The informal notion of a well posed problem captures many of the desired features of what we mean by solving a system of partial differential equations. Usually a given problem is well-posed if

  • the problem has a solution;

  • the solution is unique in a given class;

  • the solution depends continuously on the data.

The first condition is particularly important for us as we want to perform the singular limits on existing objects. It is a peculiar feature of non-linear problems that existence of solutions can be rigorously established only in the class determined by a priori estimates. Without any extra assumption concerning the magnitude of the initial data and/or the length of the existence interval (0, T), all available and known a priori bounds on solutions to the Navier-Stokes-Fourier System have been collected in Chap. 2 Accordingly, the existence theory to be developed in the forthcoming chapter necessarily uses the framework of the weak solutions introduced in Chap. 1 and identified in Chap. 2 To begin, let us point out that the existence theory is not the main objective of this book, and, strictly speaking, all results concerning the singular limits can be stated without referring to any specific solution. On the other hand, however, it seems important to know that the class of objects we deal with is not void.

The complete proof of existence for the initial-boundary value problem associated to the Navier-Stokes-Fourier system is rather technical and considerably long. The following text aims to provide a concise and self-contained treatment starting directly with the approximate problem and avoiding completely the nowadays popular “approach” based on reducing the task of existence to showing the weak sequential stability of the set of hypothetical solutions.

The principal tools to be employed in the existence proof can be summarized as follows:

  • Nowadays “classical” arguments based on compactness of embeddings of Sobolev spaces (the Rellich-Kondrashov theorem);

  • a generalized Arzelá-Ascoli compactness result for weakly continuous functions and its variants including Lions-Aubin Lemma;

  • the Div-Curl lemma developed in the theory of compensated compactness;

  • the “weak continuity” property of the so-called effective viscous flux established by P.-L. Lions and its generalization to the case of non-constant viscosity coefficients via a commutator lemma;

  • the theory of parametrized (Young) measures, in particular, its application to compositions of weakly converging sequences with a Carathéodory function;

  • the analysis of density oscillations via oscillations defect measures in weighted Lebesgue spaces.

3.1 Hypotheses

Before formulating our main existence result, we present a concise list of hypotheses imposed on the data. To see their interpretation, the reader may consult Chap. 1 for the physical background and the relevant discussion.

(i) Initial data: The initial state of the system is determined through the choice of the quantities ϱ 0, (ϱ u)0, E 0, and (ϱs)0.

The initial density ϱ 0 is a non-negative measurable function such that

$$\displaystyle{ \varrho _{0} \in L^{\frac{5} {3} }(\Omega ),\ \int _{\Omega }\varrho _{0}\ \mathrm{d}x = M_{0}> 0. }$$
(3.1)

The initial distribution of the momentum satisfies a compatibility condition

$$\displaystyle{ (\varrho \mathbf{u})_{0} = 0\ \mbox{ a.a. on the set}\ \{x \in \Omega \ \vert \ \varrho _{0}(x) = 0\}, }$$
(3.2)

notably the total amount of the kinetic energy is finite, meaning,

$$\displaystyle{ \int _{\Omega }\frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0}} \ \mathrm{d}x <\infty. }$$
(3.3)

The initial temperature is determined by a measurable function ϑ 0 satisfying

$$\displaystyle{ \vartheta _{0}> 0\ \mbox{ a.a. in}\ \Omega,\ (\varrho s)_{0} =\varrho _{0}s(\varrho _{0},\vartheta _{0}),\ \varrho _{0}s(\varrho _{0},\vartheta _{0}) \in L^{1}(\Omega ). }$$
(3.4)

Finally, we assume that he initial energy of the system is finite, specifically,

$$\displaystyle{ E_{0} = \int _{\Omega }\Big( \frac{1} {2\varrho _{0}}\vert (\varrho \mathbf{u})_{0}\vert ^{2} +\varrho _{ 0}e(\varrho _{0},\vartheta _{0})\Big)\ \mathrm{d}x <\infty. }$$
(3.5)

(ii) Source terms: For the sake of simplicity, we suppose that

$$\displaystyle{ \mathbf{f} \in L^{\infty }((0,T) \times \Omega; \mathbb{R}^{3})),\ \mathcal{Q}\geq 0,\ \mathcal{Q}\in L^{\infty }((0,T) \times \Omega ). }$$
(3.6)

(iii) Constitutive relations: The quantities p, e, and s are continuously differentiable functions for positive values of ϱ, ϑ satisfying Gibbs’ equation

$$\displaystyle{ \vartheta Ds(\varrho,\vartheta ) = De(\varrho,\vartheta ) + p(\varrho,\vartheta )D\Big(\frac{1} {\varrho } \Big)\ \mbox{ for all}\ \varrho,\vartheta> 0. }$$
(3.7)

In addition,

$$\displaystyle{ p(\varrho,\vartheta ) = p_{M}(\varrho,\vartheta ) + p_{R}(\vartheta ),\ p_{R}(\vartheta ) = \frac{a} {3}\vartheta ^{4},\ a> 0, }$$
(3.8)

and

$$\displaystyle{ e(\varrho,\vartheta ) = e_{M}(\varrho,\vartheta ) + e_{R}(\varrho,\vartheta ),\ \varrho e_{R}(\varrho,\vartheta ) = a\vartheta ^{4}, }$$
(3.9)

where, in accordance with hypothesis of thermodynamic stability (1.44), the molecular components satisfy

$$\displaystyle{ \frac{\partial p_{M}(\varrho,\vartheta )} {\partial \varrho }> 0\ \mbox{ for all}\ \varrho,\vartheta> 0, }$$
(3.10)

and

$$\displaystyle{ 0 <\frac{\partial e_{M}(\varrho,\vartheta )} {\partial \vartheta } \leq c\ \mbox{ for all}\ \varrho,\vartheta> 0. }$$
(3.11)

Furthermore,

$$\displaystyle{ \lim _{\vartheta \rightarrow 0+}e_{M}(\varrho,\vartheta ) =\underline{ e}_{M}(\varrho )> 0\ \mbox{ for any fixed}\ \varrho> 0, }$$
(3.12)

and,

$$\displaystyle{ \Big\vert \varrho \frac{\partial e_{M}(\varrho,\vartheta )} {\partial \varrho } \Big\vert \leq c\ e_{M}(\varrho,\vartheta )\ \mbox{ for all}\ \varrho,\vartheta> 0. }$$
(3.13)

Finally, we suppose that there is a function P satisfying

$$\displaystyle{ P \in C^{1}[0,\infty ),\;P(0) = 0,\;P'(0)> 0, }$$
(3.14)

and two positive constants \(0 <\underline{ Z} <\overline{Z}\) such that

$$\displaystyle{ p_{M}(\varrho,\vartheta ) =\vartheta ^{\frac{5} {2} }P\Big( \frac{\varrho } {\vartheta ^{\frac{3} {2} }} \Big)\ \mbox{ whenever}\ 0 <\varrho \leq \underline{ Z}\vartheta ^{\frac{3} {2} },\ \mbox{ or,}\ \varrho> \overline{Z}\vartheta ^{\frac{3} {2} }, }$$
(3.15)

where, in addition,

$$\displaystyle{ p_{M}(\varrho,\vartheta ) = \frac{2} {3}\varrho e_{M}(\varrho,\vartheta )\ \mbox{ for}\ \varrho> \overline{Z}\vartheta ^{\frac{3} {2} }. }$$
(3.16)

(iv) Transport coefficients: The viscosity coefficients μ, η are continuously differentiable functions of the absolute temperature ϑ, more precisely μ,  ηC 1[0, ), satisfying

$$\displaystyle{ 0 <\underline{\mu } (1 +\vartheta ^{\alpha }) \leq \mu (\vartheta ) \leq \overline{\mu }(1 +\vartheta ^{\alpha }), }$$
(3.17)
$$\displaystyle{ \sup _{\vartheta \in [0,\infty )}\vert \mu '(\vartheta )\vert \leq \overline{m}, }$$
(3.18)
$$\displaystyle{ 0 \leq \eta (\vartheta ) \leq \overline{\eta }(1 +\vartheta ^{\alpha }). }$$
(3.19)

The heat conductivity coefficient κ can be decomposed as

$$\displaystyle{ \kappa (\vartheta ) =\kappa _{M}(\vartheta ) +\kappa _{R}(\vartheta ), }$$
(3.20)

where κ M ,  κ R C 1[0, ), and

$$\displaystyle{ 0 <\underline{\kappa } _{R}\vartheta ^{3} \leq \kappa _{ R}(\vartheta ) \leq \overline{\kappa }_{R}(1 +\vartheta ^{3}), }$$
(3.21)
$$\displaystyle{ 0 <\underline{\kappa } _{M}(1 +\vartheta ^{\alpha }) \leq \kappa _{M}(\vartheta ) \leq \overline{\kappa }_{M}(1 +\vartheta ^{\alpha }). }$$
(3.22)

In formulas (3.17)–(3.22), \(\underline{\mu }\), \(\overline{\mu }\), \(\overline{m}\), \(\overline{\eta }\), \(\underline{\kappa }_{R}\), \(\overline{\kappa }_{R}\), \(\underline{\kappa }_{M}\), \(\overline{\kappa }_{M}\) are positive constants and

$$\displaystyle{ \frac{2} {5} <\alpha \leq 1. }$$
(3.23)

Remark

Some of the above hypotheses, in particular those imposed on the thermodynamic functions, are rather technical and may seem awkward at first glance. The reader should always keep in mind the prototype example

$$\displaystyle{p(\varrho,\vartheta ) = \vartheta ^{\frac{5} {2} }P\Big( \frac{\varrho } {\vartheta ^{\frac{3} {2} }} \Big) + \frac{a} {3}\vartheta ^{4},\ P(0) = 0,\ P'(0)> 0,\ \ P(Z) \approx Z^{\frac{5} {3} }\ \mbox{ for}\ Z>> 1}$$

which meets all the hypotheses stated above. Note that if a > 0 is small and P(Z) is close to a linear function for moderate values of Z, the above formula approaches the standard Boyle-Marriot law of a perfect gas.

The present hypotheses cover, in particular, the physically reasonable case when the constitutive law for the molecular pressure is that one of the monoatomic gas, meaning

$$\displaystyle{p_{M} = \frac{2} {3}\varrho e_{M},}$$

for more details see Sect. 1.4.2.

Very roughly indeed, we can say that the pressure is regularized in the area where either ϱ or ϑ are close to zero. The radiation component p R prevents the temperature field from oscillating in the vacuum zone where ϱ vanishes, while the superlinear growth of P for large arguments guarantees strong enough a priori estimates on the density ϱ in the “cold” regime ϑ ≈ 0.

3.2 Structural Properties of Constitutive Functions

The hypotheses on constitutive relations for the pressure, the internal energy and the entropy entail further restrictions imposed on the structural properties of the functions p, e, and s. Some of them have already been identified and used in Chap. 2 For reader’s convenience, they are recorded and studied in a systematic way in the text below.

  1. (i)

    The first observation is that for (3.15), (3.16) to be compatible with the hypothesis of thermodynamic stability expressed through (3.10), (3.11), the function P must obey certain structural restrictions. In particular, relation (3.10) yields

    $$\displaystyle{P'(Z)> 0\ \mbox{ whenever}\ 0 <Z <\underline{ Z},\ \mbox{ or,}\ Z> \overline{Z},}$$

    which, together with (3.14), yields

    $$\displaystyle{ P'(Z)> 0\ \mbox{ for all}\ Z \geq 0, }$$
    (3.24)

    where P has been extended to be strictly increasing on the interval \([\underline{Z},\overline{Z}]\).

    Similarly, a direct inspection of (3.11), (3.15), (3.16) gives rise to

    $$\displaystyle{ 0 <\frac{3} {2} \frac{\frac{5} {3}P(Z) - ZP'(Z)} {Z}:= c_{v,M} <c,\;\ \mbox{ whenever}\ Z = \frac{\varrho } {\vartheta ^{3/2}} \geq \overline{Z}. }$$
    (3.25)

    In particular P(Z)∕Z 5∕3 possesses a limit for Z, specifically, in accordance with (3.15), (3.16),

    $$\displaystyle{\lim _{\vartheta \rightarrow 0+}e_{M}(\varrho,\vartheta ) = \frac{3} {2}\lim _{\vartheta \rightarrow 0+}\frac{\vartheta ^{5/2}} {\varrho } P\Big( \frac{\varrho } {\vartheta ^{3/2}}\Big) = \frac{3} {2}\varrho ^{\frac{2} {3} }\lim _{Z\rightarrow \infty }\frac{P(Z)} {Z^{5/3}} \ \mbox{ for any fixed}\ \varrho> 0.}$$

    Moreover, in agreement with (3.12),

    $$\displaystyle{ \lim _{Z\rightarrow \infty }\frac{P(Z)} {Z^{5/3}} = p_{\infty }> 0, }$$
    (3.26)

    and

    $$\displaystyle{ \lim _{\vartheta \rightarrow 0+}e_{M}(\varrho,\vartheta ) =\underline{ e}_{M}(\varrho ) = \frac{3} {2}\varrho ^{2/3}p_{ \infty }. }$$
    (3.27)
  2. (ii)

    By virtue of (3.11), the function ϑe M (ϱ, ϑ) is strictly increasing on the whole interval (0, ) for any fixed ϱ > 0. This fact, together with (3.9), (3.27), gives rise to the lower bound

    $$\displaystyle{ \varrho e(\varrho,\vartheta ) \geq \frac{3p_{\infty }} {2} \varrho ^{\frac{5} {3} } + a\vartheta ^{4}. }$$
    (3.28)

    On the other hand,

    $$\displaystyle{ e_{M}(\varrho,\vartheta ) =\underline{ e}_{M}(\varrho ) +\int _{ 0}^{\vartheta }\frac{\partial e_{M}} {\partial \vartheta } (\varrho,\tau )\,\mathrm{d}\tau, }$$
    (3.29)

    which, together with (3.11) and (3.27), yields

    $$\displaystyle{ 0 \leq e_{M}(\varrho,\vartheta ) \leq c(\varrho ^{\frac{2} {3} }+\vartheta ). }$$
    (3.30)

    Similarly, relation (3.24), together with (3.14)–(3.16), and (3.26), yield the following bounds on the molecular pressure p M :

    $$\displaystyle{ \underline{c}\varrho \vartheta \leq p_{M}(\varrho,\vartheta ) \leq \overline{c}\varrho \vartheta \quad \mbox{ }\mbox{ if}\ \varrho <\overline{Z}\vartheta ^{\frac{3} {2} }, }$$
    (3.31)

    and

    $$\displaystyle{ \underline{c}\varrho ^{\frac{5} {3} } \leq p_{M}(\varrho,\vartheta ) \leq \overline{c}\left \{\begin{array}{c} \vartheta ^{\frac{5} {2} }\ \mbox{ if}\ \varrho <\overline{Z}\vartheta ^{\frac{3} {2} }\\ \\ \varrho ^{ \frac{5} {3} }\ \mbox{ if}\ \varrho> \overline{Z}\vartheta ^{\frac{3} {2} }.\end{array} \right \} }$$
    (3.32)

    Here, we have used the monotonicity of p M in ϱ in order to control the behavior of the pressure in the region

    $$\displaystyle{\underline{Z}\vartheta ^{\frac{3} {2} } \leq \varrho \leq \overline{Z}\vartheta ^{\frac{3} {2} }.}$$

    Moreover, in accordance with (3.30), (3.32), it is easy to observe that

    $$\displaystyle{ e_{M},p_{M}\quad \mbox{ are bounded on bounded sets of }[0,\infty )^{2}. }$$
    (3.33)
  3. (iii)

    In agreement with Gibbs’ relation (3.7), the specific entropy s can be written as

    $$\displaystyle{ s = s_{M} + s_{R},\quad \frac{\partial s_{M}} {\partial \vartheta } = \frac{1} {\vartheta } \frac{\partial e_{M}} {\partial \vartheta },\quad \varrho s_{R}(\varrho,\vartheta ) = \frac{4} {3}a\vartheta ^{3}, }$$
    (3.34)

    where the molecular component s M satisfies

    $$\displaystyle{ s_{M}(\varrho,\vartheta ) = S(Z),\,Z = \frac{\varrho } {\vartheta ^{3/2}},\,S'(Z) = -\frac{3} {2} \frac{\frac{5} {3}P(Z) - ZP'(Z)} {Z^{2}} <0 }$$
    (3.35)

    in the degenerate area \(\varrho> \overline{Z}\vartheta ^{\frac{3} {2} }\). Note that the function S is determined up to an additive constant.

    On the other hand, due to (3.11), the function ϑs M (ϱ, ϑ) is increasing on (0, ) for any fixed ϑ. Accordingly,

    $$\displaystyle{ s_{M}(\varrho,\vartheta ) \leq \left \{\begin{array}{c} s_{M}(\varrho,1)\quad \mbox{ if }\vartheta \leq 1\\ \\ s_{M}(\varrho,1) +\int _{ 1}^{\vartheta }\frac{\partial s_{M}} {\partial \vartheta } (\varrho,\tau )\,\mathrm{d}\tau \leq s_{M}(\varrho,1) + c\log \vartheta \quad \mbox{ if }\vartheta> 1 \end{array} \right \}, }$$
    (3.36)

    where we have exploited (3.11) combined with (3.34) in order to control

    $$\displaystyle{ \left \vert \int _{1}^{\vartheta }\frac{\partial s_{M}} {\partial \vartheta } (\varrho,\tau )\ \mathrm{d}\tau \right \vert \leq c\vert \log \vartheta \vert \ \mbox{ for all}\ \vartheta> 0. }$$
    (3.37)

    Another application of Gibbs’ relation (3.7) yields

    $$\displaystyle{\frac{\partial s_{M}} {\partial \varrho } = -\frac{1} {\varrho ^{2}} \frac{\partial p_{M}} {\partial \vartheta },}$$

    see also (1.3); therefore

    $$\displaystyle{s_{M}(\varrho,1) = s_{M}(1,1) +\int _{ 1}^{\varrho }\frac{1} {\tau ^{2}} \frac{\partial p_{M}} {\partial \vartheta } (\tau,1)\,\mathrm{d}\tau.}$$

    By virtue of (3.15) and (3.25),

    $$\displaystyle{\frac{\partial p_{M}} {\partial \vartheta } (\rho,1) = \frac{5} {2}P(\varrho ) -\frac{3} {2}\varrho P'(\varrho ) \leq c\varrho \ \mbox{ for all}\ \varrho \in (0,\underline{Z}] \cup [\overline{Z},\infty ),}$$

    whereas

    $$\displaystyle{\vert \frac{\partial p_{M}} {\partial \vartheta } (\rho,1)\vert \quad \mbox{ is bounded in }[\underline{Z},\overline{Z}].}$$

    Consequently,

    $$\displaystyle{ \vert s_{M}(\varrho,1)\vert \leq c(1 + \vert \log \varrho \vert )\ \mbox{ for all}\ \varrho \in (0,\infty ). }$$
    (3.38)

    Writing

    $$\displaystyle{s_{M}(\varrho,\vartheta ) = s_{M}(\varrho,1) +\int _{ 1}^{\varrho }\frac{\partial s_{M}} {\partial \vartheta } (\varrho,\tau )\ \mathrm{d}\tau }$$

    and resuming the previous estimates, we conclude that

    $$\displaystyle{ \vert s_{M}(\varrho,\vartheta )\vert \leq c(1 + \vert \log \varrho \vert + \vert \log \vartheta \vert )\ \mbox{ for all}\ \varrho,\vartheta> 0. }$$
    (3.39)
  4. (iv)

    It follows from (3.35) that

    $$\displaystyle{ \begin{array}{c} \lim _{Z\rightarrow \infty }S(Z) = s_{\infty } = \left \{\begin{array}{c} -\infty \\ \\ 0\end{array} \right \};\\ \\ \mbox{ whence}\\ \\ \lim _{\vartheta \rightarrow 0+}s_{M}(\varrho,\vartheta ) = s_{\infty }\ \mbox{ for any fixed}\ \varrho> 0.\end{array} }$$
    (3.40)

    where, in the latter case, we have fixed the free additive constant in the definition of S in (3.35) to obtain s = 0.

