚺sT):𝜼i]glmfsm+l,s[Ω(𝚺l𝚺s)𝜼i]flmgsm,ddtfim=l,s(Ω(𝜼l)𝚺s:𝚺i+Ω(𝜼l)𝚺s:𝚺i)dlmfsmγgim,gim=Ω𝐅mwR(sasmξs,lflm𝚺l):𝚺i+λγifim,ddtaim=l,s[Ω(𝜼lξs)ξi]dlmasml[Ωb(sasmξs)ξlξi]clm,cim=αiaim+Ωψ(sasmξs)ξi+ΩϕmwR(sasmξs,lflm𝚺l)ξi,aim(0)=(ϕ0,ξi),fim(0)=(𝐅0,𝚺𝒊),i=0,,m.\begin{cases}\displaystyle\nu\beta_{i}d_{i}^{m}&=\sum_{l,s}\left[\int_{\Omega}\xi_{l}\nabla\xi_{s}\cdot\boldsymbol{\eta}_{i}\right]b_{l}^{m}a_{s}^{m}-\sum_{l,s}\left[\int_{\Omega}\left(\boldsymbol{\Sigma}_{l}\boldsymbol{\Sigma}_{s}^{T}\right)\colon\nabla\boldsymbol{\eta}_{i}\right]g_{l}^{m}f_{s}^{m}\\ &+\sum_{l,s}\left[\int_{\Omega}\left(\nabla\boldsymbol{\Sigma}_{l}\odot\boldsymbol{\Sigma}_{s}\right)\cdot\boldsymbol{\eta}_{i}\right]f_{l}^{m}g_{s}^{m},\\ \displaystyle\frac{d}{dt}f_{i}^{m}&=\sum_{l,s}\biggl{(}-\int_{\Omega}\left(\boldsymbol{\eta}_{l}\cdot\nabla\right)\boldsymbol{\Sigma}_{s}\colon\boldsymbol{\Sigma}_{i}+\int_{\Omega}\left(\nabla\boldsymbol{\eta}_{l}\right)\boldsymbol{\Sigma}_{s}\colon\boldsymbol{\Sigma}_{i}\biggr{)}d_{l}^{m}f_{s}^{m}-\gamma g_{i}^{m},\\ \displaystyle g_{i}^{m}&=\int_{\Omega}\partial_{\mathbf{F}_{m}}w_{R}\left(\sum_{s}a_{s}^{m}\xi_{s},\sum_{l}f_{l}^{m}\boldsymbol{\Sigma}_{l}\right)\colon\boldsymbol{\Sigma}_{i}+\lambda\gamma_{i}f_{i}^{m},\\ \displaystyle\frac{d}{dt}a_{i}^{m}&=-\sum_{l,s}\left[\int_{\Omega}\left(\boldsymbol{\eta}_{l}\cdot\nabla\xi_{s}\right)\cdot\xi_{i}\right]d_{l}^{m}a_{s}^{m}-\sum_{l}\left[\int_{\Omega}b\left(\sum_{s}a_{s}^{m}\xi_{s}\right)\nabla\xi_{l}\cdot\nabla\xi_{i}\right]c_{l}^{m},\\ \displaystyle c_{i}^{m}&=\alpha_{i}a_{i}^{m}+\int_{\Omega}\psi^{\prime}\left(\sum_{s}a_{s}^{m}\xi_{s}\right)\xi_{i}+\int_{\Omega}\partial_{\phi_{m}}w_{R}\left(\sum_{s}a_{s}^{m}\xi_{s},\sum_{l}f_{l}^{m}\boldsymbol{\Sigma}_{l}\right)\xi_{i},\\ \displaystyle a_{i}^{m}(0)&=\left(\phi_{0},\xi_{i}\right),\;f_{i}^{m}(0)=\left(\mathbf{F}_{0},\boldsymbol{\Sigma_{i}}\right),\quad i=0,\dots,m.\end{cases} (32)

Due to the Assumptions A1, A2Bis, A3 on the regularity of the functions b,ψ,wRb,\psi,w_{R} and the regularity in space of the functions ξi,𝜼i,𝚺i,\xi_{i},\boldsymbol{\eta}_{i},\boldsymbol{\Sigma}_{i},, the right hand side of (32) depends continuously on the independent variables and we can apply the Peano existence theorem to infer that there exist a sufficiently small t1t_{1} with 0<t1T0<t_{1}\leq T and a local solution (dim,fim,gim,aim,cim)(d_{i}^{m},f_{i}^{m},g_{i}^{m},a_{i}^{m},c_{i}^{m}) of (32), for i=0,,mi=0,\dots,m.

4.2 A priori estimates

We now deduce a priori estimates, uniform in the discretization parameter mm, for the solutions of system (30), which can be rewritten, combining the equations over i=0,,mi=0,\dots,m, as

{νΩ𝐯m:𝜼=Ωμmϕm𝜼Ω(𝐌m𝐅mT):𝜼+Ω(𝐅m𝐌m)𝜼,Ωt𝐅m:𝚺+Ω(𝐯m)𝐅m:𝚺Ω(𝐯m)𝐅m:𝚺+γΩ𝐌m:𝚺=0,Ω𝐌m:𝚪=Ω𝐅mwR(ϕm,𝐅m):𝚪+λΩ𝐅m\setstackgapS0.4ex\Shortstack𝚪,Ωtϕmξ+Ω(𝐯mϕm)ξ+Ωb(ϕm)μmξ=0,Ωμmχ=Ωϕmχ+Ωψ(ϕm)χ+ΩϕmwR(ϕm,𝐅m)χ,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{m}\colon\nabla\boldsymbol{\eta}=\int_{\Omega}\mu_{m}\nabla\phi_{m}\cdot\boldsymbol{\eta}-\int_{\Omega}\left(\mathbf{M}_{m}\mathbf{F}_{m}^{T}\right)\colon\nabla\boldsymbol{\eta}+\int_{\Omega}\left(\nabla\mathbf{F}_{m}\odot\mathbf{M}_{m}\right)\cdot\boldsymbol{\eta},\\ \displaystyle\int_{\Omega}\partial_{t}\mathbf{F}_{m}\colon\boldsymbol{\Sigma}+\int_{\Omega}\left({\mathbf{v}}_{m}\cdot\nabla\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}-\int_{\Omega}\left(\nabla{\mathbf{v}}_{m}\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}+\gamma\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Sigma}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Gamma}=\int_{\Omega}\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\colon\boldsymbol{\Gamma}+\lambda\int_{\Omega}\nabla\mathbf{F}_{m}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Gamma},\\ \displaystyle\int_{\Omega}\partial_{t}\phi_{m}\xi+\int_{\Omega}\left(\mathbf{v}_{m}\cdot\nabla\phi_{m}\right)\xi+\int_{\Omega}b\left(\phi_{m}\right)\nabla\mu_{m}\cdot\nabla\xi=0,\\ \displaystyle\int_{\Omega}\mu_{m}\chi=\int_{\Omega}\nabla\phi_{m}\cdot\nabla\chi+\int_{\Omega}\psi^{\prime}\left(\phi_{m}\right)\chi+\int_{\Omega}\partial_{\phi_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\chi,\end{cases} (33)

for a.e. t[0,t1]t\in[0,t_{1}], with 𝜼span{𝜼0,𝜼1,,𝜼m}\boldsymbol{\eta}\in\text{span}\left\{\boldsymbol{\eta}_{0},\boldsymbol{\eta}_{1},\dots,\boldsymbol{\eta}_{m}\right\}, 𝚺,𝚪span{𝚺𝟎,𝚺𝟏,,𝚺𝒎}\boldsymbol{\Sigma},\boldsymbol{\Gamma}\in\text{span}\{\boldsymbol{\Sigma_{0}},\boldsymbol{\Sigma_{1}},\dots,\boldsymbol{\Sigma_{m}}\} and ξ,χspan{ξ0,ξ1,,ξm}\xi,\chi\in\text{span}\left\{\xi_{0},\xi_{1},\dots,\xi_{m}\right\}, with initial conditions defined in (31). We take 𝜼=𝐯m\boldsymbol{\eta}=\mathbf{v}_{m}, 𝚺=𝐌m\boldsymbol{\Sigma}=\mathbf{M}_{m}, 𝚪=t𝐅m\boldsymbol{\Gamma}=-\partial_{t}\mathbf{F}_{m}, ξ=μm\xi=\mu_{m}, χ=tϕm\chi=-\partial_{t}\phi_{m} in (33), sum all the equations and integrate in time between 0 and t[0,t1]t\in[0,t_{1}]. We get, for any t[0,t1]t\in[0,t_{1}],

ν0tΩ|𝐯m|2+γ0tΩ|𝐌m|2+0tΩb(ϕm)|μm|2\displaystyle\nu\int_{0}^{t}\int_{\Omega}\left|\nabla\mathbf{v}_{m}\right|^{2}+\gamma\int_{0}^{t}\int_{\Omega}\left|\mathbf{M}_{m}\right|^{2}+\int_{0}^{t}\int_{\Omega}b(\phi_{m})\left|\nabla\mu_{m}\right|^{2}
+Ω(12|ϕm(t)|2+ψ(ϕm(t))+wR(ϕm(t),𝐅m(t))+λ2|𝐅m(t)|2)\displaystyle+\int_{\Omega}\left(\frac{1}{2}|\nabla\phi_{m}(t)|^{2}+\psi(\phi_{m}(t))+w_{R}(\phi_{m}(t),\mathbf{F}_{m}(t))+\frac{\lambda}{2}|\nabla\mathbf{F}_{m}(t)|^{2}\right)
=Ω(12|ϕm(0)|2+ψ(ϕm(0))+wR(ϕm(0),𝐅m(0))+λ2|𝐅m(0)|2).\displaystyle=\int_{\Omega}\left(\frac{1}{2}|\nabla\phi_{m}(0)|^{2}+\psi(\phi_{m}(0))+w_{R}(\phi_{m}(0),\mathbf{F}_{m}(0))+\frac{\lambda}{2}|\nabla\mathbf{F}_{m}(0)|^{2}\right). (34)

Hence, (4.2) becomes

supt[0,t1](12|ϕm|2+λ2|𝐅m|2)+ν0t1Ω|𝐯m|2+γ0t1Ω|𝐌m|2\displaystyle\sup_{t\in[0,t_{1}]}\left(\frac{1}{2}|\nabla\phi_{m}|^{2}+\frac{\lambda}{2}|\nabla\mathbf{F}_{m}|^{2}\right)+\nu\int_{0}^{t_{1}}\int_{\Omega}\left|\nabla\mathbf{v}_{m}\right|^{2}+\gamma\int_{0}^{t_{1}}\int_{\Omega}\left|\mathbf{M}_{m}\right|^{2}
+0t1Ωb(ϕm)|μm|2C(𝐅0,ϕ0)+C,\displaystyle+\int_{0}^{t_{1}}\int_{\Omega}b(\phi_{m})\left|\nabla\mu_{m}\right|^{2}\leq C(\mathbf{F}_{0},\phi_{0})+C, (35)

where we used Assumptions A2Bis, A3 and A4. The constants in the right hand side of (4.2) depends only on the initial data and on the domain Ω\Omega and not on the discretization parameter mm and on the truncation parameter RR. Thanks to the a priori estimate (4.2), we may extend by continuity the local solution of system (33) to the interval [0,T][0,T].

From the Poincaré inequality and from (4.2), we have that

𝐯mis uniformly bounded inL2(0,T;H0,div1(Ω;3))L2(0,T;Ldiv6(Ω;3)).\mathbf{v}_{m}\;\text{is uniformly bounded in}\;L^{2}(0,T;H_{0,\operatorname{div}}^{1}(\Omega;\mathbb{R}^{3}))\hookrightarrow L^{2}(0,T;L_{\operatorname{div}}^{6}(\Omega;\mathbb{R}^{3})). (36)

Moreover, from (4.2) we have that

𝐌mis uniformly bounded inL2(0,T;L2(Ω;3×3)).\mathbf{M}_{m}\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (37)

We take 𝚺=𝐞l,r\boldsymbol{\Sigma}=\mathbf{e}_{l,r}, l,r=1,2,3l,r=1,2,3, which are proportional to the eigentensors associated to γ0,,γ8\gamma_{0},\dots,\gamma_{8}. Hence, in (33)2, integrating by parts in the second and third terms, we obtain that

tΩ𝐅m,lr+Ω𝐅m,lr𝐯m𝐧Ω𝐅m,lrdiv𝐯mΩ𝐯m,l𝐅m[r]𝐧\displaystyle\displaystyle\partial_{t}\int_{\Omega}\mathbf{F}_{m,lr}+\int_{\partial\Omega}\mathbf{F}_{m,lr}\mathbf{v}_{m}\cdot\mathbf{n}-\int_{\Omega}\mathbf{F}_{m,lr}\operatorname{div}\mathbf{v}_{m}-\int_{\partial\Omega}\mathbf{v}_{m,l}\mathbf{F}_{m}^{[r]}\cdot\mathbf{n}
+Ω𝐯m,ldiv𝐅m[r]+γΩ𝐌m,lr=𝟎.\displaystyle\displaystyle+\int_{\Omega}\mathbf{v}_{m,l}\operatorname{div}\mathbf{F}_{m}^{[r]}+\gamma\int_{\Omega}\mathbf{M}_{m,lr}=\mathbf{0}. (38)

Integrating in time the latter equations over the interval [0,T][0,T], using the facts that div𝐯m=0\operatorname{div}\mathbf{v}_{m}=0, 𝐯m=𝟎\mathbf{v}_{m}=\mathbf{0} on Ω\partial\Omega, the Cauchy-Schwarz and Young inequalities, (4.2) and Assumption A4, we obtain that

supt[0,T]|(Ω𝐅m,lr)(t)|C(𝐅0,ϕ0,T),l,r=1,2,3.\sup_{t\in[0,T]}\left|\left(\int_{\Omega}\mathbf{F}_{m,lr}\right)(t)\right|\leq C(\mathbf{F}_{0},\phi_{0},T),\quad\forall l,r=1,2,3. (39)

Hence, from (4.2), (39) and the Poincaré–Wirtinger inequality we deduce that and

𝐅mis uniformly bounded inL(0,T;H1(Ω;3×3))L(0,T;L6(Ω;3×3)).\mathbf{F}_{m}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3})). (40)

Next, taking ξ=1\xi=1 (which is a multiple of ξ0\xi_{0}) in (33)4, we obtain, using the facts that div𝐯m=0\operatorname{div}\mathbf{v}_{m}=0, 𝐯m=𝟎\mathbf{v}_{m}=\mathbf{0} on Ω\partial\Omega and integrating by parts in the second term, that

(tϕm,1)=0.\left(\partial_{t}\phi_{m},1\right)=0. (41)

Hence, from (4.2), (41) and the Poincaré–Wirtinger inequality we deduce that

ϕmis uniformly bounded inL(0,T;H1(Ω))L(0,T;L6(Ω)).\phi_{m}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega)). (42)

4.3 Higher order regularity for 𝐅m\mathbf{F}_{m} and ϕm\phi_{m}

We observe that, from (40) and A2Bis, 𝐅mL(0,T;L6(Ω;3×3))\mathbf{F}_{m}\in L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3})) implies that

𝐅mwR(ϕm,𝐅m)L(0,T;L2(Ω;3×3)).\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\in L^{\infty}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (43)

Taking 𝚪=Δ𝐅m\boldsymbol{\Gamma}=-\Delta\mathbf{F}_{m} in (33)3 we get, using the Cauchy–Schwarz and Young inequalities, that

λΔ𝐅mL2(Ω;3×3)2λ2Δ𝐅mL2(Ω;3×3)2+1λ𝐌mL2(Ω;3×3)2+1λ𝐅mwR(ϕm,𝐅m)L2(Ω;3×3)2,\lambda||\Delta\mathbf{F}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}\leq\frac{\lambda}{2}||\Delta\mathbf{F}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\frac{1}{\lambda}||\mathbf{M}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\frac{1}{\lambda}||\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2},

which, after integration in time over the interval [0,T][0,T], gives, thanks to (4.2), (43) and elliptic regularity theory, that

𝐅mL2(0,T;H2(Ω;3×3))C.||\mathbf{F}_{m}||_{L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))}\leq C. (44)

From (40), (44) and (15) (applied to 𝐅m\nabla\mathbf{F}_{m}) with p=4,h=43p=4,h=\frac{4}{3}, we get that

𝐅mis uniformly bounded in\displaystyle\mathbf{F}_{m}\;\text{is uniformly bounded in}
L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3))L103(0,T;W1,103(Ω;3×3)).\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\frac{10}{3}}(0,T;W^{1,\frac{10}{3}}(\Omega;\mathbb{R}^{3\times 3})). (45)

From (16) with p=4,h=2p=4,h=2 we then have that

𝐅mis uniformly bounded in\displaystyle\mathbf{F}_{m}\;\text{is uniformly bounded in}
L(0,T;L6(Ω;3×3))L103(0,T;W1,103(Ω;3×3))L16(0,T;L8(Ω;3×3)).\displaystyle L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{\frac{10}{3}}(0,T;W^{1,\frac{10}{3}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{16}(0,T;L^{8}(\Omega;\mathbb{R}^{3\times 3})). (46)

We also observe that, taking p=4,h=1p=4,h=1 in (15) (applied to 𝐅m\nabla\mathbf{F}_{m}), we have that

𝐅mL4(0,T;W1,3(Ω;3×3))C,\displaystyle||\mathbf{F}_{m}||_{L^{4}(0,T;W^{1,3}(\Omega;\mathbb{R}^{3\times 3}))}\leq C, (47)

uniformly in mm, and moreover, taking p=4,h=1+ϵp=4,h=1+\epsilon, with ϵ(0,3]\epsilon\in(0,3] in (15), we have that

𝐅mL48ϵ3(1+ϵ)(0,T;W1,3+ϵ(Ω;3×3))C,\displaystyle||\mathbf{F}_{m}||_{L^{4-\frac{8\epsilon}{3(1+\epsilon)}}(0,T;W^{1,3+\epsilon}(\Omega;\mathbb{R}^{3\times 3}))}\leq C, (48)

which implies, employing the Sobolev embedding theorem and defining k:=8ϵ3(1+ϵ)(0,2]k:=\frac{8\epsilon}{3(1+\epsilon)}\in(0,2], that

𝐅mL4k(0,T;C(Ω¯;3×3))C,\displaystyle||\mathbf{F}_{m}||_{L^{4-k}(0,T;C(\bar{\Omega};\mathbb{R}^{3\times 3}))}\leq C, (49)

uniformly in mm.

From Assumption A2Bis and (4.3), we have that

ϕmwR(ϕm,𝐅m)L4(0,T;L2(Ω)).\partial_{\mathbf{\phi}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\in L^{4}(0,T;L^{2}(\Omega)). (50)

Taking χ=Δϕm\chi=-\Delta\phi_{m} in (33)5, using the Cauchy–Schwarz and Young inequalities and Assumption A3, we get

Δϕm2+Ωψ+′′(ϕm)|ϕm|2μmL2(Ω;3)ϕmL2(Ω;3)Ωψ′′(ϕm)|ϕm|2\displaystyle\displaystyle||\Delta\phi_{m}||^{2}+\int_{\Omega}\psi_{+}^{\prime\prime}(\phi_{m})|\nabla\phi_{m}|^{2}\leq||\nabla\mu_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}-\int_{\Omega}\psi_{-}^{\prime\prime}(\phi_{m})|\nabla\phi_{m}|^{2}
+Ωϕmw(ϕm,𝐅m)ΔϕmμmL2(Ω;3)ϕmL2(Ω;3)+12ϕmL2(Ω;3)2\displaystyle\displaystyle+\int_{\Omega}\partial_{\mathbf{\phi}_{m}}w(\phi_{m},\mathbf{F}_{m})\Delta\phi_{m}\leq||\nabla\mu_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}+\frac{1}{2}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}
+CΩ(1+|ϕm|q)|ϕm|2+14Δϕm2+ϕmw(ϕm,𝐅m)2.\displaystyle\displaystyle+C\int_{\Omega}\left(1+|\phi_{m}|^{q}\right)|\nabla\phi_{m}|^{2}+\frac{1}{4}||\Delta\phi_{m}||^{2}+||\partial_{\mathbf{\phi}_{m}}w(\phi_{m},\mathbf{F}_{m})||^{2}. (51)

We use (19), elliptic regularity theory and the Young inequality, observing that 4q>1\frac{4}{q}>1 when q<4q<4, to write

Ω|ϕm|q|ϕm|2ϕmL(Ω)qϕmL2(Ω;3)2CϕmH1(Ω;3)4+q2(ϕmq2+Δϕmq2)\displaystyle\int_{\Omega}|\phi_{m}|^{q}|\nabla\phi_{m}|^{2}\leq||\phi_{m}||_{L^{\infty}(\Omega)}^{q}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}\leq C||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{\frac{4+q}{2}}\left(||\phi_{m}||^{\frac{q}{2}}+||\Delta\phi_{m}||^{\frac{q}{2}}\right)
CϕmH1(Ω;3)2+q+ϕmH1(Ω;3)2(q+4)4q+14Δϕm2.\displaystyle\leq C||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{2+q}+||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{\frac{2(q+4)}{4-q}}+\frac{1}{4}||\Delta\phi_{m}||^{2}.

Using this inequality in (4.3), together with the Young inequality, we obtain that

Δϕm2μmL2(Ω;3)ϕmL2(Ω;3)+CϕmH1(Ω;3)2+CϕmH1(Ω;3)2+q\displaystyle\displaystyle||\Delta\phi_{m}||^{2}\leq||\nabla\mu_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3})}+C||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{2}+C||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{2+q}
+ϕmH1(Ω;3)2(q+4)4q+ϕmw(ϕm,𝐅m)2+12Δϕm2.\displaystyle\displaystyle+||\phi_{m}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{\frac{2(q+4)}{4-q}}+||\partial_{\mathbf{\phi}_{m}}w(\phi_{m},\mathbf{F}_{m})||^{2}+\frac{1}{2}||\Delta\phi_{m}||^{2}. (52)

Taking the square of (4.3) and integrating in time over the interval (0,T)(0,T), using (4.2), Assumption A1 and (50), we infer that

ϕmis uniformly bounded inL(0,T;H1(Ω))L4(0,T;H2(Ω)).\displaystyle\phi_{m}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega))\cap L^{4}(0,T;H^{2}(\Omega)). (53)

Since L(0,T;H1(Ω))L4(0,T;H2(Ω))L(0,T;L6(Ω))L4(0,T;W1,6(Ω))L^{\infty}(0,T;H^{1}(\Omega))\cap L^{4}(0,T;H^{2}(\Omega))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega))\cap L^{4}(0,T;W^{1,6}(\Omega)), using (17) with s=h=4s=h=4 we get that

ϕmL20(0,T;L10(Ω))C.||\phi_{m}||_{L^{20}(0,T;L^{10}(\Omega))}\leq C. (54)

Taking moreover χ=1\chi=1 in (33)5, we obtain, using Assumptions A2Bis, A3 and (40), that

|(μm,1)|=|Ωψ(ϕm)+ΩϕmwR(ϕm,𝐅m)|\displaystyle\displaystyle\left|(\mu_{m},1)\right|=\left|\int_{\Omega}\psi^{\prime}\left(\phi_{m}\right)+\int_{\Omega}\partial_{\phi_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\right|
CϕmLl(Ω)l+C𝐅mLp(Ω;3×3)p+CCϕmLl(Ω)l+C(𝐅0,ϕ0)+C,\displaystyle\displaystyle\leq C||\phi_{m}||_{L^{l}(\Omega)}^{l}+C||\mathbf{F}_{m}||_{L^{p}(\Omega;\mathbb{R}^{3\times 3})}^{p}+C\leq C||\phi_{m}||_{L^{l}(\Omega)}^{l}+C(\mathbf{F}_{0},\phi_{0})+C, (55)

where p[0,4]p\in[0,4], l[0,10)l\in[0,10). Taking the square of (4.3) and integrating in time over the interval [0,T][0,T], using also (54), we obtain that |(μm,1)||(\mu_{m},1)| is bounded in L2(0,T)L^{2}(0,T) and consequently, by the Poincaré–Wirtinger inequality and (4.2),

μmis uniformly bounded inL2(0,T;H1(Ω))L2(0,T;L6(Ω)).\mu_{m}\;\text{is uniformly bounded in}\;L^{2}(0,T;H^{1}(\Omega))\hookrightarrow L^{2}(0,T;L^{6}(\Omega)). (56)

4.4 Dual estimates

We now deduce a priori estimates, uniform in mm, for the time derivatives of 𝐅m\mathbf{F}_{m} and ϕm\phi_{m} in (33). Multiplying (33)2 by a time function ζL2(4k)2k(0,T)\zeta\in L^{\frac{2(4-k)}{2-k}}(0,T), with k(0,2)k\in(0,2), choosing 𝚺=PmL,Σ𝚷\boldsymbol{\Sigma}=P_{m}^{L,\Sigma}\boldsymbol{\Pi}, with a generic 𝚷L2(Ω;3×3)\boldsymbol{\Pi}\in L^{2}(\Omega;\mathbb{R}^{3\times 3}), and integrating in time over the interval (0,T)(0,T), we get

0TΩt𝐅m:𝚷ζ0T𝐯mL6(Ω;3)𝐅mL3(Ω;3×3×3)PmL,Σ𝚷L2(Ω;3×3)|ζ|\displaystyle\int_{0}^{T}\int_{\Omega}\partial_{t}\mathbf{F}_{m}\colon\boldsymbol{\Pi}\zeta\leq\int_{0}^{T}||\mathbf{v}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3})}||\nabla\mathbf{F}_{m}||_{L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3})}||P_{m}^{L,\Sigma}\boldsymbol{\Pi}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}|\zeta|
+0T𝐯mL2(Ω;3×3)𝐅mL(Ω;3×3)PmL,Σ𝚷L2(Ω;3×3)|ζ|\displaystyle+\int_{0}^{T}||\nabla\mathbf{v}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}||\mathbf{F}_{m}||_{L^{\infty}(\Omega;\mathbb{R}^{3\times 3})}||P_{m}^{L,\Sigma}\boldsymbol{\Pi}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}|\zeta|
+γ0T𝐌mL2(Ω;3×3)𝚷L2(Ω;3×3)|ζ|\displaystyle+\gamma\int_{0}^{T}||\mathbf{M}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}||\boldsymbol{\Pi}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}|\zeta|
C(||𝐯m||L2(0,T;L6(Ω;3))||𝐅m||L4(0,T;L3(Ω;3×3×3))\displaystyle\leq C\biggl{(}||\mathbf{v}_{m}||_{L^{2}(0,T;L^{6}(\Omega;\mathbb{R}^{3}))}||\nabla\mathbf{F}_{m}||_{L^{4}(0,T;L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3}))}
+||𝐯m||L2(0,T;L2(Ω;3×3))||𝐅m||L4k(0,T;C(Ω¯;3×3))+||𝐌m||L2(0,T;L2(Ω;3×3)))\displaystyle+||\nabla\mathbf{v}_{m}||_{L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3}))}||\mathbf{F}_{m}||_{L^{4-k}(0,T;C(\bar{\Omega};\mathbb{R}^{3\times 3}))}+||\mathbf{M}_{m}||_{L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3}))}\biggr{)}
×||𝚷||L2(Ω;3×3)||ζ||L2(4k)2k(0,T)C||𝚷||L2(Ω;3×3)||ζ||L2(4k)2k(0,T),\displaystyle\times||\boldsymbol{\Pi}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}||\zeta||_{L^{\frac{2(4-k)}{2-k}}(0,T)}\leq C||\boldsymbol{\Pi}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}||\zeta||_{L^{\frac{2(4-k)}{2-k}}(0,T)}, (57)

where, in the last step, we used (4.2), (47) and (49). Hence, we deduce that

t𝐅mL43s(0,T;L2(Ω;3×3))C,||\partial_{t}\mathbf{F}_{m}||_{L^{\frac{4}{3}-s}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})\right)}\leq C, (58)

where we set s:=2k3(6k)(0,13)s:=\frac{2k}{3(6-k)}\in\left(0,\frac{1}{3}\right). Indeed, it’s easy to check that

2k2(4k)+343s=1\frac{2-k}{2(4-k)}+\frac{3}{4-3s}=1

is verified if s:=2k3(6k)s:=\frac{2k}{3(6-k)}.

