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

On various models describing behaviour of thermoviscoelastic rate-type fluids

Miroslav Bulíček Mathematical Institute, Faculty of Mathematics and Physics, Charles University, Sokolovská 83, 186 75, Prague, Czech Republic mbul8060@karlin.mff.cuni.cz  and  Jakub Woźnicki Faculty of Mathematics, Informatics and Mechanics, University of Warsaw, Stefana Banacha 2, 02-097 Warsaw, Poland; Institute of Mathematics of Polish Academy of Sciences, Jana i Jędrzeja Śniadeckich 8, 00-656 Warsaw, Poland jw.woznicki@student.uw.edu.pl
Abstract.

Viscoelastic rate-type fluid models are essential for describing the behavior of a wide range of complex materials, with applications in fields such as engineering, biomaterials, and medicine. These models are particularly useful for understanding the rheological properties of materials that exhibit both elastic and viscous behavior under deformation. However, many real-world applications involve significant thermal effects, where heat conduction and the temperature dependence of material properties must also be considered. In this paper, we introduce a thermodynamically consistent model for heat-conducting viscoelastic rate-type fluids and establish the existence of a global weak solution in a two-dimensional setting. The result holds under the condition that the initial energy and entropy are controlled in appropriate natural norms.

Key words and phrases:
non-Newtonian fluids, Giesekus model, thermo-visco-elasticity, global-in-time and large-data existence theory
1991 Mathematics Subject Classification:
35A01, 35Q35, 76A10, 76D03
Miroslav Bulíček was supported by the project No. 20-11027X financed by GAČR. He is a member of the Nečas Center for Mathematical Modeling. ORCID: 0000-0003-2380-3458; corresponding author
Jakub Woźnicki was supported by National Science Center, Poland through the project no. 2023/49/N/ST1/02737. ORCID: 0000-0002-7720-5261.

1. Introduction

Our main goal in this paper is to study the equations describing the motion of incompressible viscoelastic and heat conducting fluids and in particular to establish the long-time and the large-data theory. We focus here only on the planar case, i.e., in what follows, the fluid occupies the Lipschitz domain Ω2\Omega\subset\mathbb{R}^{2} and T>0T>0 always denotes the length of time interval. The motion of incompressible fluid is in general described by the system of partial differential equations of the form

(1.1) t𝒗+divx(𝒗𝒗)divx𝕋=𝟎,divxu=0,\begin{split}\partial_{t}\boldsymbol{v}+\operatorname{div}_{x}(\boldsymbol{v}\otimes\boldsymbol{v})-\operatorname{div}_{x}\mathbb{T}=\boldsymbol{0},\\ \operatorname{div}_{x}u=0,\end{split}

that is supposed to be satisfied in ΩT:=(0,T)×Ω\Omega_{T}:=(0,T)\times\Omega. Here, 𝒗:ΩT2\boldsymbol{v}:\Omega_{T}\to\mathbb{R}^{2} is the unknown velocity field and 𝕋:ΩTsym2×2\mathbb{T}:\Omega_{T}\to\mathbb{R}^{2\times 2}_{\text{sym}} denotes the Cauchy stress tensor, and the density of the fluid is set to be equal to one for simplicity. The above system must be completed with the initial and boundary conditions, but most importantly, the constitutive relationship for 𝕋\mathbb{T} must be prescribed. We shall consider the classical form

(1.2) 𝕋:=p𝕀+2ν𝔻𝒗+2g(𝔹𝕀),\mathbb{T}:=-p\mathbb{I}+2\nu\mathbb{D}\boldsymbol{v}+2g(\mathbb{B}-\mathbb{I}),

where pp is the unknown pressure, 𝔻𝒗:=12((𝒗)+(𝒗)T)\mathbb{D}\boldsymbol{v}:=\frac{1}{2}((\nabla\boldsymbol{v})+(\nabla\boldsymbol{v})^{T}) is the symmetric velocity gradient and 𝔹:ΩTsym2×2\mathbb{B}:\Omega_{T}\to\mathbb{R}^{2\times 2}_{\text{sym}} is the so-called extra stress. The material parameter ν\nu - the viscosity, and gg - the shear modulus are in what follows the functions of the temperature. We will specify the particular assumptions below.

The first two terms on the right-hand side of (1.2) describe the classical Newtonian fluid, but the last term, extra stress, represents the elastic effects. Hence, we need to add an equation for 𝔹\mathbb{B} and we shall assume the so-called rate-type models. This means that we complete the system (1.1) by the following system of equations

(1.3) 𝔹+δ(𝔹2𝔹)+γ(𝔹𝕀)=𝕆,\overset{\nabla}{\mathbb{B}}+\delta(\mathbb{B}^{2}-\mathbb{B})+\gamma(\mathbb{B}-\mathbb{I})=\mathbb{O},

where δ\delta and γ\gamma are material parameters related to the relaxation times in the material and possibly depending on the temperature. The symbol 𝔹\overset{\nabla}{\mathbb{B}} denotes the Oldroyd upper convective derivative defined as

(1.4) 𝔹:=t𝔹+(𝒗x)𝔹x𝒗𝔹𝔹(x𝒗)T.\displaystyle\overset{\nabla}{\mathbb{B}}:=\partial_{t}\mathbb{B}+(\boldsymbol{v}\cdot\nabla_{x})\,\mathbb{B}-\nabla_{x}\boldsymbol{v}\,\mathbb{B}-\mathbb{B}\,(\nabla_{x}\boldsymbol{v})^{T}.

Note that this is the often-used objective derivative applied to 𝔹\mathbb{B} in the context of rate-type models.

The system (1.1)–(1.3) with δ:=0\delta:=0 and γ>0\gamma>0 is generally referred to as the Oldroyd-B model, see [25], and belongs to one of the classical models of viscoelasticity. However, any mathematical theory which would lead to the global well-posed theory is missing. The other model with γ:=0\gamma:=0 and δ>0\delta>0 proposed by Giesekus, see [19], is also widely used and, contrary to the Oldroyd-B model, it seems to have much better mathematical properties. Note also that beside the upper convective derivative (1.4), one can also consider a much more general class of objectives derivatives, see [22], but it usually does not affect the qualitative analysis. Similarly, one may also consider models of second order, see [11], but again, once we are able to deal with the classical Giesekus model, it seems that further generalisations are rather straightforward. Therefore, we do not consider them here and rather focus on canonical problems. Hence, the system we are interested in reads

(1.5) {t𝒗+divx(𝒗𝒗)divx𝕋=0,divx𝒗=0,𝔹+δ(θ)(𝔹2𝔹)=𝕆,p𝕀+2ν(θ)𝔻𝒗+2g(θ)(𝔹𝕀)=𝕋,\left\{\begin{aligned} &\partial_{t}\boldsymbol{v}+\operatorname{div}_{x}(\boldsymbol{v}\otimes\boldsymbol{v})-\operatorname{div}_{x}\mathbb{T}=0,\\ &\operatorname{div}_{x}\boldsymbol{v}=0,\\ &\overset{\nabla}{\mathbb{B}}+\delta(\theta)(\mathbb{B}^{2}-\mathbb{B})=\mathbb{O},\\ &-p\mathbb{I}+2\nu(\theta)\mathbb{D}\boldsymbol{v}+2g(\theta)(\mathbb{B}-\mathbb{I})=\mathbb{T},\end{aligned}\right.

and we assume that all material coefficients may depend on the temperature θ:ΩT\theta:\Omega_{T}\to\mathbb{R}. Once we have added thermal effects into the system (the dependence of coefficients on the temperature), we must also impose the validity of the first law of thermodynamics, i.e. the conservation of energy. One of the possible equivalent form (valid for sufficiently smooth solution) is the equation for the internal energy e:ΩTe:\Omega_{T}\to\mathbb{R}, where we assume that the heat flux is given by the Fourier law

(1.6) te+divx(𝒗e)divx(κ(θ)xθ)=𝕋:𝔻𝒗,\displaystyle\partial_{t}e+\operatorname{div}_{x}(\boldsymbol{v}e)-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}\theta)=\mathbb{T}:\mathbb{D}\boldsymbol{v},

where κ\kappa denotes the heat conductivity. Note here that the symbol 𝔸:𝔹\mathbb{A}:\mathbb{B} denotes the matrix scalar product 𝔸:𝔹=i,j=12𝔸ij𝔹ij\mathbb{A}:\mathbb{B}=\sum_{i,j=1}^{2}\mathbb{A}_{ij}\mathbb{B}_{ij}.

It remains to specify the storage and dissipative mechanisms of the fluid, or in the purely mathematical sense, it remains to specify the relation between the internal energy ee, the temperature θ\theta and the tensor 𝔹\mathbb{B}. Further, this relation must be such that one can identify an entropy of the system, which fulfils the second law of thermodynamics. Here, we closely follow the approach developed in [26] and in [20], where the Helmholtz free energy plays the crucial role. We may also refer to [15] for the first systematic approach or to [21], where the entropy estimates were derived for the first time.

Hence, we assume here that the Helmholtz free energy is given by

(1.7) ψ(θ,𝔹)=cvθ(lnθ1)+g(θ)f(𝔹),\displaystyle\psi(\theta,\mathbb{B})=-c_{v}\theta(\ln\theta-1)+g(\theta)f(\mathbb{B}),

where the function f:sym2×2f:\mathbb{R}^{2\times 2}_{\text{sym}}\to\mathbb{R} is given by the formula

(1.8) f(𝔹)=tr𝔹2lndet𝔹.\displaystyle f(\mathbb{B})=\operatorname{tr}\mathbb{B}-2-\ln\det\mathbb{B}.

The first term in (1.7) corresponds to the classical law when e=cvθe=c_{v}\theta with cvc_{v} being the heat capacity, and the second term connects the elastic and the temperature effects. Note that if gg is a constant function, then this choice is very much classical in the theory of viscoelastic rate-type fluids. However, here, we deal with a more general setting similar to that of [2]. Having introduced the free energy, we define the entropy η\eta as

η=η(θ,𝔹):=θψ(θ,𝔹)=cvlnθg(θ)f(𝔹),\eta=\eta(\theta,\mathbb{B}):=-\partial_{\theta}\psi(\theta,\mathbb{B})=c_{v}\ln\theta-g^{\prime}(\theta)f(\mathbb{B}),

and the internal energy is then related by the formula

(1.9) e=e(θ,𝔹):=ψ(θ,𝔹)+θη(θ,𝔹)=cvθ+(g(θ)θg(θ))f(𝔹).\displaystyle e=e(\theta,\mathbb{B}):=\psi(\theta,\mathbb{B})+\theta\eta(\theta,\mathbb{B})=c_{v}\theta+(g(\theta)-\theta\,g^{\prime}(\theta))f(\mathbb{B}).

Using these definitions, we see that

te\displaystyle\partial_{t}e =tψ+tθη+θtη=tθ(θψ+η)+𝔹ψt𝔹+θtη=𝔹ψt𝔹+θtη\displaystyle=\partial_{t}\psi+\partial_{t}\theta\,\eta+\theta\,\partial_{t}\eta=\partial_{t}\theta(\partial_{\theta}\psi+\eta)+\partial_{\mathbb{B}}\psi\partial_{t}\mathbb{B}+\theta\partial_{t}\eta=\partial_{\mathbb{B}}\psi\partial_{t}\mathbb{B}+\theta\partial_{t}\eta
tη=te𝔹ψt𝔹θ=tet𝔹:g(θ)(𝕀𝔹1)θ,\displaystyle\implies\partial_{t}\eta=\frac{\partial_{t}e-\partial_{\mathbb{B}}\psi\partial_{t}\mathbb{B}}{\theta}=\frac{\partial_{t}e-\partial_{t}\mathbb{B}:g(\theta)(\mathbb{I}-\mathbb{B}^{-1})}{\theta},

where we used the following identity

𝔹ψ(θ,𝔹)=g(θ)𝔹f(𝔹)=g(θ)(𝕀𝔹1).\partial_{\mathbb{B}}\psi(\theta,\mathbb{B})=g(\theta)\partial_{\mathbb{B}}f(\mathbb{B})=g(\theta)(\mathbb{I}-\mathbb{B}^{-1}).

Thus, dividing (1.6) by θ\theta, taking the scalar product of the third equation in (1.5) with g(θ)(𝕀𝔹1)θ\frac{g(\theta)(\mathbb{I}-\mathbb{B}^{-1})}{\theta} and subtracting the result, we obtain the following identity for the entropy

(1.10) tη+divx(η𝒗)divx(κ(θ)x(lnθ))=κ(θ)|xθ|2θ2+2ν(θ)|𝔻𝒗|2θ+δ(θ)|𝔹𝕀|2θ0.\displaystyle\partial_{t}\eta+\operatorname{div}_{x}(\eta\,\boldsymbol{v})-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}(\ln\theta))=\frac{\kappa(\theta)|\nabla_{x}\theta|^{2}}{\theta^{2}}+\frac{2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}}{\theta}+\frac{\delta(\theta)|\mathbb{B}-\mathbb{I}|^{2}}{\theta}\geq 0.

Here, it is essential that the right-hand side is nonnegative and the term κ(θ)x(lnθ)\kappa(\theta)\nabla_{x}(\ln\theta) represents the entropy flux. It also directly follows from the above procedure that the equation for the internal energy (1.6) and the equation for the entropy (1.10) are interchangeable in (1.5) for sufficiently regular solutions. This “equivalence” of (1.6) and (1.10) is also later used not only for the analysis of the problem but also for the definition of a notion of solution.

Finally, we always assume that 𝔹\mathbb{B} is of the form

𝔹=𝔽𝔽T,𝔽2×2,det𝔽>0.\displaystyle\mathbb{B}=\mathbb{F}\,\mathbb{F}^{T},\quad\mathbb{F}\in\mathbb{R}^{2\times 2},\quad\det\mathbb{F}>0.

The above identification can be guaranteed, by imposing the following equation for 𝔽\mathbb{F}

(1.11) t𝔽+divx(𝔽𝒗)x𝒗𝔽+12δ(θ)(𝔽𝔽T𝔽𝔽)=0.\partial_{t}\mathbb{F}+\operatorname{div}_{x}(\mathbb{F}\otimes\boldsymbol{v})-\nabla_{x}\boldsymbol{v}\,\mathbb{F}+\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F})=0.

In fact, multiplying this identity by 𝔽T\mathbb{F}^{T} on the right, then transposing the equation (1.11) and multiplying the result by 𝔽\mathbb{F} from the left, adding the resulting identity together, we finally observe that 𝔹:=𝔽𝔽T\mathbb{B}:=\mathbb{F}\mathbb{F}^{T} then satisfies (1.3) with γ=0\gamma=0.

We have already introduced the complete set of equations and we close the problem by imposing the following boundary and initial conditions

(1.12) 𝒗|Ω=0,𝒗(0,x)=𝒗0(x),𝔹(0,x)=𝔹0=𝔽0𝔽0T,xθ𝐧=0,θ(0,x)=θ0.\displaystyle\boldsymbol{v}|_{\partial\Omega}=0,\quad\boldsymbol{v}(0,x)=\boldsymbol{v}_{0}(x),\quad\mathbb{B}(0,x)=\mathbb{B}_{0}=\mathbb{F}_{0}\,\mathbb{F}_{0}^{T},\quad\nabla_{x}\theta\cdot\mathbf{n}=0,\quad\theta(0,x)=\theta_{0}.

Furthermore, the initial velocity is solenoidal and have bounded kinetic energy and the internal initial energy and that the initial entropy are also under control, i.e.,

𝒗02,𝔽02,tr𝔹02lndet𝔹01,lndet𝔹0,θ01,lnθ01<+,\displaystyle\|\boldsymbol{v}_{0}\|_{2},\,\|\mathbb{F}_{0}\|_{2},\,\|\operatorname{tr}\mathbb{B}_{0}-2-\ln\det\mathbb{B}_{0}\|_{1},\,\|\ln\det\mathbb{B}_{0}\|_{\infty},\,\|\theta_{0}\|_{1},\|\ln\theta_{0}\|_{1}<+\infty,
det𝔽0>0,divx𝒗00 a.e. in Ω.\displaystyle\det\mathbb{F}_{0}>0,\qquad\operatorname{div}_{x}\boldsymbol{v}_{0}\equiv 0\textrm{ a.e. in }\Omega.

Finally, we introduce three specific choices of the form of the Helmholtz free energy and in particular the function gg appearing in (1.7):

  1. (P1)

    g(θ)const>0g(\theta)\equiv const>0: In this case the entropy depends only on the temperature and has the form

    η=cvlnθ.\eta=c_{v}\ln\theta.

    Furthermore, one has a relatively good equation for the temperature θ\theta, namely multiplying (1.10) by θ\theta, we obtain

    (1.13) cvtθ+cvdivx(θ𝒗)divx(κ(θ)xθ)=2ν(θ)|𝔻𝒗|2+δ(θ)|𝔹𝕀|2.\displaystyle c_{v}\partial_{t}\theta+c_{v}\operatorname{div}_{x}(\theta\,\boldsymbol{v})-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}\theta)=2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{B}-\mathbb{I}|^{2}.
  2. (P2)

    g(θ)=g~θg(\theta)=\tilde{g}\theta and g~+\tilde{g}\in\mathbb{R}_{+}: In this case, the internal energy ee depends only on the temperature θ\theta and is of the form

    e=cvθ.e=c_{v}\theta.

    Therefore the equation for the internal energy (1.6) is identical to the evolutionary equation for the temperature.

  3. (P3)

    g:++g:\mathbb{R}_{+}\to\mathbb{R}_{+} is continuous, nonnegative and concave: In this case, we do not have any simple form of the equation for the internal energy and the equations connecting temperature and the elastic tensor are coupled in a non-trivial way. The evolutionary equation for θ\theta depends nonlinearly on 𝔹\mathbb{B} and also ee.

Our aim is to develop the theory related to the three cases above, in particular, to the most difficult case (P3). To simplify the presentation, we prove the weak sequential stability of solutions to (P3), see Section 3. Although it is not the detail proof of existence, it is clear that the complete proof can be established. This fact is very much supported by our second result of the paper, that is connected to the case (P1) and is presented in Section 4, where we show the existence of weak solution. Section 2 is devoted to the basic notation, and in Appendix A we state some classical auxiliary propositions and recall several useful facts.

To end this introductory part, we recall available results and emphasise the key novelty of the paper. We start by recalling the results without thermal effects. As we have already mentioned, the result for the full Oldroyd-B model is an open problem. The only available result (for global solutions) is due to Lions and Masmoudi [23], where the existence is shown even in three dimensional setting but only for the so-called co-rotational case. On the other hand, for the Giesekus model (or for even more general class), the first serious attempt for global theory was done in [24]. Although the key idea there is correct, there were still some gaps in the proof, that had to be corrected. The first rigorous result for such a class of fluids was established in [2] with one proviso, an additional stress diffusion term Δ𝔹\Delta\mathbb{B} was added to the problem, which simplified the analysis a little bit. The problem without stress diffusion was finally solved in [7] for the two-dimensional setting and later in [8] for the three-dimensional setting and for a much more complex class of models. For compressible setting, we refer also to a particular result in  [5]. For problems with temperature, the situation is more delicate. Even for problems without viscoelastic effects, the theory is relatively new, and here we refer to [12, 4, 10] for relevant results. The problems with elastic effects and temperature were first rigorously treated in [9], where, however, only spherical stresses were assumed, i.e. 𝔹=b𝕀\mathbb{B}=b\mathbb{I}. Much more general results were established in [3], where the authors again assumed a modification of the Giesekus model by adding stress diffusion into the system and focussing only on the linear case (P2). The key novelty of the paper is that we do not need any stress diffusion term in the equation or the simple form of the Helmholtz free energy. Furthermore, it is clear from the proof that we are also able to cover the three-dimensional case and also a more complex model - to such a setting will be devoted our forth-coming paper.

To end this part, it is essential to mention that the paper heavily relies not only on the results mentioned above, but also on the following two very classical techniques. First, we should mention the theory of renormalisation for the transport equation developed in [14], which is used when dealing with the equation for 𝔹\mathbb{B}. Second, the compensated compactness method applied to heat conductive fluids, which was introduced by Feireisl, see [16] and further generalised and extended to other cases, see [17, 18] and [6]. These methods are used to deduce the compactness of temperature θ\theta and tensor 𝔹\mathbb{B}.

To finish this introductory part, we recall the system in which we are interested. The above equations and constraints in (1.5)–(1.11) can be written in one form as the equations for the unknowns (𝒗,θ,𝔽,p,𝕋,e)(\boldsymbol{v},\theta,\mathbb{F},p,\mathbb{T},e)

(1.14) {t𝒗+divx(𝒗𝒗)divx𝕋=0,divx𝒗=0,t𝔽+divx(𝔽𝒗)x𝒗𝔽+12δ(θ)(𝔽𝔽T𝔽𝔽)=𝕆,te+divx(𝒗e)divx(κ(θ)xθ)=(2ν(θ)𝔻𝒗+2g(θ)𝔽𝔽T):𝔻𝒗,𝕋=p𝕀+2ν(θ)𝔻𝒗+2g(θ)(𝔽𝔽T),e=θ+(g(θ)θg(θ))f(𝔹).\left\{\begin{aligned} &\partial_{t}\boldsymbol{v}+\operatorname{div}_{x}(\boldsymbol{v}\otimes\boldsymbol{v})-\operatorname{div}_{x}\mathbb{T}=0,\qquad\operatorname{div}_{x}\boldsymbol{v}=0,\\ &\partial_{t}\mathbb{F}+\operatorname{div}_{x}(\mathbb{F}\otimes\boldsymbol{v})-\nabla_{x}\boldsymbol{v}\,\mathbb{F}+\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F})=\mathbb{O},\\ &\partial_{t}e+\operatorname{div}_{x}(\boldsymbol{v}e)-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}\theta)=(2\nu(\theta)\mathbb{D}\boldsymbol{v}+2g(\theta)\mathbb{F}\,\mathbb{F}^{T}):\mathbb{D}\boldsymbol{v},\\ &\mathbb{T}=-p\mathbb{I}+2\nu(\theta)\mathbb{D}\boldsymbol{v}+2g(\theta)(\mathbb{F}\,\mathbb{F}^{T}),\qquad e=\theta+(g(\theta)-\theta\,g^{\prime}(\theta))f(\mathbb{B}).\end{aligned}\right.

with the initial and boundary conditions

(1.15) 𝒗|Ω=0,xθ𝐧|Ω=0,𝒗(0,x)=𝒗0(x),𝔽(0,x)=𝔽0,θ(0,x)=θ0.\displaystyle\boldsymbol{v}|_{\partial\Omega}=0,\quad\nabla_{x}\theta\cdot\mathbf{n}|_{\partial\Omega}=0,\quad\boldsymbol{v}(0,x)=\boldsymbol{v}_{0}(x),\quad\mathbb{F}(0,x)=\mathbb{F}_{0},\quad\theta(0,x)=\theta_{0}.

The key result of the paper, without rigorous definitions, can then be formulated as follows. For rigorous statements, we refer the reader to Sections 3 and 4.

Theorem.

Let gg in (P3) be properly chosen. Then for the relevant initial data, there is a weak global-in-time solution to (1.14)–(1.15).

The notion of a weak global-in-time solution here is also a key concept. While, we always deal with distributional solutions to (1.14)1–(1.14)2. Equation (1.14)3 can hardly be obtained in a general setting. Therefore, we relax a notion of a solution to (1.14)3 such that, we require only the entropy inequality, i.e., we use (1.10) with the inequality sign

(1.16) tη+divx(η𝒗)divx(κ(θ)x(lnθ))κ(θ)|xθ|2θ2+2ν(θ)|𝔻𝒗|2θ+δ(θ)|𝔹𝕀|2θ\displaystyle\partial_{t}\eta+\operatorname{div}_{x}(\eta\,\boldsymbol{v})-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}(\ln\theta))\geq\frac{\kappa(\theta)|\nabla_{x}\theta|^{2}}{\theta^{2}}+\frac{2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}}{\theta}+\frac{\delta(\theta)|\mathbb{B}-\mathbb{I}|^{2}}{\theta}

and we complete the system by requiring the conservation of the total energy, which means that we require

(1.17) ddtΩ|𝒗|22+edx=0.\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\frac{|\boldsymbol{v}|^{2}}{2}+e\mathop{}\!\mathrm{d}x=0.

It is worth noticing, that when e,η,𝒗e,\eta,\boldsymbol{v} are sufficiently regular and satisfy (1.16) and (1.17) then they satisfy (1.14)3 as well. For such an approach, we refer the interested reader to [16].

2. Preliminaries

In this paper, xΩdx\in\Omega\subset\mathbb{R}^{d} always denotes the spatial variable, while t(0,T)t\in(0,T) is reserved for the time variable. Through the paper, we consider mostly the dimension d=2d=2, but in several technical tools we also recall the general case d2d\geq 2. The scalar functions are written in italics, e.g. aa\in\mathbb{R}, the vector-valued objects are written in bold face, e.g. 𝒂d\boldsymbol{a}\in\mathbb{R}^{d}, the matrices by capital bold face, e.g. 𝔸d×d\mathbb{A}\in\mathbb{R}^{d\times d} and the third order tensor as 𝔄d×d×d\mathfrak{A}\in\mathbb{R}^{d\times d\times d}. In addition, to simplify the notation, we write 𝒂𝒃\boldsymbol{a}\cdot\boldsymbol{b} for a standard scalar product whenever 𝒂,𝒃d\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{d}. Similarly, by 𝔸:𝔹\mathbb{A}:\mathbb{B}, we denote the scalar product between two matrices 𝔸,𝔹d×d\mathbb{A},\mathbb{B}\in\mathbb{R}^{d\times d}, and their classical matrix product by 𝔸𝔹\mathbb{A}\,\mathbb{B}. Finally, by the symbol 𝔄𝔅\mathfrak{A}\because\mathfrak{B}, we will denote the scalar product between 𝔄,𝔅d×d×d\mathfrak{A},\mathfrak{B}\in\mathbb{R}^{d\times d\times d}. In addition, the symbol \otimes is reserved for the tensorial product, i.e., for 𝒂,𝒃d\boldsymbol{a},\boldsymbol{b}\in\mathbb{R}^{d} we denote 𝒂𝒃d×d\boldsymbol{a}\otimes\boldsymbol{b}\in\mathbb{R}^{d\times d} as (𝒂𝒃)ij:=aibj(\boldsymbol{a}\otimes\boldsymbol{b})_{ij}:=a_{i}b_{j} for i,j=1,,di,j=1,\ldots,d. For a matrix 𝔸=(Aij)i,j=1d\mathbb{A}=(A_{ij})_{i,j=1}^{d} and a vector 𝒃=(b1,,bd)\boldsymbol{b}=(b_{1},\ldots,b_{d}) we define the third order tensorial product as

(𝔸𝒃)ijk=Aijbk,(\mathbb{A}\otimes\boldsymbol{b})_{ijk}=A_{ij}\,b_{k},

and its divergence as

divx(𝔸𝒃)=j=1dxj(bj𝔸).\operatorname{div}_{x}(\mathbb{A}\otimes\boldsymbol{b})=\sum_{j=1}^{d}\partial_{x_{j}}(b_{j}\,\mathbb{A}).