  5. (v)

    Finally, as a direct consequence of (3.15),

    $$\displaystyle{\frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta ) = \vartheta P'\left ( \frac{\varrho } {\vartheta ^{\frac{3} {2} }} \right )\mbox{ if}\ \varrho <\underline{ Z}\vartheta ^{\frac{3} {2} },\ \mbox{ or,}\ \varrho> \overline{Z}\vartheta ^{\frac{3} {2} },}$$

    where, by virtue of (3.24)–(3.26),

    $$\displaystyle{ P'(Z) \geq c(1 + Z^{\frac{2} {3} }),\ c> 0,\ \mbox{ for all}\ Z \geq 0. }$$
    (3.41)

    Thus we can write

    $$\displaystyle{p_{M}(\varrho,\vartheta ) = \vartheta ^{\frac{5} {2} }P\left ( \frac{\varrho } {\vartheta ^{\frac{3} {2} }} \right ) + p_{b}(\varrho,\vartheta ),}$$

    with

    $$\displaystyle{p_{b}(\varrho,\vartheta ) = p_{M}(\varrho,\vartheta ) -\vartheta ^{\frac{5} {2} }P\left ( \frac{\varrho } {\vartheta ^{\frac{3} {2} }} \right ).}$$

    In accordance with (3.15), (3.32), we have

    $$\displaystyle{ \vert p_{b}(\varrho,\vartheta )\vert \leq c(1 + \vartheta ^{\frac{5} {2} }). }$$
    (3.42)

    Finally, we conclude with help of (3.41) that there exists d > 0 such that

    $$\displaystyle{ p_{M}(\varrho,\vartheta ) = d\varrho ^{\frac{5} {3} } + p_{m}(\varrho,\vartheta ) + p_{b}(\varrho,\vartheta ), }$$
    (3.43)

    where

    $$\displaystyle{ \frac{\partial p_{m}} {\partial \varrho } (\varrho,\vartheta )> 0\ \mbox{ for all}\ \varrho,\vartheta> 0. }$$
    (3.44)

3.3 Main Existence Result

Having collected all the preliminary material, we are in a position to formulate our main existence result concerning the weak solutions of the Navier-Stokes-Fourier system.

      Global Existence for the Navier-Stokes-Fourier System:

Theorem 3.1

Let \(\Omega \subset \mathbb{R}^{3}\) be a bounded domain of class C 2,ν , ν ∈ (0, 1). Assume that

  • the data ϱ 0 , (ϱ u)0 , E 0 , (ϱs)0 satisfy ( 3.1 )–( 3.5 );

  • the source terms f , \(\mathcal{Q}\) are given by ( 3.6 );

  • the thermodynamic functions p, e, s, and the transport coefficients μ, η, κ obey the structural hypotheses ( 3.7 )–( 3.23 ).

Then for any T > 0 the Navier-Stokes-Fourier system admits a weak solution {ϱ, u, ϑ} on \((0,T) \times \Omega\) in the sense specified in Sect.  2.1 . More precisely, {ϱ, u, ϑ} satisfy relations ( 2.2 )–(2.6), (2.9)–(2.17), (2.22)–(2.25), (2.27)–(2.32), with (2.35)–(2.37).

The complete proof of Theorem 3.1 presented in the remaining part of this chapter is tedious, rather technical, consisting in four steps:

  • The momentum equation (2.9) is replaced by a Faedo-Galerkin approximation, the equation of continuity (2.2) is supplemented with an artificial viscosity term, and the entropy production equation (2.27) is replaced by the balance of internal energy. The approximate solutions are obtained by help of the Schauder fixed point theorem, first locally in time, and then extended on the full interval (0, T) by means of suitable uniform estimates.

  • Performing the limit in the Faedo-Galerkin approximation scheme we recover the momentum equation supplemented with an artificial pressure term. Simultaneously, the balance of internal energy is converted to the entropy production equation (2.27), together with the total energy balance (2.22) containing some extra terms depending on small parameters.

  • We pass to the limit in the regularized equation of continuity sending the artificial viscosity terms to zero.

  • Finally, the proof of Theorem 3.1 is completed letting the artificial pressure term go to zero.

3.3.1 Approximation Scheme

  1. (i)

    The equation of continuity (2.2) is regularized by means of an artificial viscosity term:

    $$\displaystyle{ \partial _{t}\varrho + \mathrm{div}_{x}(\varrho \mathbf{u}) = \varepsilon \Delta \varrho \ \mbox{ in}\ (0,T) \times \Omega, }$$
    (3.45)

    and supplemented with the homogeneous Neumann boundary condition

    $$\displaystyle{ \nabla _{x}\varrho \cdot \mathbf{n}\vert _{\partial \Omega } = 0, }$$
    (3.46)

    and the initial condition

    $$\displaystyle{ \varrho (0,\cdot ) =\varrho _{0,\delta }, }$$
    (3.47)

    where

    $$\displaystyle{ \varrho _{0,\delta } \in C^{2,\nu }(\overline{\Omega }),\ \inf _{ x\in \Omega }\varrho _{0,\delta }(x)> 0,\ \nabla _{x}\varrho _{0,\delta } \cdot \mathbf{n}\vert _{\partial \Omega } = 0. }$$
    (3.48)
  2. (ii)

    The momentum balance expressed through the integral identity (2.9) is replaced by a Faedo-Galerkin approximation:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \mathbf{u} \cdot \partial _{t}\boldsymbol{\varphi } +\varrho [\mathbf{u} \otimes \mathbf{u}]: \nabla _{x}\boldsymbol{\varphi } +\Big (\,p(\varrho,\vartheta ) +\delta (\varrho ^{\Gamma } +\varrho ^{2})\Big)\mathrm{div}_{ x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
    (3.49)
    $$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\Big(\varepsilon (\nabla _{x}\varrho \nabla _{x}\mathbf{u}) \cdot \boldsymbol{\varphi } +\mathbb{S}_{\delta }: \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}_{\delta }\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }(\varrho \mathbf{u})_{0}\cdot \boldsymbol{\varphi }\ \mathrm{d}x,}$$

    to be satisfied for any test function \(\boldsymbol{\varphi }\in C_{c}^{1}([0,T);X_{n})\), where

    $$\displaystyle{ X_{n} \subset C^{2,\nu }(\overline{\Omega }; \mathbb{R}^{3}) \subset L^{2}(\Omega; \mathbb{R}^{3}) }$$
    (3.50)

    is a finite-dimensional vector space of functions satisfying either

    $$\displaystyle{ \boldsymbol{\varphi }\cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ in the case of the complete slip boundary conditions,} }$$
    (3.51)

    or

    $$\displaystyle{ \boldsymbol{\varphi }\vert _{\partial \Omega } = 0\ \mbox{ in the case of the no-slip boundary conditions.} }$$
    (3.52)

    The space X n is endowed with the Hilbert structure induced by the scalar product of the Lebesgue space \(L^{2}(\Omega; \mathbb{R}^{3})\).

    Furthermore, we set

    $$\displaystyle{ \mathbb{S}_{\delta } = \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u})\ = (\mu (\vartheta )+\delta \vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}\Big) +\eta (\vartheta )\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}, }$$
    (3.53)

    while the function

    $$\displaystyle{ \mathbf{f}_{\delta } \in C^{1}([0,T] \times \overline{\Omega }; \mathbb{R}^{3}) }$$
    (3.54)

    is a suitable approximation of the driving force f.

  3. (iii)

    Instead of the entropy balance (2.27), we consider a modified internal energy equation in the form:

    $$\displaystyle{ \partial _{t}(\varrho e_{\delta }(\varrho,\vartheta )) + \mathrm{div}_{x}(\varrho e_{\delta }(\varrho,\vartheta )\mathbf{u}) -\mathrm{div}_{x}\nabla _{x}\mathcal{K}_{\delta }(\vartheta ) }$$
    (3.55)
    $$\displaystyle{= \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u} - p(\varrho,\vartheta )\mathrm{div}_{x}\mathbf{u} +\varrho \mathcal{Q}_{\delta } + \varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5},}$$

    supplemented with the Neumann boundary condition

    $$\displaystyle{ \nabla _{x}\vartheta \cdot \mathbf{n}\vert _{\partial \Omega } = 0, }$$
    (3.56)

    and the initial condition

    $$\displaystyle{ \vartheta (0,\cdot ) =\vartheta _{0,\delta }, }$$
    (3.57)
    $$\displaystyle{ \vartheta _{0,\delta } \in W^{1,2}(\Omega ) \cap L^{\infty }(\Omega ),\ \mathrm{ess}\inf _{ x\in \Omega }\vartheta _{0,\delta }(x)> 0. }$$
    (3.58)

    Here

    $$\displaystyle{ e_{\delta }(\varrho,\vartheta ) = e_{M,\delta }(\varrho,\vartheta ) + a\vartheta ^{4},\quad e_{ M,\delta }(\varrho,\vartheta ) = e_{M}(\varrho,\vartheta )+\delta \vartheta, }$$
    (3.59)
    $$\displaystyle{\mathcal{K}_{\delta }(\vartheta ) =\int _{ 1}^{\vartheta }\kappa _{ \delta }(z)\ \mathrm{d}z,\ \kappa _{\delta }(\vartheta ) =\kappa _{M}(\vartheta ) +\kappa _{R}(\vartheta ) +\delta \Big (\vartheta ^{\Gamma } + \frac{1} {\vartheta } \Big),}$$

    and

    $$\displaystyle{ \mathcal{Q}_{\delta }\geq 0,\ \mathcal{Q}_{\delta }\in C^{1}([0,T] \times \overline{\Omega }). }$$
    (3.60)

In problem (3.45)–(3.60), the quantities ɛ, δ are small positive parameters, while \(\Gamma> 0\) is a sufficiently large fixed number. The meaning of the extra terms will become clear in the course of the proof. Loosely speaking, the ɛ-dependent quantities provide more regularity of the approximate solutions modifying the type of the field equations, while the δ-dependent quantities prevent concentrations yielding better estimates on the amplitude of the approximate solutions. For technical reasons, the limit passage must be split up in two steps letting first ɛ → 0 and then δ → 0.

3.4 Solvability of the Approximate System

We claim the following result concerning solvability of the approximate problem (3.45)–(3.60).

      Global Existence for the Approximate System:

Proposition 3.1

Let ɛ, δ be given positive parameters.

Under the hypotheses of Theorem  3.1 , there exists \(\Gamma _{0}> 0\) such that for any \(\Gamma> \Gamma _{0}\) the approximate problem ( 3.45 )–( 3.60 ) admits a strong solution {ϱ, u, ϑ} belonging to the following regularity class:

$$\displaystyle{ \begin{array}{c} \varrho \in C([0,T];C^{2,\nu }(\overline{\Omega })),\ \partial _{t}\varrho \in C([0,T];C^{0,\nu }(\overline{\Omega })),\ \inf _{[0,T]\times \overline{\Omega }}\varrho> 0,\\ \\ \mathbf{u} \in C^{1}([0,T];X_{n}),\\ \\ \vartheta \in C([0,T];W^{1,2}(\Omega )) \cap L^{\infty }((0,T) \times \Omega ),\ \partial _{t}\vartheta,\ \Delta \mathcal{K}_{\delta }(\vartheta ) \in L^{2}((0,T) \times \Omega ),\\ \\ \mathrm{ess}\inf _{(0,T)\times \Omega }\vartheta> 0. \end{array} }$$
(3.61)

Remark

As a matter of fact, since the velocity field u is continuously differentiable, a bootstrap argument could be used in order to show that ϑ is smooth, hence a classical solution of (3.55) for t > 0, as soon as the thermodynamic functions p, e as well as the transport coefficients μ, λ, and κ are smooth functions of ϱ, ϑ on the set (0, )2.

In spite of a considerable number of technicalities, the proof of Proposition 3.1 is based on standard arguments. We adopt the following strategy:

  • The solution u of the approximate momentum equation (3.49) is looked for as a fixed point of a suitable integral operator in the Banach space C([0, T]; X n ). Consequently, the functions ϱ, ϑ have to be determined in terms of u. This is accomplished in the following manner:

  • Given u, the approximate continuity equation (3.45) is solved directly by means of the standard theory of linear parabolic equations.

  • Having solved (3.45)–(3.47) we determine the temperature ϑ as a solution of the quasilinear parabolic problem (3.55)–(3.57), where ϱ, u play a role of given data.

3.4.1 Approximate Continuity Equation

The rest of this section is devoted to the proof of Proposition 3.1. We start with a series of preparatory steps. Following the strategy delineated in the previous paragraph, we fix a vector field u and discuss solvability of the Neumann-initial value problem (3.45)–(3.47).

       Approximate Continuity Equation:

Lemma 3.1

Let \(\Omega \subset \mathbb{R}^{3}\) be a bounded domain of class C 2,ν , ν ∈ (0, 1) and let uC([0, T]; X n ) be a given vector field. Suppose that ϱ 0,δ belongs to the class of regularity specified in ( 3.48 ).

Then problem ( 3.45 )–( 3.47 ) possesses a unique classical solution ϱ = ϱ u , more specifically,

$$\displaystyle{ \varrho _{\mathbf{u}} \in V \equiv \left \{\begin{array}{c} \varrho \in C([0,T];C^{2,\nu }(\overline{\Omega })),\\ \\ \partial _{t}\varrho \in C([0,T];C^{0,\nu }(\overline{\Omega })) \end{array} \right \} }$$
(3.62)

for a certain ν > 0. Moreover, the mapping uC([0, T]; X n ) ↦ ϱ u maps bounded sets in C([0, T]; X n ) into bounded sets in V and is continuous with values in \(C^{1}([0,T] \times \overline{\Omega })\) .

Finally,

$$\displaystyle{ \underline{\varrho }_{0}\exp \Big(-\int _{0}^{\tau }\|\mathrm{div}_{ x}\mathbf{u}\|_{L^{\infty }(\Omega )}\ \mathrm{d}t\Big) \leq \varrho _{\mathbf{u}}(\tau,x) }$$
(3.63)
$$\displaystyle{\leq \overline{\varrho }_{0}\exp \Big(\int _{0}^{\tau }\|\mathrm{div}_{ x}\mathbf{u}\|_{L^{\infty }(\Omega )}\ \mathrm{d}t\Big)\ \mathit{\mbox{ for all}}\ \tau \in [0,T],\ x \in \Omega,}$$

where \(\underline{\varrho }_{0} =\inf _{\Omega }\varrho _{0,\delta }\) , \(\overline{\varrho }_{0} =\sup _{\Omega }\varrho _{0,\delta }.\)

Proof

 Step 1: The unique strong solution of problem (3.45)–(3.48)

$$\displaystyle{\varrho \in L^{2}(0,T;W^{2,2}(\Omega )) \cap C([0,T];W^{1,2}(\Omega )),\;\partial _{ t}\varrho \in L^{2}((0,T) \times \Omega ))}$$

that satisfies the estimate

$$\displaystyle{\|\varrho \|_{C([0,T];W^{1,2}(\Omega ))} +\|\varrho \| _{L^{2}(0,T;W^{2,2}(\Omega ))} +\| \partial _{t}\varrho \|_{L^{2}((0,T)\times \Omega ))} \leq c\|\varrho _{0,\delta }\|_{W^{1,2}(\Omega )},}$$

with \(c = c(\varepsilon,T,\|\mathbf{u}\|_{C[0,T];C^{\nu }(\overline{\Omega }))})> 0\), may be constructed by means of the standard Galerkin approximation within the standard L 2 theory.

The maximal L pL q regularity resumed in Theorem 11.29 in Appendix applied to the problem

$$\displaystyle{ \partial _{t}\varrho -\varepsilon \Delta _{x}\varrho = f:= -\mathrm{div}_{x}(\varrho \mathbf{u}),\;\nabla _{x}\varrho \cdot \mathbf{n}\vert _{\partial \Omega } = 0,\;\varrho (0) =\varrho _{0,\delta } }$$
(3.64)

combined with a bootstrap argument gives the bound

$$\displaystyle{\|\varrho \|_{ C([0,T];W^{2-\frac{2} {p},p }(\Omega ))} +\|\varrho \| _{L^{\,p}(0,T;W^{2,p}(\Omega ))} +\| \partial _{t}\varrho \|_{L^{\,p}((0,T)\times \Omega )} \leq c\|\varrho _{0,\delta }\|_{ W^{2-\frac{2} {p},p }(\Omega )}}$$

for any p > 3.

Since \(W^{2-\frac{2} {p},p}(\Omega )\hookrightarrow C^{1,\nu }(\overline{\Omega })\) for any sufficiently large p, we have div x (ϱ u) ∈ C([0, T]; \(C^{1,\nu }(\overline{\Omega }))\) and may employ Theorem 11.30 from Appendix to show relation (3.62) as well as boundedness of the map uϱ u : C([0, T]; X n ) → V.

Step 2: The difference \(\omega =\varrho _{\mathbf{u}_{1}} -\varrho _{\mathbf{u}_{2}}\) satisfies

$$\displaystyle{\partial _{t}\omega -\varepsilon \Delta \omega + \mathrm{div}_{x}(\omega \mathbf{u}_{1}) = f:= \mathrm{div}_{x}(\varrho _{\mathbf{u}_{2}}(\mathbf{u}_{1} -\mathbf{u}_{2})),\;\nabla _{x}\omega \cdot \mathbf{n}\vert _{\partial \Omega } = 0,\;\omega (0) = 0.}$$

Similar reasoning as in the first step applied to this equation yields the continuity of the map uϱ u from C([0, T]; X n ) to \(C^{1}([0,T] \times \overline{\Omega })\).

Step 3: The difference

$$\displaystyle{\omega (t,x) =\varrho _{\mathbf{u}}(\tau,x) -\overline{\varrho }_{0}\mathrm{exp}\left (\int _{0}^{\tau }\|\mathrm{div}_{ x}\mathbf{u}\|_{L^{\infty }(\Omega )}\,\mathrm{d}t\right )}$$

obeys a differential inequality

$$\displaystyle{\partial _{t}\omega + \mathrm{div}_{x}(\omega \mathbf{u}) -\varepsilon \Delta _{x}\omega \leq 0,\;\nabla _{x}\omega \cdot \mathbf{n}\vert _{\partial \Omega } = 0,\;\omega (0) =\varrho _{0} -\overline{\varrho }_{0} \leq 0.}$$

When multiplied on the positive part | ω |+ and integrated over \(\Omega\), the first relation gives \(\|\,\vert \omega \vert ^{+}(t)\|_{L^{2}(\Omega )} \leq 0\) which shows the right inequality in (3.63). The left inequality can be obtained in a similar way. Lemma 3.1 is thus proved. The reader may consult [102, Chap. 7.3] or [224, Sect. 7.2] for more details. □

3.4.2 Approximate Internal Energy Equation

Having fixed u, together with ϱ = ϱ u —the unique solution of problem (3.45)–(3.47)—we focus on the approximate internal energy equation (3.55) that can be viewed as a quasilinear parabolic problem for the unknown ϑ.

Comparison Principle To begin, we establish a comparison principle in the class of strong (super, sub) solutions of problem (3.55)–(3.57). We recall that a function ϑ is termed a super (sub) solution if it satisfies (3.55) with “ = ” sign replaced by “ ≥ ” (“ ≤ ”).