Moreover, multiplying (33)4 by a time function ζL2(0,T)\zeta\in L^{2}(0,T), choosing ξ=PmL(π)\xi=P_{m}^{L}(\pi), with a generic πH1(Ω)\pi\in H^{1}(\Omega), and integrating in time over the interval (0,T)(0,T), we obtain

0TΩtϕmπζ0T𝐯mL6(Ω;3)ϕmL2(Ω;3×)PmL(π)L3(Ω)|ζ|\displaystyle\int_{0}^{T}\int_{\Omega}\partial_{t}\phi_{m}\pi\zeta\leq\int_{0}^{T}||\mathbf{v}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3})}||\nabla\phi_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times})}||P_{m}^{L}(\pi)||_{L^{3}(\Omega)}|\zeta|
+0Tb(ϕm)μmL2(Ω;2)PmL(π)L2(Ω;2)|ζ|\displaystyle+\int_{0}^{T}||b\left(\phi_{m}\right)\nabla\mu_{m}||_{L^{2}(\Omega;\mathbb{R}^{2})}||\nabla P_{m}^{L}(\pi)||_{L^{2}(\Omega;\mathbb{R}^{2})}|\zeta|
C(𝐯mL2(0,T;L6(Ω;3))ϕmL(0,T;L2(Ω))+b(ϕm)μmL2(0,T;L2(Ω;2)))\displaystyle\leq C\left(||\mathbf{v}_{m}||_{L^{2}(0,T;L^{6}(\Omega;\mathbb{R}^{3}))}||\nabla\phi_{m}||_{L^{\infty}(0,T;L^{2}(\Omega))}+||b\left(\phi_{m}\right)\nabla\mu_{m}||_{L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{2}))}\right)
×πH1(Ω)ζL2(0,T).\displaystyle\times||\pi||_{H^{1}(\Omega)}||\zeta||_{L^{2}(0,T)}. (59)

Hence, using (4.2) and Assumption A1 we have that

tϕmL2(0,T;(H1(Ω)))C.||\partial_{t}\phi_{m}||_{L^{2}\left(0,T;\left(H^{1}(\Omega)\right)^{\prime}\right)}\leq C. (60)

4.5 Passage to the limit as mm\to\infty

Collecting the results (36), (37), (40), (42), (4.3), (48), (53), (56), (58) and (60), which are uniform in mm, from the Banach–Alaoglu and the Aubin–Lions lemma, we finally obtain the convergence properties, up to subsequences of the solutions, which we still label by the index mm, as follows:

𝐯m𝐯inL2(0,T;H0,div1(Ω;3)),\displaystyle\mathbf{v}_{m}\rightharpoonup\mathbf{v}\quad\text{in}\quad L^{2}\left(0,T;H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right)\right), (61)
𝐌m𝐌inL2(0,T;L2(Ω;3×3))\displaystyle\mathbf{M}_{m}{\rightharpoonup}\mathbf{M}\quad\text{in}\quad L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})) (62)
𝐅m𝐅inL(0,T;H1(Ω;3×3)),\displaystyle\mathbf{F}_{m}\overset{\ast}{\rightharpoonup}\mathbf{F}\quad\text{in}\quad L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3})), (63)
𝐅m𝐅inL4k(0,T;W1,3+ϵ(Ω;3×3))L2(0,T;H2(Ω;3×3)),ϵ(0,3],\displaystyle\mathbf{F}_{m}{\rightharpoonup}\mathbf{F}\quad\text{in}\quad L^{4-k}(0,T;W^{1,3+\epsilon}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3})),\;\;\epsilon\in\left(0,3\right], (64)
t𝐅mt𝐅inL43s(0,T;L2(Ω;3×3)),s(0,13),\displaystyle\partial_{t}\mathbf{F}_{m}\rightharpoonup\partial_{t}\mathbf{F}\quad\text{in}\quad L^{\frac{4}{3}-s}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})\right),\;\;s\in\left(0,\frac{1}{3}\right), (65)
ϕmϕinL(0,T;H1(Ω))L4(0,T;H2(Ω)),\displaystyle\phi_{m}\overset{\ast}{\rightharpoonup}\phi\quad\text{in}\quad L^{\infty}(0,T;H^{1}(\Omega))\cap L^{4}(0,T;H^{2}(\Omega)), (66)
μmμinL2(0,T;H1(Ω)),\displaystyle\mu_{m}\rightharpoonup\mu\quad\text{in}\quad L^{2}\left(0,T;H^{1}\left(\Omega\right)\right), (67)
tϕmtϕinL2(0,T;(H1(Ω))),\displaystyle\partial_{t}\phi_{m}\rightharpoonup\partial_{t}\phi\quad\text{in}\quad L^{2}(0,T;\left(H^{1}(\Omega)\right)^{\prime}), (68)
𝐅m𝐅inC0(0,T;Lp(Ω;3×3))L2(0,T;W1,p(Ω;3×3)),p[1,6),anda.e. inΩT,\displaystyle\mathbf{F}_{m}\to\mathbf{F}\quad\text{in}\quad C^{0}(0,T;L^{p}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;W^{1,p}(\Omega;\mathbb{R}^{3\times 3})),\;\;p\in[1,6),\;\;\text{and}\;\;\text{a.e. in}\;\;\Omega_{T}, (69)
𝐅m𝐅inL4k(0,T;C(Ω¯;3×3)),\displaystyle\mathbf{F}_{m}\to\mathbf{F}\quad\text{in}\quad L^{4-k}(0,T;C(\bar{\Omega};\mathbb{R}^{3\times 3})), (70)
ϕmϕinC0(0,T;Lp(Ω))L4(0,T;W1,p(Ω)),p[1,6),anda.e. inΩT,\displaystyle\phi_{m}\to\phi\quad\text{in}\quad C^{0}(0,T;L^{p}(\Omega))\cap L^{4}(0,T;W^{1,p}(\Omega)),\;\;p\in[1,6),\;\;\text{and}\;\;\text{a.e. in}\;\;\Omega_{T}, (71)

with k(0,2]k\in(0,2], as mm\to\infty. We note that (70) follows from the compact embedding W1,3+ϵC0W^{1,3+\epsilon}\subset C^{0} and (64), (65). With the convergence results (61)–(71), we can pass to the limit in the system (33) as mm\to\infty. Let’s take 𝜼=𝜼m=PmS(𝐮)\boldsymbol{\eta}=\boldsymbol{\eta}_{m}=P_{m}^{S}(\mathbf{u}), for arbitrary 𝐮H0,div1(Ω;3)\mathbf{u}\in H_{0,\operatorname{div}}^{1}(\Omega;\mathbb{R}^{3}), 𝚺=𝚺m=PmL,Σ(𝚯)\boldsymbol{\Sigma}=\boldsymbol{\Sigma}_{m}=P_{m}^{L,\Sigma}(\boldsymbol{\Theta}), for arbitrary 𝚯L2(Ω;3×3)\boldsymbol{\Theta}\in L^{2}(\Omega;\mathbb{R}^{3\times 3}), 𝚪=𝚪m=PmL,Σ(𝚷)\boldsymbol{\Gamma}=\boldsymbol{\Gamma}_{m}=P_{m}^{L,\Sigma}(\boldsymbol{\Pi}), for arbitrary 𝚷H1(Ω;3×3)\boldsymbol{\Pi}\in H^{1}(\Omega;\mathbb{R}^{3\times 3}), ξ=ξm=PmL(q)\xi=\xi_{m}=P_{m}^{L}(q), for arbitrary qH1(Ω)q\in H^{1}(\Omega), χ=χm=PmL(r)\chi=\chi_{m}=P_{m}^{L}(r), for arbitrary rH1(Ω)r\in H^{1}(\Omega), multiply the equations by a function ωC0([0,T])\omega\in C_{0}^{\infty}([0,T]) and integrate over the time interval [0,T][0,T]. This gives

{ν0TωΩ𝐯m:𝜼m=0TωΩμmϕm𝜼m0TωΩ(𝐌m𝐅mT):𝜼m+0TωΩ(𝐅m𝐌m)𝜼m,0TωΩt𝐅m:𝚺m+0TωΩ(𝐯m)𝐅m:𝚺m0TωΩ(𝐯m)𝐅m:𝚺m+γ0TωΩ𝐌m:𝚺m=0,0TωΩ𝐌m:𝚪m=0TωΩ𝐅mwR(ϕm,𝐅m):𝚪m+λ0TωΩ𝐅m\setstackgapS0.4ex\Shortstack𝚪m,0TωΩtϕmξm+0TωΩ(𝐯mϕm)ξm+0TωΩb(ϕm)μmξm=0,0TωΩμmχm=0TωΩϕmχm+0TωΩψ(ϕm)χm+0TωΩϕmwR(ϕm,𝐅m)χm.\begin{cases}\displaystyle\nu\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{v}_{m}\colon\nabla\boldsymbol{\eta}_{m}=\int_{0}^{T}\omega\int_{\Omega}\mu_{m}\nabla\phi_{m}\cdot\boldsymbol{\eta}_{m}-\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{M}_{m}\mathbf{F}_{m}^{T}\right)\colon\nabla\boldsymbol{\eta}_{m}\\ \displaystyle+\int_{0}^{T}\omega\int_{\Omega}\left(\nabla\mathbf{F}_{m}\odot\mathbf{M}_{m}\right)\cdot\boldsymbol{\eta}_{m},\\ \\ \displaystyle\int_{0}^{T}\omega\int_{\Omega}\partial_{t}\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}+\int_{0}^{T}\omega\int_{\Omega}\left({\mathbf{v}}_{m}\cdot\nabla\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}-\int_{0}^{T}\omega\int_{\Omega}\left(\nabla{\mathbf{v}}_{m}\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}\\ \displaystyle+\gamma\int_{0}^{T}\omega\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Sigma}_{m}=0,\\ \\ \displaystyle\int_{0}^{T}\omega\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Gamma}_{m}=\int_{0}^{T}\omega\int_{\Omega}\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\colon\boldsymbol{\Gamma}_{m}+\lambda\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{F}_{m}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Gamma}_{m},\\ \\ \displaystyle\int_{0}^{T}\omega\int_{\Omega}\partial_{t}\phi_{m}\xi_{m}+\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{v}_{m}\cdot\nabla\phi_{m}\right)\xi_{m}+\int_{0}^{T}\omega\int_{\Omega}b\left(\phi_{m}\right)\nabla\mu_{m}\cdot\nabla\xi_{m}=0,\\ \\ \displaystyle\int_{0}^{T}\omega\int_{\Omega}\mu_{m}\chi_{m}=\int_{0}^{T}\omega\int_{\Omega}\nabla\phi_{m}\cdot\nabla\chi_{m}+\int_{0}^{T}\omega\int_{\Omega}\psi^{\prime}\left(\phi_{m}\right)\chi_{m}\\ \displaystyle+\int_{0}^{T}\omega\int_{\Omega}\partial_{\phi_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\chi_{m}.\end{cases} (72)

We observe that

{𝜼m=PmS(𝐮)𝐮inH1(Ω;3),𝚺m=PmL,Σ(𝚯)𝚯inL2(Ω;3×3),𝚪m=PmL,Σ(𝚷)𝚷inH1(Ω;3×3),ξm=PmL(q)q,χm=PmL(r)rinH1(Ω).\displaystyle\begin{cases}\boldsymbol{\eta}_{m}=P_{m}^{S}(\mathbf{u})\rightarrow\mathbf{u}\;\;\text{in}\;\;H^{1}(\Omega;\mathbb{R}^{3}),\\ \boldsymbol{\Sigma}_{m}=P_{m}^{L,\Sigma}(\boldsymbol{\Theta})\rightarrow\boldsymbol{\Theta}\;\;\text{in}\;\;L^{2}(\Omega;\mathbb{R}^{3\times 3}),\\ \boldsymbol{\Gamma}_{m}=P_{m}^{L,\Sigma}(\boldsymbol{\Pi})\rightarrow\boldsymbol{\Pi}\;\;\text{in}\;\;H^{1}(\Omega;\mathbb{R}^{3\times 3}),\\ \xi_{m}=P_{m}^{L}(q)\rightarrow q,\chi_{m}=P_{m}^{L}(r)\rightarrow r\;\;\text{in}\;\;H^{1}(\Omega).\end{cases} (73)

Thanks to (61) and (73)1, we have

ν0TωΩ𝐯m:𝜼mν0TωΩ𝐯:𝐮,\nu\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{v}_{m}\colon\nabla\boldsymbol{\eta}_{m}\to\nu\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{v}\colon\nabla\mathbf{u}, (74)

as mm\to\infty. Owing to (71), (73)1 and the fact that 𝜼m\boldsymbol{\eta}_{m} converges in H1(Ω;3)L6(Ω;3)H^{1}(\Omega;\mathbb{R}^{3})\hookrightarrow L^{6}(\Omega;\mathbb{R}^{3}), we have that ωϕm𝜼mωϕ𝐮\omega\nabla\phi_{m}\cdot\boldsymbol{\eta}_{m}\to\omega\nabla\phi\cdot\mathbf{u} strongly in L2(0,T;L32(Ω))L^{2}(0,T;L^{\frac{3}{2}}(\Omega)). Indeed, it turns out that

0T|ω|2(Ω|ϕm𝜼mϕ𝐮|32)43=0T|ω|2(Ω|(ϕmϕ)𝜼m+ϕ(𝜼m𝐮)|32)43\displaystyle\int_{0}^{T}|\omega|^{2}\left(\int_{\Omega}|\nabla\phi_{m}\cdot\boldsymbol{\eta}_{m}-\nabla\phi\cdot\mathbf{u}|^{\frac{3}{2}}\right)^{\frac{4}{3}}=\int_{0}^{T}|\omega|^{2}\left(\int_{\Omega}|\nabla(\phi_{m}-\phi)\cdot\boldsymbol{\eta}_{m}+\nabla\phi\cdot(\boldsymbol{\eta}_{m}-\mathbf{u})|^{\frac{3}{2}}\right)^{\frac{4}{3}}
C0T|ω|2(ϕmϕ)L2(Ω;𝐑3)2𝜼mL6(Ω;𝐑3)2+C0T|ω|2ϕL2(Ω;𝐑3)2𝜼m𝐮L6(Ω;𝐑3)2\displaystyle\leq C\int_{0}^{T}|\omega|^{2}||\nabla(\phi_{m}-\phi)||_{L^{2}(\Omega;\mathbf{R}^{3})}^{2}||\boldsymbol{\eta}_{m}||_{L^{6}(\Omega;\mathbf{R}^{3})}^{2}+C\int_{0}^{T}|\omega|^{2}||\nabla\phi||_{L^{2}(\Omega;\mathbf{R}^{3})}^{2}||\boldsymbol{\eta}_{m}-\mathbf{u}||_{L^{6}(\Omega;\mathbf{R}^{3})}^{2}
CωL(0,T)2ϕmϕL2(0,T;L2(Ω;3))2𝜼mL6(Ω;3)2\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla\phi_{m}-\nabla\phi||_{L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3}))}^{2}||\boldsymbol{\eta}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3})}^{2}
+CωL(0,T)2ϕL2(0,T;L2(Ω;3))2𝜼m𝐮H1(Ω;3)20,\displaystyle+C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla\phi||_{L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3}))}^{2}||\boldsymbol{\eta}_{m}-\mathbf{u}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{2}\to 0, (75)

as mm\to\infty. Hence, using (67), by the product of weak–strong convergence we have that

0TωΩμmϕm𝜼m0TωΩμϕ𝐮,\int_{0}^{T}\omega\int_{\Omega}\mu_{m}\nabla\phi_{m}\cdot\boldsymbol{\eta}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\mu\nabla\phi\cdot\mathbf{u}, (76)

as mm\to\infty. For what concerns the second term on the right hand side of (72)1, thanks to (70) and (73)1 we have that ω𝜼m𝐅mω𝐮𝐅\omega\nabla\boldsymbol{\eta}_{m}\mathbf{F}_{m}\to\omega\nabla\mathbf{u}\mathbf{F} strongly in L2(0,T;L2(Ω;3×3))L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). Indeed, it holds that

0T|ω|2Ω|𝜼m𝐅m𝐮𝐅|2=0T|ω|2Ω|𝜼m(𝐅m𝐅)+(𝜼m𝐮)𝐅|2\displaystyle\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\nabla\boldsymbol{\eta}_{m}\mathbf{F}_{m}-\nabla\mathbf{u}\mathbf{F}|^{2}=\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\nabla\boldsymbol{\eta}_{m}(\mathbf{F}_{m}-\mathbf{F})+\nabla(\boldsymbol{\eta}_{m}-\mathbf{u})\mathbf{F}|^{2}
C0T|ω|2𝐅m𝐅L(Ω;𝐑3×3)2Ω|𝜼m|2+C0T|ω|2𝐅L(Ω;𝐑3×3)2Ω|(𝜼m𝐮)|2\displaystyle\leq C\int_{0}^{T}|\omega|^{2}||\mathbf{F}_{m}-\mathbf{F}||_{L^{\infty}(\Omega;\mathbf{R}^{3\times 3})}^{2}\int_{\Omega}|\nabla\boldsymbol{\eta}_{m}|^{2}+C\int_{0}^{T}|\omega|^{2}||\mathbf{F}||_{L^{\infty}(\Omega;\mathbf{R}^{3\times 3})}^{2}\int_{\Omega}|\nabla(\boldsymbol{\eta}_{m}-\mathbf{u})|^{2}
CωL(0,T)2𝐅m𝐅L2(0,T;L(Ω;3×3))2𝜼mL2(Ω;3×3)2\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}^{2}||\mathbf{F}_{m}-\mathbf{F}||_{L^{2}(0,T;L^{\infty}(\Omega;\mathbb{R}^{3\times 3}))}^{2}||\nabla\boldsymbol{\eta}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+CωL(0,T)2𝐅L2(0,T;L(Ω;3×3))2𝜼m𝐮H1(Ω;3×3)20,\displaystyle+C||\omega||_{L^{\infty}(0,T)}^{2}||\mathbf{F}||_{L^{2}(0,T;L^{\infty}(\Omega;\mathbb{R}^{3\times 3}))}^{2}||\boldsymbol{\eta}_{m}-\mathbf{u}||_{H^{1}(\Omega;\mathbb{R}^{3\times 3})}^{2}\to 0,

as mm\to\infty. Hence, using (62), by the product of weak–strong convergence we have that

0TωΩ(𝐌m𝐅mT):𝜼m0TωΩ(𝐌𝐅T):𝐮,\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{M}_{m}\mathbf{F}_{m}^{T}\right)\colon\nabla\boldsymbol{\eta}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{M}\mathbf{F}^{T}\right)\colon\nabla\mathbf{u}, (77)

as mm\to\infty. For what concerns the third term on the right hand side of (72)1, thanks to (69), (73)1 and the fact that 𝜼mH1(Ω;3)L6(Ω;3)\boldsymbol{\eta}_{m}\in H^{1}(\Omega;\mathbb{R}^{3})\hookrightarrow L^{6}(\Omega;\mathbb{R}^{3}), we have that ω𝐅m𝜼mω𝐅𝐮\omega\nabla\mathbf{F}_{m}\boldsymbol{\eta}_{m}\to\omega\nabla\mathbf{F}\mathbf{u} strongly in L2(0,T;L2(Ω;3×3))L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). Indeed,

0T|ω|2Ω|𝐅m𝜼m𝐅𝐮|2=0T|ω|2Ω|(𝐅m𝐅)𝜼m+𝐅(𝜼m𝐮)|2\displaystyle\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\nabla\mathbf{F}_{m}\boldsymbol{\eta}_{m}-\nabla\mathbf{F}\mathbf{u}|^{2}=\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\nabla(\mathbf{F}_{m}-\mathbf{F})\boldsymbol{\eta}_{m}+\nabla\mathbf{F}(\boldsymbol{\eta}_{m}-\mathbf{u})|^{2}
C0T|ω|2(𝐅m𝐅)L3(Ω;3×3×3)2𝜼mL6(Ω;3)2\displaystyle\leq C\int_{0}^{T}|\omega|^{2}||\nabla(\mathbf{F}_{m}-\mathbf{F})||_{L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}||\boldsymbol{\eta}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3})}^{2}
+C0T|ω|2𝐅L3(Ω;3×3×3)2𝜼m𝐮L6(Ω;3)2\displaystyle+C\int_{0}^{T}|\omega|^{2}||\nabla\mathbf{F}||_{L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}||\boldsymbol{\eta}_{m}-\mathbf{u}||_{L^{6}(\Omega;\mathbb{R}^{3})}^{2}
CωL(0,T)2(𝐅m𝐅)L2(0,T;L3(Ω;3×3×3))2𝜼mL6(Ω;3)2\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla(\mathbf{F}_{m}-\mathbf{F})||_{L^{2}(0,T;L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3}))}^{2}||\boldsymbol{\eta}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3})}^{2}
+CωL(0,T)2𝐅L2(0,T;L3(Ω;3×3×3))2𝜼m𝐮H1(Ω;3)20,\displaystyle+C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla\mathbf{F}||_{L^{2}(0,T;L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3}))}^{2}||\boldsymbol{\eta}_{m}-\mathbf{u}||_{H^{1}(\Omega;\mathbb{R}^{3})}^{2}\to 0,

as mm\to\infty. Hence, using (62), by the product of weak–strong convergence we have that

0TωΩ(𝐅m𝐌m)𝜼m0TωΩ(𝐅𝐌)𝐮,\int_{0}^{T}\omega\int_{\Omega}\left(\nabla\mathbf{F}_{m}\odot\mathbf{M}_{m}\right)\cdot\boldsymbol{\eta}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\left(\nabla\mathbf{F}\odot\mathbf{M}\right)\cdot\mathbf{u}, (78)

as mm\to\infty.

Now we consider the second equation in (72). Since

ω𝚺m,ω𝚯C0(0,T;L2(Ω;3×3))L4+9s13s(0,T;L2(Ω;3×3)),\omega\boldsymbol{\Sigma}_{m},\omega\boldsymbol{\Theta}\in C^{0}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{4+\frac{9s}{1-3s}}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})\right),

with s(0,13)s\in\left(0,\frac{1}{3}\right), we get from (65) and (73)2 that

0TωΩt𝐅m:𝚺m0TωΩt𝐅:𝚯,\int_{0}^{T}\omega\int_{\Omega}\partial_{t}\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\partial_{t}\mathbf{F}\colon\boldsymbol{\Theta},

as mm\to\infty. For what concerns the second term in (72)2, we observe, thanks to (69) and (73)2, that ω𝐅m𝚺mω𝐅𝚯\omega\nabla\mathbf{F}_{m}\odot\boldsymbol{\Sigma}_{m}\to\omega\nabla\mathbf{F}\odot\boldsymbol{\Theta} strongly in L2(0,T;L65(Ω;3×3))L^{2}(0,T;L^{\frac{6}{5}}(\Omega;\mathbb{R}^{3\times 3})). Indeed, we infer that

0T|ω|2(Ω|𝐅m𝚺m𝐅𝚯|65)53\displaystyle\int_{0}^{T}|\omega|^{2}\left(\int_{\Omega}\left|\nabla\mathbf{F}_{m}\odot\boldsymbol{\Sigma}_{m}-\nabla\mathbf{F}\odot\boldsymbol{\Theta}\right|^{\frac{6}{5}}\right)^{\frac{5}{3}}
=0T|ω|2(Ω|(𝐅m𝐅)𝚺m+𝐅(𝚺m𝚯)|65)53\displaystyle=\int_{0}^{T}|\omega|^{2}\left(\int_{\Omega}\left|\nabla(\mathbf{F}_{m}-\mathbf{F})\odot\boldsymbol{\Sigma}_{m}+\nabla\mathbf{F}\odot(\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta})\right|^{\frac{6}{5}}\right)^{\frac{5}{3}}
0T|ω|2(𝐅m𝐅)L3(Ω;3×3×3)2𝚺mL2(Ω;3×3)2\displaystyle\leq\int_{0}^{T}|\omega|^{2}||\nabla(\mathbf{F}_{m}-\mathbf{F})||_{L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}||\boldsymbol{\Sigma}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+0T|ω|2𝐅L3(Ω;3×3×3)2𝚺m𝚯L2(Ω;3×3)2\displaystyle+\int_{0}^{T}|\omega|^{2}||\nabla\mathbf{F}||_{L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}||\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
CωL(0,T)2(𝐅m𝐅)L2(0,T;L3(Ω;3×3×3))2𝚺mL2(Ω;3×3)2\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla(\mathbf{F}_{m}-\mathbf{F})||_{L^{2}(0,T;L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3}))}^{2}||\boldsymbol{\Sigma}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+CωL(0,T)2𝐅L2(0,T;L3(Ω;3×3×3))2𝚺m𝚯L2(Ω;3×3)20,\displaystyle+C||\omega||_{L^{\infty}(0,T)}^{2}||\nabla\mathbf{F}||_{L^{2}(0,T;L^{3}(\Omega;\mathbb{R}^{3\times 3\times 3}))}^{2}||\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}\to 0,

as mm\to\infty. Hence, using (61), by the product of weak–strong convergence we have that

0TωΩ(𝐯m)𝐅m:𝚺m0TωΩ(𝐯)𝐅:𝚯,\int_{0}^{T}\omega\int_{\Omega}\left({\mathbf{v}}_{m}\cdot\nabla\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\left({\mathbf{v}}\cdot\nabla\right)\mathbf{F}\colon\boldsymbol{\Theta}, (79)

as mm\to\infty. For what concerns the third term in (72)2, thanks to (70) and (73)2 we have that ω𝚺m𝐅mTω𝚯𝐅T\omega\boldsymbol{\Sigma}_{m}\mathbf{F}_{m}^{T}\to\omega\boldsymbol{\Theta}\mathbf{F}^{T} strongly in L2(0,T;L2(Ω;3×3))L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). Indeed,

0T|ω|2Ω|𝚺m𝐅mT𝚯𝐅T|2=0T|ω|2Ω|𝚺m(𝐅m𝐅)T+(𝚺m𝚯)𝐅T|2\displaystyle\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\boldsymbol{\Sigma}_{m}\mathbf{F}_{m}^{T}-\boldsymbol{\Theta}\mathbf{F}^{T}|^{2}=\int_{0}^{T}|\omega|^{2}\int_{\Omega}|\boldsymbol{\Sigma}_{m}(\mathbf{F}_{m}-\mathbf{F})^{T}+(\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta})\mathbf{F}^{T}|^{2}
C0T|ω|2𝐅m𝐅L(Ω;𝐑3×3)2Ω|𝚺m|2+C0T|ω|2𝐅L(Ω;𝐑3×3)2Ω|𝚺m𝚯|2\displaystyle\leq C\int_{0}^{T}|\omega|^{2}||\mathbf{F}_{m}-\mathbf{F}||_{L^{\infty}(\Omega;\mathbf{R}^{3\times 3})}^{2}\int_{\Omega}|\boldsymbol{\Sigma}_{m}|^{2}+C\int_{0}^{T}|\omega|^{2}||\mathbf{F}||_{L^{\infty}(\Omega;\mathbf{R}^{3\times 3})}^{2}\int_{\Omega}|\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta}|^{2}
CωL(0,T)2𝐅m𝐅L2(0,T;L(Ω;3×3))2𝚺mL2(Ω;3×3)2\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}^{2}||\mathbf{F}_{m}-\mathbf{F}||_{L^{2}(0,T;L^{\infty}(\Omega;\mathbb{R}^{3\times 3}))}^{2}||\boldsymbol{\Sigma}_{m}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+CωL(0,T)2𝐅L2(0,T;L(Ω;3×3))2𝚺m𝚯L2(Ω;3×3)20,\displaystyle+C||\omega||_{L^{\infty}(0,T)}^{2}||\mathbf{F}||_{L^{2}(0,T;L^{\infty}(\Omega;\mathbb{R}^{3\times 3}))}^{2}||\boldsymbol{\Sigma}_{m}-\boldsymbol{\Theta}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}\to 0,

as mm\to\infty. Hence, using (61), by the product of weak–strong convergence we have that

0TωΩ(𝐯m)𝐅m:𝚺m0TωΩ(𝐯)𝐅:𝚯,\int_{0}^{T}\omega\int_{\Omega}\left(\nabla{\mathbf{v}}_{m}\right)\mathbf{F}_{m}\colon\boldsymbol{\Sigma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\left(\nabla{\mathbf{v}}\right)\mathbf{F}\colon\boldsymbol{\Theta}, (80)

as mm\to\infty. Finally, thanks to (62) and to (73)2, we have the limit

0TωΩ𝐌m:𝚺m0TωΩ𝐌:𝚯,\int_{0}^{T}\omega\int_{\Omega}\mathbf{M}_{m}\colon\boldsymbol{\Sigma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\mathbf{M}\colon\boldsymbol{\Theta}, (81)

as mm\to\infty.