We use the standard notation for the Sobolev and Lebesgue function space, and we frequently do not distinguish between their scalar-, vector-, or matrix-valued variants. In addition, to shorten the notation, we frequently use the following simplifications. When fLp(Ω)f\in L^{p}(\Omega), we simplify it to fLxpf\in L^{p}_{x}. Similarly, if fLp(0,T;Lq(Ω))f\in L^{p}(0,T;L^{q}(\Omega)), fLp(0,T;W1,q(Ω))f\in L^{p}(0,T;W^{1,q}(\Omega)) or fLp(0,T;W01,q(Ω))f\in L^{p}(0,T;W^{1,q}_{0}(\Omega)), then we write fLtpLxqf\in L^{p}_{t}L^{q}_{x}, fLtpWx1,qf\in L^{p}_{t}W^{1,q}_{x} or fLtpW0,x1,qf\in L^{p}_{t}W^{1,q}_{0,x} respectively (here, W1,q(Ω)W^{1,q}(\Omega) and W01,q(Ω)W^{1,q}_{0}(\Omega) are the usual Sobolev spaces). In addition, to emphasise spaces with zero divergence, we define W0,div1,p:={𝒗𝒞1(Ω;d),divx𝒗=0}¯1,pW^{1,p}_{0,\operatorname{div}}:=\overline{\{\boldsymbol{v}\in\mathcal{C}^{1}(\Omega;\mathbb{R}^{d}),\quad\operatorname{div}_{x}\boldsymbol{v}=0\}}^{\|\cdot\|_{1,p}} and also L0,divp:={𝒗𝒞1(Ω;d),divx𝒗=0}¯pL^{p}_{0,\operatorname{div}}:=\overline{\{\boldsymbol{v}\in\mathcal{C}^{1}(\Omega;\mathbb{R}^{d}),\quad\operatorname{div}_{x}\boldsymbol{v}=0\}}^{\|\cdot\|_{p}} for any p[1,)p\in[1,\infty). Further, the dual spaces to Sobolev spaces and their subset are defined as usual, e.g. W1,p(Ω):=(W1,p(Ω))W^{-1,p^{\prime}}(\Omega):=(W^{1,p}(\Omega))^{*}, W01,p(Ω):=(W01,p(Ω))W^{-1,p^{\prime}}_{0}(\Omega):=(W^{1,p}_{0}(\Omega))^{*}, W0,div1,p(Ω):=(W0,div1,p(Ω))W^{-1,p^{\prime}}_{0,\operatorname{div}}(\Omega):=(W^{1,p}_{0,\operatorname{div}}(\Omega))^{*}, etc. Here, for p[1,+]p\in[1,+\infty] we denote by pp^{\prime} its Hölder conjugate. Also in what follows, by the symbol C>0C>0 we denote a generic constant, that can change line to line and will depend only on data. In case, there is any essential dependence on other quantities, it will be clearly described.

Now, we introduce the assumptions regarding the function ν,κ,δ,g\nu,\kappa,\delta,g appearing in equations (1.5) - (1.10). We assume that the viscosity ν:+\nu:\mathbb{R}_{+}\to\mathbb{R} and the heat conductivity κ:\kappa:\mathbb{R}\to\mathbb{R} are continuous functions fulfilling for some positive constants C1,C2>0C_{1},C_{2}>0 and for all s0s\geq 0

(2.1) C1\displaystyle C_{1}\leq κ(s)C2,\displaystyle\kappa(s)\leq C_{2},
(2.2) C1\displaystyle C_{1}\leq ν(s)C2,.\displaystyle\nu(s)\leq C_{2},.

Concerning the parameter δ:\delta:\mathbb{R}\to\mathbb{R}, we also assume that it is a continuous function, but we need to distinguish between the cases (P1) and (P3). For (P1), we assume that for all s0s\geq 0 there holds

(2.3) C1δ(s)C2,\displaystyle C_{1}\leq\delta(s)\leq C_{2},

while in case (P3), we require that for all s0s\geq 0 there holds

(2.4) C1(1+s)δ(s)C2(1+s).\displaystyle C_{1}(1+s)\leq\delta(s)\leq C_{2}(1+s).

Finally, for the most general case (P3), we need to specify the assumptions on the function gg, which read as follows: We assume that g:g:\mathbb{R}\to\mathbb{R} is a 𝒞2\mathcal{C}^{2} function, which is concave and increasing. In addition, we assume that for all s[0,)s\in[0,\infty) and all λ[0,1]\lambda\in[0,1] there holds

(2.5) C1\displaystyle C_{1} g(s)C2,\displaystyle\leq g(s)\leq C_{2},
(2.6) 0\displaystyle 0 (1+s)g(s)C2.\displaystyle\leq(1+s)g^{\prime}(s)\leq C_{2}.

We finish this introductory part by defining the auxiliary function hλh_{\lambda}, where the parameter λ[0,1]\lambda\in[0,1]. We set

(2.7) hλ(s):=0szλg′′(z)dz.h_{\lambda}(s):=\int_{0}^{s}z^{\lambda}g^{\prime\prime}(z)\mathop{}\!\mathrm{d}z.

Note that it directly follows from the concavity of gg that hλh_{\lambda} is nonpositive and fulfills

(2.8) |hλ(s)|\displaystyle|h_{\lambda}(s)| =0szλg′′(z)dz0s(1+z)g′′(z)dz=(1+s)g(s)+g(0)+0sg(z)dz\displaystyle=-\int_{0}^{s}z^{\lambda}g^{\prime\prime}(z)\mathop{}\!\mathrm{d}z\leq-\int_{0}^{s}(1+z)g^{\prime\prime}(z)\mathop{}\!\mathrm{d}z=-(1+s)g^{\prime}(s)+g^{\prime}(0)+\int_{0}^{s}g^{\prime}(z)\mathop{}\!\mathrm{d}z
=(1+s)g(s)+g(0)+g(s)g(0)2C2,\displaystyle=-(1+s)g^{\prime}(s)+g^{\prime}(0)+g(s)-g(0)\leq 2C_{2},

where we used the assumptions (2.5)–(2.6).

3. Weak sequential stability of strong solutions for the case (P3)

We first formulate a rigorous version of the main theorem of the paper. It deals with the most general case (P3) and the sequential stability of the solutions.

Theorem 3.1.

Let {𝐯n,𝔽n,θn}n=1𝒞1((0,T)×Ω;2)×𝒞1((0,T)×Ω;2×2)×𝒞1,2((0,T)×Ω;+)\{\boldsymbol{v}_{n},\mathbb{F}_{n},\theta_{n}\}_{n=1}^{\infty}\subset\mathcal{C}^{1}((0,T)\times\Omega;\mathbb{R}^{2})\times\mathcal{C}^{1}((0,T)\times\Omega;\mathbb{R}^{2\times 2})\times\mathcal{C}^{1,2}((0,T)\times\Omega;\mathbb{R}_{+}) with det𝔽n,θn>0\det\mathbb{F}_{n},\theta_{n}>0 be a sequence of solutions to (1.14)–(1.15) with initial conditions {𝐯0n,θ0n,𝔽0n}n=1\{\boldsymbol{v}_{0}^{n},\theta_{0}^{n},\mathbb{F}_{0}^{n}\}_{n=1}^{\infty} fulfilling det𝔽0n,θ0n>0\det\mathbb{F}_{0}^{n},\theta_{0}^{n}>0 and

(3.1) 𝒗0n\displaystyle\boldsymbol{v}_{0}^{n} 𝒗0\displaystyle\to\boldsymbol{v}_{0} strongly in L0,div2,\displaystyle\textrm{strongly in }L^{2}_{0,\operatorname{div}},
θ0n\displaystyle\theta_{0}^{n} θ0\displaystyle\to\theta_{0} strongly in Lx1,\displaystyle\textrm{strongly in }L^{1}_{x},
lnθ0n\displaystyle\ln\theta_{0}^{n} lnθ0\displaystyle\to\ln\theta_{0} strongly in Lx1,\displaystyle\textrm{strongly in }L^{1}_{x},
𝔽0n\displaystyle\mathbb{F}_{0}^{n} 𝔽0\displaystyle\to\mathbb{F}_{0} strongly in Lx2,\displaystyle\textrm{strongly in }L^{2}_{x},
lndetF0n\displaystyle\ln\det F_{0}^{n} lndet𝔽0\displaystyle\to\ln\det\mathbb{F}_{0} strongly in Lx2.\displaystyle\textrm{strongly in }L^{2}_{x}.

Then, there exists a subsequence (which we do not relabel) and a triple (𝐯,𝔽,θ)(\boldsymbol{v},\mathbb{F},\theta) such that: we have the following convergence results for 𝐯n\boldsymbol{v}_{n}:

(3.2) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\overset{*}{\rightharpoonup}\boldsymbol{v} weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v} weakly in Lt2W0,x1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{0,x},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v} weakly in Lt,x4,\displaystyle\text{ weakly in }L^{4}_{t,x},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightarrow\boldsymbol{v} strongly in 𝒞([0,T];L0,div2),\displaystyle\text{ strongly in }\mathcal{C}([0,T];L^{2}_{0,\operatorname{div}}),

the following convergence results for 𝔽n\mathbb{F}_{n}:

(3.3) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\overset{*}{\rightharpoonup}\mathbb{F} weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightharpoonup\mathbb{F} weakly in Lt,x4,\displaystyle\text{ weakly in }L^{4}_{t,x},
𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightarrow\mathbb{F} strongly in Lt,xp,1p<4,\displaystyle\text{ strongly in }L^{p}_{t,x},1\leq p<4,

the following convergence results for θn\theta_{n}:

(3.4) θn\displaystyle\theta_{n} θ\displaystyle\rightarrow\theta strongly in Lt,x2ε,ε(0,1),\displaystyle\text{ strongly in }L^{2-\varepsilon}_{t,x},\,\varepsilon\in(0,1),
xθn\displaystyle\nabla_{x}\theta_{n} xθ\displaystyle\rightharpoonup\nabla_{x}\theta weakly in Lt,x43ε,ε(0,13),\displaystyle\text{ weakly in }L^{\frac{4}{3}-\varepsilon}_{t,x},\,\varepsilon\in\left(0,\frac{1}{3}\right),

and the following convergence results for “entropy” quantities

(3.5) ln(θn)\displaystyle\ln(\theta_{n}) ln(θ)\displaystyle\rightharpoonup\ln(\theta) weakly in Lt2Wx1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{x},
(3.6) ln(θn)\displaystyle\ln(\theta_{n}) ln(θ)\displaystyle\rightarrow\ln(\theta) strongly in Lt,x3ε,ε(0,1),\displaystyle\text{ strongly in }L^{3-\varepsilon}_{t,x},\,\varepsilon\in(0,1),
(3.7) tr(𝔽n𝔽nT)\displaystyle\operatorname{tr}(\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}) tr(𝔽𝔽T)\displaystyle\rightharpoonup\mathrm{tr}(\mathbb{F}\,\mathbb{F}^{T}) weakly in Lt,x2,\displaystyle\text{ weakly in }L^{2}_{t,x},
(3.8) lndet(𝔽n𝔽nT)\displaystyle\ln\det(\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}) lndet(𝔽𝔽T)\displaystyle\overset{*}{\rightharpoonup}\ln\det(\mathbb{F}\,\mathbb{F}^{T}) weakly* in LtLx43ε,ε(0,13).\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{\frac{4}{3}-\varepsilon}_{x},\,\varepsilon\in\left(0,\frac{1}{3}\right).

The limiting functions (𝐯,𝔽,θ)(\boldsymbol{v},\mathbb{F},\theta) solves (1.14)–(1.15) in the following sense:

(3.9) 0TΩ𝒗t𝝋𝒗𝒗:x𝝋+ν(θ)𝔻𝒗:x𝝋+g(θ)𝔽𝔽T:x𝝋dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\boldsymbol{v}\cdot\partial_{t}\boldsymbol{\varphi}-\boldsymbol{v}\otimes\boldsymbol{v}:\nabla_{x}\boldsymbol{\varphi}+\nu(\theta)\mathbb{D}\boldsymbol{v}:\nabla_{x}\boldsymbol{\varphi}+g(\theta)\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{\varphi}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝒗0(x)𝝋(0,x)dx\displaystyle\qquad=\int_{\Omega}\boldsymbol{v}_{0}(x)\cdot\boldsymbol{\varphi}(0,x)\mathop{}\!\mathrm{d}x

for any 𝛗𝒞c1([0,T)×Ω;2)\boldsymbol{\varphi}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2}) with divx𝛗=0\operatorname{div}_{x}\boldsymbol{\varphi}=0,

(3.10) 0TΩ𝔽:t𝔾𝔽𝒗x𝔾x𝒗𝔽:𝔾+12δ(θ)(𝔽𝔽T𝔽𝔽):𝔾dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\mathbb{F}:\partial_{t}\mathbb{G}-\mathbb{F}\otimes\boldsymbol{v}\because\nabla_{x}\mathbb{G}-\nabla_{x}\boldsymbol{v}\,\mathbb{F}:\mathbb{G}+\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F}):\mathbb{G}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝔽0(x):𝔾(0,x)dx\displaystyle\qquad=\int_{\Omega}\mathbb{F}_{0}(x):\mathbb{G}(0,x)\mathop{}\!\mathrm{d}x

for any 𝔾𝒞c1([0,T)×Ω;2×2)\mathbb{G}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2\times 2}),

(3.11) 0TΩ(ln(θ)g(θ)f(𝔽𝔽T))tϕ(ln(θ)g(θ)f(𝔽𝔽T))𝒗xϕ+κ(θ)xlnθxϕdxdt\displaystyle\int_{0}^{T}\int_{\Omega}-(\ln(\theta)-g^{\prime}(\theta)f(\mathbb{F}\,\mathbb{F}^{T}))\partial_{t}\phi-(\ln(\theta)-g^{\prime}(\theta)f(\mathbb{F}\,\mathbb{F}^{T}))\boldsymbol{v}\cdot\nabla_{x}\phi+\kappa(\theta)\nabla_{x}\ln\theta\,\nabla_{x}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
0TΩ(κ(θ)|xθ|2θ2+2ν(θ)|𝔻𝒗|2θ+δ(θ)|𝔹𝕀|2θ)dxdt.\displaystyle\qquad\geq\int_{0}^{T}\int_{\Omega}\left(\frac{\kappa(\theta)|\nabla_{x}\theta|^{2}}{\theta^{2}}+\frac{2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}}{\theta}+\frac{\delta(\theta)|\mathbb{B}-\mathbb{I}|^{2}}{\theta}\right)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.

for any ϕ𝒞c1([0,T)×Ω)\phi\in\mathcal{C}^{1}_{c}([0,T)\times\Omega), ϕ0\phi\geq 0,

(3.12) 0TΩ(|𝒗|22+θ+(g(θ)θg(θ))f(𝔽𝔽T))tφdxdt\displaystyle-\int_{0}^{T}\int_{\Omega}\left(\frac{|\boldsymbol{v}|^{2}}{2}+\theta+(g(\theta)-\theta\,g^{\prime}(\theta))f(\mathbb{F}\,\mathbb{F}^{T})\right)\partial_{t}\varphi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=φ(0)Ω(|𝒗0|22+θ0+(g(θ0)θ0g(θ0))f(𝔽0𝔽0T))dx\displaystyle\qquad=\varphi(0)\int_{\Omega}\left(\frac{|\boldsymbol{v}_{0}|^{2}}{2}+\theta_{0}+(g(\theta_{0})-\theta_{0}\,g^{\prime}(\theta_{0}))f(\mathbb{F}_{0}\,\mathbb{F}_{0}^{T})\right)\mathop{}\!\mathrm{d}x

for any φ𝒞c1([0,T))\varphi\in\mathcal{C}^{1}_{c}([0,T)). Moreover θ>0\theta>0 and det𝔽>0\det\mathbb{F}>0.

The rest of this section is devoted to the proof of Theorem 3.1, which is divided into several subparts.

3.1. Auxiliary identities for 𝔹n\mathbb{B}_{n} and θn\theta_{n}

We begin by recovering the equation for 𝔹n:=𝔽n𝔽nT\mathbb{B}_{n}:=\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}. Namely, we multiply the equation (1.14)2\eqref{main_sys_for_g_theta}_{2} by 𝔽nT\mathbb{F}_{n}^{T} from the right and the transpose of the mentioned equation by 𝔽n\mathbb{F}_{n} from the left. Adding the results together we obtain

(3.13) t𝔹n+divx(𝔹n𝒗n)+δ(θn)(𝔹n2𝔹n)=x𝒗n𝔹n+𝔹n(x𝒗n)T.\displaystyle\partial_{t}\mathbb{B}_{n}+\operatorname{div}_{x}\left(\mathbb{B}_{n}\boldsymbol{v}_{n}\right)+\delta(\theta_{n})(\mathbb{B}_{n}^{2}-\mathbb{B}_{n})=\nabla_{x}\boldsymbol{v}_{n}\,\mathbb{B}_{n}+\mathbb{B}_{n}\,(\nabla_{x}\boldsymbol{v}_{n})^{T}.

Next, similarly as in (1.10), we can derive the following identity

(3.14) tηn+divx(ηn𝒗n)divx(κ(θn)x(lnθn))\displaystyle\partial_{t}\eta_{n}+\operatorname{div}_{x}(\eta_{n}\,\boldsymbol{v}_{n})-\operatorname{div}_{x}(\kappa(\theta_{n})\nabla_{x}(\ln\theta_{n}))
=κ(θn)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn+δ(θn)|𝔹n𝕀|2θn,\displaystyle\qquad=\frac{\kappa(\theta_{n})|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}},

where the entropy is given as

(3.15) ηn:=lnθng(θn)f(𝔹n).\eta_{n}:=\ln\theta_{n}-g^{\prime}(\theta_{n})f(\mathbb{B}_{n}).

We also at this point recall the definition of the internal energy

(3.16) en:=θn+(g(θn)θng(θn))f(𝔹n).e_{n}:=\theta_{n}+(g(\theta_{n})-\theta_{n}\,g^{\prime}(\theta_{n}))f(\mathbb{B}_{n}).

Before moving forward, let us prove a structural lemma connected to equation (3.13) and (LABEL:eq:entropy_equality_for_g_theta).

Lemma 3.2.

Let a triple (𝐯n,θn,𝔹n)(\boldsymbol{v}_{n},\theta_{n},\mathbb{B}_{n}) solve (LABEL:eq:entropy_equality_for_g_theta)–(3.13) with ηn:=ln(θn)g(θn)f(𝔹n)\eta_{n}:=\ln(\theta_{n})-g^{\prime}(\theta_{n})f(\mathbb{B}_{n}). Then, for any λ>0\lambda>0 there holds

(3.17) t(θnλλhλ(θn)f(𝔹n))+divx((θnλλhλ(θn)f(𝔹n))𝒗n)\displaystyle\partial_{t}\left(\frac{\theta_{n}^{\lambda}}{\lambda}-h_{\lambda}(\theta_{n})\,f(\mathbb{B}_{n})\right)+\operatorname{div}_{x}\left(\left(\frac{\theta_{n}^{\lambda}}{\lambda}-h_{\lambda}(\theta_{n})\,f(\mathbb{B}_{n})\right)\boldsymbol{v}_{n}\right)
divx(κ(θn)xθnλλ)2(𝔹n𝕀):𝔻𝒗n(g(θn)θnλhλ(θn))\displaystyle\quad-\operatorname{div}_{x}\left(\kappa(\theta_{n})\nabla_{x}\frac{\theta_{n}^{\lambda}}{\lambda}\right)-2(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}(g^{\prime}(\theta_{n})\theta_{n}^{\lambda}-h_{\lambda}(\theta_{n}))
=((1λ)κ(θn)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn+δ(θn)|𝔹n𝕀|2θn(hλ(θn)θ1λ+1g(θn)θn))θnλ\displaystyle=\left(\frac{(1-\lambda)\kappa(\theta_{n})|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}}(h_{\lambda}(\theta_{n})\theta^{1-\lambda}+1-g^{\prime}(\theta_{n})\theta_{n})\right)\theta_{n}^{\lambda}

where hλ(θ)h_{\lambda}(\theta) is defined in (2.7) and f(𝔹)f(\mathbb{B}) is given by (1.8).

Proof.

Taking the scalar product of (3.13) with 𝕀𝔹n1\mathbb{I}-\mathbb{B}_{n}^{-1} and using the fact that 𝔹nf(𝔹n)=𝕀𝔹n1\partial_{\mathbb{B}_{n}}f(\mathbb{B}_{n})=\mathbb{I}-\mathbb{B}_{n}^{-1}, we deduce that

(3.18) tf(𝔹n)+divx(f(𝔹n)𝒗n)+δ(θ)|𝔹n𝕀|2=2(𝔹n𝕀):𝔻𝒗n.\displaystyle\partial_{t}f(\mathbb{B}_{n})+\operatorname{div}_{x}(f(\mathbb{B}_{n})\boldsymbol{v}_{n})+\delta(\theta)|\mathbb{B}_{n}-\mathbb{I}|^{2}=2(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}.

Using the definition of the entropy in (LABEL:eq:entropy_equality_for_g_theta), we have

(3.19) tθn(1θng′′(θn)f(𝔹n))g(θn)tf(𝔹n)\displaystyle\partial_{t}\theta_{n}\left(\frac{1}{\theta_{n}}-g^{\prime\prime}(\theta_{n})f(\mathbb{B}_{n})\right)-g^{\prime}(\theta_{n})\partial_{t}f(\mathbb{B}_{n})
+divx(θn𝒗n)(1θng′′(θn)f(𝔹n))g(θn)divx(f(𝔹n)𝒗n)\displaystyle\quad+\operatorname{div}_{x}(\theta_{n}\,\boldsymbol{v}_{n})\left(\frac{1}{\theta_{n}}-g^{\prime\prime}(\theta_{n})f(\mathbb{B}_{n})\right)-g^{\prime}(\theta_{n})\operatorname{div}_{x}(f(\mathbb{B}_{n})\,\boldsymbol{v}_{n})
divx(κ(θn)x(lnθn))\displaystyle\quad-\operatorname{div}_{x}(\kappa(\theta_{n})\nabla_{x}(\ln\theta_{n}))
=κ(θn)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn+δ(θn)|𝔹n𝕀|2θn,\displaystyle=\frac{\kappa(\theta_{n})|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}},

Then we multiply (3.18) by g(θn)g^{\prime}(\theta_{n}) and add the result to into (LABEL:eq:entropy_equality_for_g_theta2), we obtain

tθn(1θng′′(θn)f(𝔹n))+divx(θn𝒗n)(1θng′′(θn)f(𝔹n))\displaystyle\partial_{t}\theta_{n}\left(\frac{1}{\theta_{n}}-g^{\prime\prime}(\theta_{n})f(\mathbb{B}_{n})\right)+\operatorname{div}_{x}(\theta_{n}\,\boldsymbol{v}_{n})\left(\frac{1}{\theta_{n}}-g^{\prime\prime}(\theta_{n})f(\mathbb{B}_{n})\right)
+g(θn)δ(θ)|𝔹n𝕀|2divx(κ(θn)x(lnθn))\displaystyle\qquad+g^{\prime}(\theta_{n})\delta(\theta)|\mathbb{B}_{n}-\mathbb{I}|^{2}-\operatorname{div}_{x}(\kappa(\theta_{n})\nabla_{x}(\ln\theta_{n}))
=κ(θn)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn+δ(θn)|𝔹n𝕀|2θn+2g(θn)(𝔹n𝕀):𝔻𝒗n.\displaystyle=\frac{\kappa(\theta_{n})|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}}+2g^{\prime}(\theta_{n})(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}.