Lemma 3.2

Given the quantities

$$\displaystyle{ \mathbf{u} \in C([0,T];X_{n}),\ \varrho \in C([0,T];C^{2}(\overline{\Omega })),\ \partial _{ t}\varrho \in C([0,T] \times \overline{\Omega }),\ \inf _{(0,T)\times \Omega }\varrho> 0, }$$
(3.65)

assume that \(\underline{\vartheta }\) and \(\overline{\vartheta }\) are respectively a sub and super-solution to problem ( 3.55 )–( 3.57 ) belonging to the class

$$\displaystyle{ \left \{\begin{array}{c} \underline{\vartheta },\ \overline{\vartheta } \in L^{2}(0,T;W^{1,2}(\Omega )),\ \partial _{ t}\underline{\vartheta },\ \partial _{t}\overline{\vartheta } \in L^{2}((0,T) \times \Omega ),\\ \\ \Delta \mathcal{K}_{\delta }(\underline{\vartheta }),\ \Delta \mathcal{K}_{\delta }(\overline{\vartheta }) \in L^{2}((0,T) \times \Omega ), \end{array} \right \}, }$$
(3.66)
$$\displaystyle{ \left \{\begin{array}{c} 0 <\mathrm{ess}\ \inf \nolimits _{(0,T)\times \Omega }\underline{\vartheta } \leq \mathrm{ess}\ \sup \nolimits _{(0,T)\times \Omega }\underline{\vartheta } <\infty,\\ \\ 0 <\mathrm{ess}\ \inf \nolimits _{(0,T)\times \Omega }\overline{\vartheta } \leq \mathrm{ess}\ \sup \nolimits _{(0,T)\times \Omega }\overline{\vartheta } <\infty,\end{array} \right \} }$$
(3.67)

and satisfying

$$\displaystyle{ \underline{\vartheta }(0,\cdot ) \leq \overline{\vartheta }(0,\cdot )\ \mathit{\mbox{ a.a. in}}\ \Omega. }$$
(3.68)

Then

$$\displaystyle{\underline{\vartheta }(t,x) \leq \overline{\vartheta }(t,x)\ \mathit{\mbox{ a.a. in}}\ (0,T) \times \Omega.}$$

Proof

As \(\underline{\vartheta }\), \(\overline{\vartheta }\) belong to the regularity class specified in (3.66), we can compute

$$\displaystyle{ \mathrm{sgn}^{+}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big)\Big[\Big(\partial _{ t}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big) }$$
(3.69)
$$\displaystyle{+\nabla _{x}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big) \cdot \mathbf{u}\Big]}$$
$$\displaystyle{+\Delta _{x}\Big(\mathcal{K}_{\delta }(\overline{\vartheta }) -\mathcal{K}_{\delta }(\underline{\vartheta })\Big)\mathrm{sgn}^{+}\Big(\varrho e(\varrho,\underline{\vartheta }) -\varrho e(\varrho,\overline{\vartheta })\Big)}$$
$$\displaystyle{\leq \vert F(t,x,\underline{\vartheta }) - F(t,x,\overline{\vartheta })\vert \ \mathrm{sgn}^{+}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big),}$$

where we have introduced

$$\displaystyle{\mathrm{sgn}^{+}(z) = \left \{\begin{array}{l} 0\ \mbox{ if}\ z \leq 0,\\ \\ 1\ \mbox{ if}\ z> 0, \end{array} \right.}$$

and where we have set

$$\displaystyle{F(t,x,\vartheta ) = \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}(t,x)): \nabla _{x}\mathbf{u}(t,x) +\big (\varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}\big)(t,x)}$$
$$\displaystyle{-\varrho (t,x)e_{\delta }(\varrho (t,x),\vartheta )\mathrm{div}_{x}\mathbf{u}(t,x) - p(\varrho (t,x),\vartheta )\mathrm{div}_{x}\mathbf{u}(t,x) +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5} +\varrho \mathcal{Q}_{\delta }.}$$

In accordance with our hypotheses, we may assume that F = F(t, x, ϑ) is globally Lipschitz with respect to ϑ.

Denoting | z |+ = max{z, 0} the positive part, we have

$$\displaystyle{\partial _{t}\vert w\vert ^{+} = \mathrm{sgn}^{+}(w)\partial _{ t}w,\ \nabla _{x}\vert w\vert ^{+} = \mathrm{sgn}^{+}(w)\nabla _{ x}w\ \mbox{ a.a. in}\ (0,T) \times \Omega }$$

for any \(w \in W^{1,2}((0,T) \times \Omega )\), in particular,

$$\displaystyle{\mathrm{sgn}^{+}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big)\times }$$
$$\displaystyle{\times \Big[\Big(\partial _{t}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big) + \nabla _{x}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big) \cdot \mathbf{u}\Big]}$$
$$\displaystyle{= \partial _{t}\Big\vert \varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big\vert ^{+} + \nabla _{ x}\Big\vert \varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big\vert ^{+} \cdot \mathbf{u}.}$$

Moreover, as both e δ and \(\mathcal{K}_{\delta }\) are increasing functions of ϑ, we have

$$\displaystyle{\mathrm{sgn}^{+}\Big(\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big) = \mathrm{sgn}^{+}\Big(\mathcal{K}_{\delta }(\underline{\vartheta }) -\mathcal{K}_{\delta }(\overline{\vartheta })\Big).}$$

Seeing that

$$\displaystyle{\int _{\Omega }\Delta _{x}w\ \mathrm{sgn}^{+}(w)\ \mathrm{d}x \leq 0\ \mbox{ whenever}\ w \in W^{2,2}(\Omega ),\ \nabla _{ x}w \cdot \mathbf{n}\vert _{\partial \Omega } = 0,}$$

we can integrate (3.69) in order to deduce

$$\displaystyle{\int _{\Omega }\Big\vert \varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big\vert ^{+}(\tau )\ \mathrm{d}x}$$
$$\displaystyle{\leq c\int _{0}^{\tau }\int _{ \Omega }(1 + \vert \mathrm{div}_{x}\mathbf{u}\vert )\Big\vert \varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\Big\vert ^{+}\ \mathrm{d}x\ \mathrm{d}t}$$

for any τ > 0. Here we have used Lipschitz continuity of F(t, x, ⋅ ) and the fact that \(\vert \underline{\vartheta }-\overline{\vartheta }\vert\) \(\mathrm{sgn}^{+}[\varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })]\) \(\leq c\vert \varrho e_{\delta }(\varrho,\underline{\vartheta }) -\varrho e_{\delta }(\varrho,\overline{\vartheta })\vert ^{+}\) which follows from (3.9), (3.11), (3.65), (3.67). Thus a direct application of Gronwall’s lemma, together with the monotonicity of e δ with respect to ϑ, completes the proof.

Corollary 3.1

For given data ϱ, u satisfying ( 3.65 ), and a measurable function ϑ 0,δ such that

$$\displaystyle{ 0 <\underline{\vartheta } _{0} = \mathrm{ess}\inf _{\Omega }\vartheta _{0,\delta } \leq \mathrm{ess}\sup _{\Omega }\vartheta _{0,\delta } = \overline{\vartheta }_{0} <\infty, }$$
(3.70)

problem ( 3.55 )–( 3.57 ) admits at most one (strong) solution ϑ in the class specified in ( 3.66 )–( 3.67 ).

Another application of Lemma 3.2 gives rise to uniform bounds on the function ϑ in terms of the data.

Corollary 3.2

Let ϱ, u belong to the regularity class ( 3.65 ), and let ϑ 0,δ satisfy ( 3.70 ). Suppose that ϑ is a (strong) solution of problem ( 3.55 )–( 3.57 ) belonging to the regularity class ( 3.66 ).

Then there exist two constants \(\underline{\vartheta }\) , \(\overline{\vartheta }\) depending only on the quantities

$$\displaystyle{ \|\mathbf{u}\|_{C([0,T];X_{n})},\ \|\varrho \|_{C^{1}([0,T]\times \overline{\Omega })}, }$$

satisfying

$$\displaystyle{ 0 <\underline{\vartheta }\leq \underline{\vartheta }_{0} \leq \overline{\vartheta }_{0} \leq \overline{\vartheta }, }$$
(3.71)

and

$$\displaystyle{ \underline{\vartheta }\leq \vartheta (t,x) \leq \overline{\vartheta }\ \mathit{\mbox{ for a.a.}}\ (t,x) \in (0,T) \times \Omega. }$$
(3.72)

Proof

It is a routine matter to check that a constant function \(\underline{\vartheta }\) is a subsolution of (3.55)–(3.57) as soon as

$$\displaystyle{ \frac{\delta } {\underline{\vartheta }^{2}} \geq \Big [\varepsilon \underline{\vartheta }^{5} + p_{ M}(\varrho,\underline{\vartheta })\mathrm{div}_{x}\mathbf{u} + a\underline{\vartheta }^{4}\mathrm{div}_{ x}\mathbf{u} }$$
(3.73)
$$\displaystyle{+\varrho \frac{\partial e_{M}(\varrho,\underline{\vartheta })} {\partial \varrho } \Big(\partial _{t}\varrho + \mathbf{u} \cdot \nabla _{x}\varrho \Big) +\Big (e_{M}(\varrho,\underline{\vartheta }) + a\underline{\vartheta }^{4}+\delta \underline{\vartheta }\Big)\Big(\partial _{ t}\varrho + \mathrm{div}_{x}(\varrho \mathbf{u})\Big)}$$
$$\displaystyle{-\mathbb{S}_{\delta }(\underline{\vartheta },\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u} -\varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} -\varrho Q_{\delta }\Big].}$$

Revoking (3.30) we can use hypotheses (3.65), (3.13), together with estimate (3.32), in order to see that all quantities on the right-hand side of (3.73) are bounded in terms of \(\|\varrho \|_{ C^{1}([0,T]\times \overline{\Omega })}\) and \(\|\mathbf{u}\|_{C([0,T];X_{n})}\) provided, say, \(0 <\underline{\vartheta }<1\). Note that all norms are equivalent when restricted to the finite dimensional space X n .

Consequently, a direct application of the comparison principle established in Lemma 3.2 yields the left inequality in (3.72).

Following step by step with obvious modifications the above procedure, the upper bound claimed in (3.72) can be established by help of the dominating term −ɛϑ 5 in (3.55).

Remark

Corollary 3.2 reveals the role of the extra term δϑ 2 in Eq. (3.55), namely to keep the absolute temperature ϑ bounded below away from zero at this stage of the approximation procedure. Positivity of ϑ is necessary for the passage from (3.55) to the entropy balance equation used in the weak formulation of the Navier-Stokes-Fourier system.

Priori Estimates We shall derive a priori estimates satisfied by any strong solution of problem (3.55)–(3.57).

Lemma 3.3

Let the data ϱ, u belong to the regularity class ( 3.65 ), and let \(\vartheta _{0,\delta } \in W^{1,2}(\Omega )\) satisfy ( 3.70 ).

Then any strong solution ϑ of problem ( 3.55 )–( 3.57 ) belonging to the class ( 3.66 )–( 3.67 ) satisfies the estimate

$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\|\vartheta \|_{W^{1,2}(\Omega )}^{2} +\int _{ 0}^{T}\Big(\|\partial _{ t}\vartheta \|_{L^{2}(\Omega )}^{2} +\| \Delta _{ x}\mathcal{K}_{\delta }(\vartheta )\|_{L^{2}(\Omega )}^{2}\Big)\mathrm{d}t }$$
(3.74)
$$\displaystyle{\leq h\Big(\|\varrho \|_{C^{1}([0,T]\times \overline{\Omega })},\|\mathbf{u}\|_{C([0,T;X_{n})},(\inf _{(0,T)\times \Omega }\varrho )^{-1},\|\vartheta _{ 0,\delta }\|_{W^{1,2}(\Omega )}\Big),}$$

where h is bounded on bounded sets.

Proof

Note that relation (3.74) represents the standard energy estimates for problem (3.55)–(3.57). These are easily deduced via multiplying Eq. (3.55) by ϑ and integrating the resulting expression by parts in order to obtain

$$\displaystyle{ \frac{1} {2}\int _{\Omega }\varrho \frac{\partial e_{\delta }} {\partial \vartheta } (\varrho,\vartheta )\partial _{t}\vartheta ^{2}\ \mathrm{d}x -\int _{ \Omega }\varrho e_{\delta }(\varrho,\vartheta )\nabla _{x}\vartheta \cdot \mathbf{u}\ \mathrm{d}x }$$
(3.75)
$$\displaystyle{+\int _{\Omega }\kappa _{\delta }(\vartheta )\vert \nabla _{x}\vartheta \vert ^{2}\ \mathrm{d}x =\int _{ \Omega }F_{1}(t,x)\vartheta \,dx,}$$

where

$$\displaystyle{F_{1} = -\frac{\partial (\varrho e_{\delta })} {\partial \varrho } (\varrho,\vartheta )\partial _{t}\varrho + \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u}}$$
$$\displaystyle{+\varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} - p(\varrho,\vartheta )\mathrm{div}_{ x}\mathbf{u} +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5} +\varrho \mathcal{Q}_{\delta }.}$$

In view of the uniform bounds already proved in (3.72), the function F 1 is bounded in \(L^{\infty }((0,T) \times \Omega )\) in terms of the data.

Similarly, multiplying (3.55) on \(\partial _{t}\mathcal{K}_{\delta }(\vartheta )\) gives rise to

$$\displaystyle{ \frac{\mathrm{d}} {\mathrm{d}t}\int _{\Omega }\frac{1} {2}\vert \nabla _{x}\mathcal{K}_{\delta }(\vartheta )\vert ^{2}\ \mathrm{d}x + \int _{ \Omega }\varrho \kappa _{\delta }(\vartheta )\frac{\partial e_{\delta }} {\partial \vartheta } (\varrho,\vartheta )\vert \partial _{t}\vartheta \vert ^{2}\ \mathrm{d}x }$$
(3.76)
$$\displaystyle{+\int _{\Omega }\varrho \frac{\partial e_{\delta }} {\partial \vartheta } (\varrho,\vartheta )\ \partial _{t}\vartheta \nabla _{x}\mathcal{K}_{\delta }(\vartheta ) \cdot \mathbf{u}\ \mathrm{d}x = \int _{\Omega }F_{2}(t,x)\partial _{t}\vartheta \ \mathrm{d}x}$$

where

$$\displaystyle{F_{2} = -\kappa _{\delta }(\vartheta )\Big(\partial _{\varrho }[\varrho e_{\delta }](\varrho,\vartheta )\partial _{t}\varrho - \partial _{\varrho }[\varrho e_{\delta }](\varrho,\vartheta )\nabla _{x}\varrho \cdot \mathbf{u}}$$
$$\displaystyle{-\varrho e_{\delta }(\varrho,\vartheta )\mathrm{div}_{x}\mathbf{u}\Big) + \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u} + \varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}}$$
$$\displaystyle{-p(\varrho,\vartheta )\mathrm{div}_{x}\mathbf{u} +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5} +\varrho \mathcal{Q}_{\delta }}$$

is bounded in \(L^{\infty }((0,T) \times \Omega )\) in terms of the data.

Taking the sum of (3.75), (3.76), and using Young’s inequality and Gronwall’s lemma, we conclude that

$$\displaystyle{\mathrm{ess}\sup _{t\in (0,T)}\|\nabla _{x}\mathcal{K}_{\delta }(\vartheta )\|_{L^{2}(\Omega;R^{3})}^{2} +\int _{ 0}^{T}\|\partial _{ t}\vartheta \|_{L^{2}(\Omega )}^{2}\ \mathrm{d}t}$$
$$\displaystyle{\leq h\Big(\|\varrho \|_{C^{1}([0,T]\times \overline{\Omega })},\|\mathbf{u}\|_{C([0,T;X_{n})},(\inf _{(0,T)\times \Omega }\varrho )^{-1},\|\vartheta _{ 0}\|_{W^{1,2}(\Omega )}\Big).}$$

Finally, evaluating \(\Delta _{x}\mathcal{K}_{\delta }(\vartheta )\) by means of Eq. (3.55), we get (3.74).

Existence for the Approximate Internal Energy Equation Having prepared the necessary material, we are ready to show existence of strong solutions to problem (3.55)–(3.57). In fact, the a priori bounds (3.72), (3.74) imply compactness of solutions in the space \(L^{2}(0,T;W^{1,2}(\Omega ))\), in particular, any accumulation point of a family of strong solutions is another solution of the same problem. Under these circumstances, showing existence is a routine matter. Regularizing the data ϱ, u with respect to the time variable, and approximating the quantities μ, η, κ δ , e, p by smooth ones as the case may be, we can construct a family of approximate solutions to problem (3.55)–(3.57) via the classical results for quasilinear parabolic equations. Then we pass to the limit in a suitable sequence of approximate solutions to recover the (unique) solution of problem (3.55)–(3.57). The relevant theory of quasilinear parabolic equations taken over from the book (Ladyzhenskaya et al. [179, Chap. V]) is summarized in Sect. 11.16 in Appendix.

Hereafter we describe a possible way of the construction of the approximations to problem (3.55)–(3.57).

  1. (i)

    Let ν ∈ (0, 1) be the same parameter as in Lemma 3.1. To begin, we extend \(\varrho \in C([0,T];C^{2,\nu }(\overline{\Omega })) \cap C^{1}([0,T];C^{0,\nu }(\overline{\Omega }))\), uC([0, T]; X n ), continuously to \(\varrho \in C(\mathbb{R};C^{2,\nu }(\overline{\Omega })) \cap C^{1}(\mathbb{R};C^{0,\nu }(\overline{\Omega }))\), \(\mathrm{supp}\varrho \subset (-2T,2T) \times \overline{\Omega }\), \(\mathbf{u} \in C(\mathbb{R},X_{n})\), \(\mathrm{supp}\mathbf{u} \subset (-2T,2T) \times \overline{\Omega }\). We approximate \(\mathcal{Q}_{\delta }\) by smooth functions \(\mathcal{Q}_{\omega }\) on \([0,T] \times \overline{\Omega }\) and we take sequence of initial conditions

    $$\displaystyle{C^{2,\nu }(\overline{\Omega }) \ni \vartheta _{ 0,\omega } \rightarrow \vartheta _{0,\delta }\;\mbox{ in}W^{1,2}(\Omega ) \cap L^{\infty }(\Omega )}$$

    such that \(\inf _{x\in \Omega }\vartheta _{0,\omega }(x)>\underline{\vartheta } _{0}\) uniformly with respect to ω → 0+, where \(\underline{\vartheta }_{0}\) is a positive constant.

  2. (ii)

    We denote

    $$\displaystyle{E_{M}(\varrho,\vartheta ) =\varrho e_{M}(\varrho,\vartheta )}$$

    and set

    $$\displaystyle{ E_{\delta,\omega }(\varrho,\vartheta ) = [<E_{M}>]^{\omega }(\varrho,\theta _{\omega }) + a\theta _{\omega }^{4}+\delta \varrho \vartheta, }$$
    (3.77)
    $$\displaystyle{\{\partial _{\vartheta }E\}_{\delta,\omega }(\varrho,\vartheta ) = [<\partial _{\vartheta }E_{M}>]^{\omega }(\varrho,\vartheta ) + 4a\, \frac{\vartheta ^{4}} {\sqrt{\vartheta ^{2 } +\omega ^{2}}}+\delta \varrho }$$
    $$\displaystyle{\kappa _{\delta,\omega }(\vartheta ) = [<\kappa _{M}>]^{\omega }(\theta _{\omega }) + [<\kappa _{R}>]^{\omega }(\theta _{\omega }) +\delta (\theta _{\omega }^{\Gamma } + \frac{1} {\sqrt{\vartheta ^{2 } +\omega ^{2}}}),}$$
    $$\displaystyle{\mathcal{K}_{\delta,\omega }(\vartheta ) =\int _{ 1}^{\vartheta }\kappa _{ \delta,\omega }(\tau )\,d\tau,}$$
    $$\displaystyle{p_{\omega }(\varrho,\vartheta ) = [<p_{M}>]^{\omega }(\varrho,\theta _{\omega }) + \frac{a} {3}\theta _{\omega }^{4},}$$
    $$\displaystyle{G(t,x) =\Big ((\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}\Big)(t,x),\;G_{\omega }(t,x) = G^{\omega }(t,x)}$$
    $$\displaystyle{\mathbb{S}_{\delta,\omega }(\vartheta,\nabla _{x}\mathbf{u}^{\omega }) =<\mu> ^{\omega }(\theta _{\omega })\Big(\nabla \mathbf{u}^{\omega } + \nabla ^{T}\mathbf{u}^{\omega } -\frac{2} {3}\mathrm{div}\mathbf{u}^{\omega }\mathbb{I}\Big)+ <\eta> ^{\omega }(\theta _{\omega })\mathrm{div}\mathbf{u}^{\omega }\mathbb{I},}$$

    where

    $$\displaystyle{ \begin{array}{c} \theta _{\omega } =\theta _{\omega }(\vartheta ) = \frac{\sqrt{\vartheta ^{2 } +\omega ^{2}}} {1+\omega \sqrt{\vartheta ^{2 } +\omega ^{2}}},\\ \\ <a> (z) = \left \{\begin{array}{c} a(z)\;\mbox{ if }z \in (0,\infty )^{N} \\ \\ \max \{\inf _{z\in (0,\infty )^{N}}a(z)\,,\,0\}\end{array} \right \},\;N = 1,2.\end{array} }$$
    (3.78)

    The operator bb ω, ω > 0 is the standard regularizing operator, see (11.4) in Sect. 11.2, that applies to all independent variables in the case of functions < E M >, < ϑ E M >, < p >, < μ >, < η >, < κ M >, and to the variable t in the case of functions ϱ(t, x), u(t, x), G(t, x). Notice that in virtue of hypotheses (3.21)–(3.23) and (3.11)

    $$\displaystyle{ \kappa _{\delta,\omega }(\vartheta ) \geq \underline{\kappa }_{M}> 0,\quad \{\partial _{\vartheta }E\}_{\delta,\omega }(\varrho,\vartheta )>\delta \underline{\varrho }> 0 }$$
    (3.79)

    for all \((\varrho,\vartheta ) \in \mathbb{R}^{2}\), where \(\underline{\varrho }=\inf _{(0,T)\times \Omega }\varrho\).