We consider the third equation in (72). The limit of the first term in (72)3 can be studied similarly to the limit of the last term in (72)2. Concerning the second term in (72)3, we employ a similar argument as in [13, Remark 11]. Using (69), (71) and the fact that 𝐅mwRC(×3×3;3×3)\partial_{\mathbf{F}_{m}}w_{R}\in C(\mathbb{R}\times\mathbb{R}^{3\times 3};\mathbb{R}^{3\times 3}), we have that 𝐅mwR(ϕm,𝐅m)𝐅wR(ϕ,𝐅)\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\to\partial_{\mathbf{F}}w_{R}(\phi,\mathbf{F}) a.e. in ΩT\Omega_{T}. We observe that, given (69) and (17) with s=2,h=245s=2,h=\frac{24}{5}, we get that

𝐅m𝐅inLq(0,T;L65q(Ω;3×3)),q[1,9).\mathbf{F}_{m}\to\mathbf{F}\quad\text{in}\quad L^{q}(0,T;L^{\frac{6}{5}q}(\Omega;\mathbb{R}^{3\times 3})),\;\;q\in[1,9). (82)

Hence, we can prove that |𝐅m|qω𝚪m|𝐅|qω𝚷|\mathbf{F}_{m}|^{q}\omega\boldsymbol{\Gamma}_{m}\to|\mathbf{F}|^{q}\omega\boldsymbol{\Pi} strongly in L1(0,T;L1(Ω;3×3))L^{1}(0,T;L^{1}(\Omega;\mathbb{R}^{3\times 3})) for q[1,9)q\in[1,9). Indeed

0T|ω|Ω||𝐅m|q𝚪m|𝐅|q𝚷|=0T|ω|Ω|(|𝐅m|q|𝐅|q)𝚪m+|𝐅|q(𝚪m𝚷)|\displaystyle\int_{0}^{T}|\omega|\int_{\Omega}\left||\mathbf{F}_{m}|^{q}\boldsymbol{\Gamma}_{m}-|\mathbf{F}|^{q}\boldsymbol{\Pi}\right|=\int_{0}^{T}|\omega|\int_{\Omega}\left|\left(|\mathbf{F}_{m}|^{q}-|\mathbf{F}|^{q}\right)\boldsymbol{\Gamma}_{m}+|\mathbf{F}|^{q}\left(\boldsymbol{\Gamma}_{m}-\boldsymbol{\Pi}\right)\right|
=0T|ω|Ω|(|𝐅+(𝐅m𝐅)|q|𝐅|q)𝚪m+|𝐅|q(𝚪m𝚷)|\displaystyle=\int_{0}^{T}|\omega|\int_{\Omega}\left|\left(|\mathbf{F}+(\mathbf{F}_{m}-\mathbf{F})|^{q}-|\mathbf{F}|^{q}\right)\boldsymbol{\Gamma}_{m}+|\mathbf{F}|^{q}\left(\boldsymbol{\Gamma}_{m}-\boldsymbol{\Pi}\right)\right|
C0T|ω|Ω|𝐅m𝐅|q|𝚪m|+0T|ω|Ω|𝐅|q|𝚪m𝚷|\displaystyle\leq C\int_{0}^{T}|\omega|\int_{\Omega}|\mathbf{F}_{m}-\mathbf{F}|^{q}|\boldsymbol{\Gamma}_{m}|+\int_{0}^{T}|\omega|\int_{\Omega}|\mathbf{F}|^{q}|\boldsymbol{\Gamma}_{m}-\boldsymbol{\Pi}|
C0T|ω|𝐅m𝐅L65q(Ω;3×3)q𝚪mL6(Ω;3×3)+0T|ω|𝐅L65q(Ω;3×3)q𝚪m𝚷L6(Ω;3×3)\displaystyle\leq C\int_{0}^{T}|\omega|\,||\mathbf{F}_{m}-\mathbf{F}||_{L^{\frac{6}{5}q}(\Omega;\mathbb{R}^{3\times 3})}^{q}||\boldsymbol{\Gamma}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3\times 3})}+\int_{0}^{T}|\omega|\,||\mathbf{F}||_{L^{\frac{6}{5}q}(\Omega;\mathbb{R}^{3\times 3})}^{q}||\boldsymbol{\Gamma}_{m}-\boldsymbol{\Pi}||_{L^{6}(\Omega;\mathbb{R}^{3\times 3})}
CωL(0,T)𝐅m𝐅Lq(0,T;L65q(Ω;3×3))𝚪mL6(Ω;3×3)\displaystyle\leq C||\omega||_{L^{\infty}(0,T)}||\mathbf{F}_{m}-\mathbf{F}||_{L^{q}(0,T;L^{\frac{6}{5}q}(\Omega;\mathbb{R}^{3\times 3}))}||\boldsymbol{\Gamma}_{m}||_{L^{6}(\Omega;\mathbb{R}^{3\times 3})}
+CωL(0,T)𝐅Lq(0,T;L65q(Ω;3×3))𝚪m𝚷H1(Ω;3×3)0,\displaystyle+C||\omega||_{L^{\infty}(0,T)}||\mathbf{F}||_{L^{q}(0,T;L^{\frac{6}{5}q}(\Omega;\mathbb{R}^{3\times 3}))}||\boldsymbol{\Gamma}_{m}-\boldsymbol{\Pi}||_{H^{1}(\Omega;\mathbb{R}^{3\times 3})}\to 0,

as mm\to\infty, thanks to (82) and (73)3. Then, thanks to the growth behavior in A2Bis, applying a generalized form of the Lebesgue convergence theorem and using also (73)3 we have that

0TωΩ𝐅mwR(ϕm,𝐅m):𝚪m0TωΩ𝐅wR(ϕ,𝐅):𝚷,\int_{0}^{T}\omega\int_{\Omega}\partial_{\mathbf{F}_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\colon\boldsymbol{\Gamma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\partial_{\mathbf{F}}w_{R}(\phi,\mathbf{F})\colon\boldsymbol{\Pi}, (83)

as mm\to\infty. With similar arguments, thanks to the growth behavior in A2Bis, (82) and (73)4, we can also conclude that

0TωΩϕmwR(ϕm,𝐅m)χm0TωΩϕwR(ϕ,𝐅)r.\int_{0}^{T}\omega\int_{\Omega}\partial_{\phi_{m}}w_{R}(\phi_{m},\mathbf{F}_{m})\chi_{m}\to\int_{0}^{T}\omega\int_{\Omega}\partial_{\phi}w_{R}(\phi,\mathbf{F})r. (84)

We also observe that, given (71) and (17) with s=4,h=485s=4,h=\frac{48}{5}, we get that

ϕmϕinLq(0,T;L65q(Ω)),q[1,13).\phi_{m}\to\phi\quad\text{in}\quad L^{q}(0,T;L^{\frac{6}{5}q}(\Omega)),\;\;q\in[1,13). (85)

Hence, given the growth law and regularity in Assumption A3, together with (73)4, applying similarly a generalized form of the Lebesgue convergence theorem we get that

0TωΩψ(ϕm)χm0TωΩψ(ϕ)r,\int_{0}^{T}\omega\int_{\Omega}\psi^{\prime}\left(\phi_{m}\right)\chi_{m}\to\int_{0}^{T}\omega\int_{\Omega}\psi^{\prime}\left(\phi\right)r, (86)

as mm\to\infty. We also deduce, thanks to (63) and (73)3, that

0TωΩ𝐅m\setstackgapS0.4ex\Shortstack𝚪m0TωΩ𝐅\setstackgapS0.4ex\Shortstack𝚷,\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{F}_{m}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Gamma}_{m}\to\int_{0}^{T}\omega\int_{\Omega}\nabla\mathbf{F}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}, (87)

as mm\to\infty.

Considering the fourth equation in (72), since

ωξm,ωqC0(0,T;H1(Ω;3×3))L2(0,T;H1(Ω)),\omega\xi_{m},\omega q\in C^{0}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2}\left(0,T;H^{1}(\Omega)\right),

we get from (68) and (73)4 that

0TωΩtϕmξm0Tω<tϕ,q>,\int_{0}^{T}\omega\int_{\Omega}\partial_{t}\phi_{m}\xi_{m}\to\int_{0}^{T}\omega<\partial_{t}\phi,q>,

as mm\to\infty. Concerning the second term in (72)4, we obtain, with similar calculations as in (4.5), that ωϕmξmωϕq\omega\nabla\phi_{m}\xi_{m}\to\omega\nabla\phi q strongly in L2(0,T;L32(Ω))L^{2}(0,T;L^{\frac{3}{2}}(\Omega)). Hence, thanks to (61),

0TωΩ(𝐯mϕm)ξm0TωΩ(𝐯ϕ)q,\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{v}_{m}\cdot\nabla\phi_{m}\right)\xi_{m}\to\int_{0}^{T}\omega\int_{\Omega}\left(\mathbf{v}\cdot\nabla\phi\right)q, (88)

as mm\to\infty.

As for the last term in (72)4, considering (71) and Assumption A1, b(ϕm)b(ϕ)b(\phi_{m})\to b(\phi) a.e. in ΩT\Omega_{T} and is uniformly bounded, hence by applying the Lebesgue convergence theorem and (73)4 we obtain that b(ϕm)ξmb(ϕ)qb(\phi_{m})\nabla\xi_{m}\to b(\phi)\nabla q strongly in L2(0,T;L2(Ω;3))L^{2}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3})\right). Then, by (71) we have that

0TωΩb(ϕm)μmξm0TωΩb(ϕ)μq,\int_{0}^{T}\omega\int_{\Omega}b\left(\phi_{m}\right)\nabla\mu_{m}\cdot\nabla\xi_{m}\to\int_{0}^{T}\omega\int_{\Omega}b\left(\phi\right)\nabla\mu\cdot\nabla q, (89)

as mm\to\infty.

Lastly, considering (66), (67) and (73)4, we obtain that

0TωΩμmχm0TωΩμr,\int_{0}^{T}\omega\int_{\Omega}\mu_{m}\chi_{m}\to\int_{0}^{T}\omega\int_{\Omega}\mu r, (90)

and

0TωΩϕmχm0TωΩϕr,\int_{0}^{T}\omega\int_{\Omega}\nabla\phi_{m}\cdot\nabla\chi_{m}\to\int_{0}^{T}\omega\int_{\Omega}\nabla\phi\cdot\nabla r, (91)

as mm\to\infty. We also recall (84) and (86) to pass to the limit in (72)5.

We need finally to prove that the initial conditions hold. Due to (69) and (71), in particular to the facts that 𝐅m𝐅\mathbf{F}_{m}\to\mathbf{F} strongly in C0([0,T];L2(Ω;3×3))C^{0}([0,T];L^{2}(\Omega;\mathbb{R}^{3\times 3})) and ϕmϕ\phi_{m}\to\phi strongly in C0([0,T];L2(Ω))C^{0}([0,T];L^{2}(\Omega)), and the facts that 𝐅m(0)=PmL,Σ(𝐅0)𝐅0\mathbf{F}_{m}(0)=P_{m}^{L,\Sigma}(\mathbf{F}_{0})\to\mathbf{F}_{0} strongly in L2(Ω,3×3)L^{2}(\Omega,\mathbb{R}^{3\times 3}) and ϕm(0)=PmL(ϕ0)ϕ0\phi_{m}(0)=P_{m}^{L}(\phi_{0})\to\phi_{0} strongly in L2(Ω)L^{2}(\Omega), we have that 𝐅(0)=𝐅0\mathbf{F}(0)=\mathbf{F}_{0} and ϕ(0)=ϕ0\phi(0)=\phi_{0} a.e. in Ω\Omega. We have thus proved Theorem 3.1.

4.6 Passage to the limit as RR\to\infty

We recall that the limit point (𝐯R,𝐅R,𝐌R,ϕR,μR)(\mathbf{v}_{R},\mathbf{F}_{R},\mathbf{M}_{R},\phi_{R},\mu_{R}) of system (72), as mm\to\infty, satisfies the system

{νΩ𝐯R:𝐮=ΩμRϕR𝐮Ω(𝐌R𝐅RT):𝐮+Ω(𝐅R𝐌R)𝐮,Ωt𝐅R:𝚯+Ω(𝐯R)𝐅R:𝚯Ω(𝐯R)𝐅R:𝚯+γΩ𝐌R:𝚯=0,Ω𝐌R:𝚷=Ω𝐅RwR(ϕR,𝐅R):𝚷+λΩ𝐅R\setstackgapS0.4ex\Shortstack𝚷,<tϕR,q>+Ω(𝐯RϕR)q+Ωb(ϕR)μRq=0,ΩμRr=ΩϕRr+Ωψ(ϕR)r+ΩϕRwR(ϕR,𝐅R)r,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{R}\colon\nabla\mathbf{u}=\int_{\Omega}\mu_{R}\nabla\phi_{R}\cdot\mathbf{u}-\int_{\Omega}\left(\mathbf{M}_{R}\mathbf{F}_{R}^{T}\right)\colon\nabla\mathbf{u}+\int_{\Omega}\left(\nabla\mathbf{F}_{R}\odot\mathbf{M}_{R}\right)\cdot\mathbf{u},\\ \displaystyle\int_{\Omega}\partial_{t}\mathbf{F}_{R}\colon\boldsymbol{\Theta}+\int_{\Omega}\left({\mathbf{v}_{R}}\cdot\nabla\right)\mathbf{F}_{R}\colon\boldsymbol{\Theta}-\int_{\Omega}\left(\nabla{\mathbf{v}_{R}}\right)\mathbf{F}_{R}\colon\boldsymbol{\Theta}+\gamma\int_{\Omega}\mathbf{M}_{R}\colon\boldsymbol{\Theta}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}_{R}\colon\boldsymbol{\Pi}=\int_{\Omega}\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\colon\boldsymbol{\Pi}+\lambda\int_{\Omega}\nabla\mathbf{F}_{R}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi},\\ \displaystyle<\partial_{t}\phi_{R},q>+\int_{\Omega}\left(\mathbf{v}_{R}\cdot\nabla\phi_{R}\right)q+\int_{\Omega}b\left(\phi_{R}\right)\nabla\mu_{R}\cdot\nabla q=0,\\ \displaystyle\int_{\Omega}\mu_{R}r=\int_{\Omega}\nabla\phi_{R}\cdot\nabla r+\int_{\Omega}\psi^{\prime}\left(\phi_{R}\right)r+\int_{\Omega}\partial_{\phi_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})r,\end{cases} (92)

a.e. in (0,T)(0,T) and for all 𝐮H0,div1(Ω;3)\mathbf{u}\in H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right), 𝚯L2(Ω;3×3)\boldsymbol{\Theta}\in L^{2}\left(\Omega;\mathbb{R}^{3\times 3}\right), 𝚷H1(Ω;3×3)\boldsymbol{\Pi}\in H^{1}\left(\Omega;\mathbb{R}^{3\times 3}\right), q,rH1(Ω)q,r\in H^{1}(\Omega), as well as the initial conditions 𝐅R(,0)=𝐅0\mathbf{F}_{R}(\cdot,0)=\mathbf{F}_{0} a.e. in Ω\Omega and ϕR(,0)=ϕ0\phi_{R}(\cdot,0)=\phi_{0} a.e. in Ω\Omega. We note that in (92) we have restored the index RR to indicate the presence of the truncation. The aim of this section is to study the limit problem of system (92) as RR\to\infty.

We start by searching for growth laws for wR(ϕR,𝐅R)w_{R}(\phi_{R},\mathbf{F}_{R}) which are uniform in RR. From (23) and (24) we deduce that

0gR(r)Cr,0\leq g_{R}(r)\leq C\quad\forall r\in\mathbb{R}, (93)

uniformly in RR, and also that

|gR(r)|{0forrR;Cr+Crmax(1,p3)CrforRr2R;0forr2R.|g^{\prime}_{R}(r)|\leq\begin{cases}0\quad\text{for}\;r\leq R;\\ \frac{C}{r}+\frac{C}{r^{\max(1,p-3)}}\leq\frac{C}{r}\quad\text{for}\;R\leq r\leq 2R;\\ 0\quad\text{for}\,r\geq 2R.\end{cases} (94)

Then, given the definition (25) and formula (27), we obtain that

d1wR(ϕ,𝐅)C(1+|𝐅|p),-d_{1}\leq w_{R}(\phi,\mathbf{F})\leq C(1+|\mathbf{F}|^{p}), (95)

and

|𝐅wR(ϕ,𝐅)|C(1+|𝐅|p1),|ϕwR(ϕ,𝐅)|C(1+|𝐅|p),|\partial_{\mathbf{F}}w_{R}(\phi,\mathbf{F})|\leq C(1+|\mathbf{F}|^{p-1}),\;\;|\partial_{\phi}w_{R}(\phi,\mathbf{F})|\leq C(1+|\mathbf{F}|^{p}), (96)

uniformly in RR. Hence, Assumption A2 is valid for the truncated elastic energy density wRw_{R} uniformly in RR.

The a priori estimate (4.2) is uniform in the truncation parameter RR, as the weak lower semicontinuity of the involved norms translates to the limit. Hence, also arguing as in (4.2)-(42), we infer that

𝐯Ris uniformly bounded inL2(0,T;H0,div1(Ω;3))L2(0,T;Ldiv6(Ω;3)),\mathbf{v}_{R}\;\text{is uniformly bounded in}\;L^{2}(0,T;H_{0,\operatorname{div}}^{1}(\Omega;\mathbb{R}^{3}))\hookrightarrow L^{2}(0,T;L_{\operatorname{div}}^{6}(\Omega;\mathbb{R}^{3})), (97)
𝐌Ris uniformly bounded inL2(0,T;L2(Ω;3×3)),\mathbf{M}_{R}\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})), (98)
𝐅Ris uniformly bounded inL(0,T;H1(Ω;3×3))L(0,T;L6(Ω;3×3)),\mathbf{F}_{R}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3})), (99)
ϕRis uniformly bounded inL(0,T;H1(Ω))L(0,T;L6(Ω)),\phi_{R}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega)), (100)
μRis uniformly bounded inL2(0,T;L2(Ω;3)).\nabla\mu_{R}\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3})). (101)

We now derive higher order estimates for 𝐅R\mathbf{F}_{R} and ϕR\phi_{R} by employing an iterative bootstrap argument based on elliptic regularity theory. We observe that, if p4p\leq 4, the truncation operation is not active, i.e. gR(|𝐅R|)1g_{R}(|\mathbf{F}_{R}|)\equiv 1. We thus specialize to the case 4<p<64<p<6. We observe, from (99) and Assumption A2, that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL(0,T;L6p1(Ω;3×3)).\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{\infty}(0,T;L^{\frac{6}{p-1}}(\Omega;\mathbb{R}^{3\times 3})). (102)

Then, from equation (92)3, (98), elliptic regularity theory and (15) with h=2(6p)3h=\frac{2(6-p)}{3} we deduce that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω;3×3))L2(0,T;W2,6p1(Ω;3×3))L182p3(0,T;W1,182p3(Ω;3×3)).\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;W^{2,\frac{6}{p-1}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\frac{18-2p}{3}}(0,T;W^{1,\frac{18-2p}{3}}(\Omega;\mathbb{R}^{3\times 3})). (103)

From (16) with h=2(6p)h=2(6-p) and (99) we also have that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;L6(Ω;3×3))L182p3(0,T;W1,182p3(Ω;3×3))L182p(0,T;L182p(Ω;3×3)).\displaystyle L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{\frac{18-2p}{3}}(0,T;W^{1,\frac{18-2p}{3}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{18-2p}(0,T;L^{18-2p}(\Omega;\mathbb{R}^{3\times 3})). (104)

We split the argument for different sub-intervals of the interval (4,6)(4,6).

4.6.1 4<p54<p\leq 5: one bootstrap step

From (4.6) and Assumption A2 we get that

𝐅RwR(ϕR,𝐅R)is uniformly bounded in\displaystyle\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;
L182pp1(0,T;L182pp1(Ω;3×3))L2(0,T;L2(Ω;3×3)).\displaystyle L^{\frac{18-2p}{p-1}}(0,T;L^{\frac{18-2p}{p-1}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (105)

Hence, from equation (92)3, (98), and applying again elliptic regularity theory we obtain that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3))L10(0,T;L10(Ω;3×3)),\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{10}(0,T;L^{10}(\Omega;\mathbb{R}^{3\times 3})), (106)

where we used (17) with s=2,h=4s=2,h=4, which implies that

ϕRwR(ϕR,𝐅R)is uniformly bounded inL10p(0,T;L10p(Ω))L2(0,T;L2(Ω)),\displaystyle\partial_{\mathbf{\phi}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{\frac{10}{p}}(0,T;L^{\frac{10}{p}}(\Omega))\hookrightarrow L^{2}(0,T;L^{2}(\Omega)), (107)

and, proceeding as in (4.3),

ϕRis uniformly bounded inL(0,T;H1(Ω))L2(0,T;H2(Ω)).\displaystyle\phi_{R}\;\text{is uniformly bounded in}\;L^{\infty}(0,T;H^{1}(\Omega))\cap L^{2}(0,T;H^{2}(\Omega)). (108)

4.6.2 5<p1635<p\leq\frac{16}{3}: two bootstrap steps

We observe that in this interval (4.6) does not imply that 𝐅RwR(ϕR,𝐅R)\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R}) is uniformly bounded in L2(0,T;L2(Ω;3×3))L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). Hence, we modify (4.6) using (16) with h=6(6p)2p7h=\frac{6(6-p)}{2p-7}, obtaining that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;L6(Ω;3×3))L182p3(0,T;W1,182p3(Ω;3×3))L2(p1)(0,T;L6(p1)2p7(Ω;3×3)).\displaystyle L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{\frac{18-2p}{3}}(0,T;W^{1,\frac{18-2p}{3}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2(p-1)}(0,T;L^{\frac{6(p-1)}{2p-7}}(\Omega;\mathbb{R}^{3\times 3})). (109)

Hence, it holds that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL2(0,T;L62p7(Ω;3×3)),\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{\frac{6}{2p-7}}(\Omega;\mathbb{R}^{3\times 3})), (110)

and we perform a further bootstrap argument employing elliptic regularity theory in equation (92)3. Introducing the new exponent p:=2p6p^{\prime}:=2p-6, we have that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL2(0,T;L6p1(Ω;3×3)),\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{\frac{6}{p^{\prime}-1}}(\Omega;\mathbb{R}^{3\times 3})), (111)

and, proceeding as in (4.6) and (4.6),

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L182p(0,T;L182p(Ω;3×3))L304p(0,T;L304p(Ω;3×3)).\displaystyle L^{18-2p^{\prime}}(0,T;L^{18-2p^{\prime}}(\Omega;\mathbb{R}^{3\times 3}))\equiv L^{30-4p}(0,T;L^{30-4p}(\Omega;\mathbb{R}^{3\times 3})). (112)

Then finally, it turns out that

𝐅RwR(ϕR,𝐅R)is uniformly bounded in\displaystyle\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;
L304pp1(0,T;L304pp1(Ω;3×3))L2(0,T;L2(Ω;3×3)).\displaystyle L^{\frac{30-4p}{p-1}}(0,T;L^{\frac{30-4p}{p-1}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (113)

Hence, from equation (92)3, (98), and applying again elliptic regularity theory we obtain that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3))L8p2p6(0,T;L2p(Ω;3×3)),\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\frac{8p}{2p-6}}(0,T;L^{2p}(\Omega;\mathbb{R}^{3\times 3})), (114)

where we have applied (17), with s=2,h=2p6s=2,h=2p-6. Using Assumption A2, (4.6.2) implies that

ϕRwR(ϕR,𝐅R)is uniformly bounded inL82p6(0,T;L2(Ω)).\displaystyle\partial_{\mathbf{\phi}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{\frac{8}{2p-6}}(0,T;L^{2}(\Omega)). (115)

In conclusion, taking (4.3) to the power of 42p6\frac{4}{2p-6} and integrating in time over the interval (0,T)(0,T), we obtain that

ϕRis uniformly bounded in\displaystyle\phi_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω))L82p6(0,T;H2(Ω))L12p203(2p6)(0,T;W1,12p203(2p6)(Ω)),\displaystyle L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;H^{2}(\Omega))\hookrightarrow L^{\frac{12p-20}{3(2p-6)}}(0,T;W^{1,\frac{12p-20}{3(2p-6)}}(\Omega)), (116)

where we have used (18), with s=2,q=6s=2,q=6. We observe that 12p203(2p6)>2\frac{12p-20}{3(2p-6)}>2 for any p>3p>3.

4.6.3 163<p112\frac{16}{3}<p\leq\frac{11}{2}: three bootstrap steps

We observe that in this interval (4.6.2) does not imply that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL2(0,T;L2(Ω;3×3)).\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})).

Hence, we modify (4.6.2) using (16) with h=6(6p)p+p7h=\frac{6(6-p^{\prime})}{p+p^{\prime}-7}, obtaining that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;L6(Ω;3×3))L182p3(0,T;W1,182p3(Ω;3×3))L2(p1)(0,T;L6(p1)p+p7(Ω;3×3)).\displaystyle L^{\infty}(0,T;L^{6}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{\frac{18-2p^{\prime}}{3}}(0,T;W^{1,\frac{18-2p^{\prime}}{3}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2(p-1)}(0,T;L^{\frac{6(p-1)}{p+p^{\prime}-7}}(\Omega;\mathbb{R}^{3\times 3})). (117)

Hence, we infer that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL2(0,T;L6p+p7(Ω;3×3)).\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{\frac{6}{p+p^{\prime}-7}}(\Omega;\mathbb{R}^{3\times 3})). (118)

Introducing the new exponent p′′:=p+p6p^{\prime\prime}:=p+p^{\prime}-6, we have that

𝐅RwR(ϕR,𝐅R)is uniformly bounded inL2(0,T;L6p′′1(Ω;3×3)),\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;L^{2}(0,T;L^{\frac{6}{p^{\prime\prime}-1}}(\Omega;\mathbb{R}^{3\times 3})), (119)

and, proceeding as in (4.6) and (4.6),

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L182p′′(0,T;L182p′′(Ω;3×3))L426p(0,T;L426p(Ω;3×3)).\displaystyle L^{18-2p^{\prime\prime}}(0,T;L^{18-2p^{\prime\prime}}(\Omega;\mathbb{R}^{3\times 3}))\equiv L^{42-6p}(0,T;L^{42-6p}(\Omega;\mathbb{R}^{3\times 3})).