Next, we multiply the result by θnλ\theta_{n}^{\lambda} and use the definition of hλh_{\lambda} in (2.7) to observe that

(3.20) t(θnλλ)f(𝔹n)thλ(θn)+divx(θnλ𝒗nλ)f(𝔹n)divx(hλ(θn)𝒗n)\displaystyle\partial_{t}\left(\frac{\theta_{n}^{\lambda}}{\lambda}\right)-f(\mathbb{B}_{n})\partial_{t}h_{\lambda}(\theta_{n})+\operatorname{div}_{x}\left(\frac{\theta_{n}^{\lambda}\,\boldsymbol{v}_{n}}{\lambda}\right)-f(\mathbb{B}_{n})\operatorname{div}_{x}(h_{\lambda}(\theta_{n})\,\boldsymbol{v}_{n})
divx(κ(θn)xθnλλ)2g(θn)θnλ(𝔹n𝕀):𝔻𝒗n.\displaystyle\qquad-\operatorname{div}_{x}\left(\frac{\kappa(\theta_{n})\nabla_{x}\theta_{n}^{\lambda}}{\lambda}\right)-2g^{\prime}(\theta_{n})\theta_{n}^{\lambda}(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}.
=(κ(θn)(1λ)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn+δ(θn)(1g(θn)θn)|𝔹n𝕀|2θn)θnλ\displaystyle=\left(\frac{\kappa(\theta_{n})(1-\lambda)|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})(1-g^{\prime}(\theta_{n})\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}}\right)\theta_{n}^{\lambda}

Finally, we multiply (3.18) by hλ(θn)h_{\lambda}(\theta_{n}) and subtract the result from (LABEL:some_equation_for_lemma) to get

t(θnλλhλ(θn)f(𝔹n))+divx((θnλλhλ(θn)f(𝔹n))𝒗n)\displaystyle\partial_{t}\left(\frac{\theta_{n}^{\lambda}}{\lambda}-h_{\lambda}(\theta_{n})f(\mathbb{B}_{n})\right)+\operatorname{div}_{x}\left(\left(\frac{\theta_{n}^{\lambda}}{\lambda}-h_{\lambda}(\theta_{n})f(\mathbb{B}_{n})\right)\boldsymbol{v}_{n}\right)
divx(κ(θn)xθnλλ)+2(hλ(θn)g(θn)θnλ)(𝔹n𝕀):𝔻𝒗n\displaystyle\qquad-\operatorname{div}_{x}\left(\frac{\kappa(\theta_{n})\nabla_{x}\theta_{n}^{\lambda}}{\lambda}\right)+2\left(h_{\lambda}(\theta_{n})-g^{\prime}(\theta_{n})\theta_{n}^{\lambda}\right)(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}
=(κ(θn)(1λ)|xθn|2θn2+2ν(θn)|𝔻𝒗n|2θn)θnλ+δ(θn)(hλ(θn)+θnλ1g(θn)θnλ)|𝔹n𝕀|2\displaystyle=\left(\frac{\kappa(\theta_{n})(1-\lambda)|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}+\frac{2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}\right)\theta_{n}^{\lambda}+\delta(\theta_{n})(h_{\lambda}(\theta_{n})+\theta_{n}^{\lambda-1}-g^{\prime}(\theta_{n})\theta_{n}^{\lambda})|\mathbb{B}_{n}-\mathbb{I}|^{2}

and we see that (3.17) directly follows. ∎

3.2. Uniform estimates

We proceed and show some basic bounds for our sequence. Multiplying (1.14)1\eqref{main_sys_for_g_theta}_{1} by 𝒗n\boldsymbol{v}_{n} and adding the result to (1.14)3\eqref{main_sys_for_g_theta}_{3}, integrating the result over Ω\Omega, using the boundary conditions for 𝒗n\boldsymbol{v}_{n} and θn\theta_{n} in (1.15) and the fact that divx𝒗n=0\operatorname{div}_{x}\boldsymbol{v}_{n}=0, we deduce

(3.21) tΩ|𝒗n|22+endx=0.\partial_{t}\int_{\Omega}\frac{|\boldsymbol{v}_{n}|^{2}}{2}+e_{n}\mathop{}\!\mathrm{d}x=0.

First, the internal energy can be estimated as

en=θn+(g(θn)θng(θn))f(𝔹n)θn+g(0)f(𝔹n)θn+C1f(𝔹n),e_{n}=\theta_{n}+(g(\theta_{n})-\theta_{n}g^{\prime}(\theta_{n}))f(\mathbb{B}_{n})\geq\theta_{n}+g(0)f(\mathbb{B}_{n})\geq\theta_{n}+C_{1}f(\mathbb{B}_{n}),

where we used the concavity of gg and the assumption (2.5). Since ff is nonnegative function and θn>0\theta_{n}>0, we see that also en>0e_{n}>0 and using the convergence properties of the initial conditions (3.1), we have

(3.22) 𝒗nLtLx22+θnLtLx1+C1f(𝔹n)LtLx1\displaystyle\|\boldsymbol{v}_{n}\|^{2}_{L^{\infty}_{t}L^{2}_{x}}+\|\theta_{n}\|_{L^{\infty}_{t}L^{1}_{x}}+C_{1}\|f(\mathbb{B}_{n})\|_{L^{\infty}_{t}L^{1}_{x}} 𝒗nLtLx22+enLtLx1\displaystyle\leq\|\boldsymbol{v}_{n}\|^{2}_{L^{\infty}_{t}L^{2}_{x}}+\|e_{n}\|_{L^{\infty}_{t}L^{1}_{x}}
=𝒗0nLx22+e0nLx1C.\displaystyle=\|\boldsymbol{v}^{n}_{0}\|^{2}_{L^{2}_{x}}+\|e_{0}^{n}\|_{L^{1}_{x}}\leq C.

We continue, by deducing the estimates coming from the entropy. We integrate (LABEL:eq:entropy_equality_for_g_theta) over Ω\Omega, use the boundary conditions and the fact that divx𝒗n=0\operatorname{div}_{x}\boldsymbol{v}_{n}=0, and also using the assumption on the function δ\delta stated in (2.4), we get that for any τ(0,T)\tau\in(0,T)

C1\displaystyle C_{1} 0τΩ|xθn|2(θn)2+|𝔻𝒗n|2θn+(1+θn)|𝔹n𝕀|2θndxdτ\displaystyle\int_{0}^{\tau}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{(\theta_{n})^{2}}+\frac{|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{(1+\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau
0τΩtηn+divx(ηn𝒗n)divx(κ(θn)x(lnθn))dxdτ\displaystyle\leq\int_{0}^{\tau}\int_{\Omega}\partial_{t}\eta_{n}+\operatorname{div}_{x}(\eta_{n}\boldsymbol{v}_{n})-\operatorname{div}_{x}(\kappa(\theta_{n})\nabla_{x}(\ln\theta_{n}))\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau
=(Ωln(θn)g(θn)f(𝔹n)dx)|0τ.\displaystyle=\left(\int_{\Omega}\ln(\theta_{n})-g^{\prime}(\theta_{n})\,f(\mathbb{B}_{n})\mathop{}\!\mathrm{d}x\right)\Bigg{|}_{0}^{\tau}.

Since τ(0,T)\tau\in(0,T) is arbitrary, gg is nondecreasing, ff is nonnegative and lnxx1\ln x\leq x-1, we can use the assumptions (2.7) and (2.1) to deduce

(3.23) 𝔻𝒗nθnLt,x2+lnθnLtLx1+xlnθnLt,x2+(1+θn)(𝔹n𝕀)θnLt,x2\displaystyle\left\|\frac{\mathbb{D}\boldsymbol{v}_{n}}{\sqrt{\theta_{n}}}\right\|_{L^{2}_{t,x}}+\|\ln\theta_{n}\|_{L^{\infty}_{t}L^{1}_{x}}+\|\nabla_{x}\ln\theta_{n}\|_{L^{2}_{t,x}}+\left\|\frac{(1+\sqrt{\theta_{n}})(\mathbb{B}_{n}-\mathbb{I})}{\sqrt{\theta_{n}}}\right\|_{L^{2}_{t,x}} C,\displaystyle\leq C,

where we employed also the uniform estimate (3.22). In particular, we deduce from the above estimates and the fact that 𝔹n=𝔽n𝔽nT\mathbb{B}_{n}=\mathbb{F}_{n}\,\mathbb{F}_{n}^{T} that

(3.24) 𝔹nLt,x2\displaystyle\|\mathbb{B}_{n}\|_{L^{2}_{t,x}} =𝔽n𝔽nTLt,x2C,\displaystyle=\|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}\|_{L^{2}_{t,x}}\leq C,
(3.25) 𝔽nLt,x4\displaystyle\|\mathbb{F}_{n}\|_{L^{4}_{t,x}} C.\displaystyle\leq C.

We continue with uniform estimates for 𝒗n\boldsymbol{v}_{n}. We take the scalar product of (1.14)1\eqref{main_sys_for_g_theta}_{1} with 𝒗n\boldsymbol{v}_{n} and integrate the result over Ω\Omega to get for arbitrary t(0,T)t\in(0,T)

(3.26) 12Ω|𝒗n(t)|2dx+0tΩ2ν(θn)|𝔻𝒗n|2dxdτ\displaystyle\frac{1}{2}\int_{\Omega}|\boldsymbol{v}_{n}(t)|^{2}\mathop{}\!\mathrm{d}x+\int_{0}^{t}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau
=0tΩg(θn)𝔹n:𝔻𝒗ndxdτ+12Ω|𝒗0n|2dx.\displaystyle\qquad=-\int_{0}^{t}\int_{\Omega}g(\theta_{n})\mathbb{B}_{n}:\mathbb{D}\boldsymbol{v}_{n}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\frac{1}{2}\int_{\Omega}|\boldsymbol{v}_{0}^{n}|^{2}\mathop{}\!\mathrm{d}x.

Employing the Young inequality, we get

0tΩν(θn)|𝔻𝒗n|2dxdτ0tΩ(g(θn))2ν(θn)|𝔹n|2dxdτ+Ω|𝒗0n|2dx.\displaystyle\int_{0}^{t}\int_{\Omega}\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau\leq\int_{0}^{t}\int_{\Omega}\frac{(g(\theta_{n}))^{2}}{\nu(\theta_{n})}|\mathbb{B}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\int_{\Omega}|\boldsymbol{v}_{0}^{n}|^{2}\mathop{}\!\mathrm{d}x.

Hence, we may use the uniform estimate (3.24), the assumption on the parameters (2.2), (2.5) and the assumptions on the initial data (3.1) and deduce that

(3.27) 𝔻𝒗nLt,x2C(𝒗0nLx2,θ0nLx1,lnθ0nLx1,f(𝔹0n)Lx1)C.,\displaystyle\|\mathbb{D}\boldsymbol{v}_{n}\|_{L^{2}_{t,x}}\leq C(\|\boldsymbol{v}_{0}^{n}\|_{L^{2}_{x}},\|\theta_{0}^{n}\|_{L^{1}_{x}},\|\ln\theta_{0}^{n}\|_{L^{1}_{x}},\|f(\mathbb{B}_{0}^{n})\|_{L^{1}_{x}})\leq C.,

The above estimate, the assumption (2.1) and the Korn inequality imply that

(3.28) ν(θn)𝔻𝒗nLt,x2\displaystyle\|\sqrt{\nu(\theta_{n})}\mathbb{D}\boldsymbol{v}_{n}\|_{L^{2}_{t,x}} C,\displaystyle\leq C,
(3.29) x𝒗nLt,x2\displaystyle\|\nabla_{x}\boldsymbol{v}_{n}\|_{L^{2}_{t,x}} C.\displaystyle\leq C.

The Sobolev embedding, the interpolation theorem and the uniform bound (3.22) further lead to

(3.30) 𝒗nLt,x4\displaystyle\|\boldsymbol{v}_{n}\|_{L^{4}_{t,x}} C,\displaystyle\leq C,
(3.31) 𝒗nLt2Lxb\displaystyle\|\boldsymbol{v}_{n}\|_{L^{2}_{t}L^{b}_{x}} C(b) for all b<+.\displaystyle\leq C(b)\quad\textrm{ for all }b<+\infty.

To finish the uniform bounds for the velocity, we consider the estimates on t𝒗n\partial_{t}\boldsymbol{v}_{n}. It follows from (1.14) that

(3.32) t𝒗nLt2W0,div1,2𝒗n𝒗n2ν(θn)𝔻𝒗n2g(θn)𝔹nLt,x2C,\displaystyle\|\partial_{t}\boldsymbol{v}_{n}\|_{L^{2}_{t}W^{-1,2}_{0,\operatorname{div}}}\leq\left\|\boldsymbol{v}_{n}\otimes\boldsymbol{v}_{n}-2\nu(\theta_{n})\mathbb{D}\boldsymbol{v}_{n}-2g(\theta_{n})\mathbb{B}_{n}\right\|_{L^{2}_{t,x}}\leq C,

where for the second inequality, we used the assumptions (2.2) and (2.5), and the uniform estimates (3.24), (3.29) and (3.30).

We continue with improving the estimates. We start with 𝔽n\mathbb{F}_{n}. Taking the scalar product of (1.14)2 with 𝔽n\mathbb{F}_{n} and integrating the result over (0,t)×Ω(0,t)\times\Omega, we deduce the identity

(3.33) 𝔽n(t)22+0tΩδ(θn)|𝔽n𝔽nT|2dxdτ\displaystyle\|\mathbb{F}_{n}(t)\|_{2}^{2}+\int_{0}^{t}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau
=0tΩ2𝔹n:𝔻𝒗ndxdτ+0tΩδ(θn)|𝔽n|2dxdτ+𝔽0n22,\displaystyle\qquad=\int_{0}^{t}\int_{\Omega}2\mathbb{B}_{n}:\mathbb{D}\boldsymbol{v}_{n}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\int_{0}^{t}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\|\mathbb{F}_{0}^{n}\|_{2}^{2},

which by using the Hölder inequality, Young inequality and a matrix inequality |𝔽|42|𝔽𝔽T|2|\mathbb{F}|^{4}\leq 2|\mathbb{F}\,\mathbb{F}^{T}|^{2} implies

𝔽n(t)22+120tΩδ(θn)|𝔽n|4dxdτ2𝔽n𝔽nT2𝔻𝒗n2+0tΩδ(θn)dxdτ+140tΩδ(θn)|𝔽n|4dxdτ+𝔽022.\|\mathbb{F}_{n}(t)\|_{2}^{2}+\frac{1}{2}\int_{0}^{t}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{4}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau\\ \leq 2\|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}\|_{2}\|\mathbb{D}\boldsymbol{v}_{n}\|_{2}+\int_{0}^{t}\int_{\Omega}\delta(\theta_{n})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\frac{1}{4}\int_{0}^{t}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{4}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau+\|\mathbb{F}_{0}\|_{2}^{2}.

Hence, by (3.24), (3.27), the assumptions postulated for δ\delta in  (2.4), the bounds (3.22) and the fact that t(0,T)t\in(0,T) is arbitrary, we have

(3.34) 𝔽nLtLx22+0TΩδ(θn)|𝔽n|4dxdtC.\displaystyle\|\mathbb{F}_{n}\|_{L^{\infty}_{t}L^{2}_{x}}^{2}+\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{4}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C.

In particular, coming back with the obtained bound to the inequality (LABEL:some_random_inequality2) gives us

(3.35) 0TΩδ(θn)|𝔽n𝔽nT|2dxdtC.\displaystyle\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C.

To derive some compactness for θn\theta_{n} we use Lemma 3.2. We integrate (3.17) over Ω\Omega and (0,t)(0,t), neglect terms having the proper sign to deduce for any λ(0,1)\lambda\in(0,1) that

(3.36) (1λ)0tΩκ(θn)|xθn|2θn2λdxdτ1λΩθnλ(t)dxdτΩhλ(θn(t))f(𝔹n(t))dx+0tΩ(δ(θn)|𝔹n𝕀|22(𝔹n𝕀):𝔻𝒗n)(g(θn)θnλhλ(θn))dxdτ.\begin{split}&(1-\lambda)\int_{0}^{t}\int_{\Omega}\kappa(\theta_{n})\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2-\lambda}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau\\ &\quad\leq\frac{1}{\lambda}\int_{\Omega}\theta_{n}^{\lambda}(t)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau-\int_{\Omega}h_{\lambda}(\theta_{n}(t))\,f(\mathbb{B}_{n}(t))\mathop{}\!\mathrm{d}x\\ &\qquad+\int_{0}^{t}\int_{\Omega}\left(\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}-2(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}\right)\left(g^{\prime}(\theta_{n})\theta_{n}^{\lambda}-h_{\lambda}(\theta_{n})\right)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}\tau.\end{split}

Using the assumption the assumption (2.6) and the bound on hλh_{\lambda} in (2.8), we have

|g(θn)θnλ|+|hλ(θn)|C.|g^{\prime}(\theta_{n})\theta_{n}^{\lambda}|+|h_{\lambda}(\theta_{n})|\leq C.

Thus, using this bound in (3.36), combining it with the Hölder inequality and the assumption on κ\kappa in (2.1) and the already derived uniform estimates (3.22), (3.35), (3.27), (3.23), we get

(3.37) 0TΩ|xθn|2θnλdxdtC(λ), for all λ(1,2).\displaystyle\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C(\lambda),\quad\text{ for all }\lambda\in(1,2).

With help of this estimate, we continue with further estimates for θn\theta_{n}. We recall the interpolation inequality

(3.38) vLa(Ω)aC(vL22λ(Ω)a+vL22λ(Ω)a2xvL2(Ω)2),\|v\|^{a}_{L^{a}(\Omega)}\leq C\left(\|v\|^{a}_{L^{\frac{2}{2-\lambda}}(\Omega)}+\|v\|^{a-2}_{L^{\frac{2}{2-\lambda}}(\Omega)}\|\nabla_{x}v\|^{2}_{L^{2}(\Omega)}\right),

which is valid for all λ[1,2)\lambda\in[1,2) and an aa defined as

a:=2+22λ.a:=2+\frac{2}{2-\lambda}.

Hence, with using the interpolation (3.38) and the uniform bounds (3.22) and (3.37), we see that for any λ(1,2)\lambda\in(1,2)

(3.39) 0TΩθn3λdxdt=0TΩ(θn2λ2)adxdtC0Tθn2λ222λa+θn2λ222λa2x(θn2λ2)22dt=C0T(Ωθndx)3λdt+C(λ)0T(Ωθndx)(Ω|xθn|2θnλdx)dtC(λ).\begin{split}\int_{0}^{T}\int_{\Omega}\theta_{n}^{3-\lambda}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t&=\int_{0}^{T}\int_{\Omega}\left(\theta_{n}^{\frac{2-\lambda}{2}}\right)^{a}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C\int_{0}^{T}\|\theta_{n}^{\frac{2-\lambda}{2}}\|_{\frac{2}{2-\lambda}}^{a}+\|\theta^{\frac{2-\lambda}{2}}_{n}\|_{\frac{2}{2-\lambda}}^{a-2}\|\nabla_{x}(\theta_{n}^{\frac{2-\lambda}{2}})\|_{2}^{2}\mathop{}\!\mathrm{d}t\\ &=C\int_{0}^{T}\left(\int_{\Omega}\theta_{n}\mathop{}\!\mathrm{d}x\right)^{3-\lambda}\mathop{}\!\mathrm{d}t+C(\lambda)\int_{0}^{T}\left(\int_{\Omega}\theta_{n}\mathop{}\!\mathrm{d}x\right)\,\left(\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}\mathop{}\!\mathrm{d}x\right)\mathop{}\!\mathrm{d}t\\ &\leq C(\lambda).\end{split}

We also show the inhomogeneous estimate for θn\theta_{n}. Recall another interpolation inequality

(3.40) vLp(Ω)C(p,Ω)(vL2(Ω)+vL2(Ω)2pxvL2(Ω)p2p),\|v\|_{L^{p}(\Omega)}\leq C(p,\Omega)\left(\|v\|_{L^{2}(\Omega)}+\|v\|^{\frac{2}{p}}_{L^{2}(\Omega)}\|\nabla_{x}v\|_{L^{2}(\Omega)}^{\frac{p-2}{p}}\right),

which is valid for all p(2,)p\in(2,\infty). Then, for arbitrary λ(1,2)\lambda\in(1,2), we set v:=(θn)2λ2v:=(\theta_{n})^{\frac{2-\lambda}{2}} in (3.40) and by using the Hölder inequality and the uniform bounds (3.22) and (3.37), we deduce

(3.41) (θn)2λ2Lt2pp2Lxp2pp2\displaystyle\|(\theta_{n})^{\frac{2-\lambda}{2}}\|^{\frac{2p}{p-2}}_{L_{t}^{\frac{2p}{p-2}}L_{x}^{p}} =0T(θn)2λ2Lxp2pp2dt\displaystyle=\int_{0}^{T}\|(\theta_{n})^{\frac{2-\lambda}{2}}\|^{\frac{2p}{p-2}}_{L_{x}^{p}}\mathop{}\!\mathrm{d}t
C(p)0T(θn)2λ2Lx22pp2+(θn)2λ2Lx24p2x(θn)2λ2Lx22dt\displaystyle\leq C(p)\int_{0}^{T}\|(\theta_{n})^{\frac{2-\lambda}{2}}\|^{\frac{2p}{p-2}}_{L^{2}_{x}}+\|(\theta_{n})^{\frac{2-\lambda}{2}}\|^{\frac{4}{p-2}}_{L^{2}_{x}}\|\nabla_{x}(\theta_{n})^{\frac{2-\lambda}{2}}\|_{L^{2}_{x}}^{2}\mathop{}\!\mathrm{d}t
C(p,λ)(θnLtLx1p(2λ)p2+θnLtLx22(2λ)p20TΩ|xθn|2(θn)λdxdt)C(p,λ).\displaystyle\leq C(p,\lambda)\left(\|\theta_{n}\|^{\frac{p(2-\lambda)}{p-2}}_{L^{\infty}_{t}L^{1}_{x}}+\|\theta_{n}\|^{\frac{2(2-\lambda)}{p-2}}_{L^{\infty}_{t}L^{2}_{x}}\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{(\theta_{n})^{\lambda}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\right)\leq C(p,\lambda).

Next, with the use of Young’s inequality, (3.39) and (3.37) we may infer

(3.42) 0TΩ|xθn|223λdxdt=0TΩ(|xθn|2θnλ)223λ2θn(223λ)λ2dxdt0TΩ|xθn|2θnλ+θn(223λ)λ2(223λ)dxdt=0TΩ|xθn|2θnλ+θn3λdxdtC(λ),\begin{split}\int_{0}^{T}\int_{\Omega}|\nabla_{x}\theta_{n}|^{2-\frac{2}{3}\lambda}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t&=\int_{0}^{T}\int_{\Omega}\left(\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}\right)^{\frac{2-\frac{2}{3}\lambda}{2}}\theta_{n}^{\frac{(2-\frac{2}{3}\lambda)\lambda}{2}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &\leq\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}+\theta_{n}^{\frac{(2-\frac{2}{3}\lambda)\lambda}{2-(2-\frac{2}{3}\lambda)}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &=\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}+\theta_{n}^{3-\lambda}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C(\lambda),\end{split}

for any λ(1,2)\lambda\in(1,2). To summarize, it follows from (3.41) (where p>2p>2 and λ(1,2)\lambda\in(1,2) are arbitrary), from (3.39) and (3.42) give the bounds

(3.43) θnLtqLxp\displaystyle\|\theta_{n}\|_{L_{t}^{q}L_{x}^{p}} C(p,q)\displaystyle\leq C(p,q) for all p[1,) and q[1,pp1),\displaystyle\textrm{for all }p\in[1,\infty)\textrm{ and }q\in\left[1,\frac{p}{p-1}\right),
(3.44) θnLt,xq\displaystyle\|\theta_{n}\|_{L^{q}_{t,x}} C(q)\displaystyle\leq C(q) for all q[1,2),\displaystyle\textrm{for all }q\in[1,2),
(3.45) xθnLt,xq\displaystyle\|\nabla_{x}\theta_{n}\|_{L^{q}_{t,x}} C(q)\displaystyle\leq C(q) for all q[1,43).\displaystyle\textrm{for all }q\in\left[1,\frac{4}{3}\right).

Next goal is to establish the integrability of the sequence f(𝔹n)f(\mathbb{B}_{n}), which appears in the internal energy and entropy. First, due to the matrix inequality |tr𝔽|2|𝔽||\operatorname{tr}\mathbb{F}|\leq\sqrt{2}|\mathbb{F}| we have

(3.46) tr𝔹nLt,x22𝔹nLt,x2C,\displaystyle\|\operatorname{tr}\mathbb{B}_{n}\|_{L^{2}_{t,x}}\leq\sqrt{2}\|\mathbb{B}_{n}\|_{L^{2}_{t,x}}\leq C,
tr𝔹nLtLx12𝔹nLtLx1C.\displaystyle\|\operatorname{tr}\mathbb{B}_{n}\|_{L^{\infty}_{t}L^{1}_{x}}\leq\sqrt{2}\|\mathbb{B}_{n}\|_{L^{\infty}_{t}L^{1}_{x}}\leq C.

The above estimate gives us the information on the one part of f(𝔹n)f(\mathbb{B}_{n}). To get also the estimate on lndet𝔹n\ln\det\mathbb{B}_{n}, we derive the identity for this quantity. To do so, we take the scalar product of (3.13) with 𝔹n1\mathbb{B}_{n}^{-1} (note that 𝔹n\mathbb{B}_{n} is assumed to be positively definite) and get (using the fact that divx𝒗n=0\operatorname{div}_{x}\boldsymbol{v}_{n}=0)

(3.47) tlndet𝔹n+divx(𝒗nlndet𝔹n)+δ(θn)tr(𝔹n𝕀)=0.\displaystyle\partial_{t}\ln\det\mathbb{B}_{n}+\operatorname{div}_{x}(\boldsymbol{v}_{n}\ln\det\mathbb{B}_{n})+\delta(\theta_{n})\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})=0.