  3. (iii)

    We will find a solution of problem (3.55)–(3.57), as a limit of the sequence {ϑ ω } ω > 0 of solutions to the following equation

    $$\displaystyle{\{\partial _{\vartheta }E\}_{\delta,\omega }(\varrho ^{\omega },\vartheta )\partial _{t}\vartheta + \mathrm{div}\Big(E_{\delta,\omega }(\varrho ^{\omega },\vartheta )\mathbf{u}\Big) - \Delta _{x}\mathcal{K}_{\delta,\omega }(\vartheta )}$$
    $$\displaystyle{ = -\partial _{\varrho }E_{\delta,\omega }(\varrho ^{\omega },\vartheta )\partial _{t}\varrho ^{\omega } + \mathbb{S}_{\delta,\omega }(\nabla _{x}\mathbf{u}^{\omega },\vartheta ): \nabla \mathbf{u}^{\omega }+ }$$
    (3.80)
    $$\displaystyle{\varepsilon \delta G_{\omega } - p_{\omega }(\varrho ^{\omega },\vartheta ) - \frac{\delta } {\vartheta ^{2} +\omega ^{2}} +\varepsilon \theta _{ \omega }^{5} +\varrho ^{\omega }\mathcal{Q}_{\omega },}$$
    $$\displaystyle{\nabla _{x}\vartheta \cdot \mathbf{n}\vert _{\partial \Omega } = 0,\;\vartheta (0,x) =\vartheta _{0,\omega }(x).}$$

    Problem (3.80) for the unknown ϑ has the form of the following quasilinear parabolic equation

    $$\displaystyle{ \begin{array}{c} \partial _{t}\vartheta -\sum _{i,j=1}^{3}a_{ij}(t,x,\vartheta )\partial _{x_{i}}\partial _{x_{j}}\vartheta + b(t,x,\vartheta,\nabla _{x}\vartheta ) = 0\quad \mbox{ in }(0,T) \times \Omega, \\ \Big(\sum _{i,j=1}^{3}a_{ij}\partial _{x_{j}}\vartheta \,n_{i}+\psi \Big)\Big\vert _{(0,T)\times \partial \Omega } = 0, \\ \vartheta \vert _{\{0\}\times \Omega } = 0,\end{array} }$$
    (3.81)

    where

    $$\displaystyle{ a_{ij}(t,x,\vartheta ) = \frac{\kappa _{\delta,\omega }(\vartheta )} {[\partial _{\vartheta }E]_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )}\delta _{ij},\quad i,j = 1,2,3,\quad \psi = 0 }$$
    (3.82)

    and

    $$\displaystyle{ b(t,x,\vartheta,\mathbf{z}) = \frac{1} {\{\partial _{\vartheta }E\}_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )}\Big[-\kappa _{\delta,\omega }^{{\prime}}(\vartheta )\vert \mathbf{z}\vert ^{2}+ }$$
    (3.83)
    $$\displaystyle{\partial _{\varrho }E_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )\partial _{t}\varrho ^{\omega }(t,x) + \partial _{\varrho }E_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )\big(\nabla \varrho ^{\omega }\cdot \mathbf{u}^{\omega }\big)(t,x)+}$$
    $$\displaystyle{-\mathbb{S}_{\delta,\omega }(\nabla _{x}\mathbf{u}^{\omega }(t,x),\vartheta ): \nabla \mathbf{u}^{\omega }(t,x) + \partial _{\vartheta }E_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )\big(\mathbf{z} \cdot \mathbf{u}^{\omega }\big)(t,x)+}$$
    $$\displaystyle{E_{\delta,\omega }(\varrho ^{\omega }(t,x),\vartheta )\mathrm{div}_{x}\mathbf{u}^{\omega } + p_{\omega }(\varrho ^{\omega }(t,x),\vartheta )\mathrm{div}_{x}\mathbf{u}^{\omega }(t,x)}$$
    $$\displaystyle{-\epsilon \delta G_{\omega }(t,x) + \frac{\delta } {\vartheta ^{2} +\omega ^{2}} -\varepsilon \theta _{\omega }^{5}(\vartheta ) -\varrho ^{\omega }\mathcal{Q}_{\omega }(t,x)].}$$

    In accordance with the properties of mollifiers recalled in Sect. 11.2 in Appendix, a ij , b, ψ satisfy assumptions of Theorem 11.31 from Sect. 11.16. Therefore, problem (3.80) admits a (unique) solution ϑ = ϑ ω which belongs to class

    $$\displaystyle{\vartheta _{\omega } \in C([0,T];C^{2,\nu }(\overline{\Omega })) \cap C^{1}([0,T] \times \overline{\Omega }),\quad \partial _{ t}\vartheta _{\omega } \in C^{0,\nu /2}([0,T];C(\overline{\Omega })).}$$
  4. (iv)

    The proofs of Lemma 3.2, Corollary 3.2 and Lemma 3.3 apply with minor modifications to system (3.80), yielding the uniform bounds

    $$\displaystyle{\|\frac{1} {\vartheta _{\omega }} \|_{L^{\infty }((0,T)\times \Omega )} +\|\vartheta _{\omega }\|_{L^{\infty }((0,T)\times \Omega )} \leq c,}$$
    $$\displaystyle{\mathrm{ess}\sup _{t\in (0,T)}\|\vartheta _{\omega }\|_{W^{1,2}(\Omega )}^{2} +\int _{ 0}^{T}\Big(\|\partial _{ t}\vartheta _{\omega }\|_{L^{2}(\Omega )}^{2} +\| \Delta _{ x}\mathcal{K}_{\delta }(\vartheta _{\omega })\|_{L^{2}(\Omega )}^{2}\Big)\mathrm{d}t \leq c}$$

    with respect to ω → 0+. With these bounds and the properties of mollifiers recalled in Sect. 11.2 at hand, the limit passage from system (3.80) to (3.55)–(3.57) is an easy exercise.

The results achieved in this section can be stated as follows.

      Approximate Internal Energy Equation:

Lemma 3.4

Let \(\Omega \subset \mathbb{R}^{3}\) be a bounded domain of class C 2,ν , ν ∈ (0, 1). Let uC([0, T]; X n ) be a given vector field and let ϱ = ϱ u be the unique solution of the approximate problem ( 3.45 )–( 3.47 ) constructed in Lemma  3.1 . Further

  1. (i)

    let the initial datum \(\vartheta _{0,\delta } \in W^{1,2}(\Omega ) \cap L^{\infty }(\Omega )\) be bounded below away from zero as stated in hypothesis ( 3.58 ) and the source term \(\mathcal{Q}_{\delta }\) satisfies ( 3.60 );

  2. (ii)

    let the constitutive functions p, e, s and the transport coefficients μ, η, κ obey the structural assumptions ( 3.7 )–( 3.23 ).

Then problem ( 3.55 )–( 3.57 ), with e δ , \(\mathcal{K}_{\delta }\) defined in ( 3.59 ) and u , ϱ u fixed, possesses a unique strong solution ϑ = ϑ u belonging to the regularity class

$$\displaystyle{ Y = \left \{\begin{array}{c} \partial _{t}\vartheta \in L^{2}((0,T) \times \Omega ),\ \Delta _{x}\mathcal{K}_{\delta }(\vartheta ) \in L^{2}((0,T) \times \Omega ),\\ \\ \vartheta \in L^{\infty }(0,T;W^{1,2}(\Omega ) \cap L^{\infty }(\Omega )),\quad \frac{1} {\vartheta } \in L^{\infty }((0,T) \times \Omega ). \end{array} \right \} }$$
(3.84)

Moreover, the mapping uϑ u maps bounded sets in C([0, T]; X n ) into bounded sets in Y and is continuous with values in \(L^{2}(0,T;W^{1,2}(\Omega ))\) .

3.4.3 Local Solvability of the Approximate Problem

At this stage, we are ready to show the existence of approximate solutions on a possibly short time interval (0, T max). In accordance with (3.50), X n is a finite dimensional subspace of \(L^{2}(\Omega, \mathbb{R}^{3})\) endowed with the Hilbert structure induced by \(L^{2}(\Omega; \mathbb{R}^{3})\). We denote by P n the orthogonal projection of \(L^{2}(\Omega, \mathbb{R}^{3})\) onto X n . Furthermore, we set

$$\displaystyle{ \mathbf{u}_{0,\delta } = \frac{(\varrho \mathbf{u})_{0}} {\varrho _{0,\delta }},\quad \mathbf{u}_{0,\delta,n} = P_{n}[\mathbf{u}_{0,\delta }]. }$$
(3.85)

We start rewriting (3.49) as a fixed point problem:

$$\displaystyle{ \mathbf{u}(\tau ) = J\Big[\varrho (\tau ),\int _{0}^{\tau }M(t,\varrho (t),\vartheta (t),\mathbf{u}(t))\mathrm{d}t + (\varrho \mathbf{u})_{ 0}^{{\ast}}\Big] \equiv S[\mathbf{u}](\tau ),\ \tau \in [0,T], }$$
(3.86)

where we have denoted

$$\displaystyle{(\varrho \mathbf{u})_{0}^{{\ast}}\in X_{ n}^{{\ast}},\ <(\varrho \mathbf{u})_{ 0}^{{\ast}};\boldsymbol{\varphi }>\equiv \int _{ \Omega }(\varrho \mathbf{u})_{0}\cdot \boldsymbol{\varphi }\ \mathrm{d}x\ \mbox{ for all}\ \varphi \in X_{n},}$$
$$\displaystyle{M(t,\varrho,\vartheta,\mathbf{u}) \in X_{n}^{{\ast}},}$$
$$\displaystyle{<M(t,\varrho,\vartheta,\mathbf{u});\varphi>= \int _{\Omega }\Big(\varrho [\mathbf{u} \otimes \mathbf{u}]: \nabla _{x}\boldsymbol{\varphi } + (\,p +\delta (\varrho ^{\Gamma } + \varrho ^{2}))\mathrm{div}_{ x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x}$$
$$\displaystyle{-\int _{\Omega }\Big(\varepsilon (\nabla _{x}\varrho \nabla _{x}\mathbf{u}) \cdot \boldsymbol{\varphi } +\mathbb{S}_{\delta }: \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}_{\delta }(t)\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mbox{ for all}\ \varphi \in X_{n}^{{\ast}},}$$

and

$$\displaystyle{J[\varrho,\cdot ]: X_{n}^{{\ast}}\rightarrow X_{ n},\ \int _{\Omega }\varrho J[\varrho,\chi ]\cdot \boldsymbol{\varphi }\ \mathrm{d}x =<\chi;\boldsymbol{\varphi }>\ \mbox{ for all}\ \chi \in X_{n}^{{\ast}},\ \boldsymbol{\varphi }\in X_{ n}.}$$

Note that

$$\displaystyle{ \|\ J[\varrho,\chi ]\ \|_{X_{n}} \leq \frac{1} {A}\ \|\ \chi \ \|_{X_{n}^{{\ast}}},\quad A =\inf _{(t,x)\in (0,T)\times \Omega }\varrho (t,x) }$$
(3.87)

and

$$\displaystyle{ \|\ J[\varrho _{1},\chi ] - J[\varrho _{2},\chi ]\ \|_{X_{n}} \leq }$$
(3.88)
$$\displaystyle{ \frac{c} {A_{1}A_{2}}\|\varrho _{1} -\varrho _{2}\|_{L^{\infty }(\Omega )}\|\chi \|_{X_{n}^{{\ast}}},\quad A_{i} =\inf _{(t,x)\in (0,T)\times \Omega }\varrho _{i}(t,x),\;i = 1,2,}$$

where c > 0 depends solely on n, in particular, it is independent of the data specified in (2.41) and the parameters ɛ, δ, \(\Gamma\).

Given uC([0, T]; X n ), the density ϱ = ϱ u can be identified as the unique (classical) solution of the parabolic problem (3.45)–(3.48), the existence of which is guaranteed by Lemma 3.1. In particular, the (approximate) density ϱ u remains bounded below away from zero as soon as we can control div x u. Note that, at this level of approximation, the norm of div x u is dominated by that of u as the dimension of X n is finite.

With u, ϱ u at hand, the temperature ϑ = ϑ u can be determined as the unique solution of problem (3.55)–(3.57) constructed by means of Lemma 3.4, in particular, ϑ is strictly positive with a lower bound in terms of the data, see Corollary 3.2.

If \(\|\mathbf{u}\|_{C([0,T];X_{n})} \leq R\), then

$$\displaystyle{ \begin{array}{c} \quad \|J[\varrho (\tau ),\int _{0}^{\tau }M(t,\varrho (t),\mathbf{u}(t),\vartheta (t)\,\mathrm{d}t + (\varrho \mathbf{u})_{0}^{{\ast}}\|_{X_{n}} \leq \\ \\ c_{0}\frac{\overline{\varrho }_{0}} {\underline{\varrho }_{0}} \mathrm{exp}(2R\tau )\|\mathbf{u}_{0,\delta,n}\|_{X_{n}} +\tau h(R)\ \mbox{ for all}\ \tau \in [0,T], \end{array} }$$
(3.89)

where we have used Lemmas 3.13.4, specifically, bounds (3.62), (3.84). The constant c 0, determined in terms of equivalence of norms on X n , depends solely on n and h is a positive function bounded on bounded sets.

Consequently, if

$$\displaystyle{ R> 2c_{0}\frac{\overline{\varrho }_{0}} {\underline{\varrho }_{0}}\|\mathbf{u}_{0,\delta,n}\|_{X_{n}}, }$$
(3.90)

the operator uS[u] determined through (3.86) maps the ball

$$\displaystyle{ B_{R,\tau _{0}} =\Big\{ \mathbf{u} \in C([0,\tau _{0}],X_{n})\,\Big\vert \,\|\mathbf{u}\|_{C([0,\tau _{0}];X_{n})} \leq R,\,\mathbf{u}(0) = \mathbf{u}_{0,\delta,n}\Big\} }$$
(3.91)

into itself as soon as τ 0 is small enough.

Moreover, as a consequence of (3.88) and smoothness of ϱ, the image of \(B_{R,\tau _{0}}\) consists of uniformly Lipschitz functions on [0, τ 0], in particular, it belongs to a compact set in C([0, τ 0]; X n ). Thus a direct application of the Leray-Schauder fixed point theorem yields existence of a solution {ϱ, u, ϑ} of the approximate problem (3.45)–(3.57) defined on a (possibly short) time interval [0, T(n)]. Finally, taking advantage of Lemma 3.1, we deduce from (3.86) that

$$\displaystyle{ \mathbf{u} \in C^{1}([0,T(n)];X_{ n}). }$$
(3.92)

The above procedure can be iterated as many times as necessary to reach T(n) = T as long as there is a bound on u independent of T(n). The existence of such a bound is the main topic discussed in the next section.

3.4.4 Uniform Estimates and Global Existence

Let {ϱ, u, ϑ} be an approximate solution of problem (3.45)–(3.57) defined on a time interval [0, T max], T maxT. The last step in the proof of Proposition 3.1 is to establish a uniform (in time) bound on the norm \(\|\mathbf{u}(t)\|_{X_{n}}\) for t ∈ [0, T max] independent of T max. The existence of such a bound allows us to iterate the local construction described in the previous section in order to obtain an approximate solution defined on the full time interval [0, T]. To this end, the a priori estimates derived in Sect. 2.2 will be adapted in order to accommodate the extra terms arising at the actual level of approximation.

First of all, it follows from (3.45), (3.46) that the total mass remains constant in time, specifically,

$$\displaystyle{ \int _{\Omega }\varrho (t)\ \mathrm{d}x = \int _{\Omega }\varrho _{0,\delta }\ \mathrm{d}x = M_{0,\delta }\ \mbox{ for all}\ t \in [0,T_{\mathrm{max}}]. }$$
(3.93)

The next observation is that the quantity ψ u, with ψ = ψ(t), ψC c 1[0, T max), can be taken as a test function in the variational formulation of the momentum equation (3.49) to obtain

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x + \varepsilon \delta \int _{ 0}^{\tau }\int _{ \Omega }\vert \nabla _{x}\varrho \vert ^{2}(\Gamma \varrho ^{\Gamma -2} + 2)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.94)
$$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2}(\varrho \mathbf{u})_{0}\mathbf{u}(0) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\Big(p\mathrm{div}_{x}\mathbf{u} - \mathbb{S}_{\delta }: \nabla _{x}\mathbf{u}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\varrho \mathbf{f}_{\delta } \cdot \mathbf{u}\ \mathrm{d}x\ \mathrm{d}t,}$$

which, combined with (3.55), gives rise to the approximate energy balance

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} +\varrho e_{\delta }(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x }$$
(3.95)
$$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2}(\varrho \mathbf{u})_{0}\mathbf{u}(0) +\varrho _{0,\delta }e_{\delta }(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \mathcal{Q}_{\delta } +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5}\Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for all}\ \tau \in [0,T_{\mathrm{ max}}].}$$

Moreover, dividing the approximate internal energy equation (3.55) on ϑ, we obtain, after a straightforward manipulation, an approximate entropy production equation in the form

$$\displaystyle{ \partial _{t}(\varrho s_{\delta }(\varrho,\vartheta )) + \mathrm{div}_{x}(\varrho s_{\delta }(\varrho,\vartheta )\mathbf{u}) -\mathrm{div}_{x}\Big[\Big(\frac{\kappa (\vartheta )} {\vartheta } +\delta (\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\nabla _{x}\vartheta \Big] }$$
(3.96)
$$\displaystyle{= \frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } +\delta (\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} \Big]}$$
$$\displaystyle{+\frac{\varepsilon \delta } {\vartheta }(\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}+}$$
$$\displaystyle{\varepsilon \frac{\Delta _{x}\varrho } {\vartheta } \Big(\vartheta s_{\delta }(\varrho,\vartheta ) - e_{\delta }(\varrho,\vartheta ) -\frac{p(\varrho,\vartheta )} {\varrho } \Big) -\varepsilon \vartheta ^{4} + \frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }}$$

satisfied a.a. in \((0,T_{\mathrm{max}}) \times \Omega\), where

$$\displaystyle{ s_{\delta }(\varrho,\vartheta ) = s(\varrho,\vartheta )+\delta \log \vartheta, }$$
(3.97)

and

$$\displaystyle{ \vartheta s_{\delta }(\varrho,\vartheta ) - e_{\delta }(\varrho,\vartheta ) -\frac{p(\varrho,\vartheta )} {\varrho } =\vartheta s_{M,\delta }(\varrho,\vartheta ) - e_{M,\delta }(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho }. }$$
(3.98)

Relations (3.95), (3.96) give rise to uniform estimates similar to those obtained in Sect. 2.2.3. Indeed, multiplying (3.96) on \(\overline{\vartheta }\), where \(\overline{\vartheta }\) is a arbitrary positive constant, integrating over \(\Omega\), and subtracting the resulting expression from (3.96), we get

$$\displaystyle{\int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x}$$
$$\displaystyle{+\overline{\vartheta }\int _{0}^{\tau }\int _{ \Omega }\frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } +\delta (\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} }$$
$$\displaystyle{+\varepsilon \delta (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}\Big]\,\mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{\tau }\int _{ \Omega }\varepsilon \vartheta ^{5}\,\mathrm{d}x\mathrm{d}t}$$
$$\displaystyle{ = \int _{\Omega }\Big(\frac{1} {2}\varrho _{0,\delta }\vert \mathbf{u}_{0,\delta }\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x }$$
(3.99)
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \Big (1 -\frac{\overline{\vartheta }} {\vartheta }\Big)\mathcal{Q}_{\delta } + \frac{\delta } {\vartheta ^{2}} + \varepsilon \overline{\vartheta }\vartheta ^{4}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{-\varepsilon \overline{\vartheta }\int _{0}^{\tau }\int _{ \Omega }\frac{\Delta _{x}\varrho } {\vartheta } \Big(\vartheta s_{\delta }(\varrho,\vartheta ) - e_{\delta }(\varrho,\vartheta ) -\frac{p(\varrho,\vartheta )} {\varrho } \Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for all}\ \tau \in [0,T_{\mathrm{max}}],}$$

where \(H_{\delta,\overline{\vartheta }}\) is an analogue of the Helmholtz function introduced in (2.48), specifically,

$$\displaystyle{ H_{\delta,\overline{\vartheta }}(\varrho,\vartheta ) =\varrho e_{\delta }(\varrho,\vartheta ) -\overline{\vartheta }\varrho s_{\delta }(\varrho,\vartheta ) = H_{\overline{\vartheta }}(\varrho,\vartheta ) +\delta \varrho (\vartheta -\overline{\vartheta }\log \vartheta ). }$$
(3.100)