Then, we have that

𝐅RwR(ϕR,𝐅R)is uniformly bounded in\displaystyle\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;
L426pp1(0,T;L426pp1(Ω;3×3))L2(0,T;L2(Ω;3×3)).\displaystyle L^{\frac{42-6p}{p-1}}(0,T;L^{\frac{42-6p}{p-1}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (120)

Hence, from equation (92)3, (98), and applying again elliptic regularity theory we obtain, analogously to (4.6.2), that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3))L8p2p6(0,T;L2p(Ω;3×3)),\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\frac{8p}{2p-6}}(0,T;L^{2p}(\Omega;\mathbb{R}^{3\times 3})), (121)

which implies, analogously to (4.6.2), that

ϕRis uniformly bounded in\displaystyle\phi_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω))L82p6(0,T;H2(Ω))L12p203(2p6)(0,T;L12p203(2p6)(Ω)).\displaystyle L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;H^{2}(\Omega))\hookrightarrow L^{\frac{12p-20}{3(2p-6)}}(0,T;L^{\frac{12p-20}{3(2p-6)}}(\Omega)). (122)

4.6.4 Iterative argument up to p<6p<6

Based on the previous observations, we search for an iteration argument which let us apply nn bootstrap steps to obtain higher order regularity of 𝐅R\mathbf{F}_{R} and ϕR\phi_{R} in proper contracting intervals for pp, up to p<6p<6. Let n+n\in\mathbb{N}_{+}, and

10+6(n1)(n1)+2<p10+6nn+2,\frac{10+6(n-1)}{(n-1)+2}<p\leq\frac{10+6n}{n+2}, (123)

where nn is the number of bootstrap steps which must be applied in the interval defined in (123) to obtain higher order regularity of 𝐅R\mathbf{F}_{R} and ϕR\phi_{R}. We iteratively define the sequence {pn}n\{p^{n}\}_{n\in\mathbb{N}}, where nn is an index (not an exponent), with

p0=p;\displaystyle p^{0}=p; (124)
pn=p+pn16=n(p6)+p,\displaystyle p^{n}=p+p^{n-1}-6=n(p-6)+p,

and pp as in (123). Iterating nn bootstrap steps, we have that

𝐅Ris uniformly bounded inL182pn(0,T;L182pn(Ω;3×3)),\mathbf{F}_{R}\;\text{is uniformly bounded in}\;L^{18-2p^{n}}(0,T;L^{18-2p^{n}}(\Omega;\mathbb{R}^{3\times 3})),

and hence

𝐅RwR(ϕR,𝐅R)is uniformly bounded in\displaystyle\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\;\text{is uniformly bounded in}\;
L182pnp1(0,T;L182pnp1(Ω;3×3))L2(0,T;L2(Ω;3×3)).\displaystyle L^{\frac{18-2p^{n}}{p-1}}(0,T;L^{\frac{18-2p^{n}}{p-1}}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})). (125)

Note that

182pnp1=182[n(p6)+p]p1=2ifp=10+6nn+2.\frac{18-2p^{n}}{p-1}=\frac{18-2[n(p-6)+p]}{p-1}=2\;\;\text{if}\;\;p=\frac{10+6n}{n+2}.

Hence, from (16), analogously to (4.6.2), it follows that

𝐅Ris uniformly bounded in\displaystyle\mathbf{F}_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω;3×3))L2(0,T;H2(Ω;3×3))L8p2p6(0,T;L2p(Ω;3×3)),\displaystyle L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3}))\hookrightarrow L^{\frac{8p}{2p-6}}(0,T;L^{2p}(\Omega;\mathbb{R}^{3\times 3})), (126)

which implies, analogously to (4.6.2), that

ϕRis uniformly bounded in\displaystyle\phi_{R}\;\text{is uniformly bounded in}\;
L(0,T;H1(Ω))L82p6(0,T;H2(Ω))L12p203(2p6)(0,T;L12p203(2p6)(Ω)).\displaystyle L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;H^{2}(\Omega))\;\hookrightarrow L^{\frac{12p-20}{3(2p-6)}}(0,T;L^{\frac{12p-20}{3(2p-6)}}(\Omega)). (127)

We note that, thanks to (4.6.4), the bound (48) is still valid uniformly in RR.
Taking nn\to\infty, we obtain this result for p6p\to 6. We observe that for p=6p=6, pn=p=6p^{n}=p=6, 𝐅RwR(ϕR,𝐅R)L65(0,T;L65(Ω;3×3))\partial_{\mathbf{F}_{R}}w_{R}(\phi_{R},\mathbf{F}_{R})\in L^{\frac{6}{5}}(0,T;L^{\frac{6}{5}}(\Omega;\mathbb{R}^{3\times 3})) and the argument breaks down.

We highlight that (4.6.4) and (4.6.4) are valid for any p(4,6)p\in(4,6). Since L(0,T;H1(Ω))L82p6(0,T;H2(Ω))L(0,T;L6(Ω))L82p6(0,T;W1,6(Ω))L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;H^{2}(\Omega))\hookrightarrow L^{\infty}(0,T;L^{6}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;W^{1,6}(\Omega)), using (17) with s=h=82p6s=h=\frac{8}{2p-6} we get that

ϕRL2q(0,T;Lq(Ω))C,q=6+82p6.||\phi_{R}||_{L^{2q}(0,T;L^{q}(\Omega))}\leq C,\quad q=6+\frac{8}{2p-6}. (128)

Then, with similar calculations as in (4.3) and using Assumptions A2, A3, we obtain that |(μR,1)||(\mu_{R},1)| is bounded in L2(0,T)L^{2}(0,T) and consequently, by the Poincaré–Wirtinger inequality and (4.2),

μRis uniformly bounded inL2(0,T;H1(Ω))L2(0,T;L6(Ω)).\mu_{R}\;\text{is uniformly bounded in}\;L^{2}(0,T;H^{1}(\Omega))\;\hookrightarrow L^{2}(0,T;L^{6}(\Omega)). (129)

We finally observe that, thanks to (97)–(101), (4.6.4), (4.6.4) and (129) the estimates (58) and (60) translate to the limit.

Collecting the obtained estimates, which are uniform in RR, from the Banach–Alaoglu and the Aubin–Lions lemma, we finally obtain the convergence properties, up to subsequences of the solutions, which we still label by the index RR,

𝐯R𝐯inL2(0,T;H0,div1(Ω;3)),\displaystyle\mathbf{v}_{R}\rightharpoonup\mathbf{v}\quad\text{in}\quad L^{2}\left(0,T;H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right)\right), (130)
𝐌R𝐌inL2(0,T;L2(Ω;3×3)),\displaystyle\mathbf{M}_{R}{\rightharpoonup}\mathbf{M}\quad\text{in}\quad L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})), (131)
𝐅R𝐅inL(0,T;H1(Ω;3×3)),\displaystyle\mathbf{F}_{R}\overset{\ast}{\rightharpoonup}\mathbf{F}\quad\text{in}\quad L^{\infty}(0,T;H^{1}(\Omega;\mathbb{R}^{3\times 3})), (132)
𝐅R𝐅inL4k(0,T;W1,3+ϵ(Ω;3×3))L2(0,T;H2(Ω;3×3)),ϵ(0,3],\displaystyle\mathbf{F}_{R}{\rightharpoonup}\mathbf{F}\quad\text{in}\quad L^{4-k}(0,T;W^{1,3+\epsilon}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;H^{2}(\Omega;\mathbb{R}^{3\times 3})),\;\;\epsilon\in\left(0,3\right], (133)
t𝐅Rt𝐅inL43s(0,T;L2(Ω;3×3)),s(0,13),\displaystyle\partial_{t}\mathbf{F}_{R}\rightharpoonup\partial_{t}\mathbf{F}\quad\text{in}\quad L^{\frac{4}{3}-s}\left(0,T;L^{2}(\Omega;\mathbb{R}^{3\times 3})\right),\;\;s\in\left(0,\frac{1}{3}\right), (134)
ϕRϕinL(0,T;H1(Ω))L82p6(0,T;H2(Ω)),\displaystyle\phi_{R}\overset{\ast}{\rightharpoonup}\phi\quad\text{in}\quad L^{\infty}(0,T;H^{1}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;H^{2}(\Omega)), (135)
μRμinL2(0,T;H1(Ω)),\displaystyle\mu_{R}\rightharpoonup\mu\quad\text{in}\quad L^{2}\left(0,T;H^{1}\left(\Omega\right)\right), (136)
tϕRtϕinL2(0,T;(H1(Ω))),\displaystyle\partial_{t}\phi_{R}\rightharpoonup\partial_{t}\phi\quad\text{in}\quad L^{2}(0,T;\left(H^{1}(\Omega)\right)^{\prime}), (137)
𝐅R𝐅inC0(0,T;Lp(Ω;3×3))L2(0,T;W1,p(Ω;3×3)),p[1,6),anda.e. inΩT,\displaystyle\mathbf{F}_{R}\to\mathbf{F}\quad\text{in}\quad C^{0}(0,T;L^{p}(\Omega;\mathbb{R}^{3\times 3}))\cap L^{2}(0,T;W^{1,p}(\Omega;\mathbb{R}^{3\times 3})),\;\;p\in[1,6),\;\;\text{and}\;\;\text{a.e. in}\;\;\Omega_{T}, (138)
𝐅R𝐅inL4k(0,T;C(Ω¯;3×3)),\displaystyle\mathbf{F}_{R}\to\mathbf{F}\quad\text{in}\quad L^{4-k}(0,T;C(\bar{\Omega};\mathbb{R}^{3\times 3})), (139)
ϕRϕinC0(0,T;Lp(Ω))L82p6(0,T;W1,p(Ω)),p[1,6),anda.e. inΩT,\displaystyle\phi_{R}\to\phi\quad\text{in}\quad C^{0}(0,T;L^{p}(\Omega))\cap L^{\frac{8}{2p-6}}(0,T;W^{1,p}(\Omega)),\;\;p\in[1,6),\;\;\text{and}\;\;\text{a.e. in}\;\;\Omega_{T}, (140)

with k(0,2]k\in(0,2], as RR\to\infty. With the convergence results (130)–(140), we can pass to the limit in the system (92) as RR\to\infty, with essentially the same calculations as the ones employed to pass to the limit in (33) as mm\to\infty, with fixed test functions 𝐮H0,div1(Ω;3)\mathbf{u}\in H_{0,\operatorname{div}}^{1}\left(\Omega;\mathbb{R}^{3}\right), 𝚯L2(Ω;3×3)\boldsymbol{\Theta}\in L^{2}\left(\Omega;\mathbb{R}^{3\times 3}\right), 𝚷H1(Ω;3×3)\boldsymbol{\Pi}\in H^{1}\left(\Omega;\mathbb{R}^{3\times 3}\right), q,rH1(Ω)q,r\in H^{1}(\Omega). We only point out the fact that, thanks to (140) and to (18), with s=2,q=6s=2,q=6, we have that

ϕRϕinLq(0,T;Lq(Ω;3)),q[1,12p203(2p6)),\nabla\phi_{R}\to\nabla\phi\quad\text{in}\quad L^{q}(0,T;L^{q}(\Omega;\mathbb{R}^{3})),\quad q\in\left[1,\frac{12p-20}{3(2p-6)}\right), (141)

and, since, as already observed, 12p203(2p6)>2\frac{12p-20}{3(2p-6)}>2 for any p>3p>3,

ϕRϕinL2(0,T;L2(Ω;3)).\nabla\phi_{R}\to\nabla\phi\quad\text{in}\quad L^{2}(0,T;L^{2}(\Omega;\mathbb{R}^{3})). (142)

Hence, thanks to (142), we can pass to the limit as RR\to\infty e.g. in the first term on the right hand side of (92)1 with similar calculations to the ones employed in (76) (with fixed test functions). Also, given (140) and (17) with s=82p6,h=485(p3)s=\frac{8}{2p-6},h=\frac{48}{5(p-3)}, we get that

ϕmϕinLq(0,T;L65q(Ω)),q[1,5p7p3).\phi_{m}\to\phi\quad\text{in}\quad L^{q}(0,T;L^{\frac{6}{5}q}(\Omega)),\;\;q\in\left[1,\frac{5p-7}{p-3}\right). (143)

Hence, given the growth law and regularity in Assumption A3, and observing that

5p7p3>6+82p6,p<7,\frac{5p-7}{p-3}>6+\frac{8}{2p-6},\quad\forall p<7,

applying a generalized form of the Lebesgue convergence theorem as in (86) we can pass to the limit as RR\to\infty in the second term on the right hand side of (92)5. We have thus proved Theorem 3.1 also for p(4,6)p\in(4,6).

5 Finite element approximations of the model

In this section we propose two unconditionally gradient stable fully discrete finite element approximations of system (9)-(10). The gradient stability of the approximation schemes guarantees that the discrete system maintains the dissipative nature of the continuous system, with the possibility of defining a discrete energy which is a decreasing Lyapunov functional in time for the discrete solutions.

Remark 5.1

We observe that an existence result for System (29), i.e. the system with the truncated elastic energy density with polynomial growth up to p4p\leq 4, could be in principle obtained by studying the convergence of one of the finite element approximations introduced in this section, instead of using the Faedo–Galerkin approximation introduced in Section 4.1. Since the existence result obtained in Section 4.6 with polynomial growth of the elastic energy density up to p<6p<6 is obtained using elliptic regularity theory techniques, which are not available in the discrete systems associated to standard finite element approximations, the general existence result expressed in Theorem 3.1 cannot be obtained by studying the convergence of the forthcoming finite element approximations.

Let h>0h>0 be a discretization parameter and let 𝒯h\mathcal{T}_{h} be a quasi-uniform conforming decomposition of the domain Ω3\Omega\subset\mathbb{R}^{3} into 33-simplices KK, with hK=diam(K)h_{K}=\mathrm{diam}(K) and h=maxK𝒯hhKh=\max_{K\in\mathcal{T}_{h}}h_{K}. We introduce the following finite-element spaces for scalar, vector and matrix valued functions:

Sh\displaystyle S_{h} :={shC0(Ω¯):sh|K1(K),K𝒯h}H1(Ω),\displaystyle:=\{s_{h}\in C^{0}(\bar{\Omega}):s_{h}|_{K}\in\mathbb{P}_{1}(K),\forall K\in\mathcal{T}_{h}\}\subset H^{1}(\Omega),
Vh\displaystyle V_{h} :={𝐯hC0(Ω¯;3)H01(Ω;3):𝐯h|K[2(K)]3,K𝒯h},\displaystyle:=\{\mathbf{v}_{h}\in C^{0}(\bar{\Omega};\mathbb{R}^{3})\cap H_{0}^{1}(\Omega;\mathbb{R}^{3}):\mathbf{v}_{h}|_{K}\in[\mathbb{P}_{2}(K)]^{3},\forall K\in\mathcal{T}_{h}\},
Xh\displaystyle X_{h} :={𝐀hC0(Ω¯;3×3):𝐀h|K[1(K)]3×3,K𝒯h}H1(Ω;3×3),\displaystyle:=\{\mathbf{A}_{h}\in C^{0}(\bar{\Omega};\mathbb{R}^{3\times 3}):\mathbf{A}_{h}|_{K}\in[\mathbb{P}_{1}(K)]^{3\times 3},\forall K\in\mathcal{T}_{h}\}\subset H^{1}(\Omega;\mathbb{R}^{3\times 3}),

where l(K)\mathbb{P}_{l}(K), ll\in\mathbb{N}, stands for the space of polynomials of total order ll in KK. The spaces VhV_{h} and Rh:=ShL02(Ω)R_{h}:=S_{h}\cap L_{0}^{2}(\Omega), with L02(Ω):={fL2(Ω)|Ωf=0}L_{0}^{2}(\Omega):=\{f\in L^{2}(\Omega)|\int_{\Omega}f=0\}, constitute the lowest order Taylor-Hood elements [22], which are stable elements (i.e. satisfy the discrete Ladyzhenskaya-Babuška-Brezzi stability condition) to approximate the velocity and pressure variables respectively in the Stokes system. We also introduce the L2L^{2}-projection operators PhS:L2(Ω)ShP_{h}^{S}:L^{2}(\Omega)\rightarrow S_{h}, PhV:L2(Ω;3)VhP_{h}^{V}:L^{2}(\Omega;\mathbb{R}^{3})\rightarrow V_{h} and PhX:L2(Ω;3×3)XhP_{h}^{X}:L^{2}(\Omega;\mathbb{R}^{3\times 3})\rightarrow X_{h}. We set Δt=T/N\Delta t=T/N for a NN\in\mathbb{N}, and tn=nΔt,n=0,,Nt_{n}=n\Delta t,n=0,\dots,N. A finite element approximation of a continuous field f(𝐱,tn)f(\mathbf{x},t_{n}) at time tnt_{n} will be indicated by fhnf_{h}^{n}. Given the initial data ϕ0H1(Ω)\phi_{0}\in H^{1}(\Omega), 𝐅0H1(Ω;3×3)\mathbf{F}_{0}\in H^{1}(\Omega;\mathbb{R}^{3\times 3}), we set ϕh0=PhS(ϕ0)\phi_{h}^{0}=P_{h}^{S}(\phi_{0}) and 𝐅h0=PhX(𝐅0)\mathbf{F}_{h}^{0}=P_{h}^{X}(\mathbf{F}_{0}).

The first fully discrete approximation scheme is a generalization to system (9) of the convex splitting methods which are widely used to derive unconditionally gradient stable schemes for the Cahn–Hilliard equations [8, 9]. In order to derive such a scheme, we need to assume a particular form for the elastic energy density w(ϕ,𝐅)w(\phi,\mathbf{F}), i.e. we make the following assumption:

  • 𝐀𝟐𝐡\bf{A2_{h}}

    There exist functions fC1()f\in C^{1}(\mathbb{R}), with kf1f(r)kf2-k_{f1}\leq f(r)\leq k_{f2}, |f(r)|kf3|f^{\prime}(r)|\leq k_{f3}, kf1,kf2,kf30k_{f1},k_{f2},k_{f3}\geq 0, for all rr\in\mathbb{R}, functions g,hC1(3×3)g,h\in C^{1}(\mathbb{R}^{3\times 3}), with 0g(𝐓)kg(1+|𝐓|p)0\leq g(\mathbf{T})\leq k_{g}(1+|\mathbf{T}|^{p}), kg>0k_{g}>0, kh1h(𝐓)kh2(1+|𝐓|p)-k_{h1}\leq h(\mathbf{T})\leq k_{h2}(1+|\mathbf{T}|^{p}) and h(𝐓)kh2|𝐓|ph(\mathbf{T})\sim k_{h2}|\mathbf{T}|^{p} for |𝐓|+|\mathbf{T}|\to+\infty, kh10,kh2>0k_{h1}\geq 0,k_{h2}>0 and kh2kf1kgk_{h2}\geq k_{f1}k_{g}, p[0,6)p\in[0,6), for all 𝐓3×3\mathbf{T}\in\mathbb{R}^{3\times 3}, and function mC1()m\in C^{1}(\mathbb{R}), with kmm(s)-k_{m}\leq m(s), km0k_{m}\geq 0 for all ss\in\mathbb{R}, such that

    w(ϕ,𝐅)=f(ϕ)g(𝐅)+h(𝐅)+m(ϕ),w(\phi,\mathbf{F})=f(\phi)g(\mathbf{F})+h(\mathbf{F})+m(\phi), (144)

where m()m^{\prime}(\cdot) satisfies the same assumptions as ψ()\psi^{\prime}(\cdot) in A3. We note that (144), with the assumed properties of ff, gg, hh and jj, satisfies Assumption A2 with d1=kh1+kmd_{1}=k_{h1}+k_{m}. We moreover observe that the term m(ϕ)m(\phi) in (144) can be incorporated in the term ψ(ϕ)\psi(\phi), so that in the following analysis we will consider as new Cahn–Hilliard potential ψ(ϕ)ψ(ϕ)+m(ϕ)\psi(\phi)\leftarrow\psi(\phi)+m(\phi).

Remark 5.2

Under the assumptions fC1()f\in C^{1}(\mathbb{R}), with 0f(r)k10\leq f(r)\leq k_{1}, |f(r)|k2|f^{\prime}(r)|\leq k_{2}, k1,k20k_{1},k_{2}\geq 0, for all rr\in\mathbb{R}, and gC1(3×3)g\in C^{1}(\mathbb{R}^{3\times 3}), with 0g(𝐓)kg(1+|𝐓|p)0\leq g(\mathbf{T})\leq k_{g}(1+|\mathbf{T}|^{p}), kg>0k_{g}>0, for all 𝐓3×3\mathbf{T}\in\mathbb{R}^{3\times 3}, Hypothesis A2 could be satisfied also in the case w(ϕ,𝐅)=f(ϕ)g(𝐅)w(\phi,\mathbf{F})=f(\phi)g(\mathbf{F}). This means that, when the function ff in (144) is positive, we could consider in the present framework also an elastic energy density which degenerates with the variable ϕ\phi.

We introduce the following convex splittings

ψ(ϕ)=ψ+(ϕ)+ψ(ϕ),\displaystyle\psi(\phi)=\psi_{+}(\phi)+\psi_{-}(\phi), (145)
f(ϕ)=f+(ϕ)+f(ϕ),\displaystyle f(\phi)=f_{+}(\phi)+f_{-}(\phi),
g(𝐅)=g+(𝐅)+g(𝐅),\displaystyle g(\mathbf{F})=g_{+}(\mathbf{F})+g_{-}(\mathbf{F}),
h(𝐅)=h+(𝐅)+h(𝐅),\displaystyle h(\mathbf{F})=h_{+}(\mathbf{F})+h_{-}(\mathbf{F}),

into convex (indicated by the ++ index) and concave (indicated by the - index) parts.

Remark 5.3

The existence of a convex splitting for a generic function gC1(3×3)g\in C^{1}(\mathbb{R}^{3\times 3}) is guaranteed if e.g. we make the non restrictive assumption that, for any 𝐅1,𝐅23×3\mathbf{F}_{1},\mathbf{F}_{2}\in\mathbb{R}^{3\times 3}, there exists a c0>0c_{0}>0 such that

(g(𝐅2)g(𝐅1)):(𝐅2𝐅1)c0|𝐅2𝐅1|2.\left(g^{\prime}(\mathbf{F}_{2})-g^{\prime}(\mathbf{F}_{1})\right)\colon(\mathbf{F}_{2}-\mathbf{F}_{1})\geq-c_{0}|\mathbf{F}_{2}-\mathbf{F}_{1}|^{2}. (146)

Indeed, let us choose

g+(𝐅):=g(𝐅)+c02|𝐅|2.g_{+}(\mathbf{F}):=g(\mathbf{F})+\frac{c_{0}}{2}|\mathbf{F}|^{2}.

Hence,

(g+(𝐅2)g+(𝐅1)):(𝐅2𝐅1)=(g(𝐅2)g(𝐅1)):(𝐅2𝐅1)+c0|𝐅2𝐅1|20,\left(g^{\prime}_{+}(\mathbf{F}_{2})-g^{\prime}_{+}(\mathbf{F}_{1})\right)\colon(\mathbf{F}_{2}-\mathbf{F}_{1})=\left(g^{\prime}(\mathbf{F}_{2})-g^{\prime}(\mathbf{F}_{1})\right)\colon(\mathbf{F}_{2}-\mathbf{F}_{1})+c_{0}|\mathbf{F}_{2}-\mathbf{F}_{1}|^{2}\geq 0,

which implies that g+g^{\prime}_{+} is a monotone function, and therefore g+g_{+} is convex. We then set

g(𝐅):=c02|𝐅|2,g_{-}(\mathbf{F}):=-\frac{c_{0}}{2}|\mathbf{F}|^{2},

whith gg_{-} a concave function.

We also introduce the notation [f()]+:=max(0,f())[f(\cdot)]_{+}:=\max(0,f(\cdot)) for the positive part of a given function fC0()f\in C^{0}(\mathbb{R}). We then consider the following fully discretized approximation of system (9):

Problem 𝐏𝐡𝐈\mathbf{P_{h}^{I}}: for n=0,,N1n=0,\dots,N-1, given ϕhnSh\phi_{h}^{n}\in S_{h}, 𝐅hnXh\mathbf{F}_{h}^{n}\in X_{h},
find (𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h} such that,
for all (𝐮h,ph,𝚯h,𝚷h,ξh,χh)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{u}_{h},p_{h},\boldsymbol{\Theta}_{h},\boldsymbol{\Pi}_{h},\xi_{h},\chi_{h})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h},

{νΩ𝐯hn+1:𝐮hΩshn+1div𝐮h=Ωμhn+1ϕhn𝐮hΩ(𝐌hn+1(𝐅hn)T):𝐮h+Ω(𝐅hn𝐌hn+1)𝐮h,Ωphdiv𝐯hn+1=0,Ω(𝐅hn+1𝐅hn):𝚯h+ΔtΩ(𝐯hn+1)𝐅hn:𝚯hΔtΩ(𝐯hn+1)𝐅hn:𝚯h+ΔtγΩ𝐌hn+1:𝚯h=0,Ω𝐌hn+1:𝚷h=Ω(h+(𝐅hn+1)+h(𝐅hn)):𝚷h+Ω[f(ϕhn)]+(g+(𝐅hn+1)+g(𝐅hn)):𝚷h+Ω(f(ϕhn)[f(ϕhn)]+)(g+(𝐅hn)+g(𝐅hn+1)):𝚷h+λΩ𝐅hn+1\setstackgapS0.4ex\Shortstack𝚷h,Ω(ϕhn+1ϕhn)ξh+ΔtΩ(𝐯hn+1ϕhn)ξh+ΔtΩb(ϕhn)μhn+1ξh=0,Ωμhn+1χh=Ωϕhn+1χh+Ω(ψ+(ϕhn+1)+ψ(ϕhn))χh+Ω(f+(ϕhn+1)g(𝐅hn+1)+f(ϕhn)g(𝐅hn+1))χh.\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h}^{n+1}\colon\nabla\mathbf{u}_{h}-\int_{\Omega}s_{h}^{n+1}\operatorname{div}\mathbf{u}_{h}=\int_{\Omega}\mu_{h}^{n+1}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}-\int_{\Omega}\left(\mathbf{M}_{h}^{n+1}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}\\ \displaystyle+\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n+1}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}p_{h}\operatorname{div}\mathbf{v}_{h}^{n+1}=0,\\ \\ \displaystyle\int_{\Omega}\left(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}\right)\colon\boldsymbol{\Theta}_{h}+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n+1}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n+1}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Pi}_{h}=\int_{\Omega}\left(h_{+}^{\prime}(\mathbf{F}_{h}^{n+1})+h_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}+\int_{\Omega}[f(\phi_{h}^{n})]_{+}\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n+1})+g_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}\\ \displaystyle+\int_{\Omega}(f(\phi_{h}^{n})-[f(\phi_{h}^{n})]_{+})\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{F}_{h}^{n+1})\right)\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\mathbf{F}_{h}^{n+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\left(\phi_{h}^{n+1}-\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n+1}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mu_{h}^{n+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h}^{n+1}\cdot\nabla\chi_{h}+\int_{\Omega}\left(\psi_{+}^{\prime}\left(\phi_{h}^{n+1}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)\chi_{h}\\ \displaystyle+\int_{\Omega}\left(f_{+}^{\prime}(\phi_{h}^{n+1})g(\mathbf{F}_{h}^{n+1})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{F}_{h}^{n+1})\right)\chi_{h}.\end{cases} (147)
Remark 5.4

The approximation scheme (147) could be straightforwardly generalized to the case in which (144) has the form

w(ϕ,𝐅)=i=1mfi(ϕ)gi(𝐅)+h(𝐅)+m(ϕ),w(\phi,\mathbf{F})=\sum_{i=1}^{m}f_{i}(\phi)g_{i}(\mathbf{F})+h(\mathbf{F})+m(\phi), (148)

where mm\in\mathbb{N}, with the functions in (148) satisfying the Assumption 𝐀𝟐𝐡\bf{A2_{h}}. Assumption 𝐀𝟐𝐡\bf{A2_{h}} could also be relaxed by assuming the functions gig_{i} to satisfy the property that |gi(𝐓)|kh2,i(1+|𝐓|p)|g_{i}(\mathbf{T})|\leq k_{h2,i}(1+|\mathbf{T}|^{p}), p[0,6)p\in[0,6), for all 𝐓3×3\mathbf{T}\in\mathbb{R}^{3\times 3}, without constraints on their positivity, by employing a separation of the positive and negative parts of gig_{i} in (147)6 as done for the function ff in (147)4.