To deduce the information for the last term on the right hand side, we first recall the assumption on δ\delta in (2.4) and then it directly follows from (3.22) and (3.44) that

(3.48) δ(θn)LtrLxp+δ(θn)LtLx1C(q,p,r)\displaystyle\|\delta(\theta_{n})\|_{L_{t}^{r}L_{x}^{p}}+\|\delta(\theta_{n})\|_{L^{\infty}_{t}L^{1}_{x}}\leq C(q,p,r)

for all q[1,2)q\in[1,2), all p[1,)p\in[1,\infty) and all r[1,pp1)r\in[1,\frac{p}{p-1}). With the help of this identity, we derive two estimates, the first one homogeneous with respect to space and time variables and the second one to get the optimal control in (3.47). Thus, by the Hölder inequality, the bounds (3.48), (3.35), and the matrix inequality |tr𝔽|2|𝔽||\operatorname{tr}\mathbb{F}|\leq\sqrt{2}|\mathbb{F}| we obtain for all ε(0,1)\varepsilon\in(0,1)

(3.49) 0TΩ|δ(θn)tr(𝔹n𝕀)|42ε3εdxdt=0TΩ(δ(θn))2ε3ε(δ(θn)|tr(𝔹n𝕀)|2)2ε3εdxdt(0TΩ(δ(θn))2εdxdt)13ε(0TΩδ(θn)|tr(𝔹n𝕀)|2dxdt)2ε3εC(ε),\begin{split}&\int_{0}^{T}\int_{\Omega}|\delta(\theta_{n})\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{\frac{4-2\varepsilon}{3-\varepsilon}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\int_{0}^{T}\int_{\Omega}\left(\delta(\theta_{n})\right)^{\frac{2-\varepsilon}{3-\varepsilon}}\left(\delta(\theta_{n})|\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{2}\right)^{\frac{2-\varepsilon}{3-\varepsilon}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &\quad\leq\left(\int_{0}^{T}\int_{\Omega}(\delta(\theta_{n}))^{2-\varepsilon}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\right)^{\frac{1}{3-\varepsilon}}\left(\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathrm{tr}(\mathbb{B}_{n}-\mathbb{I})|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\right)^{\frac{2-\varepsilon}{3-\varepsilon}}\leq C(\varepsilon),\end{split}

and similarly, for any q[1,2)q\in[1,2) we can use the Hölder inequality and the uniform bound (3.48) to deduce

(3.50) 0T(Ω|δ(θn)tr(𝔹n𝕀)|qdx)1qdt=0T(Ω(δ(θn))q2(δ(θn)|tr(𝔹n𝕀)|2)q2dx)1qdt0T(Ω(δ(θn))q2qdx)2q2q(Ωδ(θn)|tr(𝔹n𝕀)|2dx)12dtδ(θn)Lt1Lxq2q12(0TΩδ(θn)|tr(𝔹n𝕀)|2dxdt)12C(q).\begin{split}&\int_{0}^{T}\left(\int_{\Omega}|\delta(\theta_{n})\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{q}\mathop{}\!\mathrm{d}x\right)^{\frac{1}{q}}\mathop{}\!\mathrm{d}t=\int_{0}^{T}\left(\int_{\Omega}\left(\delta(\theta_{n})\right)^{\frac{q}{2}}\left(\delta(\theta_{n})|\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{2}\right)^{\frac{q}{2}}\mathop{}\!\mathrm{d}x\right)^{\frac{1}{q}}\mathop{}\!\mathrm{d}t\\ &\quad\leq\int_{0}^{T}\left(\int_{\Omega}\left(\delta(\theta_{n})\right)^{\frac{q}{2-q}}\mathop{}\!\mathrm{d}x\right)^{\frac{2-q}{2q}}\left(\int_{\Omega}\delta(\theta_{n})|\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{2}\mathop{}\!\mathrm{d}x\right)^{\frac{1}{2}}\mathop{}\!\mathrm{d}t\\ &\quad\leq\|\delta(\theta_{n})\|^{\frac{1}{2}}_{L_{t}^{1}L_{x}^{\frac{q}{2-q}}}\left(\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\right)^{\frac{1}{2}}\leq C(q).\end{split}

In addition, by the very similar manipulation, we deduce that for any q[1,87)q\in[1,\frac{8}{7})

(3.51) 0TΩ|δ(θn)𝔽n𝔽nT𝔽n|q+|δ(θn)𝔽n|qdxdtC0TΩ1+(δ(θn))q4(δ(θn)|𝔽n|4)3q4dxdtC+C(0TΩ(δ(θn))q43q)443q(0TΩδ(θn)|𝔽n|4dxdt)3q4C(q),\begin{split}&\int_{0}^{T}\int_{\Omega}|\delta(\theta_{n})\mathbb{F}_{n}\mathbb{F}_{n}^{T}\mathbb{F}_{n}|^{q}+|\delta(\theta_{n})\mathbb{F}_{n}|^{q}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C\int_{0}^{T}\int_{\Omega}1+(\delta(\theta_{n}))^{\frac{q}{4}}\left(\delta(\theta_{n})|\mathbb{F}_{n}|^{4}\right)^{\frac{3q}{4}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &\quad\leq C+C\left(\int_{0}^{T}\int_{\Omega}(\delta(\theta_{n}))^{\frac{q}{4-3q}}\right)^{\frac{4}{4-3q}}\left(\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{4}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\right)^{\frac{3q}{4}}\leq C(q),\end{split}

where we used the estimates (3.34) and (3.48). Note that condition q<87q<\frac{8}{7} is equivalent to condition q43q<2\frac{q}{4-3q}<2, which is required in (3.48). For the entropy estimates, we multiply (3.47) by q|lndet𝔹n|q2lndet𝔹nq|\ln\det\mathbb{B}_{n}|^{q-2}\ln\det\mathbb{B}_{n}, integrate over Ω\Omega and use the Hölder inequality to get

ddtΩ|lndet𝔹n|qdx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}|\ln\det\mathbb{B}_{n}|^{q}\mathop{}\!\mathrm{d}x =qΩδ(θn)tr(𝔹n𝕀)|lndet𝔹n|q2lndet𝔹ndx\displaystyle=-q\int_{\Omega}\delta(\theta_{n})\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})|\ln\det\mathbb{B}_{n}|^{q-2}\ln\det\mathbb{B}_{n}\mathop{}\!\mathrm{d}x
(Ω|δ(θn)tr(𝔹n𝕀)|qdx)1qlndet𝔹nLxqq1.\displaystyle\leq\left(\int_{\Omega}\left|\delta(\theta_{n})\operatorname{tr}(\mathbb{B}_{n}-\mathbb{I})\right|^{q}\mathop{}\!\mathrm{d}x\right)^{\frac{1}{q}}\|\ln\det\mathbb{B}_{n}\|^{q-1}_{L^{q}_{x}}.

The Grönwall inequality and the uniform estimate (3.50) then implies

(3.52) lndet𝔹nLtLxqC(q,lndet𝔹0nLxq),\displaystyle\|\ln\det\mathbb{B}_{n}\|_{L^{\infty}_{t}L^{q}_{x}}\leq C(q,\|\ln\det\mathbb{B}_{0}^{n}\|_{L^{q}_{x}}),

for any q[1,2)q\in[1,2). Thus, with the use of (LABEL:bounds_on_trace_B_n), (3.52) and due to bounds on initial conditions lndet𝔽n\ln\det\mathbb{F}_{n} in (3.1), we obtain

(3.53) f(𝔹n)Lt,x2εC(ε) for any ε(0,1).\displaystyle\|f(\mathbb{B}_{n})\|_{L^{2-\varepsilon}_{t,x}}\leq C(\varepsilon)\textrm{ for any }\varepsilon\in(0,1).

We finish this part by deriving the uniform estimates for all terms that appear in the entropy identity (LABEL:eq:entropy_equality_for_g_theta). Recall the definition of the entropy ηn\eta_{n} in (3.15)

ηn=lnθng(θn)f(𝔹n).\eta_{n}=\ln\theta_{n}-g^{\prime}(\theta_{n})f(\mathbb{B}_{n}).

It follows from the uniform bound (3.23) and Poincaré–Wirtinger inequality that

lnθnLt2Wx1,2C\displaystyle\|\ln\theta_{n}\|_{L^{2}_{t}W^{1,2}_{x}}\leq C

and it follows from the uniform estimate (3.53), the above inequality and the assumption (2.6) that

(3.54) ηnLt,x2εCfor all ε(0,1).\displaystyle\|\eta_{n}\|_{L^{2-\varepsilon}_{t,x}}\leq C\qquad\textrm{for all }\varepsilon\in(0,1).

Using this inequality, the uniform estimate (3.30) for 𝒗n\boldsymbol{v}_{n} and the classical Hölder inequality, we see

(3.55) ηn𝒗nLt,x43εC for all ε(0,13].\displaystyle\|\eta_{n}\,\boldsymbol{v}_{n}\|_{L^{\frac{4}{3}-\varepsilon}_{t,x}}\leq C\quad\textrm{ for all }\varepsilon\in\left(0,\frac{1}{3}\right].

Next, defining the flux

(3.56) 𝒒n=ηn𝒗nκ(θn)x(lnθn),\boldsymbol{q}_{n}=\eta_{n}\,\boldsymbol{v}_{n}-\kappa(\theta_{n})\nabla_{x}(\ln\theta_{n}),

it follow from (3.55) and (3.23) that

(3.57) 𝒒nLt,x43εC for all ε(0,13].\displaystyle\|\boldsymbol{q}_{n}\|_{L^{\frac{4}{3}-\varepsilon}_{t,x}}\leq C\qquad\textrm{ for all }\varepsilon\in\left(0,\frac{1}{3}\right].

3.3. Convergence results based on the uniform estimates

In this section we use the reflexivity of underlying spaces and the uniform bounds to get the weak convergence results for the sequence of solutions. First, we focus on the convergence results for 𝒗n\boldsymbol{v}_{n}. Using (3.22), (3.29), (3.32) and the Aubin–Lions lemma (see Lemma A.1) we can find a subsequence that we do not relabel and 𝒗\boldsymbol{v} such that

(3.58) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\overset{*}{\rightharpoonup}\boldsymbol{v} weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v} weakly in Lt2W0,div1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{0,\operatorname{div}},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v} weakly in Lt,x4,\displaystyle\text{ weakly in }L^{4}_{t,x},
𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightarrow\boldsymbol{v} strongly in Lt,xp for all p[1,4),\displaystyle\text{ strongly in }L^{p}_{t,x}\textrm{ for all }p\in[1,4),
t𝒗n\displaystyle\partial_{t}\boldsymbol{v}_{n} t𝒗\displaystyle\rightharpoonup\partial_{t}\boldsymbol{v} weakly in Lt2W0,div1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{-1,2}_{0,\operatorname{div}},

which directly implies (3.2)1–(3.2)3. Similarly, due to the uniform bounds (3.34), we can find a subsequence that we do not relabel and 𝔽\mathbb{F} such that

(3.59) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\overset{*}{\rightharpoonup}\mathbb{F} weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightharpoonup\mathbb{F} weakly in Lt,x4,\displaystyle\text{ weakly in }L^{4}_{t,x},

that gives  (3.3)1–(3.3)2. Next, for the temperature, we use (3.43)–(3.45) and wee that for a subsequence

(3.60) θn\displaystyle\theta_{n} θ\displaystyle\rightarrow\theta weakly in LtqLxp for all p[1,) and q[1,pp1),\displaystyle\text{ weakly in }L_{t}^{q}L_{x}^{p}\textrm{ for all }p\in[1,\infty)\textrm{ and }q\in\left[1,\frac{p}{p-1}\right),
θn\displaystyle\theta_{n} θ\displaystyle\rightarrow\theta weakly in LtqWx1,q for all q[1,43)\displaystyle\text{ weakly in }L_{t}^{q}W^{1,q}_{x}\textrm{ for all }q\in\left[1,\frac{4}{3}\right)

that is (3.4). For nonlinear term, where we cannot a priori identify the limit due to the missing compactness of the temperature θn\theta_{n} and the extra stress tensor 𝔽n\mathbb{F}_{n}, we simply use the symbol a¯\overline{a} for a weak limit of the sequence ana_{n}. Therefore, using (3.58)–(3.60), the assumptions (2.2), (2.4), (2.5) and the uniform estimates (3.49)–(3.51) and the fact that 𝔹n=𝔽n𝔽nT\mathbb{B}_{n}=\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}, we deduce

(3.61) x𝒗n𝔽n\displaystyle\nabla_{x}\boldsymbol{v}_{n}\,\mathbb{F}_{n} x𝒗𝔽¯\displaystyle\rightharpoonup\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}} weakly in Lt,x43,\displaystyle\text{ weakly in }L^{\frac{4}{3}}_{t,x},
𝔽n𝔽nT\displaystyle\mathbb{F}_{n}\,\mathbb{F}_{n}^{T} 𝔽𝔽T¯\displaystyle\rightharpoonup\overline{\mathbb{F}\,\mathbb{F}^{T}} weakly in Lt,x2,\displaystyle\text{ weakly in }L^{2}_{t,x},
|𝔽n|2\displaystyle|\mathbb{F}_{n}|^{2} |𝔽|2¯\displaystyle\rightharpoonup\overline{|\mathbb{F}|^{2}} weakly in Lt,x2,\displaystyle\text{ weakly in }L^{2}_{t,x},
𝔽n𝔽nT𝔽n\displaystyle\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}\,\mathbb{F}_{n} 𝔽𝔽T𝔽¯\displaystyle\rightharpoonup\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}} weakly in Lt,x43,\displaystyle\text{ weakly in }L^{\frac{4}{3}}_{t,x},
g(θn)𝔽n𝔽nT\displaystyle g(\theta_{n})\mathbb{F}_{n}\,\mathbb{F}_{n}^{T} g(θ)𝔽𝔽T¯\displaystyle\rightharpoonup\overline{g(\theta)\mathbb{F}\,\mathbb{F}^{T}} weakly in Lt,x2,\displaystyle\text{ weakly in }L^{2}_{t,x},
δ(θn)𝔽n𝔽nT𝔽n\displaystyle\delta(\theta_{n})\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}\,\mathbb{F}_{n} δ(θ)𝔽𝔽T𝔽¯\displaystyle\rightharpoonup\overline{\delta(\theta)\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}} weakly in Lt,xq for all q[1,87),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in\left[1,\frac{8}{7}\right),
δ(θn)𝔽n\displaystyle\delta(\theta_{n})\mathbb{F}_{n} δ(θ)𝔽¯\displaystyle\rightharpoonup\overline{\delta(\theta)\mathbb{F}} weakly in Lt,xq for all q[1,43),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in\left[1,\frac{4}{3}\right),
ν(θn)𝔻𝒗n\displaystyle\nu(\theta_{n})\mathbb{D}\boldsymbol{v}_{n} ν(θ)𝔻𝒗¯\displaystyle\rightharpoonup\overline{\nu(\theta)\mathbb{D}\boldsymbol{v}} weakly in Lt,x2.\displaystyle\text{ weakly in }L^{2}_{t,x}.

Hence, we may let nn\to\infty in (1.14)1–(1.14)2 and using the convergence results for initial data (3.1) we can obtain by very classical procedure that

(3.62) 0TΩ𝒗t𝝋𝒗𝒗:x𝝋+ν(θ)𝔻𝒗¯:x𝝋+g(θ)𝔽𝔽T¯:x𝝋dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\boldsymbol{v}\cdot\partial_{t}\boldsymbol{\varphi}-\boldsymbol{v}\otimes\boldsymbol{v}:\nabla_{x}\boldsymbol{\varphi}+\overline{\nu(\theta)\mathbb{D}\boldsymbol{v}}:\nabla_{x}\boldsymbol{\varphi}+\overline{g(\theta)\mathbb{F}\,\mathbb{F}^{T}}:\nabla_{x}\boldsymbol{\varphi}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝒗0(x)𝝋(0,x)dx\displaystyle\qquad=\int_{\Omega}\boldsymbol{v}_{0}(x)\cdot\boldsymbol{\varphi}(0,x)\mathop{}\!\mathrm{d}x

for any 𝝋𝒞c1([0,T)×Ω;2)\boldsymbol{\varphi}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2}) with divx𝝋=0\operatorname{div}_{x}\boldsymbol{\varphi}=0,

(3.63) 0TΩ𝔽:t𝔾𝔽𝒗x𝔾x𝒗𝔽¯:𝔾+12(δ(θ)𝔽𝔽T𝔽¯δ(θ)𝔽¯):𝔾dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\mathbb{F}:\partial_{t}\mathbb{G}-\mathbb{F}\otimes\boldsymbol{v}\because\nabla_{x}\mathbb{G}-\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}:\mathbb{G}+\frac{1}{2}\left(\overline{\delta(\theta)\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}}-\overline{\delta(\theta)\mathbb{F}}\right):\mathbb{G}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝔽0(x):𝔾(0,x)dx\displaystyle\qquad=\int_{\Omega}\mathbb{F}_{0}(x):\mathbb{G}(0,x)\mathop{}\!\mathrm{d}x

for any 𝔾𝒞c1([0,T)×Ω;2×2)\mathbb{G}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2\times 2}). Note that (LABEL:almost_weak_formulation_u)–(LABEL:almost_weak_formulation_F) implies (LABEL:weak_formulation_u_g_theta)– (LABEL:weak_formulation_F_g_theta) provided that we show the point-wise convergence results

(3.64) θn\displaystyle\theta_{n} θ\displaystyle\to\theta almost everywhere in (0,T)×Ω,\displaystyle\textrm{almost everywhere in }(0,T)\times\Omega,
𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\to\mathbb{F} almost everywhere in (0,T)×Ω.\displaystyle\textrm{almost everywhere in }(0,T)\times\Omega.

In the remaining parts of this section we focus mainly on the proof of (3.64).

3.4. Compactness of the temperature θn\theta_{n}

For the proof of the compactness of the temperature, we use the new variant of the entropy method adapted to the setting of the paper. To do this, we apply the div–curl lemma on the properly chosen quantities that lead to the desired compactness. We recall the definition of entropy ηn\eta_{n} in (3.15) and the corresponding estimate (3.54)–(3.55) and also the definition of the flux 𝒒n\boldsymbol{q}_{n} in (3.56) and the related estimate (3.57). Therefore, we can extract a subsequence and η¯\overline{\eta} and 𝒒¯\overline{\boldsymbol{q}} such that

(3.65) ηn\displaystyle\eta_{n} η¯\displaystyle\rightharpoonup\overline{\eta} weakly in Lt,xq for all q[1,2),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in[1,2),
𝒒n\displaystyle\boldsymbol{q}_{n} 𝒒¯\displaystyle\rightharpoonup\overline{\boldsymbol{q}} weakly in Lt,xq for all q[0,43).\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in\left[0,\frac{4}{3}\right).

Furthermore, it follows from (LABEL:eq:entropy_equality_for_g_theta) and that the time-space vector (ηn,𝒒n)(\eta_{n},\boldsymbol{q}_{n}) and its time-space divergence fulfill

(3.66) divt,x(ηn,𝒒n)Lt,x1\displaystyle\|\operatorname{div}_{t,x}(\eta_{n},\boldsymbol{q}_{n})\|_{L^{1}_{t,x}} =tηn+divx𝒒nLt,x1\displaystyle=\|\partial_{t}\eta_{n}+\operatorname{div}_{x}\boldsymbol{q}_{n}\|_{L^{1}_{t,x}}
=κ(θn)|xθn|2θn2+ν(θn)|𝔻𝒗n|2θn+δ(θn)|𝔹n𝕀|2θnLt,x1C,\displaystyle=\left\|\frac{\kappa(\theta_{n})|\nabla_{x}\theta_{n}|^{2}}{\theta^{2}_{n}}+\frac{\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}}{\theta_{n}}+\frac{\delta(\theta_{n})|\mathbb{B}_{n}-\mathbb{I}|^{2}}{\theta_{n}}\right\|_{L^{1}_{t,x}}\leq C,

where for the last inequality we used (3.23). Consequently, using the Sobolev embedding we see that

(3.67) {divt,x(ηn,𝒒n)}n=1 is pre-compact in (W01,4((0,T)×Ω)).\displaystyle\{\operatorname{div}_{t,x}(\eta_{n},\boldsymbol{q}_{n})\}_{n=1}^{\infty}\qquad\text{ is pre-compact in }(W^{1,4}_{0}((0,T)\times\Omega))^{*}.

In a very similar way, we use the estimates (3.53) and (3.58) to get

(3.68) f(𝔹n)\displaystyle f(\mathbb{B}_{n}) f(𝔹)¯\displaystyle\rightharpoonup\overline{f(\mathbb{B})} weakly in Lt,xq for all q[1,2),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in[1,2),
f(𝔹n)𝒗n\displaystyle f(\mathbb{B}_{n})\,\boldsymbol{v}_{n} f(𝔹)¯𝒗\displaystyle\rightharpoonup\overline{f(\mathbb{B})}\,\boldsymbol{v} weakly in Lt,xqfor all q[1,43).\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{for all }q\in\left[1,\frac{4}{3}\right).\

Next, we deduce from (3.18) that

(3.69) divt,x(f(𝔹n),f(𝔹n)𝒗n)Lt,x1\displaystyle\|\operatorname{div}_{t,x}(f(\mathbb{B}_{n}),f(\mathbb{B}_{n})\,\boldsymbol{v}_{n})\|_{L^{1}_{t,x}} =tf(𝔹n)+divx(f(𝔹n)𝒗n)Lt,x1\displaystyle=\|\partial_{t}f(\mathbb{B}_{n})+\operatorname{div}_{x}\left(f(\mathbb{B}_{n})\,\boldsymbol{v}_{n}\right)\|_{L^{1}_{t,x}}
=2(𝔹n𝕀):𝔻𝒗nδ(θ)|𝔹n𝕀|2Lt,x1C,\displaystyle=\left\|2(\mathbb{B}_{n}-\mathbb{I}):\mathbb{D}\boldsymbol{v}_{n}-\delta(\theta)|\mathbb{B}_{n}-\mathbb{I}|^{2}\right\|_{L^{1}_{t,x}}\leq C,

where the last inequality follows from the uniform estimates (3.35) and (3.29). Therefore, using again the Sobolev embedding we get that

(3.70) {divt,x(f(𝔹n),f(𝔹n)𝒗n)}n=1 is pre-compact in (W01,4((0,T)×Ω)).\displaystyle\{\operatorname{div}_{t,x}(f(\mathbb{B}_{n}),f(\mathbb{B}_{n})\,\boldsymbol{v}_{n})\}_{n=1}^{\infty}\qquad\text{ is pre-compact in }(W^{1,4}_{0}((0,T)\times\Omega))^{*}.

We have constructed two weakly convergent time-space vector fields, namely (f(𝔹n),f(𝔹n)𝒗n)n=1(f(\mathbb{B}_{n}),f(\mathbb{B}_{n})\boldsymbol{v}_{n})_{n=1}^{\infty} and (ηn,𝒒n)n=1(\eta_{n},\boldsymbol{q}_{n})_{n=1}^{\infty}, whose space-time divergence divt,x\operatorname{div}_{t,x} is precompact in W0,t,x1,43W^{-1,\frac{4}{3}}_{0,t,x}. Next, we focus on finding a weakly convergent vector field whose time-space curlt,x\operatorname{curl}_{t,x} is precompact in W0,t,x1,43W^{-1,\frac{4}{3}}_{0,t,x}. To do so, we consider the sequence {θn13}n=1\{\theta_{n}^{\frac{1}{3}}\}_{n=1}^{\infty} and by (3.44) we know that there exists θ13¯\overline{\theta^{\frac{1}{3}}} such that

(3.71) θn13\displaystyle\theta_{n}^{\frac{1}{3}} θ13¯\displaystyle\rightharpoonup\overline{\theta^{\frac{1}{3}}} weakly in Lt,xq for all q[1,6).\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in[1,6).

Next, it follows from (3.37) that

xθn13Lt,x2=13xθnθn23Lt,x2C.\displaystyle\|\nabla_{x}\theta_{n}^{\frac{1}{3}}\|_{L^{2}_{t,x}}=\frac{1}{3}\left\|\frac{\nabla_{x}\theta_{n}}{\theta_{n}^{\frac{2}{3}}}\right\|_{L^{2}_{t,x}}\leq C.

Therefore, for the vector field given as (θn13,0,0)(\theta_{n}^{\frac{1}{3}},0,0), we get

curlt,x(θn13,0,0)Lt,x2Cxθn13Lt,x2C.\displaystyle\|\operatorname{curl}_{t,x}\,(\theta_{n}^{\frac{1}{3}},0,0)\|_{L^{2}_{t,x}}\leq C\|\nabla_{x}\theta_{n}^{\frac{1}{3}}\|_{L^{2}_{t,x}}\leq C.

Consequently, by the Sobolev embedding, we have

(3.72) {curlt,x(θn13,0,0)}n=1 is pre-compact in (W01,2((0,T)×Ω)).\displaystyle\{\operatorname{curl}_{t,x}\,(\theta_{n}^{\frac{1}{3}},0,0)\}_{n=1}^{\infty}\qquad\text{ is pre-compact in }(W^{1,2}_{0}((0,T)\times\Omega))^{*}.

Hence, we are able to apply the div–curl lemma (see Lemma A.2) on the sequences

{(θn13,0,0)}n=1,{(f(𝔹n),f(𝔹n)𝒗n)}n=1 and {(ηn,𝒒n)}n=1\{(\theta_{n}^{\frac{1}{3}},0,0)\}_{n=1}^{\infty},\quad\{(f(\mathbb{B}_{n}),f(\mathbb{B}_{n})\,\boldsymbol{v}_{n})\}_{n=1}^{\infty}\quad\textrm{ and }\quad\{(\eta_{n},\boldsymbol{q}_{n})\}_{n=1}^{\infty}

and thanks to (3.65), (3.67), (3.68), (3.70), (3.71) and (3.72) we observe that

(θn13,0,0)(f(𝔹n),f(𝔹n)𝒗n)\displaystyle(\theta_{n}^{\frac{1}{3}},0,0)\cdot(f(\mathbb{B}_{n}),f(\mathbb{B}_{n})\,\boldsymbol{v}_{n}) (θ13¯,0,0)(f(𝔹)¯,f(𝔹)¯𝒗)\displaystyle\rightharpoonup(\overline{\theta^{\frac{1}{3}}},0,0)\cdot(\overline{f(\mathbb{B})},\overline{f(\mathbb{B})}\,\boldsymbol{v}) weakly in Lt,x1,\displaystyle\textrm{weakly in }L^{1}_{t,x},
(θn13,0,0)(ηn,𝒒n)\displaystyle(\theta_{n}^{\frac{1}{3}},0,0)\cdot(\eta_{n},\boldsymbol{q}_{n}) (θ13¯,0,0)(η¯,𝒒¯)\displaystyle\rightharpoonup(\overline{\theta^{\frac{1}{3}}},0,0)\cdot(\overline{\eta},\overline{\boldsymbol{q}}) weakly in Lt,x1.\displaystyle\text{ weakly in }L^{1}_{t,x}.

This in particular gives that

(3.73) ηnθn13\displaystyle\eta_{n}\,\theta_{n}^{\frac{1}{3}} η¯θ13¯\displaystyle\rightharpoonup\overline{\eta}\,\overline{\theta^{\frac{1}{3}}} weakly in Lt,x1,\displaystyle\text{ weakly in }L^{1}_{t,x},
(3.74) f(𝔹n)θn13\displaystyle f(\mathbb{B}_{n})\,\theta_{n}^{\frac{1}{3}} f(𝔹)¯θ13¯\displaystyle\rightharpoonup\overline{f(\mathbb{B})}\,\overline{\theta^{\frac{1}{3}}} weakly in Lt,x1.\displaystyle\text{ weakly in }L^{1}_{t,x}.

Since the function ff is nonnegative and gg^{\prime} is nonincreasing (due to the fact that gg is concave), we have that for any wLt,xqw\in L^{q}_{t,x} with q>5q>5

0f(𝔹n)(g(θn)+g(w))(θn13w13) almost everywhere in (0,T)×Ω.0\leq f(\mathbb{B}_{n})(-g^{\prime}(\theta_{n})+g^{\prime}(w))(\theta_{n}^{\frac{1}{3}}-w^{\frac{1}{3}})\qquad\textrm{ almost everywhere in }(0,T)\times\Omega.