Here, in accordance with (3.98),

$$\displaystyle{ \int _{0}^{\tau }\int _{ \Omega }\frac{\Delta _{x}\varrho } {\vartheta } \Big(\vartheta s_{\delta }(\varrho,\vartheta ) - e_{\delta }(\varrho,\vartheta ) -\frac{p(\varrho,\vartheta )} {\varrho } \Big)\ \mathrm{d}x\ \mathrm{d}t = }$$
(3.101)
$$\displaystyle{-\int _{0}^{\tau }\int _{ \Omega }\frac{\partial } {\partial \varrho }\Big(\vartheta s_{M}(\varrho,\vartheta ) - e_{M}(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big)\frac{\vert \nabla _{x}\varrho \vert ^{2}} {\vartheta } \ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{-\int _{0}^{\tau }\int _{ \Omega }\frac{\partial } {\partial \vartheta }\Big(s_{M,\delta }(\varrho,\vartheta ) -\frac{e_{M,\delta }(\varrho,\vartheta )} {\vartheta } -\frac{p_{M}(\varrho,\vartheta )} {\varrho \vartheta } \Big)\nabla _{x}\varrho \cdot \nabla _{x}\vartheta \ \mathrm{d}x\ \mathrm{d}t,}$$

where, by virtue of Gibbs’ relation (3.7),

$$\displaystyle{ \frac{\partial } {\partial \varrho }\Big(\vartheta s_{M}(\varrho,\vartheta ) - e_{M}(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big) = -\frac{1} {\varrho } \frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta ), }$$
(3.102)
$$\displaystyle{ \frac{\partial } {\partial \vartheta }\Big(s_{M,\delta }(\varrho,\vartheta ) -\frac{e_{M,\delta }(\varrho,\vartheta )} {\vartheta } -\frac{p_{M}(\varrho,\vartheta )} {\varrho \vartheta } \Big) = \frac{1} {\vartheta ^{2}} \Big(e_{M,\delta }(\varrho,\vartheta ) +\varrho \frac{\partial e_{M}(\varrho,\vartheta )} {\partial \varrho } \Big). }$$
(3.103)

Equality (3.99) therefore transforms to

$$\displaystyle{\int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x}$$
$$\displaystyle{+\overline{\vartheta }\int _{0}^{\tau }\int _{ \Omega }\sigma _{\varepsilon,\delta }\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{\tau }\int _{ \Omega }\varepsilon \vartheta ^{5}\,\mathrm{d}x\mathrm{d}t}$$
$$\displaystyle{ = \int _{\Omega }\Big(\frac{1} {2}\varrho _{0,\delta }\vert \mathbf{u}_{0,\delta }\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x }$$
(3.104)
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \Big (1 -\frac{\overline{\vartheta }} {\vartheta }\Big)\mathcal{Q}_{\delta } + \frac{\delta } {\vartheta ^{2}} + \varepsilon \overline{\vartheta }\vartheta ^{4}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{+\varepsilon \int _{0}^{\tau }\int _{ \Omega } \frac{\overline{\vartheta }} {\vartheta ^{2}}\Big(e_{M,\delta }(\varrho,\vartheta ) +\varrho \frac{\partial e_{M}} {\partial \varrho } (\varrho,\vartheta )\Big)\nabla _{x}\varrho \nabla _{x}\vartheta \,\mathrm{d}x\mathrm{d}t\ \mbox{ for all}\ \tau \in [0,T_{\mathrm{max}}],}$$

where

$$\displaystyle{ \sigma _{\varepsilon,\delta } = \frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } +\delta (\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} }$$
(3.105)
$$\displaystyle{+\frac{\varepsilon \delta } {\vartheta }(\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} + \varepsilon \frac{\overline{\vartheta }} {\varrho \vartheta } \frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta )\vert \nabla _{x}\varrho \vert ^{2}.}$$

Similarly to Sect. 2.2.3, relation (3.104) provides all the necessary uniform estimates as soon as we check that the terms on the right hand side can be controlled by the positive quantities on the left hand side. In order to see that, observe that the term δϑ 2 on the right-hand side of (3.104) is dominated by its counterpart δϑ 3 in the entropy production term σ ɛ, δ . Analogously, the quantity \(\varepsilon \overline{\vartheta }\vartheta ^{4}\) at the right hand side is “absorbed” by the term ɛϑ 5 at the left hand side of (3.104). Finally, the term \(\varrho (1 -\frac{\overline{\vartheta }}{\vartheta })Q_{\delta }\) can be written as a sum \(\varrho (1 -\frac{\overline{\vartheta }}{\vartheta })Q_{\delta }1_{\{\vartheta \leq 1\}} +\varrho (1 -\frac{\overline{\vartheta }}{\vartheta })Q_{\delta }1_{\{\vartheta>1\}}\), where \(\int _{0}^{\tau }\int _{\Omega }\varrho (1 -\frac{\overline{\vartheta }}{\vartheta })Q_{\delta }1_{\{\vartheta \leq 1\}}\,\mathrm{d}x\,\mathrm{d}t \leq 0\), while \(\vert \int _{0}^{\tau }\int _{\Omega }\varrho (1 -\frac{\overline{\vartheta }}{\vartheta })Q_{\delta }1_{\{\vartheta>1\}}\,\mathrm{d}x\,\mathrm{d}t\vert\) is bounded by \(\overline{\varrho }T\vert \Omega \vert \|Q_{\delta }\|_{L^{\infty }((0,T)\times \Omega )}\).

Consequently, it remains to handle the quantity

$$\displaystyle{\varepsilon \int _{\Omega }\frac{1} {\vartheta ^{2}} \Big(e_{M}(\varrho,\vartheta ) +\varrho \frac{\partial e_{M}(\varrho,\vartheta )} {\partial \varrho } \Big)\nabla _{x}\varrho \cdot \nabla _{x}\vartheta \ \mathrm{d}x}$$

appearing on the right-hand side of (3.104). To this end, we first use hypothesis (3.13), together with (3.30), in order to obtain

$$\displaystyle{\Big\vert \frac{1} {\vartheta ^{2}} \Big(e_{M}(\varrho,\vartheta ) +\varrho \frac{\partial e_{M}(\varrho,\vartheta )} {\partial \varrho } \Big)\nabla _{x}\varrho \cdot \nabla _{x}\vartheta \Big\vert \leq c\Big(\frac{\varrho ^{\frac{2} {3} }+\vartheta } {\vartheta ^{2}} \Big)\vert \nabla _{x}\varrho \vert \vert \nabla _{x}\vartheta \vert,}$$

where, furthermore,

$$\displaystyle{\frac{\vert \nabla _{x}\varrho \vert \vert \nabla _{x}\vartheta \vert } {\vartheta } \leq \omega \frac{\vert \nabla _{x}\varrho \vert ^{2}} {\vartheta } + c(\omega )\frac{\vert \nabla _{x}\vartheta \vert ^{2}} {\vartheta } \ \mbox{ for any}\ \omega> 0,}$$

and, similarly,

$$\displaystyle{\frac{\varrho ^{\frac{2} {3} }\vert \nabla _{x}\varrho \vert \vert \nabla _{x}\vartheta \vert } {\vartheta ^{2}} \leq \omega \frac{\varrho ^{\frac{4} {3} }\vert \nabla _{x}\varrho \vert ^{2}} {\vartheta } + c(\omega )\frac{\vert \nabla _{x}\vartheta \vert ^{2}} {\vartheta ^{3}}.}$$

Thus we infer that

$$\displaystyle{ \varepsilon \int _{\Omega }\frac{1} {\vartheta ^{2}} \Big\vert e_{M}(\varrho,\vartheta ) +\varrho \frac{\partial e_{M}(\varrho,\vartheta )} {\partial \varrho } \Big\vert \vert \nabla _{x}\varrho \vert \vert \nabla _{x}\vartheta \vert \ \mathrm{d}x }$$
(3.106)
$$\displaystyle{\leq \frac{1} {2}\int _{\Omega }\Big[\delta \Big(\vartheta ^{\Gamma -2} + \frac{1} {\vartheta ^{3}} \Big)\vert \nabla _{x}\vartheta \vert ^{2} + \frac{\varepsilon \delta } {\vartheta }\Big(\Gamma \varrho ^{\Gamma -2} + 2\Big)\vert \nabla _{ x}\varrho \vert ^{2}\Big]\ \mathrm{d}x}$$

provided ɛ = ɛ(δ) > 0 is small enough.

Taking into account the properties of the function \(H_{\delta,\overline{\vartheta }}\) (see (2.49)–(2.50) in Sect. 2.2.3), we are ready to summarize the so far obtained estimates as follows:

$$\displaystyle{ \left \{\begin{array}{c} \mathrm{ess}\sup _{t\in (0,T_{\mathrm{max}})}\int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma -1} +\varrho ^{2})\Big)\ \mathrm{d}x \leq c, \\ \\ \int _{0}^{T_{\mathrm{max}}}\int _{\Omega }\frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u}\Big]\ \mathrm{d}x\ \mathrm{d}t \leq c\\ \\ \int _{0}^{T_{\mathrm{max}}}\int _{\Omega }\frac{1} {\vartheta } \Big(\frac{\kappa (\vartheta )} {\vartheta } +\delta (\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2}\Big)\ \mathrm{d}x\ \mathrm{d}t \leq c\\ \\ \varepsilon \int _{0}^{T_{\mathrm{max}}}\int _{\Omega }\Big(\delta \frac{1} {\vartheta ^{3}} +\vartheta ^{5}\Big)\ \mathrm{d}x\ \mathrm{d}t \leq c\\ \\ \varepsilon \delta \int _{0}^{T_{\mathrm{max}}}\int _{\Omega }\frac{1} {\vartheta } (\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c,\\ \\ \int _{0}^{T_{\mathrm{max}}}\int _{\Omega }\varepsilon \frac{\overline{\vartheta }} {\varrho \vartheta } \frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta )\vert \nabla _{x}\varrho \vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, \end{array} \right \} }$$
(3.107)

where c is a positive constant depending on the data specified in (2.41) but independent of T max, n, ɛ, and δ.

At this stage, following the line of arguments presented in Sect. 2.2.3, we can use the bounds listed in (3.107) in order to deduce uniform estimates on the approximate solutions defined on the time interval [0, T max] independent of T max. Indeed it follows from (3.107) that

$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T_{\mathrm{max}})}\|\sqrt{\varrho }\mathbf{u}\|_{L^{2}(\Omega;\mathbb{R}^{3})}^{2}+ }$$
(3.108)
$$\displaystyle{\int _{0}^{T_{\mathrm{max}} }\int _{\Omega }\frac{1} {\vartheta } \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t \leq c(\mathrm{data},\varepsilon,\delta ),}$$

in particular, by means of hypothesis (3.53) and Proposition 2.1,

$$\displaystyle{\int _{0}^{T_{\mathrm{max}} }\int _{\Omega }\Big(\vert \mathbf{u}\vert ^{2} + \vert \nabla _{ x}\mathbf{u}\vert ^{2}\Big)\ \mathrm{d}x\ \mathrm{d}t \leq c(\mathrm{data},\varepsilon,\delta ).}$$

Consequently, by virtue of (3.63), the density ϱ is bounded below away from zero uniformly on [0, T max], and we conclude

$$\displaystyle{ \sup _{[0,T_{\mathrm{max}}]}\|\mathbf{u}\|_{X_{n}} \leq c(\mathrm{data},\varepsilon,\delta ). }$$
(3.109)

As already pointed out, bound (3.109) and the local construction described in the previous section give rise to an approximate solution {ϱ, u, ϑ} defined on [0, T]. We have proved Proposition 3.1.

3.5 Faedo-Galerkin Limit

In the previous section, we constructed a family of approximate solutions to the Navier-Stokes-Fourier system satisfying (3.45)–(3.60), see Proposition 3.1. Our goal in the remaining part of this chapter is to examine successively the asymptotic limit for n, ɛ → 0, and, finally, δ → 0. The first step of this rather long procedure consists in performing the limit n.

We recall that the spaces X n introduced in Sect. 3.3.1 are formed by sufficiently smooth functions \(\boldsymbol{\varphi }\) (belonging at least to \(C^{2,\nu }(\overline{\Omega })\)) satisfying either the complete slip boundary condition (3.51) or the no-slip boundary conditions (3.52) as the case may be. Clearly, the approximate velocity field uC 1([0, T]; X n ) belongs to the same class for each fixed t ∈ [0, T]. In the remaining part of the chapter, we make an extra hypothesis that the vector space X,

$$\displaystyle{X \equiv \cup _{n=1}^{\infty }X_{ n}\ \mbox{ is dense in}\ W_{\mathbf{n}}^{1,p}(\Omega; \mathbb{R}^{3}),\ W_{ 0}^{1,p}(\Omega; \mathbb{R}^{3}),\ \mbox{ respectively,}}$$

for any 1 ≤ p < , where

$$\displaystyle{\begin{array}{c} W_{\mathbf{n}}^{1,p}(\Omega; \mathbb{R}^{3}) =\Big\{ \mathbf{v}\ \Big\vert \mathbf{v} \in L^{\,p}(\Omega; \mathbb{R}^{3}),\ \nabla _{x}\mathbf{v} \in L^{\,p}(\Omega; \mathbb{R}^{3\times 3}),\\ \\ \mathbf{v} \cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ in the sense of traces}\Big\}, \end{array} }$$
$$\displaystyle{\begin{array}{c} W_{0}^{1,p}(\Omega; \mathbb{R}^{3}) =\Big\{ \mathbf{v}\ \Big\vert \mathbf{v} \in L^{\,p}(\Omega; \mathbb{R}^{3}),\ \nabla _{x}\mathbf{v} \in L^{\,p}(\Omega; \mathbb{R}^{3\times 3}),\\ \\ \mathbf{v}\vert _{\partial \Omega } = 0\ \mbox{ in the sense of traces}\Big\}.\end{array} }$$

Such a choice of X n is possible provided \(\Omega\) belongs to the regularity class C 2,ν required by Theorem 3.1. The interested reader may consult Sect. 11.8 in Appendix for technical details.

3.5.1 Estimates Independent of the Dimension of Faedo-Galerkin Approximations

For ɛ > 0, δ > 0 fixed, let {ϱ n , u n , ϑ n } n = 1 be a sequence of approximate solutions constructed in Sect. 3.4. In accordance with (3.107), this sequence admits the following uniform estimates:

$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\int _{\Omega }\Big(\frac{1} {2}\varrho _{n}\vert \mathbf{u}_{n}\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho _{n},\vartheta _{n}) + \delta ( \frac{\varrho _{n}^{\Gamma }} {\Gamma - 1} +\varrho _{ n}^{2})\Big)(t)\ \mathrm{d}x \leq c, }$$
(3.110)
$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\Big\{\frac{1} {\vartheta _{n}}\Big[\mathbb{S}_{\delta }(\vartheta _{n},\nabla _{x}\mathbf{u}_{n}): \nabla _{x}\mathbf{u}_{n}+}$$
$$\displaystyle{ \Big(\frac{\kappa (\vartheta _{n})} {\vartheta _{n}} +\delta (\vartheta _{n}^{\Gamma -1} + \frac{1} {\vartheta _{n}^{2}})\Big)\vert \nabla _{x}\vartheta _{n}\vert ^{2}\Big] +\delta \frac{1} {\vartheta _{n}^{3}} + \varepsilon \vartheta _{n}^{5}\Big\}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.111)
$$\displaystyle{ \varepsilon \delta \int _{0}^{T}\int _{ \Omega }\frac{1} {\vartheta _{n}}(\Gamma \varrho _{n}^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho _{n}\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.112)

and

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varepsilon \frac{\overline{\vartheta }} {\varrho _{n}\vartheta _{n}} \frac{\partial p_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\vert \nabla _{x}\varrho _{n}\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.113)

where c denotes a generic constant depending only on the data specified in (2.41), in particular, c is independent of the parameters n, ɛ, and δ.

By virtue of the coercivity properties of \(H_{\delta,\overline{\vartheta }}\) established in (2.49), (2.50), the uniform bound (3.110) implies that

$$\displaystyle{ \{\varrho _{n}\}_{n=1}^{\infty }\ \mbox{ is bounded in }L^{\infty }(0,T;L^{\Gamma }(\Omega )), }$$
(3.114)

therefore we can assume

$$\displaystyle{ \varrho _{n} \rightarrow \varrho \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{\Gamma }(\Omega )). }$$
(3.115)

On the other hand, estimate (3.111), together with hypothesis (3.53) and Proposition 2.1, yield

$$\displaystyle{ \{\mathbf{u}_{n}\}_{n=1}^{\infty }\;\mbox{ bounded in }L^{2}(0,T;W^{1,2}(\Omega; \mathbb{R}^{3})), }$$
(3.116)

in particular

$$\displaystyle{ \mathbf{u}_{n} \rightarrow \mathbf{u}\ \mbox{ weakly in}\ L^{2}(0,T;W^{1,2}(\Omega; \mathbb{R}^{3})), }$$
(3.117)

at least for a suitable subsequence.

At this point it is worth noting that the limit density ϱ is still a non-negative quantity albeit not necessarily strictly positive as this important property stated in (3.63) is definitely lost in the limit passage due to the lack of suitable uniform estimates for div x u n . The fact that the class of weak solutions admits cavities (vacuum regions) seems rather embarrassing from the point of view of the model derived for non-dilute fluids, but still physically acceptable.

Convergence (3.115) can be improved to

$$\displaystyle{ \varrho _{n} \rightarrow \varrho \quad \mbox{ in }C_{\mathrm{weak}}([0,T];L^{\Gamma }(\Omega )) }$$
(3.118)

as ϱ n , u n solve Eq. (3.45). Indeed we check easily that for all \(\varphi \in C_{c}^{\infty }(\Omega )\), the functions \(t \rightarrow [\int _{\Omega }\varrho _{n}\varphi \,\mathrm{d}x](t)\) form a bounded and equi-continuous sequence in C[0, T]. Consequently, the standard Arzelà-Ascoli theorem (Theorem 1) yields

$$\displaystyle{\int _{\Omega }\varrho _{n}\varphi \,\mathrm{d}x \rightarrow \int _{\Omega }\varrho \varphi \,\mathrm{d}x\quad \mbox{ in }C[0,T]\ \mbox{ for any}\ \varphi \in C_{c}^{\infty }(\Omega ).}$$

Since ϱ n satisfy (3.114), the convergence extends easily to each \(\varphi \in L^{\Gamma '}(\Omega )\) via density.

In order to deduce uniform estimates on the approximate temperature ϑ n , we exploit the structural properties of the Helmholtz function \(H_{\overline{\vartheta }}\). Note that these follow directly from the hypothesis of thermodynamics stability and as such may be viewed as a direct consequence of natural physical principles. The following assertion will be amply used in future considerations.

     Coercivity of the Helmholtz Function:

Proposition 3.2

Let the functions p, e, and s be interrelated through Gibbs’ equation ( 1.2 ), where p and e comply with hypothesis of thermodynamic stability (1.44).

Then for any fixed \(\overline{\varrho }> 0\) , \(\overline{\vartheta }> 0\) , the Helmholtz function

$$\displaystyle{H_{\overline{\vartheta }}(\varrho,\vartheta ) =\varrho e(\varrho,\vartheta ) -\overline{\vartheta }\varrho s(\varrho,\vartheta )}$$

satisfies

$$\displaystyle{H_{\overline{\vartheta }}(\varrho,\vartheta ) \geq \frac{1} {4}\Big(\varrho e(\varrho,\vartheta ) + \overline{\vartheta }\varrho \vert s(\varrho,\vartheta )\vert \Big) -\Big\vert (\varrho -\overline{\varrho })\frac{\partial H_{2\overline{\vartheta }}} {\partial \varrho } (\overline{\varrho },2\overline{\vartheta }) + H_{2\overline{\vartheta }}(\overline{\varrho },2\overline{\vartheta })\Big\vert }$$

for all positive ϱ, ϑ.