The existence of a solution to (147) and the gradient stability of the approximation scheme are given in the following lemma.

Lemma 5.1

For all n=0,,N1n=0,\dots,N-1, given ϕhnSh\phi_{h}^{n}\in S_{h}, 𝐅hnXh\mathbf{F}_{h}^{n}\in X_{h}, with ϕh0=PhS(ϕ0)\phi_{h}^{0}=P_{h}^{S}(\phi_{0}) and 𝐅h0=PhX(𝐅0)\mathbf{F}_{h}^{0}=P_{h}^{X}(\mathbf{F}_{0}), there exists a solution (𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h} to system (147), which satisfies the following stability bound:

maxn=0N1(12ϕhn+1L2(Ω;3)2+λ2𝐅hn+1L2(Ω;3×3×3)2)+(Δt)22n=0N1(ϕhn+1ϕhn)(Δt)2L2(Ω;3)2\displaystyle\max_{n=0\to N-1}\left(\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}\right)+\frac{(\Delta t)^{2}}{2}\sum_{n=0}^{N-1}\left|\left|\frac{\nabla(\phi_{h}^{n+1}-\phi_{h}^{n})}{(\Delta t)^{2}}\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3})}^{2} (149)
+λ(Δt)22n=0N1(𝐅hn+1𝐅hn)(Δt)2L2(Ω;3×3×3)2+Δtνn=0N1𝐯hn+1L2(Ω;3×3)2\displaystyle+\frac{\lambda(\Delta t)^{2}}{2}\sum_{n=0}^{N-1}\left|\left|\frac{\nabla(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n})}{(\Delta t)^{2}}\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\Delta t\nu\sum_{n=0}^{N-1}||\nabla\mathbf{v}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+Δtγn=0N1𝐌hn+1L2(Ω;3×3)2+Δtn=0N1Ωb(ϕhn)μhn+1μhn+1C(ϕh0,𝐅h0)+C.\displaystyle+\Delta t\gamma\sum_{n=0}^{N-1}||\mathbf{M}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\sum_{n=0}^{N-1}\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\mu_{h}^{n+1}\leq C(\phi_{h}^{0},\mathbf{F}_{h}^{0})+C.

Moreover, the function

Ω(ψ(ϕhn+1)+h(𝐅hn+1)+f(ϕhn+1)g(𝐅hn+1))+12ϕhn+1L2(Ω;3)2+λ2𝐅hn+1L2(Ω;3×3×3)2\int_{\Omega}\left(\psi(\phi_{h}^{n+1})+h(\mathbf{F}_{h}^{n+1})+f(\phi_{h}^{n+1})g(\mathbf{F}_{h}^{n+1})\right)+\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (150)

is a decreasing (Lyapunov) function for the discrete solutions.

Proof. Following [10], we are going to prove the existence of a solution to System (147) by applying [23, Lemma II.1.4], which relies on proper stability bounds for the discrete system. In order to proceed, we first consider a reduced expression of System (147), where the incompressibility constraint (9)2 is enforced in the definition of the finite element space for the variable 𝐯hn+1\mathbf{v}_{h}^{n+1}, i.e. we consider 𝐯hn+1,𝐮hVh,div\mathbf{v}_{h}^{n+1},\mathbf{u}_{h}\in V_{h,\text{div}}, where

Vh,div:={𝐯hVh:Ωdiv𝐯hph=0phSh}.V_{h,\text{div}}:=\{\mathbf{v}_{h}\in V_{h}:\int_{\Omega}\text{div}\mathbf{v}_{h}p_{h}=0\;\forall p_{h}\in S_{h}\}.

We also introduce the variables

𝐆hn+1:=𝐅hn+1𝐅hn,\mathbf{G}_{h}^{n+1}:=\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n},
zhn+1:=ϕhn+1ϕhn,z_{h}^{n+1}:=\phi_{h}^{n+1}-\phi_{h}^{n},
yhn+1:=μhn+11|Ω|Ωμhn+1.y_{h}^{n+1}:=\mu_{h}^{n+1}-\frac{1}{|\Omega|}\int_{\Omega}\mu_{h}^{n+1}.

We observe that, taking ξh1\xi_{h}\equiv 1 in (147)5, after integration by parts in the second term on the left hand side and using 𝐯hn+1Vh,div\mathbf{v}_{h}^{n+1}\in V_{h,\text{div}}, we have that

Ωϕhn+1=Ωϕhn.\int_{\Omega}\phi_{h}^{n+1}=\int_{\Omega}\phi_{h}^{n}.

Hence, the variables zhn+1z_{h}^{n+1} and yhn+1y_{h}^{n+1} belong to the space

Sh,0:={qhSh:Ωqh=0}.S_{h,0}:=\{q_{h}\in S_{h}:\int_{\Omega}q_{h}=0\}.

We then require equations (147)5 and (147)6 to be valid only for test functions ξh,χhSh,0\xi_{h},\chi_{h}\in S_{h,0}, substituting ϕhn+1=zhn+1+ϕhn\phi_{h}^{n+1}=z_{h}^{n+1}+\phi_{h}^{n} and μhn+1=yhn+1+1|Ω|Ωμhn+1\mu_{h}^{n+1}=y_{h}^{n+1}+\frac{1}{|\Omega|}\int_{\Omega}\mu_{h}^{n+1} in their expressions. We further substitute 𝐅hn+1=𝐆hn+1+𝐅hn\mathbf{F}_{h}^{n+1}=\mathbf{G}_{h}^{n+1}+\mathbf{F}_{h}^{n} in (147)4 and (147)5 and we add to (147)4 a regularizing term αΩ𝐆hn+1:𝚷h\alpha\int_{\Omega}\mathbf{G}_{h}^{n+1}\colon\boldsymbol{\Pi}_{h}, with α>0\alpha>0. The latter term is introduced to be able to control, in the forthcoming stability estimates, the mass of 𝐆hn+1\mathbf{G}_{h}^{n+1}, which is not equal to zero. In a second step we will study the limit problem α0\alpha\to 0. The reduced system which we are going to study for the time being is thus the following: given ϕhnSh\phi_{h}^{n}\in S_{h}, 𝐅hnXh\mathbf{F}_{h}^{n}\in X_{h}, find (𝐯h,αn+1,𝐆h,αn+1,𝐌h,αn+1,zh,αn+1,yh,αn+1)Vh,div×Xh×Xh×Sh,0×Sh,0(\mathbf{v}_{h,\alpha}^{n+1},\mathbf{G}_{h,\alpha}^{n+1},\mathbf{M}_{h,\alpha}^{n+1},z_{h,\alpha}^{n+1},y_{h,\alpha}^{n+1})\in V_{h,\text{div}}\times X_{h}\times X_{h}\times S_{h,0}\times S_{h,0} such that, for all (𝐮h,𝚯h,𝚷h,ξh,χh)Vh,div×Xh×Xh×Sh,0×Sh,0(\mathbf{u}_{h},\boldsymbol{\Theta}_{h},\boldsymbol{\Pi}_{h},\xi_{h},\chi_{h})\in V_{h,\text{div}}\times X_{h}\times X_{h}\times S_{h,0}\times S_{h,0},

{νΩ𝐯h,αn+1:𝐮h=Ωyh,αn+1ϕhn𝐮hΩ(𝐌h,αn+1(𝐅hn)T):𝐮h+Ω(𝐅hn𝐌h,αn+1)𝐮h,Ω𝐆h,αn+1:𝚯h+ΔtΩ(𝐯h,αn+1)𝐅hn:𝚯hΔtΩ(𝐯h,αn+1)𝐅hn:𝚯h+ΔtγΩ𝐌h,αn+1:𝚯h=0,Ω𝐌h,αn+1:𝚷h=αΩ𝐆h,αn+1:𝚷h+Ω(h+(𝐆h,αn+1+𝐅hn)+h(𝐅hn)):𝚷h+Ω[f(ϕhn)]+(g+(𝐆h,αn+1+𝐅hn)+g(𝐅hn)):𝚷h+Ω(f(ϕhn)[f(ϕhn)]+)(g+(𝐅hn)+g(𝐆h,αn+1+𝐅hn)):𝚷h+λΩ(𝐆h,αn+1+𝐅hn)\setstackgapS0.4ex\Shortstack𝚷h,Ωzh,αn+1ξh+ΔtΩ(𝐯h,αn+1ϕhn)ξh+ΔtΩb(ϕhn)yh,αn+1ξh=0,Ωyh,αn+1χh=Ω(zh,αn+1+ϕhn)χh+Ω(ψ+(zh,αn+1+ϕhn)+ψ(ϕhn))χh+Ω(f+(zh,αn+1+ϕhn)g(𝐆h,αn+1+𝐅hn)+f(ϕhn)g(𝐆h,αn+1+𝐅hn))χh,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h,\alpha}^{n+1}\colon\nabla\mathbf{u}_{h}=\int_{\Omega}y_{h,\alpha}^{n+1}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}-\int_{\Omega}\left(\mathbf{M}_{h,\alpha}^{n+1}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}\\ \displaystyle+\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h,\alpha}^{n+1}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}\mathbf{G}_{h,\alpha}^{n+1}\colon\boldsymbol{\Theta}_{h}+\Delta t\int_{\Omega}\left({\mathbf{v}_{h,\alpha}^{n+1}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h,\alpha}^{n+1}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h,\alpha}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mathbf{M}_{h,\alpha}^{n+1}\colon\boldsymbol{\Pi}_{h}=\alpha\int_{\Omega}\mathbf{G}_{h,\alpha}^{n+1}\colon\boldsymbol{\Pi}_{h}\\ +\int_{\Omega}\left(h_{+}^{\prime}(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n})+h_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}+\int_{\Omega}[f(\phi_{h}^{n})]_{+}\left(g_{+}^{\prime}(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}\\ \displaystyle+\int_{\Omega}(f(\phi_{h}^{n})-[f(\phi_{h}^{n})]_{+})\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\left(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n}\right)\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}z_{h,\alpha}^{n+1}\xi_{h}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h,\alpha}^{n+1}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla y_{h,\alpha}^{n+1}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}y_{h,\alpha}^{n+1}\chi_{h}=\int_{\Omega}\nabla\left(z_{h,\alpha}^{n+1}+\phi_{h}^{n}\right)\cdot\nabla\chi_{h}+\int_{\Omega}\left(\psi_{+}^{\prime}\left(z_{h,\alpha}^{n+1}+\phi_{h}^{n}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)\chi_{h}\\ \displaystyle+\int_{\Omega}\left(f_{+}^{\prime}(z_{h,\alpha}^{n+1}+\phi_{h}^{n})g(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{G}_{h,\alpha}^{n+1}+\mathbf{F}_{h}^{n})\right)\chi_{h},\end{cases} (151)

where the subscript α\alpha indicates the dependence of the solutions of (151) on the parameter α\alpha. We now introduce the finite dimensional Hilbert space

X:=Vh,div×Xh×Xh×Sh,0×Sh,0,X:=V_{h,\text{div}}\times X_{h}\times X_{h}\times S_{h,0}\times S_{h,0},

endowed with the inner product

((𝐯1,𝐆1,𝐌1,z1,y1),(𝐯2,𝐆2,𝐌2,z2,y2))X\displaystyle((\mathbf{v}_{1},\mathbf{G}_{1},\mathbf{M}_{1},z_{1},y_{1}),(\mathbf{v}_{2},\mathbf{G}_{2},\mathbf{M}_{2},z_{2},y_{2}))_{X}
:=Ω𝐯1:𝐯2+Ω𝐆1:𝐆2+Ω𝐌1:𝐌2+Ωz1z2+Ωy1y2,\displaystyle:=\int_{\Omega}\nabla\mathbf{v}_{1}\colon\nabla\mathbf{v}_{2}+\int_{\Omega}\mathbf{G}_{1}\colon\mathbf{G}_{2}+\int_{\Omega}\mathbf{M}_{1}\colon\mathbf{M}_{2}+\int_{\Omega}\nabla z_{1}\cdot\nabla z_{2}+\int_{\Omega}\nabla y_{1}\cdot\nabla y_{2},

and the corresponding induced norm ||||X2:=(,)X||\cdot||_{X}^{2}:=(\cdot,\cdot)_{X}.
We define the continuous map 𝒫:XX\mathcal{P}:X\rightarrow X which associates to (𝐯,𝐆,𝐌,z,y)X(\mathbf{v},\mathbf{G},\mathbf{M},z,y)\in X an element 𝒫(X)X\mathcal{P}(X)\in X such that, for all (𝐯¯,𝐆¯,𝐌¯,z¯,y¯)X(\bar{\mathbf{v}},\bar{\mathbf{G}},\bar{\mathbf{M}},\bar{z},\bar{y})\in X,

(𝒫(𝐯,𝐆,𝐌,z,y),(𝐯¯,𝐆¯,𝐌¯,z¯,y¯))X=ΔtνΩ𝐯:𝐯¯ΔtΩyϕhn𝐯¯+ΔtΩ(𝐌(𝐅hn)T):𝐯¯\displaystyle\displaystyle(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\bar{\mathbf{v}},\bar{\mathbf{G}},\bar{\mathbf{M}},\bar{z},\bar{y}))_{X}=\Delta t\nu\int_{\Omega}\nabla\mathbf{v}\colon\nabla\bar{\mathbf{v}}-\Delta t\int_{\Omega}y\nabla\phi_{h}^{n}\cdot\bar{\mathbf{v}}+\Delta t\int_{\Omega}\left(\mathbf{M}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\bar{\mathbf{v}} (152)
ΔtΩ(𝐅hn)T𝐌𝐯¯+Ω𝐆:𝐌¯+ΔtΩ(𝐯)𝐅hn:𝐌¯ΔtΩ(𝐯)𝐅hn:𝐌¯\displaystyle\displaystyle-\Delta t\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\right)^{T}\mathbf{M}\cdot\bar{\mathbf{v}}+\int_{\Omega}\mathbf{G}\colon\bar{\mathbf{M}}+\Delta t\int_{\Omega}\left({\mathbf{v}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\bar{\mathbf{M}}-\Delta t\int_{\Omega}\left(\nabla\mathbf{v}\right)\mathbf{F}_{h}^{n}\colon\bar{\mathbf{M}}
+ΔtγΩ𝐌:𝐌¯Ω𝐌:𝐆¯+αΩ𝐆:𝐆¯+Ω(h+(𝐆+𝐅hn)+h(𝐅hn)):𝐆¯\displaystyle\displaystyle+\Delta t\gamma\int_{\Omega}\mathbf{M}\colon\bar{\mathbf{M}}-\int_{\Omega}\mathbf{M}\colon\bar{\mathbf{G}}+\alpha\int_{\Omega}\mathbf{G}\colon\bar{\mathbf{G}}+\int_{\Omega}\left(h_{+}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})+h_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\bar{\mathbf{G}}
+Ω[f(ϕhn)]+(g+(𝐆+𝐅hn)+g(𝐅hn)):𝐆¯+Ω(f(ϕhn)[f(ϕhn)]+)(g+(𝐅hn)+g(𝐆+𝐅hn)):𝐆¯\displaystyle\displaystyle+\int_{\Omega}[f(\phi_{h}^{n})]_{+}\left(g_{+}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\bar{\mathbf{G}}+\int_{\Omega}(f(\phi_{h}^{n})-[f(\phi_{h}^{n})]_{+})\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})\right)\colon\bar{\mathbf{G}}
+λΩ(𝐆+𝐅hn)\setstackgapS0.4ex\Shortstack𝐆¯+Ωzy¯+ΔtΩ(𝐯ϕhn)y¯+ΔtΩb(ϕhn)yy¯\displaystyle\displaystyle+\lambda\int_{\Omega}\nabla\left(\mathbf{G}+\mathbf{F}_{h}^{n}\right)\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\bar{\mathbf{G}}+\int_{\Omega}z\,\bar{y}+\Delta t\int_{\Omega}\left(\mathbf{v}\cdot\nabla\phi_{h}^{n}\right)\bar{y}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla y\cdot\nabla\bar{y}
Ωyz¯+Ω(z+ϕhn)z¯+Ω(ψ+(z+ϕhn)+ψ(ϕhn))z¯\displaystyle\displaystyle-\int_{\Omega}y\bar{z}+\int_{\Omega}\nabla\left(z+\phi_{h}^{n}\right)\cdot\nabla\bar{z}+\int_{\Omega}\left(\psi_{+}^{\prime}\left(z+\phi_{h}^{n}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)\bar{z}
+Ω(f+(z+ϕhn)g(𝐆+𝐅hn)+f(ϕhn)g(𝐆+𝐅hn))z¯.\displaystyle\displaystyle+\int_{\Omega}\left(f_{+}^{\prime}(z+\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})\right)\bar{z}.

A zero of the map 𝒫\mathcal{P}, if it exists, is a solution to System (151). In the following we are going to prove that, if (𝐯,𝐆,𝐌,z,y)X=R>0||(\mathbf{v},\mathbf{G},\mathbf{M},z,y)||_{X}=R>0 is large enough, then (𝒫(𝐯,𝐆,𝐌,z,y),(𝐯,𝐆,𝐌,z,y))X>0(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\mathbf{v},\mathbf{G},\mathbf{M},z,y))_{X}>0. Then, [23, Lemma II.1.4] implies that 𝒫\mathcal{P} admits a root (𝐯,𝐆,𝐌,z,y)X(\mathbf{v}^{*},\mathbf{G}^{*},\mathbf{M}^{*},z^{*},y^{*})\in X, with (𝐯,𝐆,𝐌,z,y)XR||(\mathbf{v}^{*},\mathbf{G}^{*},\mathbf{M}^{*},z^{*},y^{*})||_{X}\leq R. Indeed, we have that

(𝒫(𝐯,𝐆,𝐌,z,y),(𝐯,𝐆,𝐌,z,y))X=ΔtνΩ𝐯:𝐯+ΔtγΩ𝐌:𝐌+αΩ𝐆:𝐆\displaystyle\displaystyle(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\mathbf{v},\mathbf{G},\mathbf{M},z,y))_{X}=\Delta t\nu\int_{\Omega}\nabla\mathbf{v}\colon\nabla{\mathbf{v}}+\Delta t\gamma\int_{\Omega}\mathbf{M}\colon{\mathbf{M}}+\alpha\int_{\Omega}\mathbf{G}\colon{\mathbf{G}}
+Ω(h+(𝐆+𝐅hn)+h(𝐅hn)):(𝐆+𝐅hn𝐅hn)\displaystyle\displaystyle+\int_{\Omega}\left(h_{+}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})+h_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\left({\mathbf{G}}+\mathbf{F}_{h}^{n}-\mathbf{F}_{h}^{n}\right)
+Ω[f(ϕhn)]+(g+(𝐆+𝐅hn)+g(𝐅hn)):(𝐆+𝐅hn𝐅hn)\displaystyle\displaystyle+\int_{\Omega}[f(\phi_{h}^{n})]_{+}\left(g_{+}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\left({\mathbf{G}}+\mathbf{F}_{h}^{n}-\mathbf{F}_{h}^{n}\right)
+Ω(f(ϕhn)[f(ϕhn)]+)(g+(𝐅hn)+g(𝐆+𝐅hn)):(𝐆+𝐅hn𝐅hn)\displaystyle\displaystyle+\int_{\Omega}(f(\phi_{h}^{n})-[f(\phi_{h}^{n})]_{+})\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{G}+\mathbf{F}_{h}^{n})\right)\colon\left({\mathbf{G}}+\mathbf{F}_{h}^{n}-\mathbf{F}_{h}^{n}\right)
+λΩ(𝐆+𝐅hn)\setstackgapS0.4ex\Shortstack𝐆+ΔtΩb(ϕhn)yy+Ω(z+ϕhn)z\displaystyle\displaystyle+\lambda\int_{\Omega}\nabla\left(\mathbf{G}+\mathbf{F}_{h}^{n}\right)\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla{\mathbf{G}}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla y\cdot\nabla{y}+\int_{\Omega}\nabla\left(z+\phi_{h}^{n}\right)\cdot\nabla{z}
+Ω(ψ+(z+ϕhn)+ψ(ϕhn))(z+ϕhnϕhn)\displaystyle\displaystyle+\int_{\Omega}\left(\psi_{+}^{\prime}\left(z+\phi_{h}^{n}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)(z+\phi_{h}^{n}-\phi_{h}^{n})
+Ω(f+(z+ϕhn)g(𝐆+𝐅hn)+f(ϕhn)g(𝐆+𝐅hn))(z+ϕhnϕhn).\displaystyle\displaystyle+\int_{\Omega}\left(f_{+}^{\prime}(z+\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})\right)\left(z+\phi_{h}^{n}-\phi_{h}^{n}\right).

Given the facts that f+,g+,ψ+f_{+},g_{+},\psi_{+} are convex functions and f,g,ψf_{-},g_{-},\psi_{-} are concave functions of their arguments, given the assumption of the positivity of gg, using Assumption A1 and the relation a(a+b)=12|a|2+12|a+b|212|b|2a\star(a+b)=\frac{1}{2}|a|^{2}+\frac{1}{2}|a+b|^{2}-\frac{1}{2}|b|^{2} (with a,ba,b any scalar, vector or matrix elements with the corresponding scalar product indicated by the symbol \star), we get that

Δtν𝐯L2(Ω;3×3)2+Δtγ𝐌L2(Ω;3×3)2+α𝐆L2(Ω;3×3)2\displaystyle\displaystyle\Delta t\nu||\nabla\mathbf{v}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\mathbf{M}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\alpha||\mathbf{G}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2} (153)
+Ω(h(𝐆+𝐅hn)+f(ϕhn)g(𝐆+𝐅hn)+f(z+ϕhn)g(𝐆+𝐅hn)+ψ(z+ϕhn))\displaystyle\displaystyle+\int_{\Omega}\left(h(\mathbf{G}+\mathbf{F}_{h}^{n})+f(\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})+f(z+\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})+\psi\left(z+\phi_{h}^{n}\right)\right)
+λ2𝐆L2(Ω;3×3×3)2+Δtb0yL2(Ω;3)2+12zL2(Ω;3)2(𝒫(𝐯,𝐆,𝐌,z,y),(𝐯,𝐆,𝐌,z,y))X\displaystyle\displaystyle+\frac{\lambda}{2}||\nabla\mathbf{G}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\Delta tb_{0}||\nabla y||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{1}{2}||\nabla z||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}\leq(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\mathbf{v},\mathbf{G},\mathbf{M},z,y))_{X}
+Ω(h(𝐅hn)+f(ϕhn)g(𝐅hn)+f(ϕhn)g(𝐆+𝐅hn)+ψ(ϕhn))+λ2𝐅hnL2(Ω;3×3×3)2+12ϕhnL2(Ω;3)2.\displaystyle\displaystyle+\int_{\Omega}\left(h(\mathbf{F}_{h}^{n})+f(\phi_{h}^{n})g(\mathbf{F}_{h}^{n})+f(\phi_{h}^{n})g(\mathbf{G}+\mathbf{F}_{h}^{n})+\psi\left(\phi_{h}^{n}\right)\right)+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\frac{1}{2}||\nabla\phi_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}.

Now, using Assumptions 𝐀𝟐𝐡\bf{A2_{h}} and A3, we get that

Δtν𝐯L2(Ω;3×3)2+Δtγ𝐌L2(Ω;3×3)2+α𝐆L2(Ω;3×3)2+λ2𝐆L2(Ω;3×3×3)2\displaystyle\displaystyle\Delta t\nu||\nabla\mathbf{v}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\mathbf{M}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\alpha||\mathbf{G}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{G}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (154)
+Δtb0yL2(Ω;3)2+12zL2(Ω;3)2C(𝐅hn,ϕhn)(𝒫(𝐯,𝐆,𝐌,z,y),(𝐯,𝐆,𝐌,z,y))X,\displaystyle\displaystyle+\Delta tb_{0}||\nabla y||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{1}{2}||\nabla z||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}-C\left(\mathbf{F}_{h}^{n},\phi_{h}^{n}\right)\leq(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\mathbf{v},\mathbf{G},\mathbf{M},z,y))_{X},

where C(𝐅hn,ϕhn)C\left(\mathbf{F}_{h}^{n},\phi_{h}^{n}\right) is a positive constant which depends on 𝐅hn\mathbf{F}_{h}^{n} and ϕhn\phi_{h}^{n}. Hence,

(𝒫(𝐯,𝐆,𝐌,z,y),(𝐯,𝐆,𝐌,z,y))XC(𝐯,𝐆,𝐌,z,y)XC(𝐅hn,ϕhn)>0(\mathcal{P}(\mathbf{v},\mathbf{G},\mathbf{M},z,y),(\mathbf{v},\mathbf{G},\mathbf{M},z,y))_{X}\geq C||(\mathbf{v},\mathbf{G},\mathbf{M},z,y)||_{X}-C\left(\mathbf{F}_{h}^{n},\phi_{h}^{n}\right)>0

if (𝐯,𝐆,𝐌,z,y)XR||(\mathbf{v},\mathbf{G},\mathbf{M},z,y)||_{X}\geq R and RR is large enough, and [23, Lemma II.1.4] implies that 𝒫\mathcal{P} admits a root (𝐯,𝐆,𝐌,z,y)X(\mathbf{v}^{*},\mathbf{G}^{*},\mathbf{M}^{*},z^{*},y^{*})\in X, which is a solution of (151).