In particular, for w:=(θ13¯)3w:=\left(\overline{\theta^{\frac{1}{3}}}\right)^{3} we obtain

0f(𝔹n)(g(θn)+g((θ13¯)3))(θn13θ13¯),\displaystyle 0\leq f(\mathbb{B}_{n})\left(-g^{\prime}(\theta_{n})+g^{\prime}\left(\left(\overline{\theta^{\frac{1}{3}}}\right)^{3}\right)\right)\left(\theta_{n}^{\frac{1}{3}}-\overline{\theta^{\frac{1}{3}}}\right),

which by simple algebraic manipulation leads to

(3.75) f(𝔹n)g(θn)θ13¯f(𝔹n)g(θn)θn13+f(𝔹n)g((θ13¯)3)(θn13θ13¯)\displaystyle-f(\mathbb{B}_{n})g^{\prime}(\theta_{n})\overline{\theta^{\frac{1}{3}}}\leq-f(\mathbb{B}_{n})g^{\prime}(\theta_{n})\theta_{n}^{\frac{1}{3}}+f(\mathbb{B}_{n})g^{\prime}\left(\left(\overline{\theta^{\frac{1}{3}}}\right)^{3}\right)\left(\theta_{n}^{\frac{1}{3}}-\overline{\theta^{\frac{1}{3}}}\right)

almost everywhere in (0,T)×Ω(0,T)\times\Omega. Next, since gg^{\prime} is bounded, see (2.6), we can also deduce

(3.76) g(θn)f(𝔹n)\displaystyle g^{\prime}(\theta_{n})f(\mathbb{B}_{n}) g(θ)f(𝔹)¯\displaystyle\rightharpoonup\overline{g^{\prime}(\theta)f(\mathbb{B})} weakly in Lt,xq for all q[1,2),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in[1,2),
(3.77) f(𝔹n)g(θn)θn13\displaystyle f(\mathbb{B}_{n})g^{\prime}(\theta_{n})\theta_{n}^{\frac{1}{3}} f(𝔹)g(θ)θ13¯\displaystyle\rightharpoonup\overline{f(\mathbb{B})g^{\prime}(\theta)\theta^{\frac{1}{3}}} weakly in Lt,x1.\displaystyle\text{ weakly in }L^{1}_{t,x}.

Thanks to (3.74), and due to the fact that g(θ13¯)Lt,xg^{\prime}\left(\overline{\theta^{\frac{1}{3}}}\right)\in L^{\infty}_{t,x}, we have

limn0TΩf(𝔹n)g((θ13¯)3)(θn13θ13¯)dxdt=0.\lim_{n\to\infty}\int_{0}^{T}\int_{\Omega}f(\mathbb{B}_{n})g^{\prime}\left(\left(\overline{\theta^{\frac{1}{3}}}\right)^{3}\right)\left(\theta_{n}^{\frac{1}{3}}-\overline{\theta^{\frac{1}{3}}}\right)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0.

Letting nn\to\infty, using the above identity and also (3.76) we get

(3.78) 0TΩg(θ)f(𝔹)¯θ13¯dxdt0TΩg(θ)f(𝔹)θ13¯dxdt.\displaystyle\int_{0}^{T}\int_{\Omega}-\overline{g^{\prime}(\theta)f(\mathbb{B})}\,\overline{\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq\int_{0}^{T}\int_{\Omega}-\overline{g^{\prime}(\theta)f(\mathbb{B})\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.

Denoting also lnθ¯\overline{\ln\theta} as the weak limit

lnθnlnθ¯ weakly in Lt,x2,\ln\theta_{n}\rightharpoonup\overline{\ln\theta}\qquad\textrm{ weakly in }L^{2}_{t,x},

using (3.78), the definition of ηn\eta_{n}, see (3.15), and applying (3.73), and (3.78) we obtain

(3.79) 0TΩlnθ¯θ13¯dxdt\displaystyle\int_{0}^{T}\int_{\Omega}\overline{\ln\theta}\,\overline{\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t =(3.15)0TΩη¯θ13¯dxdt+0TΩg(θ)f(𝔹)¯θ13¯dxdt\displaystyle\overset{\eqref{eq:entropyn}}{=}\int_{0}^{T}\int_{\Omega}\overline{\eta}\,\overline{\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{0}^{T}\int_{\Omega}\overline{g^{\prime}(\theta)f(\mathbb{B})}\,\overline{\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
(3.78)limn+0TΩηnθn13dxdt+0TΩg(θ)f(𝔹)θ13¯dxdt\displaystyle\overset{\eqref{ineq:for_monotonicity_trick_f_of_B}}{\geq}\lim_{n\to+\infty}\int_{0}^{T}\int_{\Omega}\eta_{n}\,\theta_{n}^{\frac{1}{3}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{0}^{T}\int_{\Omega}\overline{g^{\prime}(\theta)f(\mathbb{B})\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=(3.15)limn0TΩlnθnθn13dxdtlimn0TΩg(θn)f(𝔹n)θn13dxdt\displaystyle\overset{\eqref{eq:entropyn}}{=}\lim_{n\to\infty}\int_{0}^{T}\int_{\Omega}\ln\theta_{n}\theta_{n}^{\frac{1}{3}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t-\lim_{n\to\infty}\int_{0}^{T}\int_{\Omega}g^{\prime}(\theta_{n})f(\mathbb{B}_{n})\theta_{n}^{\frac{1}{3}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
+0TΩg(θ)f(𝔹)θ13¯dxdt\displaystyle\qquad{}\qquad+\int_{0}^{T}\int_{\Omega}\overline{g^{\prime}(\theta)f(\mathbb{B})\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
0TΩ(lnθ)θ13¯dxdt.\displaystyle\;\geq\int_{0}^{T}\int_{\Omega}\overline{(\ln\theta)\theta^{\frac{1}{3}}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.

Now, notice that wln(w)w\mapsto\ln(w) and ww13w\mapsto w^{\frac{1}{3}} are increasing functions. Thus, for any nonnegative function wL1((0,T)×Ω)w\in L^{1}((0,T)\times\Omega), fulfilling lnwLt,x2\ln w\in L^{2}_{t,x}, there holds

(3.80) 0TΩ(lnθnlnw)(θn13w13)dxdt0.\displaystyle\int_{0}^{T}\int_{\Omega}(\ln\theta_{n}-\ln w)(\theta_{n}^{\frac{1}{3}}-w^{\frac{1}{3}})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\geq 0.

Due to (3.79) we may let nn\to\infty and deduce

(3.81) 0TΩ(lnθ¯lnw)(θ13¯w13)dxdt0.\displaystyle\int_{0}^{T}\int_{\Omega}(\overline{\ln\theta}-\ln w)(\overline{\theta^{\frac{1}{3}}}-w^{\frac{1}{3}})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\geq 0.

Moving forward, we repeat the Minty method. Assume that hLt,xh\in L^{\infty}_{t,x} is arbitrary and λ>0\lambda>0 and set

w:=elnθ¯λhw:=e^{\overline{\ln\theta}-\lambda h}

in (3.81). The fact that such ww is admissible follows from the following. Since hLt,xh\in L^{\infty}_{t,x}, then eλhLt,xe^{-\lambda h}\in L^{\infty}_{t,x}. Moreover, since the exponential is a convex function, we can use the weak lower semicontinuity to deduce

exp(lnθ¯)Lt,x1lim infnexp(lnθn)Lt,x1=lim infnθnLt,x1C.\|\exp(\overline{\ln\theta})\|_{L^{1}_{t,x}}\leq\liminf_{n\to\infty}\|\exp(\ln\theta_{n})\|_{L^{1}_{t,x}}=\liminf_{n\to\infty}\|\theta_{n}\|_{L^{1}_{t,x}}\leq C.

Thus, ww is an admissible function in (3.81) and with this choice after division by λ\lambda it follows that

0TΩ(θ13¯e13(lnθ¯λh))hdxdt0.\int_{0}^{T}\int_{\Omega}\left(\overline{\theta^{\frac{1}{3}}}-e^{\frac{1}{3}\left(\overline{\ln\theta}-\lambda h\right)}\right)\,h\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\geq 0.

Finally, letting λ0+\lambda\to 0_{+} and using the fact that hh is arbitrary we get

(3.82) e13lnθ¯=θ13¯=e13lnθ¯ a.e. in (0,T)×Ω.\overline{e^{\frac{1}{3}\ln\theta}}=\overline{\theta^{\frac{1}{3}}}=e^{\frac{1}{3}\overline{\ln\theta}}\text{ a.e. in }(0,T)\times\Omega.

We show that the above identity implies the strong convergence of the temperature claimed in (3.64)1. To do so, we set w:=θnw:=\theta_{n} in (3.81). Then, using (3.79), (3.82) and the fact that the exponential is the increasing function, we deduce

limn+0TΩ|(lnθ¯lnθn)(θ13¯θn13)|dxdt=limn+0TΩ|(lnθ¯lnθn)(e13lnθ¯e13lnθn)|dxdt\displaystyle\lim_{n\to+\infty}\int_{0}^{T}\int_{\Omega}\left|(\overline{\ln\theta}-\ln\theta_{n})\left(\overline{\theta^{\frac{1}{3}}}-\theta_{n}^{\frac{1}{3}}\right)\right|\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\lim_{n\to+\infty}\int_{0}^{T}\int_{\Omega}\left|(\overline{\ln\theta}-\ln\theta_{n})\left(e^{\frac{1}{3}\overline{\ln\theta}}-e^{\frac{1}{3}\ln\theta_{n}}\right)\right|\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=limn+0TΩ(lnθ¯lnθn)(θ13¯θn13)dxdt=0.\displaystyle\quad=\lim_{n\to+\infty}\int_{0}^{T}\int_{\Omega}(\overline{\ln\theta}-\ln\theta_{n})(\overline{\theta^{\frac{1}{3}}}-\theta_{n}^{\frac{1}{3}})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0.

Hence, up to the subsequence that we do not relabel,

(lnθ¯lnθn)(e13lnθ¯e13lnθn)=(lnθ¯lnθn)(θ13¯θn13)0 almost everywhere in (0,T)×Ω.\displaystyle(\overline{\ln\theta}-\ln\theta_{n})(e^{\frac{1}{3}\overline{\ln\theta}}-e^{\frac{1}{3}\ln\theta_{n}})=(\overline{\ln\theta}-\ln\theta_{n})(\overline{\theta^{\frac{1}{3}}}-\theta_{n}^{\frac{1}{3}})\rightarrow 0\text{ almost everywhere in }(0,T)\times\Omega.

Since xe13xx\mapsto e^{\frac{1}{3}x} is a strictly increasing function, the above convergence result is possible only if

lnθnlnθ¯ almost everywhere in (0,T)×Ω,\displaystyle\ln\theta_{n}\rightarrow\overline{\ln\theta}\quad\text{ almost everywhere in }(0,T)\times\Omega,

which in turn implies that

(3.83) θnθ:=elnθ¯ almost everywhere in (0,T)×Ω,\displaystyle\theta_{n}\rightarrow\theta:=e^{\overline{\ln\theta}}\text{ almost everywhere in }(0,T)\times\Omega,

that is (3.64)1.

3.5. Compactness of 𝔽n\mathbb{F}_{n}

In this part we show (3.64)2 and even more we prove the following convergence result

(3.84) 𝔽n𝔽 strongly in L2((0,T)×Ω).\displaystyle\mathbb{F}_{n}\rightarrow\mathbb{F}\text{ strongly in }L^{2}((0,T)\times\Omega).

We closely follow [7, Subsection 6.4] and [8, Subsection 3.4]. Although the two-dimensional setting could indicate that the primary choice would be the use of [7] the opposite is true. In fact, if the viscosity ν\nu and the shear modules gg were independent of the temperature, the proof in [7] could easily be adapted to our setting. However, the temperature dependence brings additional difficulties111The key problem is that even if we have the almost everywhere convergence of the temperature θnθ\theta_{n}\to\theta, we cannot prove the following identification g(θ)(𝔽𝔽T):𝔻𝒗¯\displaystyle\overline{g(\theta)(\mathbb{F}\,\mathbb{F}^{T}):\mathbb{D}\boldsymbol{v}} =g(θ)(𝔽𝔽T):𝔻𝒗¯,\displaystyle=g(\theta)\overline{(\mathbb{F}\,\mathbb{F}^{T}):\mathbb{D}\boldsymbol{v}}, ν(θ)|𝔻𝒗|2¯\displaystyle\overline{\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}} =ν(θ)|𝔻𝒗|2¯\displaystyle=\nu(\theta)\overline{|\mathbb{D}\boldsymbol{v}|^{2}} in sense of measures, which is true in case of constant ν\nu and gg. and therefore we must proceed differently, while clearly clearly indicating the corresponding differences.

In [7, 8], it is shown, that (3.84) holds true provided that there exists LL2((0,T)×Ω)L\in L^{2}((0,T)\times\Omega) such that

(3.85) 0TΩ(|𝔽|2¯|𝔽|2)tϕ𝒗(|𝔽|2¯|𝔽|2)xϕdxdt0TΩL(|𝔽|2¯|𝔽|2)ϕdxdt\displaystyle\int_{0}^{T}\int_{\Omega}-(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\partial_{t}\phi-\boldsymbol{v}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\,\nabla_{x}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq\int_{0}^{T}\int_{\Omega}L(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t

holds for any nonnegative ϕ𝒞01((,T)×Ω)\phi\in\mathcal{C}^{1}_{0}((-\infty,T)\times\Omega). Such a setting would be insufficient for the paper, but we can rather straightforwardly generalise the above result as follows. The convergence properties (3.84) follows from (3.85) by using the renormalization procedure (see (cf. [7, (6.79)] for details) and for that it is enough to require only that

(3.86) LL1((0,T)×Ω),L(|𝔽|2¯|𝔽|2)L1((0,T)×Ω),\begin{split}&L\in L^{1}((0,T)\times\Omega),\\ &L(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\in L^{1}((0,T)\times\Omega),\end{split}

which is the setting we can obtain.

We proceed here only formally, and for rigorous proof we refer to [8]. We also need to use here the concept of the biting limit and the Chacon biting lemma, see [1]. According to that, we know that for any sequence ana_{n} fulfilling

anLt,x1C,\|a_{n}\|_{L^{1}_{t,x}}\leq C,

there exists aL1((0,T)×Ω)a\in L^{1}((0,T)\times\Omega) and there exists nondecreasing sequence of measurable sets Ej(0,T)×ΩE_{j}\subset(0,T)\times\Omega, such that

(3.87) an\displaystyle a_{n} a\displaystyle\rightharpoonup a weakly in L1(Ej) for all j,\displaystyle\textrm{weakly in }L^{1}(E_{j})\quad\textrm{ for all }j\in\mathbb{N},
|((0,T)×Ω)Ej|\displaystyle|((0,T)\times\Omega)\setminus E_{j}| 0\displaystyle\to 0 as jj\to\infty.

We call aa the biting limit. It is clear that in case the classical weak limit exits, the biting and the weak limit must coincide in the space L1L^{1} and therefore in what follows we primarily work with the biting limits.

Thanks to Egorov’s theorem and (3.83) we may identify the weak limits in (3.61) and to observe

(3.88) g(θ)𝔽𝔽T¯\displaystyle\overline{g(\theta)\mathbb{F}\,\mathbb{F}^{T}} =g(θ)𝔽𝔽T¯=g(θ)𝔹,\displaystyle=g(\theta)\overline{\mathbb{F}\,\mathbb{F}^{T}}=g(\theta)\mathbb{B},
δ(θ)|𝔽|2¯\displaystyle\overline{\delta(\theta)|\mathbb{F}|^{2}} =δ(θ)|𝔽|2¯,\displaystyle=\delta(\theta)\overline{|\mathbb{F}|^{2}},
δ(θ)𝔽𝔽T𝔽¯\displaystyle\overline{\delta(\theta)\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}} =δ(θ)𝔽𝔽T𝔽¯,\displaystyle=\delta(\theta)\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}},
ν(θ)𝔻𝒗n¯\displaystyle\overline{\nu(\theta)\mathbb{D}\boldsymbol{v}_{n}} =ν(θ)𝔻𝒗\displaystyle=\nu(\theta)\mathbb{D}\boldsymbol{v}

almost everywhere in (0,T)×Ω(0,T)\times\Omega. In addition, due to the uniform estimates (3.34)–(3.35) and also (3.27), by using the Egorov theorem and the point-wise convergence of the temperature θn\theta_{n} in (3.83), we may conclude that the biting limits fulfills the following

(3.89) δ(θ)|𝔽𝔽T|2¯\displaystyle\overline{\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}} =δ(θ)|𝔽𝔽T|2¯,\displaystyle=\delta(\theta)\overline{|\mathbb{F}\,\mathbb{F}^{T}|^{2}},
ν(θ)|𝔻𝒗|2¯\displaystyle\overline{\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}} =ν(θ)|𝔻𝒗|2¯,\displaystyle=\nu(\theta)\overline{|\mathbb{D}\boldsymbol{v}|^{2}},
g(θ)𝔽𝔽T:𝔻𝒗¯\displaystyle\overline{g(\theta)\mathbb{F}\,\mathbb{F}^{T}:\mathbb{D}\boldsymbol{v}} =g(θ)𝔽𝔽T:𝔻𝒗¯\displaystyle=g(\theta)\overline{\mathbb{F}\,\mathbb{F}^{T}:\mathbb{D}\boldsymbol{v}}

almost everywhere in (0,T)×Ω(0,T)\times\Omega.

Next, take the scalar product of (1.14)2 with 2𝔽2\mathbb{F} and obtain

t|𝔽n|2+divx(|𝔽n|2𝒗n)2x𝒗n:(𝔽n𝔽nT)+δ(θn)(|𝔽n𝔽nT|2|𝔽n|2)=0.\begin{split}\partial_{t}|\mathbb{F}_{n}|^{2}+\operatorname{div}_{x}(|\mathbb{F}_{n}|^{2}\boldsymbol{v}_{n})-2\nabla_{x}\boldsymbol{v}_{n}:(\mathbb{F}_{n}\,\mathbb{F}_{n}^{T})+\delta(\theta_{n})(|\mathbb{F}_{n}\,\mathbb{F}^{T}_{n}|^{2}-|\mathbb{F}_{n}|^{2})=0.\end{split}

Letting nn\to\infty we gain with the help of the above convergence results (see [8] for rigorous justification) the following

(3.90) t|𝔽|2¯+divx(|𝔽|2¯𝒗)2x𝒗:(𝔽𝔽T)¯+δ(θ)(|𝔽𝔽T|2¯|𝔽|2¯)=0.\begin{split}\partial_{t}\overline{|\mathbb{F}|^{2}}+\operatorname{div}_{x}(\overline{|\mathbb{F}|^{2}}\boldsymbol{v})-2\overline{\nabla_{x}\boldsymbol{v}:(\mathbb{F}\,\mathbb{F}^{T})}+\delta(\theta)(\overline{|\mathbb{F}\,\mathbb{F}^{T}|^{2}}-\overline{|\mathbb{F}|^{2}})=0.\end{split}

Here, we want to set 𝔾:=𝔽ϕ\mathbb{G}:=\mathbb{F}\phi in (LABEL:almost_weak_formulation_F), however such a setting is not possible due to the low integrability of the term δ(θ)𝔽𝔽T𝔽¯\delta(\theta)\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}}, which does not belong to L43L^{\frac{4}{3}}, and therefore we cannot test by functions which are only in L4L^{4}. To solve this issue, we use as a test function

𝔽1+ε|𝔽|,\frac{\mathbb{F}}{1+\varepsilon|\mathbb{F}|},

which is bounded. Thanks to this choice, and after the classical renormalisation procedure (see [14]), and with the help of already obtain convergence results, we deduce

2tε|𝔽|2ln(1+ε|𝔽|2)ε2+2divx(ε|𝔽|2ln(1+ε|𝔽|2)ε2𝒗)2x𝒗𝔽¯:𝔽1+ε|𝔽|+δ(θ)(𝔽𝔽T𝔽¯𝔽):𝔽1+ε|𝔽|=0.\begin{split}&2\partial_{t}\frac{\varepsilon|\mathbb{F}|^{2}-\ln(1+\varepsilon|\mathbb{F}|^{2})}{\varepsilon^{2}}+2\operatorname{div}_{x}\left(\frac{\varepsilon|\mathbb{F}|^{2}-\ln(1+\varepsilon|\mathbb{F}|^{2})}{\varepsilon^{2}}\boldsymbol{v}\right)\\ &\quad-2\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}:\frac{\mathbb{F}}{1+\varepsilon|\mathbb{F}|}+\delta(\theta)\left(\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}}-\mathbb{F}\right):\frac{\mathbb{F}}{1+\varepsilon|\mathbb{F}|}=0.\end{split}

Thanks to the fact we work with the biting limits, we can now easily let ε0+\varepsilon\to 0_{+} and conclude

(3.91) t|𝔽|2+divx(|𝔽|2𝒗)2x𝒗𝔽¯:𝔽+δ(θ)(𝔽𝔽T𝔽¯𝔽):𝔽=0.\begin{split}&\partial_{t}|\mathbb{F}|^{2}+\operatorname{div}_{x}\left(|\mathbb{F}|^{2}\boldsymbol{v}\right)-2\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}:\mathbb{F}+\delta(\theta)\left(\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}}-\mathbb{F}\right):\mathbb{F}=0.\end{split}

Subtracting (3.91) from (3.90), we see that

(3.92) t(|𝔽|2¯|𝔽|2)+divx((|𝔽|2¯|𝔽|2)𝒗)+δ(θ)(|𝔽𝔽T|2¯𝔽𝔽T𝔽¯:𝔽)=2(x𝒗:(𝔽𝔽T)¯x𝒗𝔽¯:𝔽)+δ(θ)(|𝔽|2¯|𝔽|2)\begin{split}&\partial_{t}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+\operatorname{div}_{x}\left((\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\boldsymbol{v}\right)+\delta(\theta)\left(\overline{|\mathbb{F}\,\mathbb{F}^{T}|^{2}}-\overline{\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}}:\mathbb{F}\right)\\ &\quad=2\left(\overline{\nabla_{x}\boldsymbol{v}:(\mathbb{F}\,\mathbb{F}^{T})}-\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}:\mathbb{F}\right)+\delta(\theta)(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\end{split}

Here, the last term on the left hand side is nonnegative since the mapping 𝔽𝔽𝔽T𝔽\mathbb{F}\mapsto\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F} is monotone, see [7, Lemma 4.2]. The last term on the right hand side is in the required form and we need to focus on the first term on the right hand side. To do so, we use [8, Theorem 1.7] and it follows from the equation (1.5)1, the convergence results (3.58) and (3.59), and from the assumptions (2.2) and (2.5) that

ν(θ)|𝔻𝒗|2¯+g(θ)𝔽𝔽T:x𝒗¯=ν(θ)𝔻𝒗¯:𝔻𝒗+g(θ)(𝔽𝔽T)¯:x𝒗,\overline{\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}}+\overline{g(\theta)\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{v}}=\overline{\nu(\theta)\mathbb{D}\boldsymbol{v}}:\mathbb{D}\boldsymbol{v}+\overline{g(\theta)(\mathbb{F}\,\mathbb{F}^{T})}:\nabla_{x}\boldsymbol{v},

almost everywhere in (0,T)×Ω(0,T)\times\Omega. Recall, that we consider her the biting limits. Using the strong convergence of the temperature (3.83), we deduce from the above identity that

(3.93) ν(θ)g(θ)|𝔻𝒗|2¯+𝔽𝔽T:x𝒗¯=ν(θ)g(θ)|𝔻𝒗|2+(𝔽𝔽T)¯:x𝒗.\frac{\nu(\theta)}{g(\theta)}\overline{|\mathbb{D}\boldsymbol{v}|^{2}}+\overline{\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{v}}=\frac{\nu(\theta)}{g(\theta)}|\mathbb{D}\boldsymbol{v}|^{2}+\overline{(\mathbb{F}\,\mathbb{F}^{T})}:\nabla_{x}\boldsymbol{v}.

Inserting (3.93) into (3.92), where we also neglect the last term on the left hand side, leads to the following inequality

(3.94) t(|𝔽|2¯|𝔽|2)+divx((|𝔽|2¯|𝔽|2)𝒗)+2ν(θ)g(θ)(|𝔻𝒗|2¯|𝔻𝒗|2)\displaystyle\partial_{t}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+\operatorname{div}_{x}\left((\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\boldsymbol{v}\right)+\frac{2\nu(\theta)}{g(\theta)}\left(\overline{|\mathbb{D}\boldsymbol{v}|^{2}}-|\mathbb{D}\boldsymbol{v}|^{2}\right)
2((𝔽𝔽T)¯:x𝒗x𝒗𝔽¯:𝔽)+δ(θ)(|𝔽|2¯|𝔽|2)\displaystyle\quad\leq 2\left(\overline{(\mathbb{F}\,\mathbb{F}^{T})}:\nabla_{x}\boldsymbol{v}-\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}:\mathbb{F}\right)+\delta(\theta)(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})
=2((𝔽𝔽T)¯𝔽𝔽T):𝔻𝒗+2(x𝒗𝔽x𝒗𝔽¯):𝔽+δ(θ)(|𝔽|2¯|𝔽|2)\displaystyle\quad=2\left(\overline{(\mathbb{F}\,\mathbb{F}^{T})}-\mathbb{F}\,\mathbb{F}^{T}\right):\mathbb{D}\boldsymbol{v}+2\left(\nabla_{x}\boldsymbol{v}\,\mathbb{F}-\overline{\nabla_{x}\boldsymbol{v}\,\mathbb{F}}\right):\mathbb{F}+\delta(\theta)(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})
2|𝔻𝒗|(|𝔽|2¯|𝔽|2)+2|𝔽|(|x𝒗|2¯|x𝒗|2)12(|𝔽|2¯|𝔽|2)12+δ(θ)(|𝔽|2¯|𝔽|2)\displaystyle\quad\leq 2|\mathbb{D}\boldsymbol{v}|(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+2|\mathbb{F}|(\overline{|\nabla_{x}\boldsymbol{v}|^{2}}-|\nabla_{x}\boldsymbol{v}|^{2})^{\frac{1}{2}}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})^{\frac{1}{2}}+\delta(\theta)(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})

Using the assumptions (2.2) and (2.5) and also the localised version of the Korn inequality, see Appendix in [8], we deduce from (3.94)

(3.95) t(|𝔽|2¯|𝔽|2)+divx((|𝔽|2¯|𝔽|2)𝒗)+2ν(θ)g(θ)(|𝔻𝒗|2¯|𝔻𝒗|2)\displaystyle\partial_{t}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+\operatorname{div}_{x}\left((\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})\boldsymbol{v}\right)+\frac{2\nu(\theta)}{g(\theta)}\left(\overline{|\mathbb{D}\boldsymbol{v}|^{2}}-|\mathbb{D}\boldsymbol{v}|^{2}\right)
2|𝔻𝒗|(|𝔽|2¯|𝔽|2)+C|𝔽|2(|𝔽|2¯|𝔽|2)+2ν(θ)g(θ)(|𝔻𝒗|2¯|𝔻𝒗|2)+δ(θ)(|𝔽|2¯|𝔽|2).\displaystyle\quad\leq 2|\mathbb{D}\boldsymbol{v}|(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+C|\mathbb{F}|^{2}(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2})+\frac{2\nu(\theta)}{g(\theta)}\left(\overline{|\mathbb{D}\boldsymbol{v}|^{2}}-|\mathbb{D}\boldsymbol{v}|^{2}\right)+\delta(\theta)(\overline{|\mathbb{F}|^{2}}-|\mathbb{F}|^{2}).