Proof

As the result obviously holds if s(ϱ, ϑ) ≤ 0, we focus on the case s(ϱ, ϑ) > 0. It follows from (2.49), (2.50) that

$$\displaystyle{H_{2\overline{\vartheta }}(\varrho,\vartheta ) \geq (\varrho -\overline{\varrho })\frac{\partial H_{2\overline{\vartheta }}} {\partial \varrho } (\overline{\varrho },2\overline{\vartheta }) + H_{2\overline{\vartheta }}(\overline{\varrho },2\overline{\vartheta });}$$

whence

$$\displaystyle{H_{\overline{\vartheta }}(\varrho,\vartheta ) = \frac{1} {2}\varrho e(\varrho,\vartheta ) + \frac{1} {2}H_{2\overline{\vartheta }}(\varrho,\vartheta ) \geq \frac{1} {2}\varrho e(\varrho,\vartheta )}$$
$$\displaystyle{+\frac{1} {2}\Big((\varrho -\overline{\varrho })\frac{\partial H_{2\overline{\vartheta }}} {\partial \varrho } (\overline{\varrho },2\overline{\vartheta }) + H_{2\overline{\vartheta }}(\overline{\varrho },2\overline{\vartheta })\Big),}$$

and, similarly,

$$\displaystyle{H_{\overline{\vartheta }}(\varrho,\vartheta ) = \overline{\vartheta }\varrho s(\varrho,\vartheta ) + H_{2\overline{\vartheta }}(\varrho,\vartheta ) \geq \overline{\vartheta }\varrho s(\varrho,\vartheta )}$$
$$\displaystyle{+(\varrho -\overline{\varrho })\frac{\partial H_{2\overline{\vartheta }}} {\partial \varrho } (\overline{\varrho },2\overline{\vartheta }) + H_{2\overline{\vartheta }}(\overline{\varrho },2\overline{\vartheta }).}$$

Summing up the last two inequalities we obtain the desired conclusion.

On the basis of Proposition 3.2, we can deduce from hypothesis (3.9) and the total energy estimate (3.110) that

$$\displaystyle{ \{\vartheta _{n}\}_{n=1}^{\infty }\;\mbox{ is bounded in }L^{\infty }(0,T;L^{4}(\Omega )), }$$
(3.119)

therefore we may assume

$$\displaystyle{ \vartheta _{n} \rightarrow \vartheta \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{4}(\Omega )). }$$
(3.120)

In addition, using boundedness of the entropy production rate stated in (3.111) we get

$$\displaystyle{ \{\nabla _{x}\vartheta _{n}^{\frac{\Gamma } {2} }\}_{n=1}^{\infty },\ \Big\{\nabla _{x}\Big( \frac{1} {\sqrt{\vartheta _{n}}}\Big)\Big\}_{n=1}^{\infty }\ \mbox{ bounded in}\ L^{2}(0,T;L^{2}(\Omega; \mathbb{R}^{3})). }$$
(3.121)

Estimates (3.119), (3.121), together with Poincare’s inequality formulated in terms of Proposition 2.2, yield

$$\displaystyle{ \{\vartheta _{n}\}_{n=1}^{\infty },\ \{\vartheta _{ n}^{\frac{\Gamma } {2} }\}_{n+1}^{\infty }\ \mbox{ bounded in }L^{2}(0,T;W^{1,2}(\Omega )), }$$
(3.122)

in particular,

$$\displaystyle{ \vartheta _{n} \rightarrow \vartheta \ \mbox{ weakly in}\ L^{2}(0,T;W^{1,2}(\Omega )). }$$
(3.123)

Moreover, by virtue of estimate (3.111), we have

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega } \frac{1} {\vartheta _{n}^{3}}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.124)

notably the limit function ϑ is positive almost everywhere in \((0,T) \times \Omega\) and satisfies

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\frac{1} {\vartheta ^{3}} \ \mathrm{d}x\ \mathrm{d}t \leq \liminf _{n\rightarrow \infty }\int _{0}^{T}\int _{ \Omega } \frac{1} {\vartheta _{n}^{3}}\ \mathrm{d}x\ \mathrm{d}t, }$$
(3.125)

where we have used convexity of the function zz −3 on (0, ), see Theorem 11.27 in Appendix.

Finally, the standard embedding relation \(W^{1,2}(\Omega )\hookrightarrow L^{6}(\Omega )\), together with (3.121), can be used in order to derive higher integrability estimates of ϑ n , namely

$$\displaystyle{ \{\vartheta _{n}\}_{n=1}^{\infty }\ \mbox{ bounded in}\ L^{\Gamma }(0,T;L^{3\Gamma }(\Omega )). }$$
(3.126)

Note that, as a byproduct of (3.125), (3.126),

$$\displaystyle{ \{\log (\vartheta _{n})\}_{n=1}^{\infty }\ \mbox{ is bounded in}\ L^{q}((0,T) \times \Omega )\ \mbox{ for any finite}\ q \geq 1. }$$
(3.127)

3.5.2 Limit Passage in the Approximate Continuity Equation

At this stage, we are ready to show strong (pointwise) convergence of the approximate densities and to let n in equation (3.45). To this end, we need to control the term pdiv x u in the approximate energy balance (3.94).

A direct application of (3.32) yields

$$\displaystyle{\Big\vert \int _{0}^{T}\int _{ \Omega }p(\varrho _{n},\vartheta _{n})\mathrm{div}_{x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t\Big\vert \leq c\int _{0}^{T}\int _{ \Omega }(\varrho _{n}^{\frac{5} {3} } +\vartheta _{ n}^{\frac{5} {2} } +\vartheta _{ n}^{4})\vert \mathrm{div}_{x}\mathbf{u}_{n}\vert \ \mathrm{d}x\ \mathrm{d}t,}$$

where, by virtue of (3.114), (3.116), (3.119), and (3.126), the last integral is bounded provided \(\Gamma> 5\). Accordingly, relation (3.94) gives rise to

$$\displaystyle{ \varepsilon \delta \int _{0}^{T}\int _{ \Omega }(\Gamma \varrho _{n}^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho _{n}\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.128)

with c independent of n. Applying the Poincaré inequality (see Proposition 2.2) we get

$$\displaystyle{ \{\varrho _{n}\}_{n=1}^{\infty },\ \{\varrho _{ n}^{\frac{\Gamma } {2} }\}_{n=1}^{\infty }\ \mbox{ bounded in}\ L^{2}(0,T;W^{1,2}(\Omega )), }$$
(3.129)

and

$$\displaystyle{ \{\varrho _{n}\}_{n=1}^{\infty }\ \mbox{ bounded in}\ L^{\Gamma }(0,T;L^{3\Gamma }(\Omega )). }$$
(3.130)

The next step is to obtain uniform estimates on t ϱ n , \(\Delta \varrho _{n}\). This is a delicate task as

$$\displaystyle{(\partial _{t} -\varepsilon \Delta )[\varrho _{n}] = -\nabla _{x}\varrho _{n} \cdot \mathbf{u}_{n} -\varrho _{n}\mathrm{div}_{x}\mathbf{u}_{n},}$$

where, in accordance with (3.116), (3.129), ∇ x ϱ n ⋅ u n is bounded in \(L^{1}(0,T;L^{\frac{3} {2} }(\Omega ))\), notably this quantity is merely integrable with respect to time. To overcome this difficulty, multiply Eq. (3.45) on G′(ϱ n ) and integrate by parts to obtain

$$\displaystyle{ \partial _{t}\int _{\Omega }G(\varrho _{n})\ \mathrm{d}x + \varepsilon \int _{\Omega }G''(\varrho _{n})\vert \nabla _{x}\varrho _{n}\vert ^{2}\ \mathrm{d}x = \int _{ \Omega }\Big(G(\varrho _{n}) - G'(\varrho _{n})\varrho _{n}\Big)\mathrm{div}_{x}\mathbf{u}_{n}\ \mathrm{d}x. }$$
(3.131)

This is of course nothing other than an integrated “parabolic” version of the renormalized continuity equation (2.2). Taking G(ϱ n ) = ϱ n log(ϱ n ) we easily deduce

$$\displaystyle{ \varepsilon \int _{0}^{T}\int _{ \Omega }\frac{\vert \nabla _{x}\varrho _{n}\vert ^{2}} {\varrho _{n}} \ \mathrm{d}x\ \mathrm{d}t \leq c. }$$
(3.132)

As a consequence of (3.110), the kinetic energy is bounded, specifically,

$$\displaystyle{ \mathrm{ess}\ \sup \nolimits _{t\in (0,T)}\int _{\Omega }\varrho _{n}\vert \mathbf{u}_{n}\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c; }$$
(3.133)

whence estimate (3.132) can be used to obtain

$$\displaystyle{\|\nabla _{x}\varrho _{n} \cdot \mathbf{u}_{n}\|_{L^{1}(\Omega )} \leq \Big\| \frac{\nabla _{x}\varrho _{n}} {\sqrt{\varrho _{n}}} \Big\|_{L^{2}(\Omega;\mathbb{R}^{3})}\|\sqrt{\varrho _{n}}\mathbf{u}_{n}\|_{L^{2}(\Omega;\mathbb{R}^{3})},}$$

where the product on the right-hand side is bounded in L 2(0, T). Then a standard interpolation argument implies

$$\displaystyle{ \left \{\begin{array}{c} \{\nabla _{x}\varrho _{n} \cdot \mathbf{u}_{n}\}_{n=1}^{\infty }\ \mbox{ bounded in}\ L^{q}(0,T;L^{\,p}(\Omega ))\\ \\ \mbox{ for any}\ p \in (1, \frac{3} {2}),\ \ \mbox{ where}\ q = q(\,p) \in (1,2).\end{array} \right \} }$$
(3.134)

Applying the L pL q theory to the parabolic equation (3.45) (see Sect. 11.15 in Appendix) we conclude that

$$\displaystyle{ \begin{array}{c} \{\partial _{t}\varrho _{n}\}_{n=1}^{\infty },\ \{\partial _{x_{i}}\partial _{x_{j}}\varrho _{n}\}_{n=1}^{\infty },\ i,j = 1,\ldots,3\ \mbox{ are bounded in}\ L^{q}(0,T;L^{\,p}(\Omega ))\\ \\ \mbox{ for any}\ p \in (1, \frac{3} {2}),\ \ \mbox{ where}\ q = q(\,p) \in (1,2). \end{array} }$$
(3.135)

Now we are ready to carry out the limit passage n in the approximate continuity Eq. (3.45). To begin, the uniform bounds established (3.135), together with the standard compactness embedding relations for Sobolev spaces, imply

$$\displaystyle{ \varrho _{n} \rightarrow \varrho \ \mbox{ a.a. in}\ (0,T) \times \Omega. }$$
(3.136)

Moreover, in view of (3.99), (3.117), (3.134), (3.135), and (3.136), it is easy to let n in the approximate continuity Eq. (3.45) to obtain

$$\displaystyle{ \partial _{t}\varrho + \mathrm{div}_{x}(\varrho \mathbf{u}) = \varepsilon \Delta \varrho \ \mbox{ a.a. in}\ (0,T) \times \Omega, }$$
(3.137)

where ϱ is a non-negative function satisfying

$$\displaystyle{ \nabla _{x}\varrho (t,\cdot ) \cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ for a.a.}\ t \in (0,T)\ \mbox{ in the sense of traces}, }$$
(3.138)

together with the initial condition

$$\displaystyle{ \varrho (0,\cdot ) =\varrho _{0,\delta }, }$$
(3.139)

where ϱ 0,δ has been specified in (3.48).

Our next goal is to show strong convergence of the gradients ∇ x ϱ n . To this end, we use the “renormalized” identity (3.131) with G(z) = z 2, together with the pointwise convergence established in (3.136), to deduce

$$\displaystyle{\int _{\Omega }\varrho _{n}^{2}(\tau )\ \mathrm{d}x + 2\varepsilon \int _{ 0}^{\tau }\int _{ \Omega }\vert \nabla _{x}\varrho _{n}\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \rightarrow \int _{ \Omega }\varrho _{0,\delta }^{2}\ \mathrm{d}x -\int _{ 0}^{\tau }\int _{ \Omega }\varrho ^{2}\mathrm{div}_{ x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t}$$

for any 0 < τT. On the other hand, multiplying Eq. (3.137) on ϱ and integrating by parts yields

$$\displaystyle{\int _{\Omega }\varrho _{0,\delta }^{2}\ \mathrm{d}x -\int _{ 0}^{\tau }\int _{ \Omega }\varrho ^{2}\mathrm{div}_{ x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t = \int _{\Omega }\varrho ^{2}(\tau )\ \mathrm{d}x + 2\varepsilon \int _{ 0}^{\tau }\int _{ \Omega }\vert \nabla _{x}\varrho \vert ^{2}\ \mathrm{d}x\ \mathrm{d}t;}$$

whence

$$\displaystyle{ \nabla _{x}\varrho _{n} \rightarrow \nabla _{x}\varrho \ \mbox{ (strongly) in}\ L^{2}(0,T;L^{2}(\Omega; \mathbb{R}^{3})). }$$
(3.140)

3.5.3 Strong Convergence of the Approximate Temperatures and the Limit in the Entropy Equation

Strong Convergence of the Approximate Temperatures The next step is to perform the limit in the approximate entropy balance (3.96). Here the main problem is to show strong (pointwise) convergence of the temperature. Indeed all estimates on {ϑ n } n = 1 established above concern only the spatial derivatives leaving open the question of possible time oscillations. Probably the most elegant way to overcome this difficulty is based on the celebrated Div-Curl lemma discovered by Tartar [254].

     Div-Curl Lemma:

Proposition 3.3

Let \(Q \subset \mathbb{R}^{N}\) be an open set. Assume

$$\displaystyle{\begin{array}{c} \mathbf{U}_{n} \rightarrow \mathbf{U}\ \mathit{\mbox{ weakly in}}\ L^{\,p}(Q; \mathbb{R}^{N}),\\ \\ \mathbf{V}_{n} \rightarrow \mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{q}(Q; \mathbb{R}^{N}), \end{array} }$$

where

$$\displaystyle{\frac{1} {p} + \frac{1} {q} = \frac{1} {r} <1.}$$

In addition, let

$$\displaystyle{\left.\begin{array}{r} \mathrm{div}\ \mathbf{U}_{n} \equiv \nabla \cdot \mathbf{U}_{n},\\ \\ \mathbf{curl}\ \mathbf{V}_{n} \equiv (\nabla \mathbf{V}_{n} -\nabla ^{T}\mathbf{V}_{n}) \end{array} \right \}\mathit{\mbox{ be precompact in}}\left \{\begin{array}{l} W^{-1,s}(Q), \\ W^{-1,s}(Q, \mathbb{R}^{N\times N}), \end{array} \right.}$$

for a certain s > 1.

Then

$$\displaystyle{\mathbf{U}_{n} \cdot \mathbf{V}_{n} \rightarrow \mathbf{U} \cdot \mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{r}(Q).}$$

Proposition 3.3 is proved in Sect. 11.14 in Appendix for reader’s convenience.

The basic idea is to apply Proposition 3.3 to the pair of functions

$$\displaystyle{ \begin{array}{c} \mathbf{U}_{n} = [\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n}),\mathbf{r}_{n}^{(1)}],\\ \\ \mathbf{V}_{n} = [\vartheta _{n},0,0,0],\end{array} }$$
(3.141)

defined on the set \(Q = (0,T) \times \Omega \subset \mathbb{R}^{4}\), where the term r n (1), together with the necessary piece of information concerning div t, x U n , are provided by Eq. (3.96).

To see this, we observe first that the only problematic term on the right-hand side of (3.96) can be handled as

$$\displaystyle{ \frac{\Delta _{x}\varrho _{n}} {\vartheta _{n}} \Big(\vartheta _{n}s_{\delta }(\varrho _{n},\vartheta _{n}) - e_{\delta }(\varrho _{n},\vartheta _{n}) -\frac{p(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big) }$$
(3.142)
$$\displaystyle{= \mathrm{div}_{x}\Big[\Big(\vartheta _{n}s_{M,\delta }(\varrho _{n},\vartheta _{n}) - e_{M,\delta }(\varrho _{n},\vartheta _{n}) -\frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big)\frac{\nabla _{x}\varrho _{n}} {\vartheta _{n}} \Big]}$$
$$\displaystyle{+\frac{\partial p_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\frac{\vert \nabla _{x}\varrho _{n}\vert ^{2}} {\varrho _{n}\vartheta _{n}} -\Big (e_{M,\delta }(\varrho _{n},\vartheta _{n}) +\varrho _{n}\frac{\partial e_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\Big)\frac{\nabla _{x}\varrho _{n} \cdot \nabla _{x}\vartheta _{n}} {\vartheta _{n}^{2}} }$$

(cf. (3.101)–(3.103)). Indeed, in accordance with the uniform estimates (3.106), (3.111), the approximate entropy balance equation (3.96) can be now written in the form

$$\displaystyle{ \partial _{t}(\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})) + \mathrm{div}_{x}(\mathbf{r}_{n}^{(1)}) = r_{ n}^{(2)} + r_{ n}^{(3)}, }$$
(3.143)

where

$$\displaystyle{\mathbf{r}_{n}^{(1)} =\varrho _{ n}s_{\delta }(\varrho _{n},\vartheta _{n})\mathbf{u}_{n} -\frac{\kappa _{\delta }(\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\vartheta _{n}}$$
$$\displaystyle{-\varepsilon \Big(\vartheta _{n}s_{M,\delta }(\varrho _{n},\vartheta _{n}) - e_{M,\delta }(\varrho _{n},\vartheta _{n}) -\frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big)\frac{\nabla _{x}\varrho _{n}} {\vartheta _{n}},}$$
$$\displaystyle{r_{n}^{(2)} = \frac{1} {\vartheta _{n}}\Big[\mathbb{S}_{\delta }(\vartheta _{n},\nabla _{x}\mathbf{u}_{n}): \nabla _{x}\mathbf{u}_{n} +\Big (\frac{\kappa (\vartheta _{n})} {\vartheta _{n}} +\delta (\vartheta _{n}^{\Gamma -1} + \frac{1} {\vartheta _{n}^{2}})\Big)\vert \nabla _{x}\vartheta _{n}\vert ^{2} +\delta \frac{1} {\vartheta _{n}^{2}}\Big]+}$$
$$\displaystyle{ \frac{\varepsilon \delta } {\vartheta _{n}}(\Gamma \varrho _{n}^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho _{n}\vert ^{2} + \varepsilon \frac{1} {\varrho _{n}\vartheta _{n}} \frac{\partial p_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\vert \nabla _{x}\varrho _{n}\vert ^{2} \geq 0,}$$

and

$$\displaystyle{r_{n}^{(3)} = -\varepsilon \Big(e_{ M,\delta }(\varrho _{n},\vartheta _{n}) +\varrho _{n}\frac{\partial e_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\Big)\frac{\nabla _{x}\varrho _{n} \cdot \nabla _{x}\vartheta _{n}} {\vartheta _{n}^{2}} -\varepsilon \vartheta _{n}^{4} + \frac{\varrho _{n}} {\vartheta _{n}}\mathcal{Q}_{\delta }.}$$

Hence, by virtue of the uniform estimates (3.106), (3.111)–(3.113), and (3.119),

$$\displaystyle{\mathrm{div}_{t,x}\mathbf{U}_{n} = r_{n}^{(2)} + r_{ n}^{(3)}}$$

is bounded in \(L^{1}((0,T) \times \Omega )\), therefore precompact in \(W^{-1,s}((0,T) \times \Omega )\) provided \(s \in [1, \frac{4} {3})\) (cf. Sect. 7). On the other hand, due to (3.116), curl t, x V n is obviously bounded in \(L^{2}((0,T) \times \Omega; \mathbb{R}^{4})\) which is compactly embedded into \(W^{-1,2}((0,T) \times \Omega; \mathbb{R}^{4})\). Let us remark that the “space-time” operator curl t, x applied to the vector field [ϑ n , 0, 0, 0] involves only the partial derivatives in the spatial variable x.

Consequently, in order to apply Proposition 3.3 in the situation described in (3.141), we have to show that ϱ n s(ϱ n , ϑ n ) and r n (1) are bounded in a Lebesgue space “better” than only L 1.