With the aim of recovering a solution for the original System (147), we take the limit in (151) as α0\alpha\to 0. We thus need to obtain a-priori estimates for system (151) which are uniform in the parameter α\alpha. Let us take 𝐮h=Δt𝐯h,αn+1\mathbf{u}_{h}=\Delta t\mathbf{v}_{h,\alpha}^{n+1}, 𝚯h=𝐌h,αn+1\boldsymbol{\Theta}_{h}=\mathbf{M}_{h,\alpha}^{n+1}, 𝚷h=𝐆h,αn+1\boldsymbol{\Pi}_{h}=-\mathbf{G}_{h,\alpha}^{n+1}, ξh=yh,αn+1\xi_{h}=y_{h,\alpha}^{n+1} and χh=zh,αn+1\chi_{h}=-z_{h,\alpha}^{n+1} in (151). Summing all the equations, using again the identity a(a+b)=12|a|2+12|a+b|212|b|2a\star(a+b)=\frac{1}{2}|a|^{2}+\frac{1}{2}|a+b|^{2}-\frac{1}{2}|b|^{2} (with a,ba,b any scalar, vector or matrix elements with the corresponding scalar product indicated by the symbol \star), the convexity of the functions f+,g+,ψ+f_{+},g_{+},\psi_{+} and the concavity of the functions f,g,ψf_{-},g_{-},\psi_{-}, given also the assumption of the positivity of gg, we get, similarly to (153), that

Δtν𝐯h,αn+1L2(Ω;3×3)2+Δtγ𝐌h,αn+1L2(Ω;3×3)2+α𝐆h,αn+1L2(Ω;3×3)2+λ2𝐆h,αn+1L2(Ω;3×3×3)2\displaystyle\displaystyle\Delta t\nu||\nabla\mathbf{v}_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\mathbf{M}_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\alpha||\mathbf{G}_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{G}_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (155)
+Δtb0yh,αn+1L2(Ω;3)2+12zh,αn+1L2(Ω;3)2C(ϕhn,𝐅hn),\displaystyle\displaystyle+\Delta tb_{0}||\nabla y_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{1}{2}||\nabla z_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}\leq C(\phi_{h}^{n},\mathbf{F}_{h}^{n}),

uniformly in α\alpha. Moreover, similarly to (4.2), taking 𝚯h=𝐞l,r\boldsymbol{\Theta}_{h}=\mathbf{e}_{l,r}, l,r=1,2,3l,r=1,2,3, in (151)2 and using the fact that 𝐯h,αn+1Vh,div\mathbf{v}_{h,\alpha}^{n+1}\in V_{h,\text{div}} and (155), we obtain that

|Ω𝐆h,αlrn+1|=|ΔtγΩ𝐌h,αlrn+1|C𝐌h,αn+1L2(Ω;3×3)C(ϕhn,𝐅hn),\left|\int_{\Omega}\mathbf{G}_{h,\alpha lr}^{n+1}\right|=\left|\Delta t\gamma\int_{\Omega}\mathbf{M}_{h,\alpha lr}^{n+1}\right|\leq C||\mathbf{M}_{h,\alpha}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}\leq C(\phi_{h}^{n},\mathbf{F}_{h}^{n}), (156)

uniformly in α\alpha. From (155) and (156) we conclude, thanks to the Poincaré–Wirtinger inequality, that

𝐆h,αn+1H1(Ω;3×3)2C(ϕhn,𝐅hn),||\mathbf{G}_{h,\alpha}^{n+1}||_{H^{1}(\Omega;\mathbb{R}^{3\times 3})}^{2}\leq C(\phi_{h}^{n},\mathbf{F}_{h}^{n}), (157)

uniformly in α\alpha. Then, thanks to the Bolzano–Weierstrass theorem and (155)-(157), given any sequence α0\alpha\to 0, we can identify a subsequence α0\alpha^{\prime}\to 0 and a limit point

(𝐯h,αn+1,𝐆h,αn+1,𝐌h,αn+1,zh,αn+1,yh,αn+1)(𝐯hn+1,𝐆hn+1,𝐌hn+1,zhn+1,yhn+1)(\mathbf{v}_{h,\alpha^{\prime}}^{n+1},\mathbf{G}_{h,\alpha^{\prime}}^{n+1},\mathbf{M}_{h,\alpha^{\prime}}^{n+1},z_{h,\alpha^{\prime}}^{n+1},y_{h,\alpha^{\prime}}^{n+1})\to(\mathbf{v}_{h}^{n+1},\mathbf{G}_{h}^{n+1},\mathbf{M}_{h}^{n+1},z_{h}^{n+1},y_{h}^{n+1})

in XX which satisfies (151) as α0\alpha^{\prime}\to 0.

Finally, we define

(𝐯hn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1):=(𝐯hn+1,𝐆hn+1+𝐅hn,𝐌hn+1,zhn+1+ϕhn,yhn+1+ρ¯),(\mathbf{v}_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1}):=(\mathbf{v}_{h}^{n+1},\mathbf{G}_{h}^{n+1}+\mathbf{F}_{h}^{n},\mathbf{M}_{h}^{n+1},z_{h}^{n+1}+\phi_{h}^{n},y_{h}^{n+1}+\bar{\rho}),

with

ρ¯:=1|Ω|(Ω(ψ+(ϕhn+1)+ψ(ϕhn))+Ω(f+(ϕhn+1)g(𝐅hn+1)+f(ϕhn)g(𝐅hn+1))).\bar{\rho}:=\frac{1}{|\Omega|}\left(\int_{\Omega}\left(\psi_{+}^{\prime}\left(\phi_{h}^{n+1}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)+\int_{\Omega}\left(f_{+}^{\prime}(\phi_{h}^{n+1})g(\mathbf{F}_{h}^{n+1})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{F}_{h}^{n+1})\right)\right).

Observing moreover that (151)1 defines a linear functional from VhV_{h} to \mathbb{R} which vanishes on Vh,divV_{h,\text{div}}, we conclude by standard arguments (see e.g. [14, Section 1.2, Chapter III]) that there exists a unique shn+1Rhs_{h}^{n+1}\in R_{h} such that

(𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1)Vh×Rh×Xh×Xh×Sh×Xh(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times X_{h}

is a solution of System (147).

Taking 𝐮h=Δt𝐯hn+1\mathbf{u}_{h}=\Delta t\mathbf{v}_{h}^{n+1}, ph=Δtshn+1p_{h}=\Delta ts_{h}^{n+1}, 𝚯h=𝐌hn+1\boldsymbol{\Theta}_{h}=\mathbf{M}_{h}^{n+1}, 𝚷h=(𝐅hn+1𝐅hn)\boldsymbol{\Pi}_{h}=-(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}), ξh=μhn+1\xi_{h}=\mu_{h}^{n+1} and χh=(ϕhn+1ϕhn)\chi_{h}=-(\phi_{h}^{n+1}-\phi_{h}^{n}) in (147), with similar calculations as in (155) we obtain that

Ω(f(ϕhn+1)g(𝐅hn+1)+ψ(ϕhn+1)+h(𝐅hn+1))+12ϕhn+1L2(Ω;3)2+λ2𝐅hn+1L2(Ω;3×3×3)2\displaystyle\displaystyle\int_{\Omega}\left(f(\phi_{h}^{n+1})g(\mathbf{F}_{h}^{n+1})+\psi\left(\phi_{h}^{n+1}\right)+h\left(\mathbf{F}_{h}^{n+1}\right)\right)+\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (158)
+12(ϕhn+1ϕhn)L2(Ω;3)2+λ2(𝐅hn+1𝐅hn)L2(Ω;3×3×3)2\displaystyle\displaystyle+\frac{1}{2}\left|\left|\nabla\left(\phi_{h}^{n+1}-\phi_{h}^{n}\right)\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}\left|\left|\nabla\left(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}\right)\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}
+Δtν𝐯hn+1L2(Ω;3×3)2+Δtγ𝐌hn+1L2(Ω;3×3)2+ΔtΩb(ϕhn)μhn+1μhn+1\displaystyle\displaystyle+\Delta t\nu||\nabla\mathbf{v}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\mathbf{M}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\mu_{h}^{n+1}
Ω(f(ϕhn)g(𝐅hn)+ψ(ϕhn)+h(𝐅hn))+12ϕhnL2(Ω;3)2+λ2𝐅hnL2(Ω;3×3×3)2,\displaystyle\displaystyle\leq\int_{\Omega}\left(f(\phi_{h}^{n})g(\mathbf{F}_{h}^{n})+\psi\left(\phi_{h}^{n}\right)+h\left(\mathbf{F}_{h}^{n}\right)\right)+\frac{1}{2}||\nabla\phi_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2},

which gives that (150) is a decreasing (Lyapunov) function for the discrete solutions. Summing (158) from n=0mn=0\to m, for m=0N1m=0\to N-1, and using Assumptions A2,A3, we finally get (149). \Box

Remark 5.5

We observe that System (147) admits a solution and is unconditionally gradient stable for any value of Δt>0\Delta t>0, with no requirements on the smallness of Δt\Delta t (with respect to hh and the model parameters).

Remark 5.6

System (147) is fully coupled and nonlinear, and can be solved e.g. by means of a Newton method, which at each iteration requires the solution of the full tangent algebraic system. In order to decrease the computational demand for the solution of System (147), we practically solve a fixed-point iteration scheme which decouples (147)1-(147)2, (147)3-(147)4 and (147)5-(147)6, i.e. we solve the following problem:

Problem 𝐏𝐡𝐈𝐃\mathbf{P_{h}^{ID}}: for n=0,,N1n=0,\dots,N-1, given ϕhnSh\phi_{h}^{n}\in S_{h}, 𝐅hnXh\mathbf{F}_{h}^{n}\in X_{h}, and for k=0,,K1k=0,\dots,K-1, given ϕhn+1,0=ϕhn\phi_{h}^{n+1,0}=\phi_{h}^{n}, 𝐅hn+1,0=𝐅hn\mathbf{F}_{h}^{n+1,0}=\mathbf{F}_{h}^{n}, μhn+1,0=μhn\mu_{h}^{n+1,0}=\mu_{h}^{n}, 𝐌hn+1,0=𝐌hn\mathbf{M}_{h}^{n+1,0}=\mathbf{M}_{h}^{n}, find

(𝐯hn+1,k+1,shn+1,k+1,𝐅hn+1,k+1,𝐌hn+1,k+1,ϕhn+1,k+1,μhn+1,k+1)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{v}_{h}^{n+1,k+1},s_{h}^{n+1,k+1},\mathbf{F}_{h}^{n+1,k+1},\mathbf{M}_{h}^{n+1,k+1},\phi_{h}^{n+1,k+1},\mu_{h}^{n+1,k+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h}

such that, for all (𝐮h,ph,𝚯h,𝚷h,ξh,χh)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{u}_{h},p_{h},\boldsymbol{\Theta}_{h},\boldsymbol{\Pi}_{h},\xi_{h},\chi_{h})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h},

{νΩ𝐯hn+1,k+1:𝐮hΩshn+1,k+1div𝐮h=Ωμhn+1,kϕhn𝐮hΩ(𝐌hn+1,k(𝐅hn)T):𝐮h+Ω(𝐅hn𝐌hn+1,k)𝐮h,Ωphdiv𝐯hn+1,k+1=0,Ω(𝐅hn+1,k+1𝐅hn):𝚯h+ΔtΩ(𝐯hn+1,k+1)𝐅hn:𝚯hΔtΩ(𝐯hn+1,k+1)𝐅hn:𝚯h+ΔtγΩ𝐌hn+1,k+1:𝚯h=0,Ω𝐌hn+1,k+1:𝚷h=Ω(h+(𝐅hn+1,k+1)+h(𝐅hn)):𝚷h+Ω[f(ϕhn)]+(g+(𝐅hn+1,k+1)+g(𝐅hn)):𝚷h+Ω(f(ϕhn)[f(ϕhn)]+)(g+(𝐅hn)+g(𝐅hn+1,k+1)):𝚷h+λΩ𝐅hn+1,k+1\setstackgapS0.4ex\Shortstack𝚷h,Ω(ϕhn+1,k+1ϕhn)ξh+ΔtΩ(𝐯hn+1,k+1ϕhn)ξh+ΔtΩb(ϕhn)μhn+1,k+1ξh=0,Ωμhn+1,k+1χh=Ωϕhn+1,k+1χh+Ω(ψ+(ϕhn+1,k+1)+ψ(ϕhn))χh+Ω(f+(ϕhn+1,k+1)g(𝐅hn+1,k+1)+f(ϕhn)g(𝐅hn+1,k+1))χh,\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h}^{n+1,k+1}\colon\nabla\mathbf{u}_{h}-\int_{\Omega}s_{h}^{n+1,k+1}\operatorname{div}\mathbf{u}_{h}=\int_{\Omega}\mu_{h}^{n+1,k}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}-\int_{\Omega}\left(\mathbf{M}_{h}^{n+1,k}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}\\ \displaystyle+\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n+1,k}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}p_{h}\operatorname{div}\mathbf{v}_{h}^{n+1,k+1}=0,\\ \\ \displaystyle\int_{\Omega}\left(\mathbf{F}_{h}^{n+1,k+1}-\mathbf{F}_{h}^{n}\right)\colon\boldsymbol{\Theta}_{h}+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n+1,k+1}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n+1,k+1}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h}^{n+1,k+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mathbf{M}_{h}^{n+1,k+1}\colon\boldsymbol{\Pi}_{h}=\int_{\Omega}\left(h_{+}^{\prime}(\mathbf{F}_{h}^{n+1,k+1})+h_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}\\ \displaystyle+\int_{\Omega}[f(\phi_{h}^{n})]_{+}\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n+1,k+1})+g_{-}^{\prime}(\mathbf{F}_{h}^{n})\right)\colon\boldsymbol{\Pi}_{h}\\ \displaystyle+\int_{\Omega}(f(\phi_{h}^{n})-[f(\phi_{h}^{n})]_{+})\left(g_{+}^{\prime}(\mathbf{F}_{h}^{n})+g_{-}^{\prime}(\mathbf{F}_{h}^{n+1,k+1})\right)\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\mathbf{F}_{h}^{n+1,k+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\left(\phi_{h}^{n+1,k+1}-\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n+1,k+1}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1,k+1}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mu_{h}^{n+1,k+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h}^{n+1,k+1}\cdot\nabla\chi_{h}+\int_{\Omega}\left(\psi_{+}^{\prime}\left(\phi_{h}^{n+1,k+1}\right)+\psi_{-}^{\prime}\left(\phi_{h}^{n}\right)\right)\chi_{h}\\ \displaystyle+\int_{\Omega}\left(f_{+}^{\prime}(\phi_{h}^{n+1,k+1})g(\mathbf{F}_{h}^{n+1,k+1})+f_{-}^{\prime}(\phi_{h}^{n})g(\mathbf{F}_{h}^{n+1,k+1})\right)\chi_{h},\end{cases} (159)

and set

𝐯hn+1=𝐯hn+1,K+1,shn+1=shn+1,K+1,𝐅hn+1=𝐅hn+1,K+1,𝐌hn+1=𝐌hn+1,K+1,\displaystyle\displaystyle\mathbf{v}_{h}^{n+1}=\mathbf{v}_{h}^{n+1,K+1},\,s_{h}^{n+1}=s_{h}^{n+1,K+1},\,\mathbf{F}_{h}^{n+1}=\mathbf{F}_{h}^{n+1,K+1},\,\mathbf{M}_{h}^{n+1}=\mathbf{M}_{h}^{n+1,K+1}, (160)
ϕhn+1=ϕhn+1,K+1,μhn+1=μhn+1,K+1.\displaystyle\displaystyle\phi_{h}^{n+1}=\phi_{h}^{n+1,K+1},\,\mu_{h}^{n+1}=\mu_{h}^{n+1,K+1}.

The value of KK is chosen as the first value at which

𝐯hn+1,K+1𝐯hn+1,KL(Ω,3)+shn+1,K+1shn+1,KL(Ω)\displaystyle||\mathbf{v}_{h}^{n+1,K+1}-\mathbf{v}_{h}^{n+1,K}||_{L^{\infty}(\Omega,\mathbb{R}^{3})}+||s_{h}^{n+1,K+1}-s_{h}^{n+1,K}||_{L^{\infty}(\Omega)}
+𝐅hn+1,K+1𝐅hn+1,KL(Ω,3×3)+𝐅hn+1,K+1𝐅hn+1,KL(Ω,3×3)\displaystyle+||\mathbf{F}_{h}^{n+1,K+1}-\mathbf{F}_{h}^{n+1,K}||_{L^{\infty}(\Omega,\mathbb{R}^{3\times 3})}+||\mathbf{F}_{h}^{n+1,K+1}-\mathbf{F}_{h}^{n+1,K}||_{L^{\infty}(\Omega,\mathbb{R}^{3\times 3})}
+ϕhn+1,K+1ϕhn+1,KL(Ω)+μhn+1,K+1μhn+1,KL(Ω)<tol,\displaystyle+||\phi_{h}^{n+1,K+1}-\phi_{h}^{n+1,K}||_{L^{\infty}(\Omega)}+||\mu_{h}^{n+1,K+1}-\mu_{h}^{n+1,K}||_{L^{\infty}(\Omega)}<\text{tol},

where tol>0\text{tol}>0 is a small tolerance for the convergence of the algorithm. System (159) is decoupled and can be solved, at each iteration kk, in the following order:

  • - Step 1

    Given μhn+1,k\mu_{h}^{n+1,k} and 𝐌hn+1,k\mathbf{M}_{h}^{n+1,k}, solve (159)1-(159)2, which is a linear saddle-point problem;

  • - Step 2

    Given 𝐯hn+1,k+1\mathbf{v}_{h}^{n+1,k+1} from Step 1, solve (159)3-(159)4 by means of a Newton method;

  • - Step 3

    Given 𝐅hn+1,k+1\mathbf{F}_{h}^{n+1,k+1} from Step 2, solve (159)5-(159)6 by means of a Newton method.

System (159) defines a map :Sh×XhSh×Xh\mathcal{M}:S_{h}\times X_{h}\rightarrow S_{h}\times X_{h}, (μhn+1,k,𝐌hn+1,k)=(μhn+1,k+1,𝐌hn+1,k+1)\mathcal{M}(\mu_{h}^{n+1,k},\mathbf{M}_{h}^{n+1,k})=(\mu_{h}^{n+1,k+1},\mathbf{M}_{h}^{n+1,k+1}). A fixed point of \mathcal{M} is a solution of System (147), which exists thanks to Lemma 5.1. The convergence of the iterations (159), for k0k\geq 0, together with the uniqueness of the fixed point, could be proved under proper smallness conditions on Δt\Delta t, with respect to hh and to the model parameters, which guarantee that \mathcal{M} is a contraction.

Problems 𝐏𝐡𝐈\mathbf{P_{h}^{I}} and 𝐏𝐡𝐈𝐃\mathbf{P_{h}^{ID}} are gradient stable only if Assumption (144) is satisfied. In order to derive a more general approximation scheme, which is also decoupled as 𝐏𝐡𝐈𝐃\mathbf{P_{h}^{ID}}, we design a second approximation scheme, based on the fully decoupled scalar auxiliary variable scheme recently introduced in [16] for the Cahn–Hilliard Navier–Stokes system and generalized here to our model. The latter scheme will be unconditionally gradient stable given a general form of w(ϕ,𝐅)w(\phi,\mathbf{F}). We thus introduce the auxiliary variable

β:=Ω(j(ϕ,𝐅))+k,j(ϕ,𝐅):=ψ(ϕ)+w(ϕ,𝐅)\beta:=\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k},\quad j(\phi,\mathbf{F}):=\psi(\phi)+w(\phi,\mathbf{F})

with k>c2+d1k>c_{2}+d_{1}, so that β>0\beta>0. We rewrite system (9) as

{νΔ𝐯+s=βΩ(j(ϕ,𝐅))+kμϕ+βΩ(j(ϕ,𝐅))+kdiv(𝐌𝐅T)+βΩ(j(ϕ,𝐅))+k𝐅𝐌,div𝐯=0,𝐅t+βΩ(j(ϕ,𝐅))+k(𝐯)𝐅βΩ(j(ϕ,𝐅))+k(𝐯)𝐅+γ𝐌=0,𝐌=βΩ(j(ϕ,𝐅))+k𝐅w(ϕ,𝐅)λΔ𝐅,ϕt+βΩ(j(ϕ,𝐅))+k𝐯ϕdiv(b(ϕ)μ)=0,μ=βΩ(j(ϕ,𝐅))+kψ(ϕ)Δϕ+βΩ(j(ϕ,𝐅))+kϕw(ϕ,𝐅),βt=12Ω(j(ϕ,𝐅))+k(Ωϕj(ϕ,𝐅)ϕt+Ω𝐅w(ϕ,𝐅):𝐅t).\begin{cases}\displaystyle-\nu\Delta\mathbf{v}+\nabla s=\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\mu\nabla\phi+\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\operatorname{div}\left(\mathbf{M}\mathbf{F}^{T}\right)\\ \displaystyle+\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\nabla\mathbf{F}\odot\mathbf{M},\\ \\ \operatorname{div}\mathbf{v}=0,\\ \\ \displaystyle\frac{\partial\mathbf{F}}{\partial t}+\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\left(\mathbf{v}\cdot\nabla\right)\mathbf{F}-\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}(\nabla\mathbf{v})\mathbf{F}+\gamma\mathbf{M}=0,\\ \\ \displaystyle\mathbf{M}=\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\partial_{\mathbf{F}}w(\phi,\mathbf{F})-\lambda\Delta\mathbf{F},\\ \\ \displaystyle\frac{\partial\phi}{\partial t}+\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\mathbf{v}\cdot\nabla\phi-\operatorname{div}(b(\phi)\nabla\mu)=0,\\ \\ \displaystyle\mu=\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\psi^{\prime}(\phi)-\Delta\phi+\frac{\beta}{\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\partial_{\phi}w(\phi,\mathbf{F}),\\ \displaystyle\frac{\partial\beta}{\partial t}=\frac{1}{2\sqrt{\int_{\Omega}(j(\phi,\mathbf{F}))+k}}\left(\int_{\Omega}\partial_{\phi}j(\phi,\mathbf{F})\frac{\partial\phi}{\partial t}+\int_{\Omega}\partial_{\mathbf{F}}w(\phi,\mathbf{F})\colon\frac{\partial\mathbf{F}}{\partial t}\right).\end{cases} (161)

Setting β0=Ω(j(ϕ0,𝐅0))+k\beta^{0}=\sqrt{\int_{\Omega}(j(\phi_{0},\mathbf{F}_{0}))+k}, we then consider the following fully discretized approximation of system (161):

Problem 𝐏𝐡𝐈𝐈\mathbf{P_{h}^{II}}: for n=0,,N1n=0,\dots,N-1, given 𝐯hnVh\mathbf{v}_{h}^{n}\in V_{h}, ϕhn,μhnSh,βhn,𝐅hn,𝐌hnXh\phi_{h}^{n},\mu_{h}^{n}\in S_{h},\beta_{h}^{n}\in\mathbb{R},\mathbf{F}_{h}^{n},\mathbf{M}_{h}^{n}\in X_{h},
find (𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h} and βhn+1\beta_{h}^{n+1}\in\mathbb{R} such that, for all (𝐮h,ph,𝚯h,𝚷h,ξh,χh)Vh×Rh×Xh×Xh×Sh×Sh(\mathbf{u}_{h},p_{h},\boldsymbol{\Theta}_{h},\boldsymbol{\Pi}_{h},\xi_{h},\chi_{h})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times S_{h},

{νΩ𝐯hn+1:𝐮hΩshn+1div𝐮h=βhn+1Ω(j(ϕhn,𝐅hn))+kΩμhnϕhn𝐮hβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐌hn(𝐅hn)T):𝐮h+βhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐅hn𝐌hn)𝐮h,Ωphdiv𝐯hn+1=0,Ω(𝐅hn+1𝐅hn):𝚯h+Δtβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯hΔtβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯h+ΔtγΩ𝐌hn+1:𝚯h=0,Ω𝐌hn+1:𝚷h=βhn+1Ω(j(ϕhn,𝐅hn))+kΩ𝐅w(ϕhn,𝐅hn):𝚷h+λΩ𝐅hn+1\setstackgapS0.4ex\Shortstack𝚷h,Ω(ϕhn+1ϕhn)ξh+βhn+1Ω(j(ϕhn,𝐅hn))+kΔtΩ(𝐯hnϕhn)ξh+ΔtΩb(ϕhn)μhn+1ξh=0,Ωμhn+1χh=Ωϕhn+1χh+βhn+1Ω(j(ϕhn,𝐅hn))+kΩϕj(ϕhn,𝐅hn)χh,(βhn+1βhn)=12Ω(j(ϕhn,𝐅hn))+k[(Ωϕj(ϕhn,𝐅hn)(ϕhn+1ϕhn))+(Ω𝐅w(ϕhn,𝐅hn):(𝐅hn+1𝐅hn))+ΔtΩ(𝐯hn)𝐅hn:𝐌hn+1ΔtΩ(𝐅hn)T𝐌hn𝐯hn+1ΔtΩ(𝐯hn)𝐅hn:𝐌hn+1+ΔtΩ(𝐌hn(𝐅hn)T):𝐯hn+1+ΔtΩ(𝐯hnϕhn)μhn+1ΔtΩμhnϕhn𝐯hn+1].\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h}^{n+1}\colon\nabla\mathbf{u}_{h}-\int_{\Omega}s_{h}^{n+1}\operatorname{div}\mathbf{u}_{h}=\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}\\ \displaystyle-\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}p_{h}\operatorname{div}\mathbf{v}_{h}^{n+1}=0,\\ \\ \displaystyle\int_{\Omega}\left(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}\right)\colon\boldsymbol{\Theta}_{h}+\Delta t\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle-\Delta t\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Pi}_{h}=\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\mathbf{F}_{h}^{n+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\left(\phi_{h}^{n+1}-\phi_{h}^{n}\right)\xi_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mu_{h}^{n+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h}^{n+1}\cdot\nabla\chi_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\chi_{h},\\ \\ \displaystyle\left(\beta_{h}^{n+1}-\beta_{h}^{n}\right)=\frac{1}{2\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\biggl{[}\left(\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})(\phi_{h}^{n+1}-\phi_{h}^{n})\right)\\ \displaystyle+\left(\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n})\right)+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h}^{n+1}-\Delta t\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\right)^{T}\mathbf{M}_{h}^{n}\cdot\mathbf{v}_{h}^{n+1}\\ \displaystyle-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h}^{n+1}+\Delta t\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{v}_{h}^{n+1}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\mu_{h}^{n+1}\\ \displaystyle-\Delta t\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{v}_{h}^{n+1}\biggr{]}.\end{cases} (162)

We observe that (162) is a linear coupled system. The existence and uniqueness of a solution to (162) and its gradient stability are given in the following lemma.

Lemma 5.2

For all n=0,,N1n=0,\dots,N-1, given 𝐯hnVh\mathbf{v}_{h}^{n}\in V_{h}, ϕhn,μhnSh\phi_{h}^{n},\,\mu_{h}^{n}\in S_{h}, 𝐅hn,𝐌hnXh\mathbf{F}_{h}^{n},\,\mathbf{M}_{h}^{n}\in X_{h}, βhn+\beta_{h}^{n}\in\mathbb{R}_{+}, with ϕh0=PhS(ϕ0)\phi_{h}^{0}=P_{h}^{S}(\phi_{0}), 𝐅h0=PhX(𝐅0)\mathbf{F}_{h}^{0}=P_{h}^{X}(\mathbf{F}_{0}), βh0=Ω(j(ϕh0,𝐅h0))+k\beta_{h}^{0}=\sqrt{\int_{\Omega}(j(\phi_{h}^{0},\mathbf{F}_{h}^{0}))+k}, 𝐯h0=𝟎\mathbf{v}_{h}^{0}=\boldsymbol{0}, μh0=0\mu_{h}^{0}=0, 𝐌h0=𝟎\mathbf{M}_{h}^{0}=\boldsymbol{0}, there exists a unique solution (𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1,βhn+1)(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1},\beta_{h}^{n+1}) of system (162), which satisfies the following stability bound:

maxn=0N1(12ϕhn+1L2(Ω;3)2+λ2𝐅hn+1L2(Ω;3×3×3)2+(βhn+1)2)\displaystyle\max_{n=0\to N-1}\left(\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\left(\beta_{h}^{n+1}\right)^{2}\right) (163)
+(Δt)22n=0N1(ϕhn+1ϕhn)(Δt)2L2(Ω;3)2+λ(Δt)22n=0N1(𝐅hn+1𝐅hn)(Δt)2L2(Ω;3×3×3)2\displaystyle+\frac{(\Delta t)^{2}}{2}\sum_{n=0}^{N-1}\left|\left|\frac{\nabla(\phi_{h}^{n+1}-\phi_{h}^{n})}{(\Delta t)^{2}}\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda(\Delta t)^{2}}{2}\sum_{n=0}^{N-1}\left|\left|\frac{\nabla(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n})}{(\Delta t)^{2}}\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}
+(Δt)2n=0N1(βhn+1βhn)(Δt)2L2(Ω)2+Δtνn=0N1𝐯hn+1L2(Ω;3×3)2+Δtγn=0N1𝐌hn+1L2(Ω;3×3)2\displaystyle+(\Delta t)^{2}\sum_{n=0}^{N-1}\left|\left|\frac{(\beta_{h}^{n+1}-\beta_{h}^{n})}{(\Delta t)^{2}}\right|\right|_{L^{2}(\Omega)}^{2}+\Delta t\nu\sum_{n=0}^{N-1}||\nabla\mathbf{v}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma\sum_{n=0}^{N-1}||\mathbf{M}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}
+Δtn=0N1Ωb(ϕhn)μhn+1μhn+1C(ϕh0,𝐅h0,βh0)+C.\displaystyle+\Delta t\sum_{n=0}^{N-1}\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\mu_{h}^{n+1}\leq C(\phi_{h}^{0},\mathbf{F}_{h}^{0},\beta_{h}^{0})+C.