Finally, from the above inequality, we can deduce (3.85), where

L:=C(1+|𝔻𝒗|+|𝔽|2+δ(θ)).L:=C(1+|\mathbb{D}\boldsymbol{v}|+|\mathbb{F}|^{2}+\delta(\theta)).

Consequently, the strong convergence (3.84) follows. We want to emphasize that the above computations was rather formal, and we encourage the interested reader to [7, 8] for rigorous justifications of several steps.

The convergence (3.84) combined with the convergence results obtained previously is enough to conclude (LABEL:weak_formulation_u_g_theta), (LABEL:weak_formulation_F_g_theta) from (LABEL:almost_weak_formulation_u) and (LABEL:almost_weak_formulation_F). To show (LABEL:weak_formulation_theta_g_theta), we can use the a priori bounds (3.54)–(3.57) and let nn\to\infty in the weak formulation of (LABEL:eq:entropy_equality_for_g_theta). Using the point-wise convergence of θn\theta_{n} and 𝔽n\mathbb{F}_{n}, and the Fatou lemma for the nonnegative terms on the right hand side, we can easily conclude (LABEL:weak_formulation_theta_g_theta).

Furthermore, due to (3.83) and (3.84), and since θn>0\theta_{n}>0 and det𝔽n>0\det\mathbb{F}_{n}>0, we have

(3.96) θ0,det𝔽0, a.e. in (0,T)×Ω.\displaystyle\theta\geq 0,\quad\det\mathbb{F}\geq 0,\quad\text{ a.e. in }(0,T)\times\Omega.

Even more, the Fatou lemma together with the uniform estimates (3.23), (3.52) imply that

lnθLtLx1+lndet(𝔽𝔽T)LtLx1C,\|\ln\theta\|_{L^{\infty}_{t}L^{1}_{x}}+\|\ln\det(\mathbb{F}\,\mathbb{F}^{T})\|_{L^{\infty}_{t}L^{1}_{x}}\leq C,

which, together with (3.96), is enough to deduce

θ>0,det𝔽>0.\theta>0,\quad\det\mathbb{F}>0.

almost everywhere in (0,T)×Ω(0,T)\times\Omega.

To finish the proof, we multiply (3.21) by φ𝒞c1(,T)\varphi\in\mathcal{C}^{1}_{c}(-\infty,T) and integrate over (0,T)(0,T) to get

(3.97) 0TΩ(|𝒗n|22+en)tφdxdt=φ(0)Ω|𝒗0n|22+e0ndx.-\int_{0}^{T}\int_{\Omega}\left(\frac{|\boldsymbol{v}_{n}|^{2}}{2}+e_{n}\right)\partial_{t}\varphi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\varphi(0)\int_{\Omega}\frac{|\boldsymbol{v}_{0}^{n}|^{2}}{2}+e_{0}^{n}\mathop{}\!\mathrm{d}x.

Using the uniform bounds (3.53) and (3.44), the definition of ene_{n} in (3.16) and the strong convergence of θn\theta_{n}, 𝔽n\mathbb{F}_{n} and 𝒗n\boldsymbol{v}_{n}, we can let nn\to\infty in (3.97) and deduce (3.12), where we also use the assumptions on gg, see (2.5)–(2.6), and the strong convergence properties of the initial conditions in (3.1). The proof is complete.

4. Existence of the weak solutions for the case (P1)

We now focus on the problem (P1) and set cv=g(θ)1c_{v}=g(\theta)\equiv 1 for simplicity. For this setting, we consider the system (1.14)–(1.15), where we replace (1.14)3 by (1.13). Note that (1.14)3 and (1.13) are equivalent on the level of classical solutions. Our main result reads as follows.

Theorem 4.1.

Assume that ν\nu, κ\kappa and δ\delta are continuous functions satisfying (2.1)–(2.3). Let initial conditions {𝐯0,θ0,𝔽0}\{\boldsymbol{v}_{0},\theta_{0},\mathbb{F}_{0}\} fulfill

(4.1) 𝒗0L0,div2,θ0Lx1,lnθ0Lx1,𝔽0Lx2 and lndetF0Lx2\displaystyle\boldsymbol{v}_{0}\in L^{2}_{0,\operatorname{div}},\quad\theta_{0}\in L^{1}_{x},\quad\ln\theta_{0}\in L^{1}_{x},\quad\mathbb{F}_{0}\in L^{2}_{x}\quad\textrm{ and }\quad\ln\det F_{0}\in L^{2}_{x}

and det𝔽0>0\det\mathbb{F}_{0}>0 and θ0>0\theta_{0}>0 almost everywhere in Ω\Omega. Then there exists a triple {𝐯,𝔽,θ}\{\boldsymbol{v},\mathbb{F},\theta\} such that

(4.2) 𝒗\displaystyle\boldsymbol{v} 𝒞([0,T];L0,div2)Lt2W0,x1,2,\displaystyle\in\mathcal{C}([0,T];L^{2}_{0,\operatorname{div}})\cap L^{2}_{t}W^{1,2}_{0,x},
𝔽\displaystyle\mathbb{F} 𝒞([0,T];Lx2)Lt,x4,\displaystyle\in\mathcal{C}([0,T];L^{2}_{x})\cap L^{4}_{t,x},
θ\displaystyle\theta LtLx1Lt,xpLtqWx1,q\displaystyle\in L^{\infty}_{t}L^{1}_{x}\cap L^{p}_{t,x}\cap L^{q}_{t}W^{1,q}_{x} for any p[1,2) and q[1,43),\displaystyle\textrm{ for any }p\in[1,2)\textrm{ and }q\in\left[1,\frac{4}{3}\right),
lnθ\displaystyle\ln\theta LtLx1,lndet𝔽LtLx1,\displaystyle\in L^{\infty}_{t}L^{1}_{x},\qquad\ln\det\mathbb{F}\in L^{\infty}_{t}L^{1}_{x},

where det𝔽>0\det\mathbb{F}>0 and θ>0\theta>0 almost everywhere in (0,T)×Ω(0,T)\times\Omega. The functions (𝐯,𝔽,θ)(\boldsymbol{v},\mathbb{F},\theta) solves (1.14)–(1.15) in the following sense:

(4.3) 0TΩ𝒗t𝝋𝒗𝒗:x𝝋+2ν(θ)𝔻𝒗:x𝝋+2𝔽𝔽T:x𝝋dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\boldsymbol{v}\cdot\partial_{t}\boldsymbol{\varphi}-\boldsymbol{v}\otimes\boldsymbol{v}:\nabla_{x}\boldsymbol{\varphi}+2\nu(\theta)\mathbb{D}\boldsymbol{v}:\nabla_{x}\boldsymbol{\varphi}+2\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{\varphi}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝒗0(x)𝝋(0,x)dx\displaystyle\qquad=\int_{\Omega}\boldsymbol{v}_{0}(x)\cdot\boldsymbol{\varphi}(0,x)\mathop{}\!\mathrm{d}x

for any 𝛗𝒞c1([0,T)×Ω;2)\boldsymbol{\varphi}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2}) with divx𝛗=0\operatorname{div}_{x}\boldsymbol{\varphi}=0,

(4.4) 0TΩ𝔽:t𝔾𝔽𝒗x𝔾x𝒗𝔽:𝔾+12δ(θ)(𝔽𝔽T𝔽𝔽):𝔾dxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\mathbb{F}:\partial_{t}\mathbb{G}-\mathbb{F}\otimes\boldsymbol{v}\because\nabla_{x}\mathbb{G}-\nabla_{x}\boldsymbol{v}\,\mathbb{F}:\mathbb{G}+\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F}):\mathbb{G}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=Ω𝔽0(x):𝔾(0,x)dx\displaystyle\qquad=\int_{\Omega}\mathbb{F}_{0}(x):\mathbb{G}(0,x)\mathop{}\!\mathrm{d}x

for any 𝔾𝒞c1([0,T)×Ω;2×2)\mathbb{G}\in\mathcal{C}^{1}_{c}([0,T)\times\Omega;\mathbb{R}^{2\times 2}),

(4.5) 0TΩθtϕdxdt0TΩθ𝒗xϕdxdt+0TΩκ(θ)xθxϕdxdt\displaystyle-\int_{0}^{T}\int_{\Omega}\theta\,\partial_{t}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t-\int_{0}^{T}\int_{\Omega}\theta\,\boldsymbol{v}\cdot\nabla_{x}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{0}^{T}\int_{\Omega}\kappa(\theta)\nabla_{x}\theta\cdot\nabla_{x}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
0TΩ2ν(θ)|𝔻𝒗|2ϕdxdt0TΩδ(θ)|𝔽𝔽T𝕀|2ϕdxdt=Ωθ0(x)ϕ(0,x)dx\displaystyle\quad-\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t-\int_{0}^{T}\int_{\Omega}\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}-\mathbb{I}|^{2}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\int_{\Omega}\theta_{0}(x)\phi(0,x)\mathop{}\!\mathrm{d}x

for any ϕ𝒞c1([0,T)×Ω)\phi\in\mathcal{C}^{1}_{c}([0,T)\times\Omega).

Since the proof of Theorem 4.1 is very similar to the one conducted in [7, 2], and does not include many new techniques, compared to the proof of the Theorem 3.1 we will skip most of the technical detail and simply focus on the differences in the approach.

4.1. Approximating scheme

To prove our main theorem, we use a four-step approximation scheme. Our first goal will be to prove the existence of a solution to the system

(4.6) {t𝒗+divx(𝒗𝒗)divx𝕋=0,divx𝒗=0,t𝔽+divx(𝔽𝒗)x𝒗𝔽+12δ(θ)(𝔽𝔽T𝔽𝔽)=εΔ𝔽,p𝕀+2ν(θ)𝔻𝒗+2(𝔽𝔽T𝕀)=𝕋,tθ+divx(θ𝒗)divx(κ(θ)xθ)=2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T𝕀|2.\left\{\begin{aligned} &\partial_{t}\boldsymbol{v}+\operatorname{div}_{x}(\boldsymbol{v}\otimes\boldsymbol{v})-\operatorname{div}_{x}\mathbb{T}=0,\qquad\operatorname{div}_{x}\boldsymbol{v}=0,\\ &\partial_{t}\mathbb{F}+\operatorname{div}_{x}(\mathbb{F}\otimes\boldsymbol{v})-\nabla_{x}\boldsymbol{v}\,\mathbb{F}+\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F})=\varepsilon\Delta\mathbb{F},\\ &-p\mathbb{I}+2\nu(\theta)\mathbb{D}\boldsymbol{v}+2(\mathbb{F}\,\mathbb{F}^{T}-\mathbb{I})=\mathbb{T},\\ &\partial_{t}\theta+\operatorname{div}_{x}(\theta\boldsymbol{v})-\operatorname{div}_{x}(\kappa(\theta)\nabla_{x}\theta)=2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}-\mathbb{I}|^{2}.\end{aligned}\right.

for a fixed ε(0,1)\varepsilon\in(0,1). The initial and boundary conditions are set as

(4.7) 𝒗|Ω=0,𝒗(0,x)=𝒗0(x),𝔽(0,x)=𝔽0,x𝔽𝐧=0,xθ𝐧=0,θ(0,x)=θ0r,\displaystyle\boldsymbol{v}|_{\partial\Omega}=0,\quad\boldsymbol{v}(0,x)=\boldsymbol{v}_{0}(x),\quad\mathbb{F}(0,x)=\mathbb{F}_{0},\quad\nabla_{x}\mathbb{F}\cdot\mathbf{n}=0,\quad\nabla_{x}\theta\cdot\mathbf{n}=0,\quad\theta(0,x)=\theta^{r}_{0},

for r(0,1)r\in(0,1) and

(4.8) θ0r(x)={θ0(x) whenever rθ0(x)r1,1 otherwise.\displaystyle\theta^{r}_{0}(x)=\left\{\begin{aligned} &\theta_{0}(x)\quad&&\text{ whenever }r\leq\theta_{0}(x)\leq r^{-1},\\ &1\quad&&\text{ otherwise.}\end{aligned}\right.

To prove the existence of the ε\varepsilon-approximation, we employ the Galerkin approximation. Most importantly, we split the convergence results. The Galerkin approximative schema for θ\theta and 𝒗\boldsymbol{v} and 𝔽\mathbb{F} allows testing only by linear functions of the solutions. However, and as can be seen by the result of Lemma 3.2, the expected bounds on the temperature are obtained via testing by a nonlinear function. Therefore, the existence scheme is not completely trivial and here we follow the methods developed in [2, 10, 4]. This means that we first converge in the equation for the temperature but keep the equation for 𝒗\boldsymbol{v} and 𝔽\mathbb{F} in the Galerkin form. Then, we can deduce the optimal estimates for the temperature and consequently also for 𝒗\boldsymbol{v} and 𝔽\mathbb{F}. Next, we remove the finite-dimensional approximation for all quantities and finally, we let ε0+\varepsilon\to 0_{+} and r0+r\to 0_{+} in (4.6).

4.2. Galerkin approximation

We fix s>3s>3. Since Ω\Omega is two dimensional domain, we have

(4.9) Ws1,2(Ω)L(Ω).\displaystyle W^{s-1,2}(\Omega)\hookrightarrow L^{\infty}(\Omega).

Next, we consider {𝝎n}n\{\boldsymbol{\omega}_{n}\}_{n\in\mathbb{N}}, {𝔸n}n\{\mathbb{A}_{n}\}_{n\in\mathbb{N}}, {ϕm}m\{\phi_{m}\}_{m\in\mathbb{N}} to be the orthogonal bases of W0,divs,2(Ω;2)W^{s,2}_{0,\operatorname{div}}(\Omega;\mathbb{R}^{2}), Ws,2(Ω;2×2)W^{s,2}(\Omega;\mathbb{R}^{2\times 2}) and Ws,2(Ω;)W^{s,2}(\Omega;\mathbb{R}) respectively, that are also orthonormal with repsect to L2L^{2}, and whose projections (in L2L^{2}) are continuous. We also introduce the projection of the initial data as

(4.10) 𝒗0n\displaystyle\boldsymbol{v}_{0n} :=i=1n𝝎i(Ω𝒗0𝝎idx)\displaystyle:=\sum_{i=1}^{n}\boldsymbol{\omega}_{i}\left(\int_{\Omega}\boldsymbol{v}_{0}\boldsymbol{\omega}_{i}\mathop{}\!\mathrm{d}x\right)\qquad and we have 𝒗0nLx2𝒗0Lx2,\displaystyle\textrm{ and we have }\|\boldsymbol{v}_{0n}\|_{L^{2}_{x}}\leq\|\boldsymbol{v}_{0}\|_{L^{2}_{x}},
𝔽0n\displaystyle\mathbb{F}_{0n} :=i=1n𝔸i(Ω𝔽0𝔸idx)\displaystyle:=\sum_{i=1}^{n}\mathbb{A}_{i}\left(\int_{\Omega}\mathbb{F}_{0}\mathbb{A}_{i}\mathop{}\!\mathrm{d}x\right)\qquad and we have 𝔽0nLx2𝔽0Lx2,\displaystyle\textrm{ and we have }\|\mathbb{F}_{0n}\|_{L^{2}_{x}}\leq\|\mathbb{F}_{0}\|_{L^{2}_{x}},
θ0mr\displaystyle\theta_{0m}^{r} :=i=1mϕi(Ωθ0rϕidx)\displaystyle:=\sum_{i=1}^{m}\phi_{i}\left(\int_{\Omega}\theta_{0}^{r}\phi_{i}\mathop{}\!\mathrm{d}x\right)\qquad and we have θ0nrLx2θ0rLx2\displaystyle\textrm{ and we have }\|\theta_{0n}^{r}\|_{L^{2}_{x}}\leq\|\theta_{0}^{r}\|_{L^{2}_{x}}

and note that we have the following convergence results

(4.11) 𝒗0n\displaystyle\boldsymbol{v}_{0n} 𝒗0\displaystyle\to\boldsymbol{v}_{0} strongly in Lx2,\displaystyle\textrm{strongly in }L^{2}_{x},
𝔽0n\displaystyle\mathbb{F}_{0n} 𝔽0\displaystyle\to\mathbb{F}_{0} strongly in Lx2,\displaystyle\textrm{strongly in }L^{2}_{x},
θ0nr\displaystyle\theta_{0n}^{r} θ0r\displaystyle\to\theta_{0}^{r} strongly in Lx2,\displaystyle\textrm{strongly in }L^{2}_{x},

where θ0r\theta_{0}^{r} is defined in (4.8). We look for the Galerkin approximation of (4.6), which has the following form

(4.12) 𝒗nm(t,x)=i=1nαinm(t)𝝎i(x),𝔽nm(t,x)=i=1nβinm(t)𝔸i(x),θnm(t,x)=i=1mγinm(t)ϕi(x),\displaystyle\boldsymbol{v}_{nm}(t,x)=\sum_{i=1}^{n}\alpha^{nm}_{i}(t)\boldsymbol{\omega}_{i}(x),\quad\mathbb{F}_{nm}(t,x)=\sum_{i=1}^{n}\beta^{nm}_{i}(t)\mathbb{A}_{i}(x),\quad\theta_{nm}(t,x)=\sum_{i=1}^{m}\gamma^{nm}_{i}(t)\phi_{i}(x),

and we require the approximations to satisfy the following initial conditions

(4.13) 𝒗nm(0,x)=𝒗0n(x),𝔽nm(0,x)=𝔽0n(x),θnm(0,x)=θ0mr(x)\boldsymbol{v}_{nm}(0,x)=\boldsymbol{v}_{0n}(x),\qquad\mathbb{F}_{nm}(0,x)=\mathbb{F}_{0n}(x),\qquad\theta_{nm}(0,x)=\theta_{0m}^{r}(x)

for almost all xΩx\in\Omega, and we also require them to satisfy the following system of ordinary differential equations:

(4.14) ddtΩ𝒗nm𝝎jdxΩ𝒗nm𝒗nm:x𝝎jdx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\boldsymbol{v}_{nm}\cdot\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}\boldsymbol{v}_{nm}\otimes\boldsymbol{v}_{nm}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x
+Ω2ν(θnm)𝔻𝒗nm:x𝝎jdx+Ω2𝔽nm𝔽nmT:x𝝎jdx=0,\displaystyle\qquad+\int_{\Omega}2\nu(\theta_{nm})\mathbb{D}\boldsymbol{v}_{nm}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\mathbb{F}_{nm}\,\mathbb{F}_{nm}^{T}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x=0,

for any j=1,,nj=1,\ldots,n;

(4.15) ddtΩ𝔽nm:𝔸jdxΩ𝔽nm𝒗nmx𝔸jdxΩ(x𝒗nm𝔽nm):𝔸jdx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\mathbb{F}_{nm}:\mathbb{A}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}\mathbb{F}_{nm}\otimes\boldsymbol{v}_{nm}\because\nabla_{x}\mathbb{A}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}(\nabla_{x}\boldsymbol{v}_{nm}\,\mathbb{F}_{nm}):\mathbb{A}_{j}\mathop{}\!\mathrm{d}x
+12Ωδ(θnm)(𝔽nm𝔽nmT𝔽nm𝔽nm):𝔸jdx+εΩx𝔽nmx𝔸jdx=0,\displaystyle\quad+\frac{1}{2}\int_{\Omega}\delta(\theta_{nm})\left(\mathbb{F}_{nm}\,\mathbb{F}_{nm}^{T}\,\mathbb{F}_{nm}-\mathbb{F}_{nm}\right):\mathbb{A}_{j}\mathop{}\!\mathrm{d}x+\varepsilon\int_{\Omega}\nabla_{x}\mathbb{F}_{nm}\because\nabla_{x}\mathbb{A}_{j}\mathop{}\!\mathrm{d}x=0,

for any j=1,,nj=1,\ldots,n;

(4.16) ddtΩθnmϕidxΩ𝒗nmθnmxϕidx+Ωκ(θnm)xθnmxϕidx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\theta_{nm}\phi_{i}\mathop{}\!\mathrm{d}x-\int_{\Omega}\boldsymbol{v}_{nm}\theta_{nm}\cdot\nabla_{x}\phi_{i}\mathop{}\!\mathrm{d}x+\int_{\Omega}\kappa(\theta_{nm})\nabla_{x}\theta_{nm}\cdot\nabla_{x}\phi_{i}\mathop{}\!\mathrm{d}x
Ω2ν(θnm)|𝔻𝒗nm|2ϕidxΩδ(θnm)|𝔽nm𝔽nmT𝕀|2ϕidx=0\displaystyle\qquad-\int_{\Omega}2\nu(\theta_{nm})|\mathbb{D}\boldsymbol{v}_{nm}|^{2}\phi_{i}\mathop{}\!\mathrm{d}x-\int_{\Omega}\delta(\theta_{nm})|\mathbb{F}_{nm}\,\mathbb{F}_{nm}^{T}-\mathbb{I}|^{2}\phi_{i}\mathop{}\!\mathrm{d}x=0

for any i=1,,mi=1,\ldots,m.

The local-in-time existence of a solution to the above system follows from the Carathéodory theory and the global-in-time existence then follows from the estimates deduced in the next section.

4.3. Convergence with m+m\to+\infty

We follow very closely the argumentation in [7, Appendix B - estimate (B.15)] and therefore the proof here will be sketchy and we focus mainly onl the parts that are different. Hence, we multiply (LABEL:eq:galerkin_velocity) and (4.15) by αj\alpha_{j} and 2βj2\beta_{j} respectively, after summing over j=1,,nj=1,\ldots,n and adding both equations together, we deduce also with the help of the Young inequality and the assumptions on material parameters (2.1)–(2.3)

(4.17) 𝒗nmLtLx2+𝔽nmLtLx2+𝔻𝒗nmLt,x2+𝔽nmLt,x4+εx𝔽nmLt,x2\displaystyle\|\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\mathbb{F}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\mathbb{D}\boldsymbol{v}_{nm}\|_{L^{2}_{t,x}}+\|\mathbb{F}_{nm}\|_{L^{4}_{t,x}}+\sqrt{\varepsilon}\|\nabla_{x}\mathbb{F}_{nm}\|_{L^{2}_{t,x}}
C(𝒗0nLx2,𝔽0nLx2)C,\displaystyle\qquad\leq C(\|\boldsymbol{v}_{0n}\|_{L^{2}_{x}},\|\mathbb{F}_{0n}\|_{L^{2}_{x}})\leq C,

where for the second inequality we used the properties of initial conditions in (4.10). Using also the Korn inequality, we have

(4.18) 𝒗nmLtLx2+𝔽nmLtLx2+x𝒗nmLt,x2+𝔽nmLt,x4+εx𝔽nmLt,x2C.\displaystyle\|\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\mathbb{F}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\nabla_{x}\boldsymbol{v}_{nm}\|_{L^{2}_{t,x}}+\|\mathbb{F}_{nm}\|_{L^{4}_{t,x}}+\sqrt{\varepsilon}\|\nabla_{x}\mathbb{F}_{nm}\|_{L^{2}_{t,x}}\leq C.

It is worth noting here, that the above estimates are independent of nn, mm and rr and due to weak-lower semicontinuity are kept till the end of the proof. We follow by estimates that are nn- or rr-dependent. Due to the orthogonality of the bases {𝝎n}n\{\boldsymbol{\omega}_{n}\}_{n\in\mathbb{N}} and {𝔸n}n\{\mathbb{A}_{n}\}_{n\in\mathbb{N}}, the embedding (4.9), and (4.18) we get

(4.19) 𝒗nmLtWx1,𝒗nmLtWxs,2C(n)𝒗nmLtLx2C(n),\displaystyle\|\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}W^{1,\infty}_{x}}\leq\|\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}W^{s,2}_{x}}\leq C(n)\|\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}\leq C(n),
𝔽nmLtWx1,𝔽nmLtWxs,2C(n)𝔽nmLtLx2C(n).\displaystyle\|\mathbb{F}_{nm}\|_{L^{\infty}_{t}W^{1,\infty}_{x}}\leq\|\mathbb{F}_{nm}\|_{L^{\infty}_{t}W^{s,2}_{x}}\leq C(n)\|\mathbb{F}_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}\leq C(n).

Next, we deduce the estimates for the temperature. We multiply (4.16) by γj(t)\gamma_{j}(t) and sum the results over j=1,,mj=1,\ldots,m. Using the fact that divx𝒗nm=0\operatorname{div}_{x}\boldsymbol{v}_{nm}=0 we obtain

12ddtθnmLx22+κ(θnm)xθnmLx22=Ω2ν(θnm)|𝔻𝒗nm|2θnm+Ωδ(θnm)|𝔽nm𝔽nmT𝕀|2θnm.\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|\theta_{nm}\|^{2}_{L^{2}_{x}}+\|\sqrt{\kappa(\theta_{nm})}\nabla_{x}\theta_{nm}\|^{2}_{L^{2}_{x}}=\int_{\Omega}2\nu(\theta_{nm})|\mathbb{D}\boldsymbol{v}_{nm}|^{2}\theta_{nm}+\int_{\Omega}\delta(\theta_{nm})|\mathbb{F}_{nm}\,\mathbb{F}^{T}_{nm}-\mathbb{I}|^{2}\theta_{nm}.

Thus, with the use of Hölder’s inequality, (LABEL:infty_bounds_for_velocit_and_elastic) as well as the bounds (2.1)–(2.3), we arrive at

(4.20) 12ddtθnmLx22+C1xθnmLx22C(1+𝒗nmLtWx1,2+𝔽nmLtLx4)θnmLx2.\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|\theta_{nm}\|^{2}_{L^{2}_{x}}+C_{1}\|\nabla_{x}\theta_{nm}\|^{2}_{L^{2}_{x}}\leq C\left(1+\|\boldsymbol{v}_{nm}\|^{2}_{L^{\infty}_{t}W^{1,\infty}_{x}}+\|\mathbb{F}_{nm}\|^{4}_{L^{\infty}_{t}L^{\infty}_{x}}\right)\|\theta_{nm}\|_{L^{2}_{x}}.