To this end, write

$$\displaystyle{\varrho s_{\delta }(\varrho,\vartheta ) = \frac{4} {3}a\vartheta ^{3} +\varrho s_{ M}(\varrho,\vartheta ) +\delta \varrho \log (\vartheta ),}$$

where ϱ n s M (ϱ n , ϑ n ) satisfies (3.39), therefore

$$\displaystyle{\varrho _{n}\vert s_{\delta }(\varrho _{n},\vartheta _{n})\vert \leq c(\varrho _{n} +\vartheta _{ n}^{3} +\varrho _{ n}\vert \log \varrho _{n}\vert +\varrho _{n}\vert \log \vartheta _{n}\vert ).}$$

Consequently, thanks to estimates (3.127), (3.129),

$$\displaystyle{ \begin{array}{c} \{\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})\}_{n=1}^{\infty }\ \mbox{ is bounded in }L^{\frac{\Gamma } {3} }((0,T) \times \Omega ),\\ \\ \{\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})\mathbf{u}_{n}\}_{n=1}^{\infty }\ \mbox{ is bounded in }L^{p}((0,T) \times \Omega ), \frac{1} {p} = \frac{1} {2} + \frac{3} {\Gamma }\mbox{ provided }\Gamma> 6.\end{array} }$$
(3.144)

Next we observe that (3.111) implies in the way explained in (2.58) that

$$\displaystyle{\{\nabla \log (\vartheta _{n})\}_{n=1}^{\infty }\ \mbox{ is bounded in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3}).}$$

Furthermore, it follows from (3.111) that

$$\displaystyle{\left \{\frac{\sqrt{\kappa _{\delta }(\vartheta _{n } )}} {\vartheta _{n}} \nabla _{x}\vartheta _{n}\right \}_{n=1}^{\infty }\ \mbox{ is bounded in}\ L^{2}((0,T) \times \Omega; \mathbb{R}^{3}).}$$

Moreover, estimates (3.124), (3.126) and (3.119) combined with a simple interpolation yield

$$\displaystyle{\{\sqrt{\kappa _{\delta }(\vartheta _{n } )}\}_{n=1}^{\infty }\ \mbox{ is bounded in}\ L^{\,p}((0,T) \times \Omega )\ \mbox{ for a certain}\ p> 2,}$$

on condition that \(\Gamma> 6\). From the last two estimates, we deduce that

$$\displaystyle{ \Big\{\frac{\kappa _{\delta }(\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\vartheta _{n}\Big\}_{n=1}^{\infty }\ \mbox{ is bounded in}\ L^{\,p}((0,T) \times \Omega;R^{3})\ \mbox{ for a certain}\ p> 1. }$$
(3.145)

Finally, the ɛ-dependent quantity contained in r n (1) can be handled in the following way:

  • Similarly to the proof of formula (3.144), we conclude, by help of estimates (3.126), (3.127), (3.132), that

    $$\displaystyle{ \{s_{\delta }(\varrho _{n},\vartheta _{n})\nabla \varrho _{n}\}_{n=1}^{\infty }\ \mbox{ is bounded in }L^{ \frac{2\Gamma } {\Gamma +6} }((0,T) \times \Omega ) }$$
    (3.146)

    provided \(\Gamma> 6\).

  • Since the specific internal energy e M satisfies (3.30), we have

    $$\displaystyle{\Big\vert \frac{e_{M}(\varrho _{n},\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\varrho _{n}\Big\vert \leq c(1 + \frac{\varrho _{n}^{\frac{2} {3} }} {\vartheta _{n}} )\vert \nabla _{x}\varrho _{n}\vert;}$$

    whence, in accordance with estimates (3.114), (3.124), and (3.129),

    $$\displaystyle{ \Big\{\frac{e_{M}(\varrho _{n},\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\varrho _{n}\Big\}_{n=1}^{\infty }\ \mbox{ is bounded in}\ L^{ \frac{6\Gamma } {5\Gamma +4} }((0,T) \times \Omega; \mathbb{R}^{3}). }$$
    (3.147)
  • By virtue of (3.31) and (3.32),

    $$\displaystyle{ \Big\vert \frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}\vartheta _{n}} \nabla _{x}\varrho _{n}\Big\vert \leq c\vert \nabla _{x}\varrho _{n}\vert \Big(1 + \frac{\varrho _{n}^{\frac{2} {3} }} {\vartheta _{n}} \Big), }$$
    (3.148)

    where the right hand side can be controlled exactly as in (3.147).

Having verified the hypotheses of Proposition 3.3 for the vector fields U n , V n specified in (3.141), we are allowed to conclude that

$$\displaystyle{ \overline{\varrho s_{\delta }(\varrho,\vartheta )\vartheta } = \overline{\varrho s_{\delta }(\varrho,\vartheta )}\,\vartheta }$$
(3.149)

provided \(\Gamma> 6\). In formula (3.149) and hereafter, the symbol \(\overline{F(\mathbf{U})}\) denotes a weak L 1-limit of the sequence of composed functions {F(U n )} n = 1 (cf. Sect. 8).

Since the entropy is an increasing function of the absolute temperature, relation (3.149) can be used to deduce strong (pointwise) convergence of the sequence {ϑ n } n = 1 .

To begin, we recall (3.97), namely

$$\displaystyle{\varrho s_{\delta }(\varrho,\vartheta ) = \varrho s_{M}(\varrho,\vartheta ) +\delta \varrho \log (\vartheta ) + \frac{4} {3}a\vartheta ^{3}.}$$

As all three components of the entropy are increasing in ϑ, we observe that

$$\displaystyle{ \overline{\varrho s_{M}(\varrho,\vartheta )\vartheta } \geq \overline{\varrho s_{M}(\varrho,\vartheta )}\vartheta,\ \overline{\varrho \log (\vartheta )\vartheta } \geq \overline{\varrho \log (\vartheta )}\vartheta,\ \mbox{ and}\ \overline{\vartheta ^{4}} \geq \overline{\vartheta ^{3}}\vartheta. }$$
(3.150)

Indeed, as {ϱ n } n = 1 converges strongly (see (3.136)) we have

$$\displaystyle{\overline{\varrho s_{M}(\varrho,\vartheta )\vartheta } = \varrho \overline{s_{M}(\varrho,\vartheta )\vartheta },\ \overline{\varrho s_{M}(\varrho,\vartheta )} = \varrho \overline{s_{M}(\varrho,\vartheta )},}$$

where, as a direct consequence of monotonicity of s M in ϑ,

$$\displaystyle{\overline{s_{M}(\varrho,\vartheta )\vartheta } \geq \overline{s_{M}(\varrho,\vartheta )}\vartheta,}$$

see Theorem 11.26 in Appendix. Here, we have used (3.123), (3.136) yielding

$$\displaystyle{s_{M}(\varrho _{n},\vartheta )(\vartheta _{n} -\vartheta ) \rightarrow 0\ \mbox{ weakly in}\ L^{1}((0,T) \times \Omega ).}$$

The remaining two inequalities in (3.150) can be shown in a similar way.

Combining (3.149), (3.150) we infer that

$$\displaystyle{\overline{\vartheta ^{4}} = \overline{\vartheta ^{3}}\vartheta,}$$

in particular, at least for a suitable subsequence, we have

$$\displaystyle{ \vartheta _{n} \rightarrow \vartheta \quad \mbox{ a.e. in }(0,T) \times \Omega ) }$$
(3.151)

(cf. Theorems 11.2611.27 in Appendix).

Limit in the Approximate Entropy Equation Our ultimate goal in this section is to let n in the approximate entropy Eq. (3.143).

First of all, we estimate the term

$$\displaystyle{\varepsilon \Big(e_{M,\delta }(\varrho _{n},\vartheta _{n}) +\varrho _{n}\frac{\partial e_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\Big)\frac{\nabla _{x}\varrho _{n} \cdot \nabla _{x}\vartheta _{n}} {\vartheta _{n}^{2}} }$$

in the same way as in (3.106) transforming (3.143) to inequality

$$\displaystyle{ \partial _{t}(\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})) + \mathrm{div}_{x}\Big(\varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})\mathbf{u}_{n} -\frac{\kappa _{\delta }(\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\vartheta _{n}\Big) }$$
(3.152)
$$\displaystyle{-\varepsilon \mathrm{div}_{x}\Big[\Big(\vartheta _{n}s_{M,\delta }(\varrho _{n},\vartheta _{n}) - e_{M,\delta }(\varrho _{n},\vartheta _{n}) -\frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big)\frac{\nabla _{x}\varrho _{n}} {\vartheta _{n}} \Big]}$$
$$\displaystyle{\geq \frac{1} {\vartheta _{n}}\Big[\mathbb{S}_{\delta }(\vartheta _{n},\nabla _{x}\mathbf{u}_{n}): \nabla _{x}\mathbf{u}_{n} +\Big (\frac{\kappa (\vartheta _{n})} {\vartheta _{n}} + \frac{\delta } {2}(\vartheta _{n}^{\Gamma -1} + \frac{1} {\vartheta _{n}^{2}})\Big)\vert \nabla _{x}\vartheta _{n}\vert ^{2} +\delta \frac{1} {\vartheta _{n}^{2}}\Big]}$$
$$\displaystyle{+ \frac{\varepsilon \delta } {2\vartheta _{n}}(\Gamma \varrho _{n}^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho _{n}\vert ^{2} + \varepsilon \frac{1} {\varrho _{n}\vartheta _{n}} \frac{\partial p_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})\vert \nabla _{x}\varrho _{n}\vert ^{2} -\varepsilon \vartheta _{ n}^{4} + \frac{\varrho _{n}} {\vartheta _{n}}\mathcal{Q}_{\delta }.}$$

As a consequence of (3.136), (3.144), (3.151),

$$\displaystyle{ \varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n}) \rightarrow \varrho s_{\delta }(\varrho,\vartheta )\;\mbox{ (strongly) in }L^{2}((0,T) \times \Omega ), }$$
(3.153)

and, in accordance with (3.116),

$$\displaystyle{ \varrho _{n}s_{\delta }(\varrho _{n},\vartheta _{n})\mathbf{u}_{n} \rightarrow \varrho s_{\delta }(\varrho,\vartheta )\mathbf{u}\;\mbox{ weakly in }L^{1}((0,T) \times \Omega; \mathbb{R}^{3}). }$$
(3.154)

Since the sequence {ϑ n } n = 1 converges a.a. in \((0,T) \times \Omega\), we can use hypotheses (3.21), (3.22), together with estimates (3.119), (3.122), (3.124), (3.126), to get

$$\displaystyle{\frac{\kappa (\vartheta _{n})} {\vartheta _{n}} \rightarrow \frac{\kappa (\vartheta )} {\vartheta } \quad \mbox{ (strongly) in }L^{2}((0,T) \times \Omega )}$$

yielding, in combination with (3.123),

$$\displaystyle{ \frac{\kappa (\vartheta _{n})} {\vartheta _{n}} \nabla _{x}\vartheta _{n} \rightarrow \frac{\kappa (\vartheta )} {\vartheta } \nabla _{x}\vartheta \quad \mbox{ weakly in }L^{1}((0,T) \times \Omega; \mathbb{R}^{3}). }$$
(3.155)

On the other hand, by virtue of relations (3.121), (3.124), (3.126),

$$\displaystyle{ \Big(\vartheta _{n}^{\Gamma -1} + \frac{1} {\vartheta _{n}^{2}}\Big)\nabla _{x}\vartheta _{n} = \frac{1} {\Gamma }\nabla _{x}(\vartheta _{n}^{\Gamma }) -\nabla _{ x}(1/\vartheta _{n}) \rightarrow }$$
(3.156)
$$\displaystyle{ \frac{1} {\Gamma }\nabla _{x}(\overline{\vartheta ^{\Gamma }}) -\nabla _{x}\overline{1/\vartheta }\;\mbox{ weakly in }L^{\,p}((0,T) \times \Omega )\mbox{ for some }p> 1,}$$

where, according to (3.151),

$$\displaystyle{ \frac{1} {\Gamma }\nabla _{x}(\overline{\vartheta ^{\Gamma }}) -\nabla _{x}\overline{1/\vartheta } = \frac{1} {\Gamma }\nabla _{x}(\vartheta ^{\Gamma }) -\nabla _{ x}1/\vartheta =\vartheta ^{\Gamma -1}\nabla _{ x}\vartheta + \frac{1} {\vartheta ^{2}} \nabla _{x}\vartheta. }$$
(3.157)

In order to control the ɛ-term on the left hand side of (3.152), we first observe that

$$\displaystyle{\Big\vert \frac{1} {\vartheta } \Big(\vartheta s_{M,\delta }(\varrho,\vartheta ) - e_{M,\delta }(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big)\nabla \varrho \Big\vert \leq c(\vert \log \vartheta \vert + \vert \log \varrho \vert + \frac{\varrho ^{2/3}} {\vartheta } + 1)\vert \nabla \varrho \vert,}$$

where we have used (3.31), (3.32), (3.39).

As a next step, we apply relations (3.122), (3.129), and (3.132), together with the arguments leading to (3.147), in order to deduce boundedness of the quantity

$$\displaystyle{\frac{1} {\vartheta _{n}}\Big(\vartheta s_{M,\delta }(\varrho _{n},\vartheta _{n}) - e_{M,\delta }(\varrho _{n},\vartheta _{n}) -\frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big)\nabla \varrho _{n}}$$
$$\displaystyle{\mbox{ in }L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3})\mbox{ for some }p> 1.}$$

In particular, by virtue of (3.136), (3.140), (3.151), we obtain

$$\displaystyle{ \frac{1} {\vartheta _{n}}\Big(\vartheta _{n}s_{M,\delta }(\varrho _{n},\vartheta _{n}) - e_{M,\delta }(\varrho _{n},\vartheta _{n}) -\frac{p_{M}(\varrho _{n},\vartheta _{n})} {\varrho _{n}} \Big)\nabla \varrho _{n} \rightarrow }$$
(3.158)
$$\displaystyle{\frac{1} {\vartheta } \Big(\vartheta s_{M,\delta }(\varrho,\vartheta ) - e_{M,\delta }(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big)\nabla \varrho \quad \mbox{ weakly in }L^{1}((0,T) \times \Omega; \mathbb{R}^{3}).}$$

Finally, we identify the asymptotic limit for n of the approximate entropy production rate represented through the quantities on the right-hand side of (3.152). In accordance with (3.111), we have

$$\displaystyle{\left \{\sqrt{(\Big(\frac{\mu (\vartheta _{n } )} {\vartheta _{n}} +\delta \Big)}\Big(\nabla _{x}\mathbf{u}_{n} + \nabla _{x}^{T}\mathbf{u}_{ n} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}_{n}\Big)\right \}_{n=1}^{\infty },\ \left \{\sqrt{\frac{\eta (\vartheta _{n } )} {\vartheta _{n}}} \mathrm{div}_{x}\mathbf{u}_{n}\right \}_{n=1}^{\infty }}$$

bounded in \(L^{2}((0,T) \times \Omega; \mathbb{R}^{3\times 3})\), and in \(L^{2}((0,T) \times \Omega )\), respectively. In particular,

$$\displaystyle{ \begin{array}{c} \sqrt{\Big(\frac{\mu (\vartheta _{n } )} {\vartheta _{n}} +\delta \Big)}\Big(\nabla _{x}\mathbf{u}_{n} + \nabla _{x}^{T}\mathbf{u}_{n} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}_{n}\Big)\\ \\ \rightarrow \sqrt{\Big(\frac{\mu (\vartheta )} {\vartheta } +\delta \Big)}\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}\Big)\;\mbox{ weakly in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3\times 3}), \end{array} }$$
(3.159)

where we have used (3.117) and (3.151).

Similarly,

$$\displaystyle{ \sqrt{\frac{\eta (\vartheta _{n } )} {\vartheta _{n}}} \mathrm{div}_{x}\mathbf{u}_{n} \rightarrow \sqrt{\frac{\eta (\vartheta )} {\vartheta }} \mathrm{div}_{x}\mathbf{u}\;\mbox{ weakly in }L^{2}((0,T) \times \Omega ), }$$
(3.160)

and, by virtue of (3.111), (3.123) and (3.151),

$$\displaystyle{ \frac{\sqrt{\kappa _{\delta }(\vartheta _{n } )}} {\vartheta _{n}} \nabla _{x}\vartheta _{n} \rightarrow \frac{\sqrt{\kappa _{\delta }(\vartheta )}} {\vartheta } \nabla _{x}\vartheta \;\mbox{ weakly in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3}). }$$
(3.161)

By the same token, due to (3.112), (3.136), (3.140),

$$\displaystyle{ \sqrt{\Big(\frac{\Gamma \varrho _{n }^{\Gamma -2 } + 2} {\vartheta _{n}} \Big)}\,\nabla _{x}\varrho _{n} \rightarrow }$$
(3.162)
$$\displaystyle{\sqrt{\Big(\frac{\Gamma \varrho ^{\Gamma -2 } + 2} {\vartheta } \Big)}\,\nabla _{x}\varrho \quad \mbox{ weakly in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3}),}$$

while, by virtue of (3.113), (3.136), (3.140), (3.151),

$$\displaystyle{ \frac{1} {\sqrt{\varrho _{n } \vartheta _{n}}}\sqrt{ \frac{\partial p_{M}} {\partial \varrho } (\varrho _{n},\vartheta _{n})}\,\nabla _{x}\varrho _{n} \rightarrow }$$
(3.163)
$$\displaystyle{ \frac{1} {\sqrt{\varrho \vartheta }}\sqrt{\frac{\partial p_{M } } {\partial \varrho } (\varrho,\vartheta )}\,\nabla _{x}\varrho \quad \mbox{ weakly in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3}).}$$

Finally, as a consequence of (3.136), (3.151), and the bounds established in (3.124), (3.126), (3.130), we have

$$\displaystyle{ \varepsilon \vartheta _{n}^{4} -\frac{\varrho _{n}} {\vartheta _{n}}\mathcal{Q}_{\delta }\rightarrow \varepsilon \vartheta ^{4} -\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\;\mbox{ in }L^{\,p}((0,T) \times \Omega )\mbox{ for some }p> 1. }$$
(3.164)

The convergence results just established are sufficient in order to perform the weak limit for n in the approximate entropy balance (3.152). Although we are not able to show strong convergence of the gradients of ϱ, ϑ, and u, the inequality sign in (3.152) is preserved under the weak limit because of lower semi-continuity of convex superposition operators (cf. Theorem 11.27 in Appendix). Consequently, we are allowed to conclude that

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varrho s_{\delta }(\varrho,\vartheta )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{ \Omega }\Big(\frac{\mathbf{q}_{\delta }} {\vartheta } + \varepsilon \mathbf{r}_{\epsilon }\Big) \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t }$$
(3.165)
$$\displaystyle{+\int _{0}^{T}\int _{ \Omega }\sigma _{\varepsilon,\delta }\varphi \ \mathrm{d}x\ \mathrm{d}t \leq -\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x +\int _{ 0}^{T}\int _{ \Omega }\Big(\varepsilon \vartheta ^{4} -\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\Big)\varphi \ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{\mbox{ for any}\ \varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega }),\quad \varphi \geq 0,}$$

where we have set

$$\displaystyle{ \begin{array}{c} \mathbf{q}_{\delta } = \mathbf{q}_{\delta }(\vartheta,\nabla \vartheta ) =\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta,\ \kappa _{\delta }(\vartheta ) =\kappa (\vartheta ) +\delta \Big (\vartheta ^{\Gamma } + \frac{1} {\vartheta } \Big),\\ \\ s_{\delta }(\varrho,\vartheta ) = s(\varrho,\vartheta )+\delta \log \vartheta, \end{array} }$$
(3.166)

and

$$\displaystyle{ \sigma _{\varepsilon,\delta } = \frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } + \frac{\delta } {2}(\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} \Big]+ }$$
(3.167)
$$\displaystyle{+ \frac{\varepsilon \delta } {2\vartheta }(\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} + \varepsilon \frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta )\frac{\vert \nabla _{x}\varrho \vert ^{2}} {\varrho \vartheta },}$$
$$\displaystyle{\mathbf{r}_{\varepsilon } = -\Big(\vartheta s_{M,\delta }(\varrho,\vartheta ) - e_{M,\delta }(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big)\frac{\nabla _{x}\varrho } {\vartheta }.}$$

3.5.4 Limit in the Approximate Momentum Equation

With regard to formulas (3.32), (3.53), estimates (3.114), (3.116), (3.119), (3.126), (3.129), (3.130), and the asymptotic limits established in (3.117), (3.136), (3.140), (3.151), it is easy to identify the limit for n in all quantities appearing in the approximate momentum equation (3.49) for a fixed test function φ, with the exception of the convective term. Note that, even at this level of approximations, we have already lost compactness of the velocity field in the time variable because of the hypothetical presence of vacuum zones.