Moreover, the function

(βhn+1)2+12ϕhn+1L2(Ω;3)2+λ2𝐅hn+1L2(Ω;3×3×3)2\left(\beta_{h}^{n+1}\right)^{2}+\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (164)

is a decreasing (Lyapunov) function for the discrete solution.

Proof. As in the proof of Lemma 5.1, we start by considering a reduced expression of System (162) in which 𝐯hn,𝐯hn+1,𝐮hVh,div\mathbf{v}_{h}^{n},\mathbf{v}_{h}^{n+1},\mathbf{u}_{h}\in V_{h,\text{div}}, i.e. the following: given 𝐯hnVh,div\mathbf{v}_{h}^{n}\in V_{h,\text{div}}, ϕhn,μhnSh,βhn,𝐅hn,𝐌hnXh\phi_{h}^{n},\mu_{h}^{n}\in S_{h},\beta_{h}^{n}\in\mathbb{R},\mathbf{F}_{h}^{n},\mathbf{M}_{h}^{n}\in X_{h}, find (𝐯hn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1)Vh,div×Xh×Xh×Sh×Sh(\mathbf{v}_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1})\in V_{h,\text{div}}\times X_{h}\times X_{h}\times S_{h}\times S_{h} and βhn+1\beta_{h}^{n+1}\in\mathbb{R} such that, for all (𝐮h,𝚯h,𝚷h,ξh,χh)Vh,div×Xh×Xh×Sh×Sh(\mathbf{u}_{h},\boldsymbol{\Theta}_{h},\boldsymbol{\Pi}_{h},\xi_{h},\chi_{h})\in V_{h,\text{div}}\times X_{h}\times X_{h}\times S_{h}\times S_{h},

{νΩ𝐯hn+1:𝐮h=βhn+1Ω(j(ϕhn,𝐅hn))+kΩμhnϕhn𝐮hβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐌hn(𝐅hn)T):𝐮h+βhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐅hn𝐌hn)𝐮h,Ω(𝐅hn+1𝐅hn):𝚯h+Δtβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯hΔtβhn+1Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯h+ΔtγΩ𝐌hn+1:𝚯h=0,Ω𝐌hn+1:𝚷h=βhn+1Ω(j(ϕhn,𝐅hn))+kΩ𝐅w(ϕhn,𝐅hn):𝚷h+λΩ𝐅hn+1\setstackgapS0.4ex\Shortstack𝚷h,Ω(ϕhn+1ϕhn)ξh+βhn+1Ω(j(ϕhn,𝐅hn))+kΔtΩ(𝐯hnϕhn)ξh+ΔtΩb(ϕhn)μhn+1ξh=0,Ωμhn+1χh=Ωϕhn+1χh+βhn+1Ω(j(ϕhn,𝐅hn))+kΩϕj(ϕhn,𝐅hn)χh,(βhn+1βhn)=12Ω(j(ϕhn,𝐅hn))+k[(Ωϕj(ϕhn,𝐅hn)(ϕhn+1ϕhn))+(Ω𝐅w(ϕhn,𝐅hn):(𝐅hn+1𝐅hn))+ΔtΩ(𝐯hn)𝐅hn:𝐌hn+1ΔtΩ(𝐅hn)T𝐌hn𝐯hn+1ΔtΩ(𝐯hn)𝐅hn:𝐌hn+1+ΔtΩ(𝐌hn(𝐅hn)T):𝐯hn+1+ΔtΩ(𝐯hnϕhn)μhn+1ΔtΩμhnϕhn𝐯hn+1].\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h}^{n+1}\colon\nabla\mathbf{u}_{h}=\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}\\ \displaystyle-\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}\left(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}\right)\colon\boldsymbol{\Theta}_{h}+\Delta t\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle-\Delta t\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mathbf{M}_{h}^{n+1}\colon\boldsymbol{\Pi}_{h}=\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\mathbf{F}_{h}^{n+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\left(\phi_{h}^{n+1}-\phi_{h}^{n}\right)\xi_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}\mu_{h}^{n+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h}^{n+1}\cdot\nabla\chi_{h}+\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\chi_{h},\\ \\ \displaystyle\left(\beta_{h}^{n+1}-\beta_{h}^{n}\right)=\frac{1}{2\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\biggl{[}\left(\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})(\phi_{h}^{n+1}-\phi_{h}^{n})\right)\\ \displaystyle+\left(\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n})\right)+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h}^{n+1}-\Delta t\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\right)^{T}\mathbf{M}_{h}^{n}\cdot\mathbf{v}_{h}^{n+1}\\ \displaystyle-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h}^{n+1}+\Delta t\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{v}_{h}^{n+1}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\mu_{h}^{n+1}\\ \displaystyle-\Delta t\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{v}_{h}^{n+1}\biggr{]}.\end{cases} (165)

The existence of a solution to System (165), which is a finite dimensional algebraic system with the same number of equations and unknowns, is guaranteed by its linearity. The solution is also unique. Indeed, let us consider two solutions (𝐯1,𝐅1,𝐌1,ϕ1,μ1,β1)(\mathbf{v}_{1},\mathbf{F}_{1},\mathbf{M}_{1},\phi_{1},\mu_{1},\beta_{1}) and (𝐯2,𝐅2,𝐌2,ϕ2,μ2,β2)(\mathbf{v}_{2},\mathbf{F}_{2},\mathbf{M}_{2},\phi_{2},\mu_{2},\beta_{2}) of System (165), satisfying the same initial and boundary conditions, and let us define 𝐯¯=𝐯1𝐯2\bar{\mathbf{v}}=\mathbf{v}_{1}-\mathbf{v}_{2}, 𝐅¯=𝐅1𝐅2\bar{\mathbf{F}}=\mathbf{F}_{1}-\mathbf{F}_{2}, 𝐌¯=𝐌1𝐌2\bar{\mathbf{M}}=\mathbf{M}_{1}-\mathbf{M}_{2}, ϕ¯=ϕ1ϕ2\bar{\phi}=\phi_{1}-\phi_{2}, μ¯=μ1μ2\bar{\mu}=\mu_{1}-\mu_{2}, β¯=β1β2\bar{\beta}=\beta_{1}-\beta_{2}. Taking the difference of the equations in (165) for the variables (𝐯1,𝐅1,𝐌1,ϕ1,μ1,β1)(\mathbf{v}_{1},\mathbf{F}_{1},\mathbf{M}_{1},\phi_{1},\mu_{1},\beta_{1}) and (𝐯2,𝐅2,𝐌2,ϕ2,μ2,β2)(\mathbf{v}_{2},\mathbf{F}_{2},\mathbf{M}_{2},\phi_{2},\mu_{2},\beta_{2}), we obtain

{νΩ𝐯¯:𝐮h=β¯Ω(j(ϕhn,𝐅hn))+kΩμhnϕhn𝐮hβ¯Ω(j(ϕhn,𝐅hn))+kΩ(𝐌hn(𝐅hn)T):𝐮h+β¯Ω(j(ϕhn,𝐅hn))+kΩ(𝐅hn𝐌hn)𝐮h,Ω𝐅¯:𝚯h+Δtβ¯Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯hΔtβ¯Ω(j(ϕhn,𝐅hn))+kΩ(𝐯hn)𝐅hn:𝚯h+ΔtγΩ𝐌¯:𝚯h=0,Ω𝐌¯:𝚷h=β¯Ω(j(ϕhn,𝐅hn))+kΩ𝐅w(ϕhn,𝐅hn):𝚷h+λΩ𝐅¯\setstackgapS0.4ex\Shortstack𝚷h,Ωϕ¯ξh+β¯Ω(j(ϕhn,𝐅hn))+kΔtΩ(𝐯hnϕhn)ξh+ΔtΩb(ϕhn)μ¯ξh=0,Ωμ¯χh=Ωϕ¯χh+β¯Ω(j(ϕhn,𝐅hn))+kΩϕj(ϕhn,𝐅hn)χh,β¯=12Ω(j(ϕhn,𝐅hn))+k[(Ωϕj(ϕhn,𝐅hn)ϕ¯)+Ω𝐅w(ϕhn,𝐅hn)𝐅¯+ΔtΩ(𝐯hn)𝐅hn:𝐌¯ΔtΩ(𝐅hn)T𝐌hn𝐯¯ΔtΩ(𝐯hn)𝐅hn:𝐌¯+ΔtΩ(𝐌hn(𝐅hn)T):𝐯¯+ΔtΩ(𝐯hnϕhn)μ¯ΔtΩμhnϕhn𝐯¯].\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\bar{\mathbf{v}}\colon\nabla\mathbf{u}_{h}=\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}\\ \displaystyle-\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}+\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n}\right)\cdot\mathbf{u}_{h},\\ \\ \displaystyle\int_{\Omega}\bar{\mathbf{F}}\colon\boldsymbol{\Theta}_{h}+\Delta t\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}\\ \displaystyle-\Delta t\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}+\Delta t\gamma\int_{\Omega}\bar{\mathbf{M}}\colon\boldsymbol{\Theta}_{h}=0,\\ \\ \displaystyle\int_{\Omega}\bar{\mathbf{M}}\colon\boldsymbol{\Pi}_{h}=\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\bar{\mathbf{F}}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\bar{\phi}\,\xi_{h}+\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\bar{\mu}\cdot\nabla\xi_{h}=0,\\ \\ \displaystyle\int_{\Omega}\bar{\mu}\chi_{h}=\int_{\Omega}\nabla\bar{\phi}\cdot\nabla\chi_{h}+\frac{\bar{\beta}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\chi_{h},\\ \\ \displaystyle\bar{\beta}=\frac{1}{2\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\biggl{[}\left(\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\bar{\phi}\right)\\ \displaystyle+\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\bar{\mathbf{F}}+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\bar{\mathbf{M}}-\Delta t\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\right)^{T}\mathbf{M}_{h}^{n}\cdot\bar{\mathbf{v}}\\ \displaystyle-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\bar{\mathbf{M}}+\Delta t\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\bar{\mathbf{v}}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\bar{\mu}\\ \displaystyle-\Delta t\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\bar{\mathbf{v}}\biggr{]}.\end{cases} (166)

Taking 𝐮h=Δt𝐯¯\mathbf{u}_{h}=\Delta t\bar{\mathbf{v}}, 𝚯h=𝐌¯\boldsymbol{\Theta}_{h}=\bar{\mathbf{M}}, 𝚷h=𝐅¯\boldsymbol{\Pi}_{h}=-\bar{\mathbf{F}}, ξh=μ¯\xi_{h}=\bar{\mu}, χh=ϕ¯\chi_{h}=-\bar{\phi} in (166) and multiplying (166)6 by 2β¯2\bar{\beta}, summing all the equations and using Assumption A1, we obtain that

Δtν𝐯¯L2(Ω;3×3)2+Δtγ𝐌¯L2(Ω;3×3)2+λ𝐅¯L2(Ω;3×3×3)2\displaystyle\displaystyle\Delta t\nu||\nabla\bar{\mathbf{v}}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\bar{\mathbf{M}}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\lambda||\nabla\bar{\mathbf{F}}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (167)
+Δtb0μ¯L2(Ω;3)2+ϕ¯L2(Ω;3)2+2β¯20.\displaystyle\displaystyle+\Delta tb_{0}||\nabla\bar{\mu}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+||\nabla\bar{\phi}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+2\bar{\beta}^{2}\leq 0.

Taking moreover 𝚯h=𝐞l,r\boldsymbol{\Theta}_{h}=\mathbf{e}_{l,r}, l,r=1,2,3l,r=1,2,3 and ξh=χh=1\xi_{h}=\chi_{h}=1 in (166), using moreover the fact that 𝐯hnVh,div\mathbf{v}_{h}^{n}\in V_{h,\text{div}}, we obtain that

|Ω𝐅¯|C𝐌¯L2(Ω;3×3),Ωϕ¯=0,|Ωμ¯|C|β¯|.\left|\int_{\Omega}\bar{\mathbf{F}}\right|\leq C||\bar{\mathbf{M}}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})},\;\int_{\Omega}\bar{\phi}=0,\;\left|\int_{\Omega}\bar{\mu}\right|\leq C|\bar{\beta}|. (168)

From (167) and (168) we finally conclude that 𝐯¯=𝟎\bar{\mathbf{v}}=\mathbf{0}, 𝐅¯=𝟎\bar{\mathbf{F}}=\mathbf{0}, 𝐌¯=𝟎\bar{\mathbf{M}}=\mathbf{0}, ϕ¯=0\bar{\phi}=0, μ¯=0\bar{\mu}=0, β¯=0\bar{\beta}=0, hence the solution of System (165) is unique. Since (165)1 defines a linear functional from VhV_{h} to \mathbb{R} which vanishes on Vh,divV_{h,\text{div}}, we conclude by standard arguments (see e.g. [14, p.22]) that there exists a unique shn+1Rhs_{h}^{n+1}\in R_{h} such that

(𝐯hn+1,shn+1,𝐅hn+1,𝐌hn+1,ϕhn+1,μhn+1,βhn+1)Vh×Rh×Xh×Xh×Sh×Xh×(\mathbf{v}_{h}^{n+1},s_{h}^{n+1},\mathbf{F}_{h}^{n+1},\mathbf{M}_{h}^{n+1},\phi_{h}^{n+1},\mu_{h}^{n+1},\beta_{h}^{n+1})\in V_{h}\times R_{h}\times X_{h}\times X_{h}\times S_{h}\times X_{h}\times\mathbb{R}

is the unique solution of System (162).

Let us now take 𝐮h=Δt𝐯hn+1\mathbf{u}_{h}=\Delta t\mathbf{v}_{h}^{n+1}, ph=Δtshn+1p_{h}=\Delta ts_{h}^{n+1}, 𝚯h=𝐌hn+1\boldsymbol{\Theta}_{h}=\mathbf{M}_{h}^{n+1}, 𝚷h=(𝐅hn+1𝐅hn)\boldsymbol{\Pi}_{h}=-(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}), ξh=μhn+1\xi_{h}=\mu_{h}^{n+1}, χh=(ϕhn+1ϕhn)\chi_{h}=-(\phi_{h}^{n+1}-\phi_{h}^{n}) in (162) and let us multiply (162)7 by 2βhn+12\beta_{h}^{n+1}. Summing all the equations, and using the identity a(ab)=12|a|2+12|ab|212|b|2a\star(a-b)=\frac{1}{2}|a|^{2}+\frac{1}{2}|a-b|^{2}-\frac{1}{2}|b|^{2}, we obtain

Δtν𝐯hn+1L2(Ω;3×3)2+Δtγ𝐌hn+1L2(Ω;3×3)2+λ2𝐅hn+1L2(Ω;3×3×3)2\displaystyle\displaystyle\Delta t\nu||\nabla\mathbf{v}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\Delta t\gamma||\mathbf{M}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3})}^{2}+\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2} (169)
+λ2(𝐅hn+1𝐅hn)L2(Ω;3×3×3)2+ΔtΩb(ϕhn)μhn+1μhn+1\displaystyle\displaystyle+\frac{\lambda}{2}\left|\left|\nabla\left(\mathbf{F}_{h}^{n+1}-\mathbf{F}_{h}^{n}\right)\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\mu_{h}^{n+1}
+12ϕhn+1L2(Ω;3)2+12(ϕhn+1ϕhn)L2(Ω;3)2+(βhn+1)2+(βhn+1βhn)2\displaystyle+\frac{1}{2}||\nabla\phi_{h}^{n+1}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\frac{1}{2}\left|\left|\nabla\left(\phi_{h}^{n+1}-\phi_{h}^{n}\right)\right|\right|_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\left(\beta_{h}^{n+1}\right)^{2}+\left(\beta_{h}^{n+1}-\beta_{h}^{n}\right)^{2}
=λ2𝐅hnL2(Ω;3×3×3)2+12ϕhnL2(Ω;3)2+(βhn)2,\displaystyle\displaystyle=\frac{\lambda}{2}||\nabla\mathbf{F}_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3\times 3\times 3})}^{2}+\frac{1}{2}||\nabla\phi_{h}^{n}||_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+\left(\beta_{h}^{n}\right)^{2},

which gives that (164) is a decreasing (Lyapunov) function for the discrete solutions. Summing (169) from n=0mn=0\to m, for m=0N1m=0\to N-1, we finally get (163). \Box
Following the ideas reported [16], starting from the coupled system (162) we obtain a decoupled system by introducing the following ansatz for its solution:

𝐯hn+1=𝐯h,0n+1+Ahn+1𝐯h,1n+1,\displaystyle\mathbf{v}_{h}^{n+1}=\mathbf{v}_{h,0}^{n+1}+A_{h}^{n+1}\mathbf{v}_{h,1}^{n+1}, (170)
shn+1=sh,0n+1+Ahn+1sh,1n+1,\displaystyle s_{h}^{n+1}=s_{h,0}^{n+1}+A_{h}^{n+1}s_{h,1}^{n+1},
𝐅hn+1=𝐅h,0n+1+Ahn+1𝐅h,1n+1,\displaystyle\mathbf{F}_{h}^{n+1}=\mathbf{F}_{h,0}^{n+1}+A_{h}^{n+1}\mathbf{F}_{h,1}^{n+1},
𝐌hn+1=𝐌h,0n+1+Ahn+1𝐌h,1n+1,\displaystyle\mathbf{M}_{h}^{n+1}=\mathbf{M}_{h,0}^{n+1}+A_{h}^{n+1}\mathbf{M}_{h,1}^{n+1},
ϕhn+1=ϕh,0n+1+Ahn+1ϕh,1n+1,\displaystyle\phi_{h}^{n+1}=\phi_{h,0}^{n+1}+A_{h}^{n+1}\phi_{h,1}^{n+1},
μhn+1=μh,0n+1+Ahn+1μh,1n+1,\displaystyle\mu_{h}^{n+1}=\mu_{h,0}^{n+1}+A_{h}^{n+1}\mu_{h,1}^{n+1},

where

Ahn+1:=βhn+1Ω(j(ϕhn,𝐅hn))+k.A_{h}^{n+1}:=\frac{\beta_{h}^{n+1}}{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}.

Inserting the expansions (170) in (162) and equating the terms of the same order in Ahn+1A_{h}^{n+1}, we obtain that the variables 𝐯h,in+1,sh,in+1,𝐅h,in+1,𝐅h,in+1,ϕh,in+1,μh,in+1\mathbf{v}_{h,i}^{n+1},s_{h,i}^{n+1},\mathbf{F}_{h,i}^{n+1},\mathbf{F}_{h,i}^{n+1},\phi_{h,i}^{n+1},\mu_{h,i}^{n+1}, i=0,1i=0,1, satisfy the following decoupled systems.

Stokes subsystem:

{νΩ𝐯h,0n+1:𝐮hΩsh,0n+1div𝐮h=0,Ωphdiv𝐯h,0n+1=0νΩ𝐯h,1n+1:𝐮hΩshn+1div𝐮h=Ωμhnϕhn𝐮hΩ(𝐌hn(𝐅hn)T):𝐮h+Ω(𝐅hn𝐌hn)𝐮h,Ωphdiv𝐯h,1n+1=0.\begin{cases}\displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h,0}^{n+1}\colon\nabla\mathbf{u}_{h}-\int_{\Omega}s_{h,0}^{n+1}\operatorname{div}\mathbf{u}_{h}=0,\\ \displaystyle\int_{\Omega}p_{h}\operatorname{div}\mathbf{v}_{h,0}^{n+1}=0\\ \\ \displaystyle\nu\int_{\Omega}\nabla\mathbf{v}_{h,1}^{n+1}\colon\nabla\mathbf{u}_{h}-\int_{\Omega}s_{h}^{n+1}\operatorname{div}\mathbf{u}_{h}=\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{u}_{h}-\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{u}_{h}\\ \displaystyle+\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\odot\mathbf{M}_{h}^{n}\right)\cdot\mathbf{u}_{h},\\ \displaystyle\int_{\Omega}p_{h}\operatorname{div}\mathbf{v}_{h,1}^{n+1}=0.\end{cases} (172)

Since 𝐯h,0n+1|Ω=0\mathbf{v}_{h,0}^{n+1}|_{\partial\Omega}=0, (172)1,2 imply that 𝐯h,0n+10\mathbf{v}_{h,0}^{n+1}\equiv 0.

Elasticity subsystem:

{Ω(𝐅h,0n+1𝐅hn):𝚯h+ΔtγΩ𝐌h,0n+1:𝚯h=0,Ω𝐌h,0n+1:𝚷h=λΩ𝐅h,0n+1\setstackgapS0.4ex\Shortstack𝚷h,Ω𝐅h,1n+1:𝚯h+ΔtΩ(𝐯hn)𝐅hn:𝚯hΔtΩ(𝐯hn)𝐅hn:𝚯h+ΔtγΩ𝐌h,1n+1:𝚯h=0,Ω𝐌h,1n+1:𝚷h=Ω𝐅w(ϕhn,𝐅hn):𝚷h+λΩ𝐅h,1n+1\setstackgapS0.4ex\Shortstack𝚷h.\begin{cases}\displaystyle\int_{\Omega}\left(\mathbf{F}_{h,0}^{n+1}-\mathbf{F}_{h}^{n}\right)\colon\boldsymbol{\Theta}_{h}+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h,0}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}_{h,0}^{n+1}\colon\boldsymbol{\Pi}_{h}=\lambda\int_{\Omega}\nabla\mathbf{F}_{h,0}^{n+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h},\\ \\ \displaystyle\int_{\Omega}\mathbf{F}_{h,1}^{n+1}\colon\boldsymbol{\Theta}_{h}+\Delta t\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}-\Delta t\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\boldsymbol{\Theta}_{h}+\Delta t\gamma\int_{\Omega}\mathbf{M}_{h,1}^{n+1}\colon\boldsymbol{\Theta}_{h}=0,\\ \displaystyle\int_{\Omega}\mathbf{M}_{h,1}^{n+1}\colon\boldsymbol{\Pi}_{h}=\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\boldsymbol{\Pi}_{h}+\lambda\int_{\Omega}\nabla\mathbf{F}_{h,1}^{n+1}\mathbin{\setstackgap{S}{0.4ex}\mathrel{\Shortstack{{.}{.}{.}}}}\nabla\boldsymbol{\Pi}_{h}.\end{cases} (173)

Cahn–Hilliard subsystem:

{Ω(ϕh,0n+1ϕhn)ξh+ΔtΩb(ϕhn)μh,0n+1ξh=0,Ωμh,0n+1χh=Ωϕh,0n+1χh,Ωϕh,1n+1ξh+ΔtΩ(𝐯hnϕhn)ξh+ΔtΩb(ϕhn)μhn+1ξh=0,Ωμh,1n+1χh=Ωϕh,1n+1χh+Ωϕj(ϕhn,𝐅hn)χh.\begin{cases}\displaystyle\int_{\Omega}\left(\phi_{h,0}^{n+1}-\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h,0}^{n+1}\cdot\nabla\xi_{h}=0,\\ \displaystyle\int_{\Omega}\mu_{h,0}^{n+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h,0}^{n+1}\cdot\nabla\chi_{h},\\ \\ \displaystyle\int_{\Omega}\phi_{h,1}^{n+1}\xi_{h}+\Delta t\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\xi_{h}+\Delta t\int_{\Omega}b\left(\phi_{h}^{n}\right)\nabla\mu_{h}^{n+1}\cdot\nabla\xi_{h}=0,\\ \displaystyle\int_{\Omega}\mu_{h,1}^{n+1}\chi_{h}=\int_{\Omega}\nabla\phi_{h,1}^{n+1}\cdot\nabla\chi_{h}+\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\chi_{h}.\end{cases} (174)

Moreover,

[Ω(j(ϕhn,𝐅hn))+kΔt12Ω(j(ϕhn,𝐅hn))+k{(Ωϕj(ϕhn,𝐅hn)ϕh,1n+1Δt)\displaystyle\displaystyle\biggl{[}\frac{\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}{\Delta t}-\frac{1}{2\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\biggl{\{}\left(\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\frac{\phi_{h,1}^{n+1}}{\Delta t}\right) (175)
+(Ω𝐅w(ϕhn,𝐅hn):𝐅h,1n+1Δt)+Ω(𝐯hn)𝐅hn:𝐌h,1n+1Ω(𝐅hn)T𝐌hn𝐯h,1n+1\displaystyle\displaystyle+\left(\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\frac{\mathbf{F}_{h,1}^{n+1}}{\Delta t}\right)+\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h,1}^{n+1}-\int_{\Omega}\left(\nabla\mathbf{F}_{h}^{n}\right)^{T}\mathbf{M}_{h}^{n}\cdot\mathbf{v}_{h,1}^{n+1}
Ω(𝐯hn)𝐅hn:𝐌h1n+1+Ω(𝐌hn(𝐅hn)T):𝐯h,1n+1+Ω(𝐯hnϕhn)μh,1n+1\displaystyle\displaystyle-\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h_{1}}^{n+1}+\int_{\Omega}\left(\mathbf{M}_{h}^{n}(\mathbf{F}_{h}^{n})^{T}\right)\colon\nabla\mathbf{v}_{h,1}^{n+1}+\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\mu_{h,1}^{n+1} (176)
Ωμhnϕhn𝐯h,1n+1}]Ahn+1\displaystyle\displaystyle-\int_{\Omega}\mu_{h}^{n}\nabla\phi_{h}^{n}\cdot\mathbf{v}_{h,1}^{n+1}\biggr{\}}\biggr{]}A_{h}^{n+1} (177)
=[βhnΔt+12Ω(j(ϕhn,𝐅hn))+k{(Ωϕj(ϕhn,𝐅hn)ϕh,0nϕhnΔt)\displaystyle\displaystyle=\biggl{[}\frac{\beta_{h}^{n}}{\Delta t}+\frac{1}{2\sqrt{\int_{\Omega}(j(\phi_{h}^{n},\mathbf{F}_{h}^{n}))+k}}\biggl{\{}\left(\int_{\Omega}\partial_{\phi}j(\phi_{h}^{n},\mathbf{F}_{h}^{n})\frac{\phi_{h,0}^{n}-\phi_{h}^{n}}{\Delta t}\right) (178)
+(Ω𝐅w(ϕhn,𝐅hn):𝐅h,0n+1𝐅hnΔt)+Ω(𝐯hn)𝐅hn:𝐌h,0n+1Ω(𝐯hn)𝐅hn:𝐌h0n+1\displaystyle\displaystyle+\left(\int_{\Omega}\partial_{\mathbf{F}}w(\phi_{h}^{n},\mathbf{F}_{h}^{n})\colon\frac{\mathbf{F}_{h,0}^{n+1}-\mathbf{F}_{h}^{n}}{\Delta t}\right)+\int_{\Omega}\left({\mathbf{v}_{h}^{n}}\cdot\nabla\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h,0}^{n+1}-\int_{\Omega}\left(\nabla{\mathbf{v}_{h}^{n}}\right)\mathbf{F}_{h}^{n}\colon\mathbf{M}_{h_{0}}^{n+1}
+Ω(𝐯hnϕhn)μh,0n+1}].\displaystyle\displaystyle+\int_{\Omega}\left(\mathbf{v}_{h}^{n}\cdot\nabla\phi_{h}^{n}\right)\mu_{h,0}^{n+1}\biggr{\}}\biggr{]}. (179)

It’s easy to show, by taking 𝐮h=𝐯h,1n+1\mathbf{u}_{h}=\mathbf{v}_{h,1}^{n+1} and ph=sh,1n+1p_{h}=s_{h,1}^{n+1} in (172)3,4, 𝚯h=𝐌h,1n+1\boldsymbol{\Theta}_{h}=\mathbf{M}_{h,1}^{n+1}, 𝚷h=𝐅h,1n+1Δt\boldsymbol{\Pi}_{h}=\frac{\mathbf{F}_{h,1}^{n+1}}{\Delta t} in (173)3,4 and ξh=μh,1n+1\xi_{h}=\mu_{h,1}^{n+1}, χh=ϕh,1n+1Δt\chi_{h}=\frac{\phi_{h,1}^{n+1}}{\Delta t} in (174)3,4, that the term in the square brackets which multiplies Ahn+1A_{h}^{n+1} in (175) is strictly positive, hence Ahn+1A_{h}^{n+1} is uniquely determined.