The Grönwall inequality and the estimates (LABEL:infty_bounds_for_velocit_and_elastic) then directly leads to

(4.21) θnmLtLx2+xθnmLt,x2C(n,T)θ0mrLx2C(n,r).\displaystyle\|\theta_{nm}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\nabla_{x}\theta_{nm}\|_{L^{2}_{t,x}}\leq C(n,T)\|\theta_{0mr}\|_{L^{2}_{x}}\leq C(n,r).

Moving forward, we focus on estimates for time derivatives. First, using the equation (LABEL:eq:galerkin_velocity), the orthogonality of the basis {𝝎i}i\{\boldsymbol{\omega}_{i}\}_{i\in\mathbb{N}} and applying the bounds (LABEL:infty_bounds_for_velocit_and_elastic) together with the Hölder inequality, we get

|(αjnm)(t)|\displaystyle|(\alpha^{nm}_{j})^{\prime}(t)| =|Ωt𝒗nm𝝎jdx|\displaystyle=\left|\int_{\Omega}\partial_{t}\boldsymbol{v}_{nm}\cdot\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x\right|
=|Ω𝒗nm𝒗nm:x𝝎j(ν(θnm)𝔻𝒗nm:x𝝎j)(𝔽nm𝔽nmT):x𝝎jdx|C(n).\displaystyle=\left|\int_{\Omega}\boldsymbol{v}_{nm}\otimes\boldsymbol{v}_{nm}:\nabla_{x}\boldsymbol{\omega}_{j}-(\nu(\theta_{nm})\mathbb{D}\boldsymbol{v}_{nm}:\nabla_{x}\boldsymbol{\omega}_{j})\right.-(\mathbb{F}_{nm}\,\mathbb{F}^{T}_{nm}):\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x\Big{|}\leq C(n).

Thus, using the definition of 𝒗nm\boldsymbol{v}_{nm} in (4.12), we get

(4.22) t𝒗nmLtWx1,C(n).\displaystyle\|\partial_{t}\boldsymbol{v}_{nm}\|_{L^{\infty}_{t}W^{1,\infty}_{x}}\leq C(n).

By a very similar arguments, we can deduce from (4.15) and (LABEL:infty_bounds_for_velocit_and_elastic) that

(4.23) t𝔽nmLtWx1,C(n).\displaystyle\|\partial_{t}\mathbb{F}_{nm}\|_{L^{\infty}_{t}W^{1,\infty}_{x}}\leq C(n).

For the temperature, we proceed slightly differently. Due to the continuity (independent of mm) of the projection of the basis {ϕj}j\{\phi_{j}\}_{j\in\mathbb{N}} and from the identity (4.16), we have for all times t(0,T)t\in(0,T)

tθnmW1,2(Ω)2C(Ω|𝒗nmθnm|2+|xθnm|2dx)+C(Ω|𝔻𝒗nm|4+|𝔽nm𝔽nmT𝕀|4dx)C(n)(1+θnmWx1,22),\begin{split}\|\partial_{t}\theta_{nm}\|^{2}_{W^{-1,2}(\Omega)}&\leq C\left(\int_{\Omega}|\boldsymbol{v}_{nm}\theta_{nm}|^{2}+|\nabla_{x}\theta_{nm}|^{2}\mathop{}\!\mathrm{d}x\right)\\ &\quad+C\left(\int_{\Omega}|\mathbb{D}\boldsymbol{v}_{nm}|^{4}+|\mathbb{F}_{nm}\,\mathbb{F}_{nm}^{T}-\mathbb{I}|^{4}\mathop{}\!\mathrm{d}x\right)\\ &\leq C(n)\left(1+\|\theta_{nm}\|_{W^{1,2}_{x}}^{2}\right),\end{split}

where we also used the assumptions (2.1)–(2.2). Integration over (0,T)(0,T) and the use of (4.21) gives

(4.24) tθnmL2((0,T);(W1,2(Ω)))C(n,r).\|\partial_{t}\theta_{nm}\|_{L^{2}((0,T);(W^{1,2}(\Omega))^{*})}\leq C(n,r).

Having the estimates (LABEL:infty_bounds_for_velocit_and_elastic), (4.22), (4.23), (4.21) and (4.24), we can use the Banach–Alaoglu theorem, the classical Sobolev–Morrey embedding and the Aubin–Lions compactness lemma, see Lemma A.1, and we can find a triple {𝒗n,𝔽n,θn}\{\boldsymbol{v}_{n},\mathbb{F}_{n},\theta_{n}\} such that for a subsequence that we do not relabel we have the following convergence results as mm\to\infty: for the velocity field

(4.25) t𝒗nm\displaystyle\partial_{t}\boldsymbol{v}_{nm} t𝒗n\displaystyle\overset{*}{\rightharpoonup}\partial_{t}\boldsymbol{v}_{n}\qquad weakly* in LtW0,div1,,\displaystyle\text{ weakly* in }L^{\infty}_{t}W^{1,\infty}_{0,\operatorname{div}},
𝒗nm\displaystyle\boldsymbol{v}_{nm} 𝒗n\displaystyle\rightarrow\boldsymbol{v}_{n}\qquad strongly in 𝒞([0,T];W1,(Ω));\displaystyle\text{ strongly in }\mathcal{C}([0,T];W^{1,\infty}(\Omega));

for the elastic stress tensor

(4.26) t𝔽nm\displaystyle\partial_{t}\mathbb{F}_{nm} t𝔽n\displaystyle\overset{*}{\rightharpoonup}\partial_{t}\mathbb{F}_{n}\qquad weakly* in Lt2Wx1,,\displaystyle\text{ weakly* in }L^{2}_{t}W^{1,\infty}_{x},
𝔽nm\displaystyle\mathbb{F}_{nm} 𝔽n\displaystyle\rightarrow\mathbb{F}_{n}\qquad strongly in 𝒞([0,T];W1,(Ω));\displaystyle\text{ strongly in }\mathcal{C}([0,T];W^{1,\infty}(\Omega));

and for the temperature

(4.27) tθnm\displaystyle\partial_{t}\theta_{nm} tθn\displaystyle\rightharpoonup\partial_{t}\theta_{n}\qquad weakly in L2((0,T);W1,2(Ω)),\displaystyle\text{ weakly in }L^{2}((0,T);W^{-1,2}(\Omega)),
θnm\displaystyle\theta_{nm} θn\displaystyle\overset{*}{\rightharpoonup}\theta_{n}\qquad weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
θnm\displaystyle\theta_{nm} θn\displaystyle\rightarrow\theta_{n}\qquad strongly in Lt2Lx2.\displaystyle\text{ strongly in }L^{2}_{t}L^{2}_{x}.

In addition, we have the following form for 𝒗n\boldsymbol{v}_{n} and 𝔽n\mathbb{F}_{n}

(4.28) 𝒗n(t,x)=i=1nαin(t)𝝎i(x),𝔽n(t,x)=i=1nβin(t)𝔸i(x).\displaystyle\boldsymbol{v}_{n}(t,x)=\sum_{i=1}^{n}\alpha^{n}_{i}(t)\boldsymbol{\omega}_{i}(x),\quad\mathbb{F}_{n}(t,x)=\sum_{i=1}^{n}\beta^{n}_{i}(t)\mathbb{A}_{i}(x).

The convergence results (4.25)–(4.27) allows us to let mm\to\infty in (LABEL:eq:galerkin_velocity)–(4.16) and deduce

(4.29) ddtΩ𝒗n𝝎jdxΩ𝒗n𝒗n:x𝝎jdx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\boldsymbol{v}_{n}\cdot\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}\boldsymbol{v}_{n}\otimes\boldsymbol{v}_{n}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x
+Ω2ν(θn)𝔻𝒗n:x𝝎jdx+Ω2𝔽n𝔽nT:x𝝎jdx=0,\displaystyle\qquad+\int_{\Omega}2\nu(\theta_{n})\mathbb{D}\boldsymbol{v}_{n}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}:\nabla_{x}\boldsymbol{\omega}_{j}\mathop{}\!\mathrm{d}x=0,

for any j=1,,nj=1,\ldots,n;

(4.30) ddtΩ𝔽n:𝔸jdxΩ𝔽n𝒗nx𝔸jdxΩ(x𝒗n𝔽n):𝔸jdx\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\int_{\Omega}\mathbb{F}_{n}:\mathbb{A}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}\mathbb{F}_{n}\otimes\boldsymbol{v}_{n}\because\nabla_{x}\mathbb{A}_{j}\mathop{}\!\mathrm{d}x-\int_{\Omega}(\nabla_{x}\boldsymbol{v}_{n}\,\mathbb{F}_{n}):\mathbb{A}_{j}\mathop{}\!\mathrm{d}x
+12Ωδ(θn)(𝔽n𝔽nT𝔽n𝔽n):𝔸jdx+εΩx𝔽nx𝔸jdx=0,\displaystyle\quad+\frac{1}{2}\int_{\Omega}\delta(\theta_{n})\left(\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}\,\mathbb{F}_{n}-\mathbb{F}_{n}\right):\mathbb{A}_{j}\mathop{}\!\mathrm{d}x+\varepsilon\int_{\Omega}\nabla_{x}\mathbb{F}_{n}\because\nabla_{x}\mathbb{A}_{j}\mathop{}\!\mathrm{d}x=0,

for any j=1,,nj=1,\ldots,n;

(4.31) tθnϕΩ𝒗nθnxϕdx+Ωκ(θn)xθnxϕdx\displaystyle\langle\partial_{t}\theta_{n}\phi\rangle-\int_{\Omega}\boldsymbol{v}_{n}\theta_{n}\cdot\nabla_{x}\phi\mathop{}\!\mathrm{d}x+\int_{\Omega}\kappa(\theta_{n})\nabla_{x}\theta_{n}\cdot\nabla_{x}\phi\mathop{}\!\mathrm{d}x
Ω2ν(θn)|𝔻𝒗n|2ϕdxΩδ(θn)|𝔽n𝔽nT𝕀|2ϕdx=0\displaystyle\qquad-\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}\phi\mathop{}\!\mathrm{d}x-\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}-\mathbb{I}|^{2}\phi\mathop{}\!\mathrm{d}x=0

for any ϕW1,2(Ω)\phi\in W^{1,2}(\Omega) and almost all t(0,T)t\in(0,T). Furthermore, due to the standard parabolic embedding we also have that θn𝒞([0,T];L2(Ω))\theta_{n}\in\mathcal{C}([0,T];L^{2}(\Omega)) and that the unknowns attain the following initial conditions

𝒗n(0)=𝒗0n,𝔽n(0)=𝔽0n,θn(0)=θ0r.\boldsymbol{v}_{n}(0)=\boldsymbol{v}_{0n},\qquad\mathbb{F}_{n}(0)=\mathbb{F}_{0n},\qquad\theta_{n}(0)=\theta_{0}^{r}.

4.4. Convergence with n+n\to+\infty

First, we recall the uniform bounds for 𝒗n\boldsymbol{v}_{n} and 𝔽n\mathbb{F}_{n}. Due to the weak-lower semicontinuity, it follows from (4.18) that

(4.32) 𝒗nLtLx2+𝔽nLtLx2+x𝒗nLt,x2+𝔽nLt,x4+εx𝔽nLt,x2C.\displaystyle\|\boldsymbol{v}_{n}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\mathbb{F}_{n}\|_{L^{\infty}_{t}L^{2}_{x}}+\|\nabla_{x}\boldsymbol{v}_{n}\|_{L^{2}_{t,x}}+\|\mathbb{F}_{n}\|_{L^{4}_{t,x}}+\sqrt{\varepsilon}\|\nabla_{x}\mathbb{F}_{n}\|_{L^{2}_{t,x}}\leq C.

For the time derivatives t𝒗n\partial_{t}\boldsymbol{v}_{n} and t𝔽n\partial_{t}\mathbb{F}_{n}, the bounds (4.23), (4.22) are not uniform with respect to nn. Therefore, we must proceed differently. Based on the estimate (4.32) and on the identities (LABEL:eq:galerkin_1stconv_velocity)–(4.30), following [7, Appendix B - estimates (B.19) and (B.21)] we obtain

(4.33) t𝒗nL2((0,T);W0,div1,2(Ω))+t𝔽nL43((0,T);W1,2(Ω))C.\displaystyle\|\partial_{t}\boldsymbol{v}_{n}\|_{L^{2}((0,T);W^{-1,2}_{0,\operatorname{div}}(\Omega))}+\|\partial_{t}\mathbb{F}_{n}\|_{L^{\frac{4}{3}}((0,T);W^{-1,2}(\Omega))}\leq C.

Next, we focus on estimates for θn\theta_{n}. In this case, we want to test (4.31) by θnλ1\theta_{n}^{\lambda-1} for arbitrary 0<λ<10<\lambda<1 as was already explained in Lemma 3.2. To do so properly, we first provide the minimum principle for θn\theta_{n}. We set ϕ:=min{0,θnr}\phi:=\min\{0,\theta_{n}-r\} in (4.31) and in a very similar manner as in [3], we observe that

(4.34) θn(x,t)r>0, for a.e. (t,x)(0,T)×Ω.\displaystyle\theta_{n}(x,t)\geq r>0,\quad\text{ for a.e. }(t,x)\in(0,T)\times\Omega.

We proceed by the uniform estimate for θn\theta_{n}. We set ϕ:=1\phi:=1 in (4.31). Then, after integrating the result over (0,t)(0,t), using Lemma A.3, the condition divx𝒗n=0\operatorname{div}_{x}\boldsymbol{v}_{n}=0, and the positivity of θn\theta_{n} (4.34), we get for almost all t(0,T)t\in(0,T)

θn(t)Lx1=0tΩν(θn)|𝔻𝒗n|2+δ(θn)|𝔽n𝔽nT𝕀|2dτ+θn(0)Lx1.\|\theta_{n}(t)\|_{L^{1}_{x}}=\int_{0}^{t}\int_{\Omega}\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}+\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}-\mathbb{I}|^{2}\mathop{}\!\mathrm{d}\tau+\|\theta_{n}(0)\|_{L^{1}_{x}}.

Thus, after applying the estimate (4.32), the definition of θ0r\theta_{0}^{r} and the assumption (4.1), we get

(4.35) θnLtLx1C(1+θ0rLx1)C(1+θ0Lx1)C.\displaystyle\|\theta_{n}\|_{L^{\infty}_{t}L^{1}_{x}}\leq C(1+\|\theta_{0}^{r}\|_{L^{1}_{x}})\leq C(1+\|\theta_{0}\|_{L^{1}_{x}})\leq C.

Moving forward, we can set ϕ:=θnλ1\phi:=\theta_{n}^{\lambda-1} in (4.31). Performing similar operations as in Lemma 3.2, we obtain (using also the uniform bounds (4.32))

(4.36) 0TΩ|xθn|2θnλdxdtC(λ) for any λ(1,2).\displaystyle\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{\lambda}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\leq C(\lambda)\quad\text{ for any }\lambda\in(1,2).

In particular, setting ϕ:=θn1\phi:=\theta_{n}^{-1} in (4.31), we deduce that for all t(0,T)t\in(0,T)

(4.37) Ω|lnθn(t)|dx+0TΩ|xθn|2θn2dxdt\displaystyle\int_{\Omega}|\ln\theta_{n}(t)|\mathop{}\!\mathrm{d}x+\int_{0}^{T}\int_{\Omega}\frac{|\nabla_{x}\theta_{n}|^{2}}{\theta_{n}^{2}}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t C(1+Ωθ(t)+θ0r+|lnθ0r|dx)\displaystyle\leq C\left(1+\int_{\Omega}\theta(t)+\theta_{0}^{r}+|\ln\theta_{0}^{r}|\mathop{}\!\mathrm{d}x\right)
C(1+Ωθ0+|lnθ0|dx)C,\displaystyle\leq C\left(1+\int_{\Omega}\theta_{0}+|\ln\theta_{0}|\mathop{}\!\mathrm{d}x\right)\leq C,

where we used (4.35). Note here, we got the rr-independent estimate. With this (similarly as in the estimates below (3.37)), we may infer uniform bounds on θn\theta_{n} and xθn\nabla_{x}\theta_{n}

(4.38) θnLtpLxp\displaystyle\|\theta_{n}\|_{L^{p^{\prime}}_{t}L^{p}_{x}} C(p)\displaystyle\leq C(p) for all p[1,),\displaystyle\textrm{for all }p\in[1,\infty),
(4.39) xθnLt,xq\displaystyle\|\nabla_{x}\theta_{n}\|_{L^{q}_{t,x}} C(q)\displaystyle\leq C(q) for all q[1,43),\displaystyle\textrm{for all }q\in\left[1,\frac{4}{3}\right),
(4.40) lnθnLtLx1\displaystyle\|\ln\theta_{n}\|_{L^{\infty}_{t}L^{1}_{x}} C,\displaystyle\leq C,
(4.41) xlnθnLt,x2\displaystyle\|\nabla_{x}\ln\theta_{n}\|_{L^{2}_{t,x}} C.\displaystyle\leq C.

In order to get the compactness of the temperature, we also deduce the estimate for the time derivative of θn\theta_{n}. This is done similarly as in (4.24), but because we have much weaker estimates on 𝔻𝒗n\mathbb{D}\boldsymbol{v}_{n} and 𝔽n\mathbb{F}_{n} than in previous section, compare (LABEL:infty_bounds_for_velocit_and_elastic) and (4.32), we can deduce from (4.31) the following bound (note that W2,2𝒞(Ω¯)W^{2,2}\hookrightarrow\mathcal{C}(\overline{\Omega}))

(4.42) tθnL1((0,T);W2,2(Ω))C.\displaystyle\|\partial_{t}\theta_{n}\|_{L^{1}((0,T);W^{-2,2}(\Omega))}\leq C.

Consequently, we may apply the Banach–Alaoglu theorem and the generalized version of the Aubin–Lions lemma, see Lemma A.1, and it follows from (4.38)–(4.42) that there is a subsequence that we do not relabel such that

(4.43) θn\displaystyle\theta_{n} θ\displaystyle\rightharpoonup\theta\quad weakly in Lt,xq for all q[1,2),\displaystyle\text{ weakly in }L^{q}_{t,x}\textrm{ for all }q\in[1,2),
(4.44) θn\displaystyle\theta_{n} θ\displaystyle\rightharpoonup\theta\quad weakly in LtpWx1,p for all p[1,43),\displaystyle\text{ weakly in }L^{p}_{t}W^{1,p}_{x}\textrm{ for all }p\in\left[1,\frac{4}{3}\right),
(4.45) lnθn\displaystyle\ln\theta_{n} lnθ\displaystyle\rightharpoonup\ln\theta\quad weakly in Lt2Wx1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{x},
(4.46) θn\displaystyle\theta_{n} θ\displaystyle\rightarrow\theta\quad strongly in LtqLxs for all s[1,) and all q[1,s).\displaystyle\text{ strongly in }L^{q}_{t}L^{s}_{x}\textrm{ for all }s\in[1,\infty)\textrm{ and all }q\in[1,s^{\prime}).

Similarly, due to (4.32)–(4.33), we have

(4.47) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v}\quad weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
(4.48) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightharpoonup\boldsymbol{v}\quad weakly in Lt2W0,x1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{0,x},
(4.49) t𝒗n\displaystyle\partial_{t}\boldsymbol{v}_{n} t𝒗\displaystyle\rightharpoonup\partial_{t}\boldsymbol{v}\quad weakly in L2((0,T);W0,div1,2(Ω))\displaystyle\text{ weakly in }L^{2}((0,T);W^{-1,2}_{0,\operatorname{div}}(\Omega))
(4.50) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\rightarrow\boldsymbol{v}\quad strongly in Lt,xq, for q[1,4),\displaystyle\text{ strongly in }L^{q}_{t,x},\text{ for }q\in[1,4),
(4.51) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightharpoonup\mathbb{F}\quad weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
(4.52) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightharpoonup\mathbb{F}\quad weakly in Lt2Wx1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{x},
(4.53) t𝔽n\displaystyle\partial_{t}\mathbb{F}_{n} t𝔽\displaystyle\rightharpoonup\partial_{t}\mathbb{F}\quad weakly in L43((0,T);W1,2(Ω;2×2))\displaystyle\text{ weakly in }L^{\frac{4}{3}}((0,T);W^{-1,2}(\Omega;\mathbb{R}^{2\times 2}))
(4.54) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\rightarrow\mathbb{F}\quad strongly in Lt,xq, for q[1,4).\displaystyle\text{ strongly in }L^{q}_{t,x},\text{ for }q\in[1,4).

The convergence results (4.43)–(4.54) allow us to let nn\to\infty in all terms appearing in (LABEL:eq:galerkin_1stconv_velocity)–(4.31) except the terms involving

ν(θn)|𝔻𝒗n|2,|𝔽n𝔽nT𝕀|2 and tθn.\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2},\qquad|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}-\mathbb{I}|^{2}\qquad\textrm{ and }\qquad\partial_{t}\theta_{n}.

In particular, for the 𝒗\boldsymbol{v} and 𝔽\mathbb{F} we obtain the following identities:

(4.55) t𝒗,𝝎Ω𝒗𝒗:x𝝎dx+Ω2ν(θ)𝔻𝒗:x𝝎dx+Ω2𝔽𝔽T:x𝝎dx=0\displaystyle\langle\partial_{t}\boldsymbol{v},\boldsymbol{\omega}\rangle-\int_{\Omega}\boldsymbol{v}\otimes\boldsymbol{v}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\nu(\theta)\mathbb{D}\boldsymbol{v}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x=0

for almost all time t(0,T)t\in(0,T) and for all 𝝎W0,div1,2(Ω)\boldsymbol{\omega}\in W^{1,2}_{0,\operatorname{div}}(\Omega);

(4.56) t𝔽,𝔸Ω𝔽𝒗x𝔸dxΩ(x𝒗𝔽):𝔸dx\displaystyle\langle\partial_{t}\mathbb{F},\mathbb{A}\rangle-\int_{\Omega}\mathbb{F}\otimes\boldsymbol{v}\because\nabla_{x}\mathbb{A}\mathop{}\!\mathrm{d}x-\int_{\Omega}(\nabla_{x}\boldsymbol{v}\,\mathbb{F}):\mathbb{A}\mathop{}\!\mathrm{d}x
+12Ωδ(θ)(𝔽𝔽T𝔽𝔽):𝔸dx+εΩx𝔽x𝔸dx=0\displaystyle\qquad+\frac{1}{2}\int_{\Omega}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F}):\mathbb{A}\mathop{}\!\mathrm{d}x+\varepsilon\int_{\Omega}\nabla_{x}\mathbb{F}\because\nabla_{x}\mathbb{A}\mathop{}\!\mathrm{d}x=0

for almost all time t(0,T)t\in(0,T) and for all 𝔸W1,2(Ω;2×2)\mathbb{A}\in W^{1,2}(\Omega;\mathbb{R}^{2\times 2}). Moreover, by classical arguments, we have that 𝒗𝒞([0,T];L0,div2)\boldsymbol{v}\in\mathcal{C}([0,T];L^{2}_{0,\operatorname{div}}) and 𝔽𝒞([0,T];L2(Ω;2×2))\mathbb{F}\in\mathcal{C}([0,T];L^{2}(\Omega;\mathbb{R}^{2\times 2})) and also that 𝒗(0)=𝒗0\boldsymbol{v}(0)=\boldsymbol{v}_{0} and 𝔽(0)=𝔽0\mathbb{F}(0)=\mathbb{F}_{0}. Furthermore, thanks to the Fatou lemma and the uniform bound (4.40), we also have θLtLx1\theta\in L^{\infty}_{t}L^{1}_{x}.

To deal with the problematic terms, we employ the energy methods. We multiply the jj-th equation in (LABEL:eq:galerkin_1stconv_velocity) by αjn\alpha_{j}^{n} and sum the result over j=1,,nj=1,\ldots,n and integrate over t(0,T)t\in(0,T) to conclude (using the fact that divx𝒗n=0\operatorname{div}_{x}\boldsymbol{v}_{n}=0)

(4.57) Ω|𝒗n(T)|22dx+0TΩ2ν(θn)|𝔻𝒗n|2+2𝔽n𝔽nT:x𝒗ndxdt=Ω|𝒗0n|2dx.\displaystyle\int_{\Omega}\frac{|\boldsymbol{v}_{n}(T)|^{2}}{2}\mathop{}\!\mathrm{d}x+\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}+2\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}:\nabla_{x}\boldsymbol{v}_{n}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\int_{\Omega}\frac{|\boldsymbol{v}_{0n}|}{2}\mathop{}\!\mathrm{d}x.

Similarly, we set 𝝎:=𝒗\boldsymbol{\omega}:=\boldsymbol{v} in (4.55) and integrate the equation over (0,T)(0,T) and obtain

(4.58) Ω|𝒗(T)|22dx+0TΩ2ν(θ)|𝔻𝒗|2+2𝔽𝔽T:x𝒗dxdt=Ω|𝒗0|2dx.\displaystyle\int_{\Omega}\frac{|\boldsymbol{v}(T)|^{2}}{2}\mathop{}\!\mathrm{d}x+\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+2\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{v}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=\int_{\Omega}\frac{|\boldsymbol{v}_{0}|}{2}\mathop{}\!\mathrm{d}x.

Similarly, we multiply the jj-th equation in (4.15) by 2βjn2\beta_{j}^{n} and sum with respect to j=1,,nj=1,\ldots,n to get after integration over t(0,T)t\in(0,T)

(4.59) Ω|𝔽n(T)|2dx+0TΩδ(θn)|𝔽n𝔽nT|22x𝒗n:𝔽n𝔽nT+2ε|x𝔽n|2dxdt\displaystyle\int_{\Omega}|\mathbb{F}_{n}(T)|^{2}\mathop{}\!\mathrm{d}x+\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}-2\nabla_{x}\boldsymbol{v}_{n}:\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=0TΩδ(θn)|𝔽n|2dxdt+Ω|𝔽0n(T)|2dx\displaystyle\qquad=\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{\Omega}|\mathbb{F}_{0n}(T)|^{2}\mathop{}\!\mathrm{d}x

and by setting 𝔸:=𝔽\mathbb{A}:=\mathbb{F} in (LABEL:eq:galerkin_2ndconv_elastic), we have

(4.60) Ω|𝔽(T)|2dx+0TΩδ(θ)|𝔽𝔽T|22x𝒗:𝔽𝔽T+2ε|x𝔽n|2dxdt\displaystyle\int_{\Omega}|\mathbb{F}(T)|^{2}\mathop{}\!\mathrm{d}x+\int_{0}^{T}\int_{\Omega}\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}-2\nabla_{x}\boldsymbol{v}:\mathbb{F}\,\mathbb{F}^{T}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=0TΩδ(θ)|𝔽|2dxdt+Ω|𝔽0(T)|2dx.\displaystyle\qquad=\int_{0}^{T}\int_{\Omega}\delta(\theta)|\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{\Omega}|\mathbb{F}_{0}(T)|^{2}\mathop{}\!\mathrm{d}x.

Summing (4.57) and (LABEL:eq:galerkin_ener1), we have

(4.61) 0TΩ2ν(θn)|𝔻𝒗n|2+δ(θn)|𝔽n𝔽nT|2+2ε|x𝔽n|2dxdt\displaystyle\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}+\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=0TΩδ(θn)|𝔽n|2dxdt+Ω|𝒗0n|2+|𝔽0n|2|𝒗n(T)|22|𝔽n(T)|2dx.\displaystyle\qquad=\int_{0}^{T}\int_{\Omega}\delta(\theta_{n})|\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{\Omega}\frac{|\boldsymbol{v}_{0n}|}{2}+|\mathbb{F}_{0n}|^{2}-\frac{|\boldsymbol{v}_{n}(T)|^{2}}{2}-|\mathbb{F}_{n}(T)|^{2}\mathop{}\!\mathrm{d}x.

Similarly, summing (4.58) and (LABEL:eq:ener2) leads to

(4.62) 0TΩ2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T|2+2ε|x𝔽|2dxdt\displaystyle\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=0TΩδ(θ)|𝔽|2dxdt+Ω|𝒗0|2+|𝔽0|2|𝒗(T)|22|𝔽(T)|2dx.\displaystyle\qquad=\int_{0}^{T}\int_{\Omega}\delta(\theta)|\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{\Omega}\frac{|\boldsymbol{v}_{0}|}{2}+|\mathbb{F}_{0}|^{2}-\frac{|\boldsymbol{v}(T)|^{2}}{2}-|\mathbb{F}(T)|^{2}\mathop{}\!\mathrm{d}x.

Using (4.54), (4.46), the assumption (2.3), the convergence of initial conditions in (4.11) and also the weak lower semicontinuity, we get

(4.63) lim supn+0TΩ2ν(θn)|𝔻𝒗n|2dx+δ(θn)|𝔽n𝔽nT|2+2ε|x𝔽n|2dxdt0TΩ2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T|2+2ε|x𝔽|2dxdt.\begin{split}\limsup_{n\to+\infty}&\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}\mathop{}\!\mathrm{d}x+\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &\quad\leq\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.\end{split}

Finally, using the assumptions (2.2)–(2.3), the convergence results (4.46)–(4.54) and the inequality (4.63), we get

(4.64) C1\displaystyle C_{1} lim supn0TΩ|𝔻𝒗n𝔻𝒗|2+|𝔽n𝔽nT𝔽𝔽T|2dxdt\displaystyle\limsup_{n\to\infty}\int_{0}^{T}\int_{\Omega}|\mathbb{D}\boldsymbol{v}_{n}-\mathbb{D}\boldsymbol{v}|^{2}+|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}-\mathbb{F}\,\mathbb{F}^{T}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
lim supn0TΩ2ν(θn)|𝔻𝒗n𝔻𝒗|2+δ(θn)|𝔽n𝔽nT𝔽𝔽T|2+2ε|x𝔽nx𝔽|2dxdt\displaystyle\leq\limsup_{n\to\infty}\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}-\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}-\mathbb{F}\,\mathbb{F}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}-\nabla_{x}\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
lim supn0TΩ2ν(θn)|𝔻𝒗n|2+δ(θn)|𝔽n𝔽nT|2+2ε|x𝔽n|2dxdt\displaystyle\leq\limsup_{n\to\infty}\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})|\mathbb{D}\boldsymbol{v}_{n}|^{2}+\delta(\theta_{n})|\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}_{n}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
+limn0TΩ2ν(θn)(|𝔻𝒗|22𝔻𝒗n:𝔻𝒗)+δ(θn)(|𝔽𝔽T|22𝔽n𝔽nT:𝔽𝔽T)dxdt\displaystyle\qquad+\lim_{n\to\infty}\int_{0}^{T}\int_{\Omega}2\nu(\theta_{n})(|\mathbb{D}\boldsymbol{v}|^{2}-2\mathbb{D}\boldsymbol{v}_{n}:\mathbb{D}\boldsymbol{v})+\delta(\theta_{n})(|\mathbb{F}\,\mathbb{F}^{T}|^{2}-2\mathbb{F}_{n}\,\mathbb{F}_{n}^{T}:\mathbb{F}\,\mathbb{F}^{T})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
+limn0TΩ2ε(|x𝔽|22x𝔽nx𝔽)dxdt\displaystyle\qquad+\lim_{n\to\infty}\int_{0}^{T}\int_{\Omega}2\varepsilon(|\nabla_{x}\mathbb{F}|^{2}-2\nabla_{x}\mathbb{F}_{n}\because\nabla_{x}\mathbb{F})\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
0TΩ2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T|2+2ε|x𝔽|2dxdt\displaystyle\leq\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
0TΩ2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T|2+2ε|x𝔽|2dxdt=0\displaystyle\qquad-\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0

Consequently, due to Korn inequality and the Sobolev embedding we deduced that

(4.65) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\to\boldsymbol{v} strongly in Lt2Wx1,2,\displaystyle\textrm{ strongly in }L^{2}_{t}W^{1,2}_{x},
(4.66) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\to\mathbb{F} strongly in Lt2Wx1,2,\displaystyle\textrm{ strongly in }L^{2}_{t}W^{1,2}_{x},
(4.67) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\to\mathbb{F} strongly in Lt,x4.\displaystyle\textrm{ strongly in }L^{4}_{t,x}.

Having the above convergence results in hands, combined with (4.43)–(4.54), we deduce from (4.31) the following

(4.68) 0TΩθtϕθ𝒗xϕ+κ(θ)xθxϕdxdt\displaystyle\int_{0}^{T}\int_{\Omega}-\theta\,\partial_{t}\phi-\theta\,\boldsymbol{v}\cdot\nabla_{x}\phi+\kappa(\theta)\nabla_{x}\theta\cdot\nabla_{x}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=0TΩ2ν(θ)|𝔻𝒗|2ϕ+δ(θ)|𝔽𝔽T𝕀|2ϕdxdt+Ωθ0r(x)ϕ(0,x)dx,\displaystyle\quad=\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}\phi+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}-\mathbb{I}|^{2}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t+\int_{\Omega}\theta^{r}_{0}(x)\phi(0,x)\mathop{}\!\mathrm{d}x,

for all ϕ𝒞c1([0,T)×Ω¯)\phi\in\mathcal{C}^{1}_{c}([0,T)\times\overline{\Omega}). Moreover, it follows from the above identity that

(4.69) tθL1((0,T);W2,2(Ω)).\partial_{t}\theta\in L^{1}((0,T);W^{-2,2}(\Omega)).

4.5. Convergence with ε0+\varepsilon\to 0_{+} and r0+r\to 0_{+}.

We denote by {𝒗ε,𝔽ε,θε}ε>0\{\boldsymbol{v}_{\varepsilon},\mathbb{F}_{\varepsilon},\theta_{\varepsilon}\}_{\varepsilon>0} the solution constructed in the previous section and set r=ε0+r=\varepsilon\to 0_{+} to finish the proof of Theorem 4.1. We recall the uniform ε\varepsilon-independent estimates (4.32), (4.33), (4.35), (4.36) and (4.42), which remains valid also here due to the Fatou lemma and weak-lower semicontinuity. Therefore, we may again extract a subsequence that we do not relabel such that

θε\displaystyle\theta_{\varepsilon} θ\displaystyle\rightharpoonup\theta\quad weakly in LtpWx1,p for all p[1,43),\displaystyle\text{ weakly in }L^{p}_{t}W^{1,p}_{x}\textrm{ for all }p\in\left[1,\frac{4}{3}\right),
θε\displaystyle\theta_{\varepsilon} θ\displaystyle\rightarrow\theta\quad strongly in LtsLxq for all q[1,) and s[1,q),\displaystyle\text{ strongly in }L^{s}_{t}L^{q}_{x}\textrm{ for all }q\in[1,\infty)\textrm{ and }s\in[1,q^{\prime}),
lnθε\displaystyle\ln\theta_{\varepsilon} lnθ\displaystyle\rightarrow\ln\theta\quad weakly in Lt2Wx1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{x},
𝒗ε\displaystyle\boldsymbol{v}_{\varepsilon} 𝒗\displaystyle\rightharpoonup\boldsymbol{v}\quad weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝒗ε\displaystyle\boldsymbol{v}_{\varepsilon} 𝒗\displaystyle\rightharpoonup\boldsymbol{v}\quad weakly in Lt2W0,x1,2,\displaystyle\text{ weakly in }L^{2}_{t}W^{1,2}_{0,x},
t𝒗ε\displaystyle\partial_{t}\boldsymbol{v}_{\varepsilon} t𝒗\displaystyle\rightharpoonup\partial_{t}\boldsymbol{v}\quad weakly in L2((0,T);W0,div1,2(Ω))\displaystyle\text{ weakly in }L^{2}((0,T);W^{-1,2}_{0,\operatorname{div}}(\Omega))
𝒗ε\displaystyle\boldsymbol{v}_{\varepsilon} 𝒗\displaystyle\rightarrow\boldsymbol{v}\quad strongly in Lt,xq, for q[1,4),\displaystyle\text{ strongly in }L^{q}_{t,x},\text{ for }q\in[1,4),
𝔽ε\displaystyle\mathbb{F}_{\varepsilon} 𝔽\displaystyle\rightharpoonup\mathbb{F}\quad weakly* in LtLx2,\displaystyle\text{ weakly* in }L^{\infty}_{t}L^{2}_{x},
𝔽ε\displaystyle\mathbb{F}_{\varepsilon} 𝔽\displaystyle\rightharpoonup\mathbb{F}\quad weakly in Lt,x4,\displaystyle\text{ weakly in }L^{4}_{t,x},
t𝔽ε\displaystyle\partial_{t}\mathbb{F}_{\varepsilon} t𝔽\displaystyle\rightharpoonup\partial_{t}\mathbb{F}\quad weakly in L43((0,T);W1,2(Ω;2×2)),\displaystyle\text{ weakly in }L^{\frac{4}{3}}((0,T);W^{-1,2}(\Omega;\mathbb{R}^{2\times 2})),
εx𝔽ε\displaystyle\varepsilon\nabla_{x}\mathbb{F}_{\varepsilon} 𝕆\displaystyle\rightarrow\mathbb{O}\quad strongly in Lt,x2.\displaystyle\text{ strongly in }L^{2}_{t,x}.

Compared to the previous section, we did not get the strong convergence of 𝔽ε𝔽\mathbb{F}_{\varepsilon}\to\mathbb{F} directly from the uniform estimates. However, we can now use the convergence scheme from the proof of Theorem 3.1 (in fact, here it is easier since the function δ\delta is bounded) and deduce

𝔽ε𝔽 strongly in Lt,x2.\displaystyle\mathbb{F}_{\varepsilon}\rightarrow\mathbb{F}\quad\text{ strongly in }L^{2}_{t,x}.

Then we can deduce from (4.55) and (LABEL:eq:galerkin_2ndconv_elastic) that

(4.70) t𝒗,𝝎Ω𝒗𝒗:x𝝎dx+Ω2ν(θ)𝔻𝒗:x𝝎dx+Ω2𝔽𝔽T:x𝝎dx=0\displaystyle\langle\partial_{t}\boldsymbol{v},\boldsymbol{\omega}\rangle-\int_{\Omega}\boldsymbol{v}\otimes\boldsymbol{v}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\nu(\theta)\mathbb{D}\boldsymbol{v}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x+\int_{\Omega}2\mathbb{F}\,\mathbb{F}^{T}:\nabla_{x}\boldsymbol{\omega}\mathop{}\!\mathrm{d}x=0

for almost all time t(0,T)t\in(0,T) and for all 𝝎W0,div1,2(Ω)\boldsymbol{\omega}\in W^{1,2}_{0,\operatorname{div}}(\Omega) and

(4.71) t𝔽,𝔸Ω𝔽𝒗x𝔸+(x𝒗𝔽):𝔸12δ(θ)(𝔽𝔽T𝔽𝔽):𝔸dx=0\displaystyle\langle\partial_{t}\mathbb{F},\mathbb{A}\rangle-\int_{\Omega}\mathbb{F}\otimes\boldsymbol{v}\because\nabla_{x}\mathbb{A}+(\nabla_{x}\boldsymbol{v}\,\mathbb{F}):\mathbb{A}-\frac{1}{2}\delta(\theta)(\mathbb{F}\,\mathbb{F}^{T}\,\mathbb{F}-\mathbb{F}):\mathbb{A}\mathop{}\!\mathrm{d}x=0

for almost all time t(0,T)t\in(0,T) and for all 𝔸Wx1,2(Ω;2×2)\mathbb{A}\in W^{1,2}_{x}(\Omega;\mathbb{R}^{2\times 2}). Moreover, we also have the final formulations in the main theorem (LABEL:weak_formulation_u_constant_g) and (LABEL:weak_formulation_F_constant_g).

Next, we need to get the compactness of the terms appearing on the right-hand side of (LABEL:eq:galerkin_2ndconv_temp). We want to repeat the scheme in Section 4.4, namely the computation between (4.57)–(4.63). To do so, we need to justify the setting 𝔸:=𝔽\mathbb{A}:=\mathbb{F} in (LABEL:eq:galerkin_2ndconv_elastic_zero). This can be however justify by the renormalisation technique developed in [14], see also [7] in the context of viscoelastic fluids. Thus, similarly as in (4.63), we deduce

(4.72) lim supε0+0TΩ2ν(θε)|𝔻𝒗ε|2dx+δ(θε)|𝔽ε𝔽εT|2+2ε|x𝔽ε|2dxdt0TΩ2ν(θ)|𝔻𝒗|2+δ(θ)|𝔽𝔽T|2dxdt.\begin{split}\limsup_{\varepsilon\to 0_{+}}&\int_{0}^{T}\int_{\Omega}2\nu(\theta_{\varepsilon})|\mathbb{D}\boldsymbol{v}_{\varepsilon}|^{2}\mathop{}\!\mathrm{d}x+\delta(\theta_{\varepsilon})|\mathbb{F}_{\varepsilon}\,\mathbb{F}_{\varepsilon}^{T}|^{2}+2\varepsilon|\nabla_{x}\mathbb{F}_{\varepsilon}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t\\ &\quad\leq\int_{0}^{T}\int_{\Omega}2\nu(\theta)|\mathbb{D}\boldsymbol{v}|^{2}+\delta(\theta)|\mathbb{F}\,\mathbb{F}^{T}|^{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.\end{split}

Having this, we may repeat (4.64) and conclude

(4.73) 𝒗n\displaystyle\boldsymbol{v}_{n} 𝒗\displaystyle\to\boldsymbol{v} strongly in Lt2Wx1,2,\displaystyle\textrm{ strongly in }L^{2}_{t}W^{1,2}_{x},
(4.74) 𝔽n\displaystyle\mathbb{F}_{n} 𝔽\displaystyle\to\mathbb{F} strongly in Lt,x4.\displaystyle\textrm{ strongly in }L^{4}_{t,x}.

And finally, we can let ε0+\varepsilon\to 0_{+} in (LABEL:eq:galerkin_2ndconv_temp) and deduce (LABEL:weak_formulation_theta_constant_g). Furthermore, since lnθLt2Wx1,2\ln\theta\in L^{2}_{t}W^{1,2}_{x}, we see that θ>0\theta>0 almost everywhere in (0,T)×Ω(0,T)\times\Omega. The proof is complete.

Appendix A Auxiliary propositions

We recall here several useful tools. The first two results are about the compactness of some weakly converging sequences.

Lemma A.1.

(Generalized Aubin–Lions lemma, [27, Lemma 7.7]) Denote by

W1,p,q(I;X1,X2):={uLp(I;X1);dudtLq(I;X2)}.W^{1,p,q}(I;X_{1},X_{2}):=\left\{u\in L^{p}(I;X_{1});\frac{du}{dt}\in L^{q}(I;X_{2})\right\}.

Then if X1X_{1} is a separable, reflexive Banach space, X2X_{2} is a Banach space and X3X_{3} is a metrizable locally convex Hausdorff space, X1X_{1} embeds compactly into X2X_{2}, X2X_{2} embeds continuously into X3X_{3}, 1<p<1<p<\infty and 1q1\leq q\leq\infty, we have

W1,p,q(I;X1,X3) embeds compactly into Lp(I;X2).W^{1,p,q}(I;X_{1},X_{3})\text{ embeds compactly into }L^{p}(I;X_{2}).

In particular any bounded sequence in W1,p,q(I;X1,X3)W^{1,p,q}(I;X_{1},X_{3}) has a convergent subsequence in Lp(I;X2)L^{p}(I;X_{2}).

Lemma A.2.

(The div-curl lemma, [13]) Let Ωn\Omega\subset\mathbb{R}^{n} be an open and bounded domain with a Lipschitz boundary, and let p,q(1+)p,q\in(1+\infty) with 1p+1q=1\frac{1}{p}+\frac{1}{q}=1. Suppose 𝐮kLp(Ω;n)\boldsymbol{u}_{k}\in L^{p}(\Omega;\mathbb{R}^{n}), 𝐯kLq(Ω;n)\boldsymbol{v}_{k}\in L^{q}(\Omega;\mathbb{R}^{n}) are sequences such that

𝒖k𝒖 weakly in Lp(Ω;n),𝒗k𝒗 weakly in Lq(Ω;n),\boldsymbol{u}_{k}\rightharpoonup\boldsymbol{u}\text{ weakly in }L^{p}(\Omega;\mathbb{R}^{n}),\quad\boldsymbol{v}_{k}\rightharpoonup\boldsymbol{v}\text{ weakly in }L^{q}(\Omega;\mathbb{R}^{n}),

and

𝒖k𝒗k is equiintegrable.\boldsymbol{u}_{k}\cdot\boldsymbol{v}_{k}\text{ is equiintegrable}.

Finally, assume that

div𝒖kdiv𝒖 strongly in (W01,(Ω)),curl𝒗kcurl𝒗 strongly in (W01,(Ω;Mn×n)).\operatorname{div}\boldsymbol{u}_{k}\rightarrow\operatorname{div}\boldsymbol{u}\text{ strongly in }(W^{1,\infty}_{0}(\Omega))^{*},\quad\mathrm{curl}\,\boldsymbol{v}_{k}\rightarrow\mathrm{curl}\,\boldsymbol{v}\text{ strongly in }(W^{1,\infty}_{0}(\Omega;M^{n\times n}))^{*}.

Then,

𝒖k𝒗k𝒖𝒗 weakly in L1(Ω).\boldsymbol{u}_{k}\cdot\boldsymbol{v}_{k}\rightharpoonup\boldsymbol{u}\cdot\boldsymbol{v}\text{ weakly in }L^{1}(\Omega).

The next very classical result is about the integration by parts formula in Bochner function spaces.

Lemma A.3.

Let 1<p,q<+1<p,q<+\infty. Suppose ψ:\psi:\mathbb{R}\rightarrow\mathbb{R} is a Lipschitz function. For rr\in\mathbb{R} we define

Ψ(x)=rxψ(s)ds,x.\Psi(x)=\int_{r}^{x}\psi(s)\mathop{}\!\mathrm{d}s,\quad x\in\mathbb{R}.

Then, for any uW1,p,p(I;W1,q(Ω)):={uLp(I;W1,q(Ω));dudt(Lp(I;W1,q(Ω)))}u\in W^{1,p,p}(I;W^{1,q}(\Omega)):=\left\{u\in L^{p}(I;W^{1,q}(\Omega));\frac{du}{dt}\in(L^{p}(I;W^{1,q}(\Omega)))^{*}\right\} it holds

t1t2tu,ψ(u)dt=ΩΨ(u(t2))dxΩΨ((u(t1))dx,t1,t2I.\int_{t_{1}}^{t_{2}}\langle\partial_{t}u,\psi(u)\rangle\mathop{}\!\mathrm{d}t=\int_{\Omega}\Psi(u(t_{2}))\mathop{}\!\mathrm{d}x-\int_{\Omega}\Psi((u(t_{1}))\mathop{}\!\mathrm{d}x,\quad t_{1},t_{2}\in I.

Data Availability

Data sharing is not applicable to this article as no data sets were generated or analysed during the current study.

Declarations - Conflict of interest

The authors do not have a conflict of interest to declare that are relevant to the content of this article.

References

  • [1] J. M. Ball and F. Murat. Remarks on Chacon’s biting lemma. Proc. Amer. Math. Soc., 107(3):655–663, 1989.
  • [2] M. Bathory, M. Bulíček, and J. Málek. Large data existence theory for three-dimensional unsteady flows of rate-type viscoelastic fluids with stress diffusion. Adv. Nonlinear Anal., 10(1):501–521, 2021.
  • [3] M. Bathory, M. Bulíček, and J. Málek. Coupling the Navier-Stokes-Fourier equations with the Johnson-Segalman stress-diffusive viscoelastic model: global-in-time and large-data analysis. Math. Models Methods Appl. Sci., 34(3):417–476, 2024.
  • [4] M. Bulíček, E. Feireisl, and J. Málek. A Navier-Stokes-Fourier system for incompressible fluids with temperature dependent material coefficients. Nonlinear Anal. Real World Appl., 10(2):992–1015, 2009.
  • [5] M. Bulíček, E. Feireisl, and J. Málek. On a class of compressible viscoelastic rate-type fluids with stress-diffusion. Nonlinearity, 32(12):4665–4681, 2019.
  • [6] M. Bulíček, A. Jüngel, M. Pokorný, and N. Zamponi. Existence analysis of a stationary compressible fluid model for heat-conducting and chemically reacting mixtures. J. Math. Phys., 63(5):Paper No. 051501, 48, 2022.
  • [7] M. Bulíček, T. Los, Y. Lu, and J. Málek. On planar flows of viscoelastic fluids of Giesekus type. Nonlinearity, 35(12):6557–6604, 2022.
  • [8] M. Bulíček, T. Los, and J. Málek. On three-dimensional flows of viscoelastic fluids of Giesekus type. Nonlinearity, 38(1):015004 (42pp), 2025.
  • [9] M. Bulíček, J. Málek, V. Průša, and E. 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.
  • [10] M. Bulíček, J. Málek, and K. R. Rajagopal. Mathematical analysis of unsteady flows of fluids with pressure, shear-rate, and temperature dependent material moduli that slip at solid boundaries. SIAM J. Math. Anal., 41(2):665–707, 2009.
  • [11] J. Burgers. Mechanical considerations—model systems—phenomenological theories of relaxations and viscosity. First Report on Viscosity and Plasticity (New York: Nordemann Publishing Company), pages 5–67, 1939.
  • [12] L. Consiglieri. Weak solutions for a class of non-Newtonian fluids with energy transfer. J. Math. Fluid Mech., 2(3):267–293, 2000.
  • [13] S. Conti, G. Dolzmann, and S. Müller. The div-curl lemma for sequences whose divergence and curl are compact in W1,1W^{-1,1}. C. R. Math. Acad. Sci. Paris, 349(3-4):175–178, 2011.
  • [14] R. J. DiPerna and P.-L. Lions. Ordinary differential equations, transport theory and Sobolev spaces. Invent. Math., 98(3):511–547, 1989.
  • [15] M. Dressler, B. J. Edwards, and H. C. Öttinger. Macroscopic thermodynamics of flowing polymeric liquids. Rheologica Acta, 38(2):117–136, 1999.
  • [16] E. Feireisl. Dynamics of viscous compressible fluids, volume 26 of Oxford Lecture Series in Mathematics and its Applications. Oxford University Press, Oxford, 2004.
  • [17] E. Feireisl and A. Novotný. Singular limits in thermodynamics of viscous fluids. Advances in Mathematical Fluid Mechanics. Birkhäuser/Springer, Cham, second edition, 2017.
  • [18] E. Feireisl and A. Novotný. Mathematics of open fluid systems. Nečas Center Series. Birkhäuser/Springer, Cham, [2022] ©2022.
  • [19] H. Giesekus. A simple constitutive equation for polymer fluids based on the concept of deformation-dependent tensorial mobility. Journal of Non-Newtonian Fluid Mechanics, 11(1):69–109, 1982.
  • [20] J. Hron, V. Miloš, V. Průša, O. Souček, and K. Tůma. On thermodynamics of incompressible viscoelastic rate type fluids with temperature dependent material coefficients. International Journal of Non-Linear Mechanics, 95:193–208, 2017.
  • [21] D. Hu and T. Lelièvre. New entropy estimates for Oldroyd-B and related models. Commun. Math. Sci., 5(4):909–916, 2007.
  • [22] M. Johnson and D. Segalman. A model for viscoelastic fluid behavior which allows non-affine deformation. Journal of Non-Newtonian Fluid Mechanics, 2(3):255–270, 1977.
  • [23] P. L. Lions and N. Masmoudi. Global solutions for some Oldroyd models of non-Newtonian flows. Chinese Ann. Math. Ser. B, 21(2):131–146, 2000.
  • [24] N. Masmoudi. Global existence of weak solutions to macroscopic models of polymeric flows. Journal de Mathématiques Pures et Appliquées, 96(5):502–520, 2011.
  • [25] J. G. Oldroyd. On the formulation of rheological equations of state. Proc. Roy. Soc. London Ser. A, 200:523–541, 1950.
  • [26] K. Rajagopal and A. Srinivasa. A thermodynamic frame work for rate type fluid models. Journal of Non-Newtonian Fluid Mechanics, 88(3):207–227, 2000.
  • [27] T. Roubíček. Nonlinear partial differential equations with applications, volume 153 of International Series of Numerical Mathematics. Birkhäuser/Springer Basel AG, Basel, second edition, 2013.