6 Results

In this section we report the results of numerical simulations in two space dimensions for different test cases. We consider the following form of the free energy density:

ψ(ϕ)=β4αϕ2(1ϕ)2,\psi(\phi)=\frac{\beta}{4\alpha}\phi^{2}(1-\phi)^{2}, (180)
w(ϕ,𝐅)=ζ2(𝐅T𝐅𝐇(ϕ)):(𝐅T𝐅𝐇(ϕ)),w(\phi,\mathbf{F})=\frac{\zeta}{2}\left(\mathbf{F}^{T}\mathbf{F}-\mathbf{H}(\phi)\right)\colon\left(\mathbf{F}^{T}\mathbf{F}-\mathbf{H}(\phi)\right), (181)

where β,α,ζ>0\beta,\alpha,\zeta>0 are model parameters, and

𝐇(ϕ)=𝐅~(ϕ)T𝐅~(ϕ),\mathbf{H}(\phi)=\tilde{\mathbf{F}}(\phi)^{T}\tilde{\mathbf{F}}(\phi),

with

𝐅~(ϕ)=(1aϕ01)\tilde{\mathbf{F}}(\phi)=\begin{pmatrix}1&&a\phi\\ 0&&1\end{pmatrix}

and a>0a>0. In particular, (180) is the standard Cahn–Hilliard smooth double-well potential, with stable minima at ϕ0\phi\equiv 0 and ϕ1\phi\equiv 1, where the parameter β\beta is proportional to the surface tension and the parameter α\alpha represents the interface thickness in the Cahn–Hilliard surface term. Consequently, the term proportional to |ϕ|2|\nabla\phi|^{2} in (3) is multiplied by a factor ϵ=βα\epsilon=\beta\alpha, (ϵ,α,β\epsilon,\alpha,\beta were taken equal to one in the previous analysis for ease of notation). Moreover, (181) represents an elastic energy density of shape memory alloy type (see e.g. [21, Eq. (3.20a)]), where the pure phases associated to the stable minima of the phase field potential are characterized by different elastic properties. Indeed, we have that

ϕ(ψ(ϕ)+w(ϕ,𝐅))=ψ(ϕ)ζ𝐇(ϕ):(𝐅T𝐅𝐇(ϕ)),\partial_{\phi}\left(\psi(\phi)+w(\phi,\mathbf{F})\right)=\psi^{\prime}(\phi)-\zeta\mathbf{H}^{\prime}(\phi)\colon\left(\mathbf{F}^{T}\mathbf{F}-\mathbf{H}(\phi)\right),

and

𝐅w(ϕ,𝐅)=2ζ𝐅(𝐅T𝐅𝐇(ϕ)).\partial_{\mathbf{F}}w(\phi,\mathbf{F})=2\zeta\mathbf{F}\left(\mathbf{F}^{T}\mathbf{F}-\mathbf{H}(\phi)\right).

Hence, the global minima for ψ(ϕ)+w(ϕ,𝐅)\psi(\phi)+w(\phi,\mathbf{F}), at which it takes its minimum value equal to zero, are attained at ϕ0\phi\equiv 0, ϕ1\phi\equiv 1 and 𝐅𝐑𝐅~(ϕ)\mathbf{F}\equiv\mathbf{R}\tilde{\mathbf{F}}(\phi), where 𝐑SO(2×2)\mathbf{R}\in SO(\mathbb{R}^{2\times 2}) is a generic rotation matrix. We observe that (181) is frame indifferent but not isotropic.

In the following, we will consider two test cases to investigate the phase separation dynamics and the coarsening dynamics associated to (180) and (181), starting from different initial configurations. In Test Case 1 we will consider the phase separation dynamics starting from different initial conditions for ϕ\phi. In particular, we will consider for ϕ0\phi_{0} two different configurations given by a small uniformly distributed random perturbation around the values ϕ0=0.3\phi_{0}=0.3 or ϕ0=0.7\phi_{0}=0.7, leading to different topologies of the stationary states, determined both from the metastability of ψ()\psi(\cdot) and from the elasticity dynamics described by w(ϕ,𝐅)w(\phi,\mathbf{F}). In both cases, we will take as initial condition for the deformation gradient 𝐅0=𝐈\mathbf{F}_{0}=\mathbf{I}.

The values of the parameters for Test Case 1 are taken as ν=1\nu=1, λ=0.001\lambda=0.001, β=0.1\beta=0.1, α=0.002\alpha=0.002, ζ=10\zeta=10, a=0.5a=0.5. Moreover, we take a constant mobility b(ϕ)1b(\phi)\equiv 1 and we consider a domain Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1], with a uniform triangulation of dimension 64×6464\times 64, and Δt=0.001ϵ\Delta t=0.001\epsilon. In Test Case 1 we will vary the value of the parameter γ\gamma, considering γ=1\gamma=1 or γ=0.001\gamma=0.001, in order to observe the phase separation dynamics under different degrees of diffusive regularization of the transport equation for the deformation gradient.

In Test Case 2 we will consider the merging and coarsening dynamics of isolated circular subregions of a pure phase immersed in a bath of the opposite pure phase. In particular, we will consider two different configurations with two and four initial circular subregions placed symmetrically with respect to the centre of the domain. The values of the parameters for Test Case 2 are taken as ν=1\nu=1, λ=0.001\lambda=0.001, β=0.1\beta=0.1, α=0.02\alpha=0.02, a=0.5a=0.5 and γ=0.001\gamma=0.001, b(ϕ)1b(\phi)\equiv 1. We will vary the value of the elastic modulus, taking ζ=1\zeta=1 and ζ=10\zeta=10, in order to observe the effects of higher stiffness on the merging and coarsening dynamics. We consider a domain Ω=[0,1]×[0,1]\Omega=[0,1]\times[0,1], with a uniform triangulation of dimension 64×6464\times 64, furtherly refined in a neighborhood of the support set of ϕ0\phi_{0}. The time step is taken as Δt=0.00001ϵ\Delta t=0.00001\epsilon.

Remark 6.1

The expression (181) can be rewritten as

w(ϕ,𝐅)=ζ2𝐅T𝐅:𝐅T𝐅ζ𝐇(ϕ):𝐅T𝐅+ζ2𝐇(ϕ):𝐇(ϕ),\displaystyle w(\phi,\mathbf{F})=\frac{\zeta}{2}\mathbf{F}^{T}\mathbf{F}\colon\mathbf{F}^{T}\mathbf{F}-\zeta\mathbf{H}(\phi)\colon\mathbf{F}^{T}\mathbf{F}+\frac{\zeta}{2}\mathbf{H}(\phi)\colon\mathbf{H}(\phi), (182)

which is of the particular form (148) and satisfies Assumption 𝐀𝟐𝐡\bf{A2_{h}} in the general case of Remark 5.4 when ϕ\phi is bounded, which is always practically verified by the numerical solutions.

The fully coupled system (147) is solved through the iteration method (159), while the solution of (162) is obtained by solving the independent subsystems (172), (173), (174) and (175) and using (170).

6.1 Gradient stability of (147) and (162)

In this section we show the gradient stability of (147) and (162). As representative results, we report the solutions of (147) and (162), with the energy density defined in (180) and (181), in the case γ=1\gamma=1, ϕ0=0.3±0.5ι\phi_{0}=0.3\pm 0.5\iota and 𝐅0=𝐈\mathbf{F}_{0}=\mathbf{I}, where ι\iota is a random perturbation uniformly distributed in the interval [0,1][0,1].

In Figure 1 we report the plots of the Lyapunov functionals (150) and (164) (subtracting to it the value of kk), which we call LCSL_{CS} and LDSAVL_{DSAV} respectively, versus time, together with the functionals

ECH:=Ω(ψ(ϕ)+ϵ2|ϕ|2)E_{CH}:=\int_{\Omega}\left(\psi(\phi)+\frac{\epsilon}{2}|\nabla\phi|^{2}\right)

and

EEL:=Ω(w(ϕ,𝐅)+λ2|𝐅|2).E_{EL}:=\int_{\Omega}\left(w(\phi,\mathbf{F})+\frac{\lambda}{2}|\nabla\mathbf{F}|^{2}\right).
Refer to caption
Figure 1: Plot of the Lyapunov functionals (150) (LCSL_{CS}) and (164) (LDSAVL_{DSAV}) and of the functionals ECHE_{CH} and EELE_{EL} vs time for the schemes (147) (Convex Splitting, left panels) and (162) (DSAV, right panels).

We observe from Figure 1 that both systems (147) and (162) are gradient stable. The plots show that the two systems behave almost identically over the whole time span until the attainment of the stationary state, with the convex splitting scheme showing a steeper decrease of the Lyapunov functional at early times than the DSAV scheme. In Figure 2 we compare the plots of ϕ\phi and |𝐅||\mathbf{F}| for the two schemes at initial and at late time points, observing that the numerical solutions for ϕ\phi and 𝐅\mathbf{F} of (147) and (162) are statistically and topologically similar throughout the whole dynamics. Note that, since the initial condition ϕ0\phi_{0} is random, the numerical simulations for the two schemes do not start from the same initial conditions. In Figure 3 we also compare the line plots, along a vertical line, of |𝐅||\mathbf{F}| at time t=0.01t=0.01, i.e. at an early time where the two schemes show slight differences in the Lyapunov functional decrease, observing higher numerical oscillations for the scheme (162).

Refer to caption
Figure 2: Plots of ϕ\phi and |𝐅||\mathbf{F}| at different time points for the schemes (147) (Convex Splitting) and (162) (DSAV).
Refer to caption
Figure 3: Line plots of |𝐅||\mathbf{F}|, along a vertical line, at time t=0.01t=0.01, for the schemes (147) (Convex Splitting) and (162) (DSAV).

The computational time to solve (147) for 150000150000 time steps is 1.5e+6s\sim 1.5\mathrm{e}{+6}\,s, while computational time to solve (162) for 150000150000 time steps is 3.4e+5s\sim 3.4\mathrm{e}{+5}\,s. Hence the time needed to solve (147) is almost one order of magnitude greater than the computational time needed to solve (162).

In the following, the numerical results for Test Case 1 and Test Case 2 are obtained as solutions of (147).

6.2 Test case 1 – Phase separation

We first consider the initial conditions ϕ0=0.3±0.5ι\phi_{0}=0.3\pm 0.5\iota, where ι\iota is a random perturbation uniformly distributed in the interval [0,1][0,1], and 𝐅0=𝐈\mathbf{F}_{0}=\mathbf{I}. In Figure 4 we show the numerical results at different time points throughout the phase separation dynamics, up to late times at which we can observe the coarsening dynamics of the separated domain subregions, for both the cases with γ=1\gamma=1 and γ=0.001\gamma=0.001.

Refer to caption
Figure 4: Plot of ϕ\phi (first row) and 𝐅xx\mathbf{F}_{xx} (second row), 𝐅xy\mathbf{F}_{xy} (third row), 𝐅yx\mathbf{F}_{yx} (fourth row), 𝐅yy\mathbf{F}_{yy} (fifth row), det𝐅\text{det}\mathbf{F} (sixth row) at different time points in the case ϕ0=0.3±0.5ι\phi_{0}=0.3\pm 0.5\iota, for γ=1\gamma=1 (first and second columns) and γ=0.001\gamma=0.001 (third and fourth columns).

We observe from Figure 4 that the phase separation dynamics for the ϕ\phi variable consists in the formation of elongated circular clusters with ϕ1\phi\sim 1 along orthogonal directions determined by the separation dynamics for the 𝐅\mathbf{F} variable, immersed in a bath with ϕ0\phi\sim 0. At late times, these clusters grow and merge, forming strips oriented in orthogonal directions. The variable 𝐅\mathbf{F} assumes at late times values 𝐅𝐑𝐅(ϕ)\mathbf{F}\sim\mathbf{R}\mathbf{F}(\phi) in the regions with ϕ1\phi\sim 1, with off-diagonal components oriented along strips, and 𝐅𝐈\mathbf{F}\sim\mathbf{I} in the regions with ϕ0\phi\sim 0. For instance, checking the values taken by the components of the deformation gradient in one cluster with ϕ1\phi\sim 1 along the second column of Figure 4, we observe the values 𝐅xx0.98\mathbf{F}_{xx}\sim 0.98, 𝐅xy0.29\mathbf{F}_{xy}\sim 0.29, 𝐅yx0.20\mathbf{F}_{yx}\sim 0.20, 𝐅yy1.08\mathbf{F}_{yy}\sim 1.08, which corresponds to

(cos(θ)sin(θ)sin(θ)cos(θ))(10.501)\begin{pmatrix}\cos(\theta)&&-\sin(\theta)\\ \sin(\theta)&&\cos(\theta)\end{pmatrix}\begin{pmatrix}1&&0.5\\ 0&&1\end{pmatrix}

with θ=0.2\theta=0.2 radiants. In the case γ=1\gamma=1 we also observe higher deviations from the values det𝐅=1\text{det}\mathbf{F}=1, concentrated at the boundary regions of the clusters with ϕ1\phi\sim 1, then in the case γ=0.001\gamma=0.001, where det𝐅1\text{det}\mathbf{F}\sim 1 over the whole domain.

We then consider the initial conditions ϕ0=0.7±0.2ι\phi_{0}=0.7\pm 0.2\iota and 𝐅0=𝐈\mathbf{F}_{0}=\mathbf{I}. In Figure 5 we show the numerical results at different time points, for both the cases with γ=1\gamma=1 and γ=0.001\gamma=0.001.

Refer to caption
Figure 5: Plot of ϕ\phi (first row) and 𝐅xx\mathbf{F}_{xx} (second row), 𝐅xy\mathbf{F}_{xy} (third row), 𝐅yx\mathbf{F}_{yx} (fourth row), 𝐅yy\mathbf{F}_{yy} (fifth row), det𝐅\text{det}\mathbf{F} (sixth row) at different time points in the case ϕ0=0.7±0.2ι\phi_{0}=0.7\pm 0.2\iota, for γ=1\gamma=1 (first and second columns) and γ=0.001\gamma=0.001 (third and fourth columns).

We observe from Figure 5 that the phase separation dynamics for the ϕ\phi variable consists in the formation of elongated circular clusters with ϕ0\phi\sim 0 along orthogonal directions determined by the separation dynamics for the 𝐅\mathbf{F} variable, immersed in a bath with ϕ1\phi\sim 1. As in the case reported in Figure 4, at late times the clusters grow and merge, forming strips oriented in orthogonal directions and 𝐅\mathbf{F} assumes values 𝐅𝐑𝐅(ϕ)\mathbf{F}\sim\mathbf{R}\mathbf{F}(\phi) in the regions with ϕ1\phi\sim 1, with off-diagonal components oriented along strips, and 𝐅𝐈\mathbf{F}\sim\mathbf{I} in the regions with ϕ0\phi\sim 0. As in Figure 4, in the case γ=1\gamma=1 we observe higher deviations from the values det𝐅=1\text{det}\mathbf{F}=1, concentrated at the boundary regions of the clusters with ϕ0\phi\sim 0, then in the case γ=0.001\gamma=0.001, where det𝐅1\text{det}\mathbf{F}\sim 1 over the whole domain.

6.3 Test case 2 – Coarsening

We start by considering the initial conditions ϕ0=1.0(χB1+χB2)\phi_{0}=1.0(\chi_{B1}+\chi_{B2}), where χB1\chi_{B1} and χB2\chi_{B2} are the characteristic functions of two circular regions placed symmetrically along the xx direction, and

𝐅0=(1aϕ001).\mathbf{F}_{0}=\begin{pmatrix}1&&a\phi_{0}\\ 0&&1\end{pmatrix}.

In Figure 6 we show the numerical results at different time points, both for the cases ζ=1\zeta=1 and ζ=10\zeta=10.

Refer to caption
Figure 6: Plot of ϕ\phi (first and second rows) and |𝐯||\mathbf{v}| (third and fourth rows) at different time points, for ζ=1\zeta=1 (first and third rows) and ζ=10\zeta=10 (second and fourth rows), in the case ϕ0=1.0(χB1+χB2)\phi_{0}=1.0(\chi_{B1}+\chi_{B2}).

We observe from Figure 6 that, for a low value of the elastic modulus ζ=1\zeta=1, the initial circular clusters evolve, with interacting tips initially oriented along the bisecting directions of the plane and advected by the velocity field, finally merging into the equilibrium shape of an ellipse with the major axis oriented along the xx axis. Increasing the elastic modulus to ζ=10\zeta=10, the circular clusters interact only weakly and do not merge during the observed time window, deforming to a rectangular shape. The results in Figure 6 may be compared to the results shown in [15, Figures 6-7] for the merging of two circular precipitates in a linear isotropic inhomogeneous elastic medium. While in the latter case with the considered elastic parameters the equilibrium shape is circular, in the present case the anisotropy associated to the pure phase ϕ1\phi\equiv 1 drives an elliptic or rectangular equilibrium shape.

We finally consider the initial conditions ϕ0=1.0(χB1+χB2+χB3+χB4)\phi_{0}=1.0(\chi_{B1}+\chi_{B2}+\chi_{B3}+\chi_{B4}), where χB1\chi_{B1}, χB2\chi_{B2}, χB3\chi_{B3} and χB4\chi_{B4} are the characteristic functions of four circular regions placed symmetrically with respect to the center of the domain, and

𝐅0=(1aϕ001).\mathbf{F}_{0}=\begin{pmatrix}1&&a\phi_{0}\\ 0&&1\end{pmatrix}.

In Figure 7 we show the numerical results at different time points, both for the cases ζ=1\zeta=1 and ζ=10\zeta=10.

Refer to caption
Figure 7: Plot of ϕ\phi (first and second rows) and |𝐯||\mathbf{v}| (third and fourth rows) at different time points, for ζ=1\zeta=1 (first and third rows) and ζ=10\zeta=10 (second and fourth rows), in the case ϕ0=1.0(χB1+χB2+χB3+χB4)\phi_{0}=1.0(\chi_{B1}+\chi_{B2}+\chi_{B3}+\chi_{B4}).

We observe from Figure 7 that, for ζ=1\zeta=1, the initial circular clusters evolve, with interacting tips initially oriented along the bisecting directions of the plane and advected by the velocity field, similarly to the case in Figure 6. The clusters merge both horizontally and vertically, initially surrounding a circular region with ϕ0\phi\equiv 0 which dissolves at late times. The final equilibrium shape is given by a square. This is different from the results with linear isotropic inhomogeneous elasticity and periodic boundary conditions reported in [15, Figure 11], where the final equilibrium configuration is given by a cross. Also, in the case with ζ=10\zeta=10 the circular clusters interact only weakly and do not merge during the observed time window, deforming into a rectangular shape.

We highlight the fact that the numerical results shown in [15, Figures 6-7-11] concern interacting soft precipitates in a linear isotropic inhomogeneous elastic medium. In the case with hard precipitates, i.e. when the elastic modulus associated to the phase ϕ1\phi\equiv 1 is higher than the elastic modulus associated to the phase ϕ0\phi\equiv 0, the numerical results in [15] show a repulsion between the precipitates, which do not interact. This is comparable to what happens in our numerical simulations in the case of an high value of the elastic modulus.

7 Conclusion

In this paper we have proposed a new Cahn–Hilliard phase field model coupled to nonlinear incompressible finite viscoelasticity, where a new kind of diffusive regularization, of Allen–Cahn type, is introduced in the transport equation for the deformation gradient, together with a regularizing interface term depending on the gradient of the deformation gradient in the free energy density of the system. The designed regularization, which preserves the dissipative structure of the equations, helps in enhancing the space and time regularity of the deformation gradient. The resulting transport equation for the deformation gradient with Allen–Cahn type regularization was expressed in a dual mixed formulation, introducing a dual variable of the deformation gradient which enters also in the expression for the Cauchy stress tensor. Through a Galerkin approximation of the model equations and the introduction of truncated problems, where the polynomial growth of the elastic energy density is truncated to degree 44, we proved existence of a global in time weak solution in three space dimensions and for elastic energy densities which are coupled to the phase field variable and which possibly degenerate for some values of the phase field variable. Then, thanks to an iterative argument based on elliptic regularity bootstrap steps applied to the Allen–Cahn transport equation for the deformation gradient, we extended the existence result, passing to the limit for the truncation parameter tending to infinity, to the case of a Cahn–Hilliard potential and an elastic energy density with maximum allowed polynomial growths. In three space dimensions, we found maximum admissible polynomial growth degrees of 1010 for the Cahn–Hilliard potential and 66 for the elastic energy density, with the degree of the Cahn–Hilliard potential depending on the degree of the elastic energy density. We also proposed two kind of unconditionally energy stable and efficient finite element approximations of the model, based on convex splitting ideas and on the use of a scalar auxiliary variable, proving the existence and stability of discrete solutions. We finally showed numerical results for different test cases with shape memory alloy type free energies, characterized by different elastic properties of the pure phases of the phase field variable, which verify the gradient stability properties of the proposed schemes and show qualitatively how the topology of stationary states depends on both the phase separation and the elasticity dynamics. Future developments of the present work will investigate the cases with singular phase field potential and compressible elasticity, together with the generalization of the proposed model to models for biomathematical applications.

8 Acknowledgements

This research was supported by the Italian Ministry of Education, University and Research (MIUR): Dipartimenti di Eccellenza Program (2018–2022) – Dept. of Mathematics “F. Casorati”, University of Pavia. In addition, AA, PC and ER gratefully mention some other support from the MIUR-PRIN Grant 2020F3NCPX “Mathematics for industry 4.0 (Math4I4)” and their affiliation to the GNAMPA (Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni) of INdAM (Istituto Nazionale di Alta Matematica).

References

  • [1] S. Agmon. Lectures on elliptic boundary value problems. Providence, RI: AMS Chelsea Publishing, 2010.
  • [2] A. Agosti, P. Colli, H. Garcke, and E. Rocca. A Cahn-Hilliard model coupled to viscoelasticity with large deformations. 2022. https://https://arxiv.org/abs/2204.04951.
  • [3] J. W. Barrett and S. Boyaval. Existence and approximation of a (regularized) Oldroyd-B model. Math. Models Methods Appl. Sci., 21(9):1783–1837, 2011. doi: https://doi.org/10.1142/S0218202511005581.
  • [4] B. Benesova, J. Forster, C. Liu, and A. Schlömerkemper. Existence of weak solutions to an evolutionary model for magnetoelasticity. SIAM J. Math. Anal., 50(1):1200–1236, 2016. doi: https://doi.org/10.1137/17M1111486.
  • [5] H. Brezis and P. Mironescu. Gagliardo-Nirenberg inequalities and non-inequalities: the full story. Ann. Inst. H. Poincaré - Anal. Non Linéaire, 35:1355–1376, 2018. doi: https://doi.org/10.1016/j.anihpc.2017.11.007.
  • [6] A. Brunk and M. Lukácová-Medvidová. Global existence of weak solutions to viscoelastic phase separation: Part i regular case. Nonlinearity, 35(7):3417––3458, 2022. https://doi.org/10.1088/1361-6544/ac5920.
  • [7] Miroslav Bulíček, Josef Málek, Vít Průša, and Endre Süli. On incompressible heat-conducting viscoelastic rate-type fluids with stress-diffusion and purely spherical elastic response. SIAM J. Math. Anal., 53(4):3985–4030, 2021. doi: https://doi.org/10.1137/20M1384452.
  • [8] C. M. Elliott and A. M. Stuart. The global dynamics of discrete semilinear parabolic equations. SIAM J. Numer. Anal., 30(6):1622–1663, 1993. doi: https://doi.org/10.1137/0730084.
  • [9] D.J. Eyre. An unconditionally stable one-step scheme for gradient systems. 1998. http://www.math.utah.edu/~eyre/research/methods/stable.ps.
  • [10] H. Garcke, M. Hinze, and C. Kahle. A stable and linear time discretization for a thermodynamically consistent model for two-phase incompressible flow. Appl. Numer. Math., 99:151–171, 2016. doi: https://doi.org/10.1016/j.apnum.2015.09.002.
  • [11] H. Garcke, P. Knopf, S. Mitra, and A. Schlömerkemper. Strong well-posedness, stability and optimal control theory for a mathematical model for magneto-viscoelastic fluids. Calc. Var. Partial Differential Equations, 61(5):Paper No. 179, 52, 2022. https://doi.org/10.1007/s00526-022-02271-y.
  • [12] H. Garcke, B. Kovács, and D. Trautwein. Viscoelastic Cahn-Hilliard models for tumour growth. Math. Models Methods Appl. Sci., 2022. doi: https://doi.org/10.1142/S0218202522500634,.
  • [13] H. Garcke and K. F. Lam. Analysis of a Cahn–Hilliard system with non-zero dirichlet conditions modeling tumor growth with chemotaxis. Discrete Contin. Dyn. Syst., 37(8):42–77, 2018. doi: https://doi.org/10.3934/dcds.2017183.
  • [14] V. Girault and P. A. Raviart. Finite Element Methods for Navier-Stokes Equations. Springer-Verlag, Berlin, 1986. ISBN: 978-3-642-61623-5.
  • [15] P. H. Leo, J. S. Lowengrub, and H. J. Jou. A diffuse interface model for microstructural evolution in elastically stressed solids. Acta mater., 46(6):2113–2130, 1998. doi: https://doi.org/10.1016/S1359-6454(97)00377-7.
  • [16] X. Li and J. Shen. On fully decoupled MSAV schemes for the Cahn–Hilliard–Navier–Stokes model of two-phase incompressible flows. Math. Models Methods Appl. Sci., 32(3):457–495, 2022. doi: https://doi.org/10.1142/S0218202522500117.
  • [17] R. W. Ogden. Large deformation isotropic elasticity – on the correlation of theory and experiment for incompressible rubberlike solids. Proc. R. Soc. Lond. A, 326(1567):565–584, 1972. doi: https://doi.org/10.1098/rspa.1972.0026.
  • [18] T. Roubíček. Thermodynamics of viscoelastic solids, its Eulerian formulation, and existence of weak solutions. 2022. doi: https://arxiv.org/abs/2203.06080.
  • [19] T. Roubíček. Visco-elastodynamics at large strains Eulerian. Z. Angew. Math. Phys., 73(80), 2022. doi: https://doi.org/10.1007/s00033-022-01686-z.
  • [20] T. Roubíček and U. Stefanelli. Viscoelastodynamics of swelling porous solids at large strains by an eulerian approach. 2022. Preprint: https://arxiv.org/abs/2201.10959.
  • [21] T. Roubíček and G. Tomassetti. Thermodynamics of shape-memory alloys under electric current. Z. Angew. Math. Phys., 61:1–20, 2010. doi: https://doi.org/10.1007/s00033-009-0007-1.
  • [22] C. Taylor and P. Hood. A numerical solution of the Navier–Stokes equations using the finite element technique. Comput. & Fluids, 1(1):73–100, 1973. doi: https://doi.org/10.1016/0045-7930(73)90027-3.
  • [23] R. Temam. Navier-Stokes Equations: Theory and Numerical Analysis. North-Holland Publishing Company, Amsterdam, New York, Oxford, 1977. ISBN: 0 444 85307 3.
  • [24] Liu. Y and D. Trautwein. On a diffuse interface model for incompressible viscoelastic two-phase flows. 2022. Preprint: https://arxiv.org/abs/2212.13507.