To begin, observe that

$$\displaystyle{\varrho _{n}\mathbf{u}_{n} \otimes \mathbf{u}_{n} \rightarrow \overline{\varrho \mathbf{u} \otimes \mathbf{u}}\ \mbox{ weakly in}\ L^{q}((0,T) \times \Omega; \mathbb{R}^{3\times 3})\ \mbox{ for a certain}\ q> 1,}$$

where we have used the uniform bounds (3.110), (3.116). Thus we have to show

$$\displaystyle{ \overline{\varrho \mathbf{u} \otimes \mathbf{u}} = \varrho \mathbf{u} \otimes \mathbf{u}. }$$
(3.168)

To this end, observe first that

$$\displaystyle{\varrho _{n}\mathbf{u}_{n} \rightarrow \varrho \mathbf{u}\ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{\frac{5} {4} }(\Omega; \mathbb{R}^{3}))}$$

as a direct consequence of estimates (3.114), (3.133), and strong convergence of the density established in (3.136).

Moreover, it can be deduced from the approximate momentum equation (3.49) that the functions

$$\displaystyle{ \left \{t\mapsto \int _{\Omega }\varrho _{n}\mathbf{u}_{n}\cdot \phi \ \mathrm{d}x\right \}\ \mbox{ are equi-continuous and bounded in}\ C([0,T]) }$$
(3.169)

for any fixed ϕ ∈ ∪ n = 1 X n . Since the set ∪ n = 1 X n is dense in \(L^{5}(\Omega; \mathbb{R}^{3})\) we obtain, by means of the Arzelà-Ascoli theorem, that

$$\displaystyle{\varrho _{n}\mathbf{u}_{n} \rightarrow \varrho \mathbf{u}\;\mbox{ in }C_{\mathrm{weak}}([0,T];L^{5/4}(\Omega )).}$$

On the other hand, as the Lebesgue space \(L^{5/4}(\Omega )\) is compactly embedded into the dual \(W^{-1,2}(\Omega )\), we infer that

$$\displaystyle{ \varrho _{n}\mathbf{u}_{n} \rightarrow \varrho \mathbf{u}\ \mbox{ (strongly) in}\ C_{\mathrm{weak}}([0,T];W^{-1,2}(\Omega; \mathbb{R}^{3})). }$$
(3.170)

Relation (3.170), together with the weak convergence of the velocities in the space \(L^{2}(0,T;W^{1,2}(\Omega; \mathbb{R}^{3}))\) established in (3.117), give rise to (3.168).

3.5.5 The Limit System Resulting from the Faedo-Galerkin Approximation

Having completed the necessary preliminary steps, in particular, the strong convergence of the density in (3.140), and the strong convergence of the temperature in (3.151), we can let n in the approximate system (3.45)–(3.60) to deduce that the limit quantities {ϱ, u, ϑ} satisfy:

  1. (i)

    Approximate continuity equation:

    $$\displaystyle{ \partial _{t}\varrho + \mathrm{div}_{x}(\varrho \mathbf{u}) = \varepsilon \Delta \varrho \ \mbox{ a.a. in}\ (0,T) \times \Omega, }$$
    (3.171)

    together with the homogeneous Neumann boundary condition

    $$\displaystyle{ \nabla _{x}\varrho \cdot \mathbf{n}\vert _{\partial \Omega } = 0, }$$
    (3.172)

    and the initial condition

    $$\displaystyle{ \varrho (0,\cdot ) =\varrho _{0,\delta }. }$$
    (3.173)
  2. (ii)

    Approximate balance of momentum:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \mathbf{u} \cdot \partial _{t}\boldsymbol{\varphi } +\varrho [\mathbf{u} \otimes \mathbf{u}]: \nabla _{x}\boldsymbol{\varphi } +\Big (\,p +\delta (\varrho ^{\Gamma } +\varrho ^{2})\Big)\mathrm{div}_{ x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
    (3.174)
    $$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\Big(\varepsilon (\nabla _{x}\varrho \nabla _{x}\mathbf{u}) \cdot \boldsymbol{\varphi } +\mathbb{S}_{\delta }: \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}_{\delta }\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }(\varrho \mathbf{u})_{0}\cdot \boldsymbol{\varphi }\ \mathrm{d}x,}$$

    satisfied for any test function \(\boldsymbol{\varphi }\in C_{c}^{\infty }([0,T) \times \overline{\Omega }; \mathbb{R}^{3})\), where either

    $$\displaystyle{ \boldsymbol{\varphi }\cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ in the case of the complete slip boundary conditions,} }$$
    (3.175)

    or

    $$\displaystyle{ \boldsymbol{\varphi }\vert _{\partial \Omega } = 0\ \mbox{ in the case of the no-slip boundary conditions,} }$$
    (3.176)

    and where we have set

    $$\displaystyle{ \mathbb{S}_{\delta } = \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}) = (\mu (\vartheta )+\delta \vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{\perp }\mathbf{u} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}\Big) +\eta (\vartheta )\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}. }$$
    (3.177)
  3. (iii)

    Approximate entropy inequality:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varrho s_{\delta }(\varrho,\vartheta )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{ \Omega }\Big(\frac{\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta } {\vartheta } + \varepsilon \mathbf{r}\Big) \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t }$$
    (3.178)
    $$\displaystyle{+\int _{0}^{T}\int _{ \Omega }\sigma _{\varepsilon,\delta }\varphi \ \mathrm{d}x\ \mathrm{d}t \leq -\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x +\int _{ 0}^{T}\int _{ \Omega }\Big(\varepsilon \vartheta ^{4} -\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\Big)\varphi \ \mathrm{d}x\ \mathrm{d}t}$$

    for any test function \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\), φ ≥ 0, where we have set

    $$\displaystyle{ s_{\delta }(\varrho,\vartheta ) = s(\varrho,\vartheta )+\delta \log \vartheta,\ \kappa _{\delta }(\vartheta ) =\kappa (\vartheta ) +\delta \Big (\vartheta ^{\Gamma } + \frac{1} {\vartheta } \Big), }$$
    (3.179)

    and

    $$\displaystyle{ \sigma _{\varepsilon,\delta } = \frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } + \frac{\delta } {2}(\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} \Big]+ }$$
    (3.180)
    $$\displaystyle{+ \frac{\varepsilon \delta } {2\vartheta }(\Gamma \varrho ^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho \vert ^{2} + \varepsilon \frac{\partial p_{M}} {\partial \varrho } (\varrho,\vartheta )\frac{\vert \nabla _{x}\varrho \vert ^{2}} {\varrho \vartheta },}$$
    $$\displaystyle{\mathbf{r} = -\Big(\vartheta s_{M,\delta }(\varrho,\vartheta ) - e_{M,\delta }(\varrho,\vartheta ) -\frac{p_{M}(\varrho,\vartheta )} {\varrho } \Big)\frac{\nabla _{x}\varrho } {\vartheta }.}$$
  4. (iv)

    Approximate total energy balance:

    $$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} +\varrho e_{\delta }(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x }$$
    (3.181)
    $$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0,\delta }\vert ^{2}} {\varrho _{0,\delta }} +\varrho _{0,\delta }e_{0,\delta } + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
    $$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \mathcal{Q}_{\delta } +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5}\Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for a.a.}\ \tau \in [0,T],}$$

    where

    $$\displaystyle{ e_{\delta }(\varrho,\vartheta ) = e(\varrho,\vartheta ) +\delta \vartheta. }$$
    (3.182)

3.5.6 The Entropy Production Rate Represented by a Positive Measure

In accordance with the general ideas discussed in Sect. 1.2, the entropy inequality can be interpreted as a weak formulation of a balance law with the production rate represented by a positive measure. More specifically, writing (3.178) in the form

$$\displaystyle{\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x -\int _{0}^{T}\int _{ \Omega }\Big(\varepsilon \vartheta ^{4} -\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\Big)\varphi \ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{-\int _{0}^{T}\int _{ \Omega }\varrho s_{\delta }(\varrho,\vartheta )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{ \Omega }\Big(\frac{\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta } {\vartheta } + \varepsilon \mathbf{r}\Big) \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{\geq \int _{0}^{T}\int _{ \Omega }\sigma _{\varepsilon,\delta }\varphi \ \mathrm{d}x\ \mathrm{d}t}$$

for any \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\), φ ≥ 0, the left-hand side can be understood as a non-negative linear form defined on the space of smooth functions with compact support in \([0,T) \times \overline{\Omega }\).

Consequently, by means of the classical Riesz representation theorem, there exists a regular, non-negative Borel measure \(\Sigma _{\varepsilon,\delta }\) on the set \([0,T) \times \overline{\Omega }\), that can be trivially extended on the compact set \([0,T] \times \overline{\Omega }\) such that

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varrho s_{\delta }(\varrho,\vartheta )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{ \Omega }\Big(\frac{\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta } {\vartheta } + \varepsilon \mathbf{r}\Big) \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t }$$
(3.183)
$$\displaystyle{+ <\Sigma _{\varepsilon,\delta };\varphi> _{[\mathcal{M};C]([0,T]\times \overline{\Omega })} = -\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x +\int _{ 0}^{T}\int _{ \Omega }\Big(\varepsilon \vartheta ^{4} -\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\Big)\varphi \ \mathrm{d}x\ \mathrm{d}t}$$

for any \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\). Moreover,

$$\displaystyle{ \Sigma _{\varepsilon,\delta } \geq \sigma _{\varepsilon,\delta }, }$$
(3.184)

where we have identified the function \(\sigma _{\varepsilon,\delta } \in L^{1}((0,T) \times \Omega )\) with a non-negative measure, see (1.13)–(1.17) for more details.

3.6 Artificial Diffusion Limit

The next step in the proof of Theorem 3.1 is to let ɛ → 0 in the approximate system (3.171)–(3.181) in order to eliminate the artificial diffusion term in (3.171) as well as the other ɛ-dependent quantities in the remaining equations. Such a step is not straightforward, as we loose the uniform bound on ∇ x ϱ; whence compactness of ϱ with respect to the space variable becomes an issue. In particular, the lack of pointwise convergence of the densities has to be taken into account in the proof of pointwise convergence of the approximate temperatures; accordingly, the procedure described in the previous section relating formulas (3.150), (3.151) has to be considerably modified. Apart from these principal new difficulties a number of other rather technical issues has to be addressed. In particular, uniform bounds must be established in order to show that all ɛ-dependent quantities in the approximate continuity equation (3.171), momentum equation (3.174), and energy balance (3.181) vanish in the asymptotic limit ɛ → 0. Similarly, the non-negative quantities appearing in the approximate entropy production rate σ ɛ, δ are used to obtain uniform bounds in order to eliminate the “artificial” entropy flux r in (3.178).

In order to show pointwise convergence of the approximate temperatures, we take advantage of certain general properties of weak convergence of composed functions expressed conveniently in terms of parameterized (Young) measures (see Sect. 3.6.2). On the other hand, similarly to the recently developed existence theory for compressible viscous fluids, we use the extra regularity properties of the quantity \(\Pi:= p(\varrho,\vartheta ) - (\frac{4} {3}\mu (\vartheta ) +\eta (\vartheta ))\mathrm{div}_{x}\mathbf{u}\) called effective viscous flux in order to establish pointwise convergence of the approximate densities. Such an approach requires a proper description of possible oscillations of the densities provided by the renormalized continuity equation (cf. Sect. 11.19 in Appendix).

3.6.1 Uniform Estimates and Limit in the Approximate Continuity Equation

Let {ϱ ɛ , u ɛ , ϑ ɛ } ɛ > 0 be a family of solutions to the approximate system (3.171)–(3.181) constructed in Sect. 3.5. Similarly to Sect. 2.2.3, the total energy balance (3.181), together with the entropy inequality represented through (3.183), give rise to the dissipation balance

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho _{\varepsilon }\vert \mathbf{u}_{\varepsilon }\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho _{\varepsilon },\vartheta _{\varepsilon }) + \delta ( \frac{\varrho _{\varepsilon }^{\Gamma }} {\Gamma - 1} + \varrho _{\varepsilon }^{2})\Big)(\tau )\ \mathrm{d}x }$$
(3.185)
$$\displaystyle{+\overline{\vartheta }\Sigma _{\varepsilon,\delta }\Big[[0,\tau ] \times \overline{\Omega }\Big] +\int _{ 0}^{\tau }\int _{ \Omega }\varepsilon \vartheta ^{5}\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0,\delta }} + H_{\delta,\overline{\vartheta }}(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho _{\varepsilon }\mathbf{f}_{\delta } \cdot \mathbf{u}_{\varepsilon } +\varrho \Big (1 -\frac{\overline{\vartheta }} {\vartheta _{\varepsilon }}\Big)\mathcal{Q}_{\delta } + \frac{\delta } {\vartheta _{\varepsilon }^{2}} + \varepsilon \overline{\vartheta }\vartheta ^{4}\Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for a.a.}\ \tau \in [0,T],}$$

where \(\Sigma _{\varepsilon,\delta } \in \mathcal{M}^{+}([0,T] \times \overline{\Omega })\) is the entropy production rate introduced in Sect. 3.5.6, and the “approximate Helmholtz function” \(H_{\delta,\overline{\vartheta }}\) is given through (3.100).

Repeating the arguments used after formula (3.104) we obtain

$$\displaystyle{ \sup _{\varepsilon>0}\left \{\mathrm{ess}\sup _{t\in (0,T)}\int _{\Omega }\Big(\frac{1} {2}\varrho _{\varepsilon }\vert \mathbf{u}_{\varepsilon }\vert ^{2} + H_{\delta,\overline{\vartheta }}(\varrho _{\varepsilon },\vartheta _{\varepsilon }) + \delta ( \frac{\varrho _{\varepsilon }^{\Gamma }} {\Gamma - 1} +\varrho _{ \varepsilon }^{2})\Big)(t)\ \mathrm{d}x\right \} <\infty, }$$
(3.186)

together with

$$\displaystyle{ \sup _{\varepsilon>0}\left \{\Sigma _{\varepsilon,\delta }\Big[[0,T] \times \overline{\Omega }\Big] +\int _{ 0}^{T}\int _{ \Omega }\varepsilon \vartheta _{\varepsilon }^{5}\ \mathrm{d}x\ \mathrm{d}t\right \} <\infty, }$$
(3.187)

where, in accordance with (3.180), (3.184), estimate (3.187) further implies

$$\displaystyle{\sup _{\varepsilon>0}\left \{\int _{0}^{T}\int _{ \Omega }\Big\{\frac{1} {\vartheta _{\varepsilon }} \Big[\mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \nabla _{x}\mathbf{u}_{\varepsilon }+\right.}$$
$$\displaystyle{ \left.\Big(\frac{\kappa (\vartheta _{\varepsilon })} {\vartheta _{\varepsilon }} +\delta (\vartheta _{\varepsilon }^{\Gamma -1} + \frac{1} {\vartheta _{\varepsilon }^{2}})\Big)\vert \nabla _{x}\vartheta _{\varepsilon }\vert ^{2}\Big] +\delta \frac{1} {\vartheta _{\varepsilon }^{3}} + \varepsilon \vartheta _{\varepsilon }^{5}\Big\}\ \mathrm{d}x\ \mathrm{d}t\right \} <\infty, }$$
(3.188)
$$\displaystyle{ \sup _{\varepsilon>0}\left \{\varepsilon \delta \int _{0}^{T}\int _{ \Omega }\frac{1} {\vartheta _{\varepsilon }} (\Gamma \varrho _{\varepsilon }^{\Gamma -2} + 2)\vert \nabla _{ x}\varrho _{\varepsilon }\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t\right \} <\infty, }$$
(3.189)

and

$$\displaystyle{ \sup _{\varepsilon>0}\left \{\int _{0}^{T}\int _{ \Omega }\varepsilon \frac{\overline{\vartheta }} {\varrho _{\varepsilon }\vartheta _{\varepsilon }} \frac{\partial p_{M}} {\partial \varrho } (\varrho _{\varepsilon },\vartheta _{\varepsilon })\vert \nabla _{x}\varrho _{\varepsilon }\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t\right \} <\infty. }$$
(3.190)

Exactly as in Sect. 3.5, the above estimates can be used to deduce that

$$\displaystyle{ \varrho _{\varepsilon } \rightarrow \varrho \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{\Gamma }(\Omega )), }$$
(3.191)
$$\displaystyle{ \mathbf{u}_{\varepsilon } \rightarrow \mathbf{u}\ \mbox{ weakly in}\ L^{2}(0,T;W^{1,2}(\Omega; \mathbb{R}^{3})), }$$
(3.192)

and

$$\displaystyle{ \vartheta _{\varepsilon } \rightarrow \vartheta \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{4}(\Omega )), }$$
(3.193)

at least for suitable subsequences. Moreover, we have \(\mathbf{u}(t,\cdot ) \in W_{\mathbf{n}}^{1,2}(\Omega; \mathbb{R}^{3})\) for a.a. t ∈ (0, T) in the case of the complete slip boundary conditions, while \(\mathbf{u}(t,\cdot ) \in W_{0}^{1,2}(\Omega; \mathbb{R}^{3})\) for a.a. t ∈ (0, T), if the no-slip boundary conditions are imposed.

Multiplying Eq. (3.171) on ϱ ɛ and integrating by parts we get

$$\displaystyle{\frac{1} {2}\int _{\Omega }\varrho _{\varepsilon }^{2}(\tau )\ \mathrm{d}x + \varepsilon \int _{ 0}^{\tau }\int _{ \Omega }\vert \nabla _{x}\varrho _{\varepsilon }\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{= \frac{1} {2}\int _{\Omega }\varrho _{0,\delta }^{2}\ \mathrm{d}x -\frac{1} {2}\int _{0}^{\tau }\int _{ \Omega }\varrho _{\varepsilon }^{2}\mathrm{div}_{ x}\mathbf{u}_{\varepsilon }\ \mathrm{d}x\ \mathrm{d}t;}$$

whence, taking (3.191)–(3.193) into account, we can see that

$$\displaystyle{\{\sqrt{\varepsilon }\nabla _{x}\varrho _{\varepsilon }\}_{\varepsilon>0}\ \mbox{ is bounded in}\ L^{2}(0,T;L^{2}(\Omega; \mathbb{R}^{3})),}$$

in particular,

$$\displaystyle{ \varepsilon \nabla _{x}\varrho _{\varepsilon } \rightarrow 0\ \mbox{ in}\ L^{2}(0,T;L^{2}(\Omega; \mathbb{R}^{3})). }$$
(3.194)

As the time derivative t ϱ ɛ can be expressed by means of Eq. (3.171), convergence in (3.191) can be, similarly to (3.118), strengthened to

$$\displaystyle{ \varrho _{\varepsilon } \rightarrow \varrho \ \mbox{ in}\ C_{\mathrm{weak}}([0,T];L^{\Gamma }(\Omega )). }$$
(3.195)

Relation (3.195), combined with (3.192) and boundedness of the kinetic energy, yields

$$\displaystyle{ \varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \rightarrow \varrho \mathbf{u}\ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{ \frac{2\Gamma } {\Gamma +1} }(\Omega; \mathbb{R}^{3})). }$$
(3.196)

Thus we conclude that the limit functions ϱ, u satisfy the integral identity

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \partial _{t}\varphi +\varrho \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t + \int _{\Omega }\varrho _{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x = 0 }$$
(3.197)

for any test function \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\). Moreover, since the boundary \(\partial \Omega\) is regular (Lipschitz) we can extend continuously the velocity field u outside \(\Omega\) in such a way that the resulting vector field belongs to \(W^{1,2}(\mathbb{R}^{3}; \mathbb{R}^{3})\). (In the case of no-slip boundary conditions one can take trivial extension, where u = 0 outside \(\Omega\).) Accordingly, setting ϱ ≡ 0 in \(\mathbb{R}^{3}\setminus \Omega\) we can assume that ϱ, u solve the equation of continuity

$$\displaystyle{ \partial _{t}\varrho + \mathrm{div}_{x}(\varrho \mathbf{u}) = 0\ \mbox{ in}\ \mathcal{D}'((0,T) \times \mathbb{R}^{3}). }$$
(3.198)

3.6.2 Entropy Balance and Strong Convergence of the Approximate Temperatures

Our principal objective is to show strong (pointwise) convergence of the family {ϑ ɛ } ɛ > 0. Following the same strategy as in Sect. 3.5.3, we divide the proof into three steps:

  1. (i)

    Div-Curl lemma (Proposition 3.3) is applied to show that

    $$\displaystyle{\overline{\varrho s_{\delta }(\varrho,\vartheta )G(\vartheta )} = \overline{\varrho s_{\delta }(\varrho,\vartheta )}\;\overline{G(\vartheta )}}$$

    for any GW 1,(0, ). This relation is reminiscent of formula (3.149); the quantity G playing a role of a cut-off function is necessary because of the low integrability of ϑ. The proof uses the same arguments as in Sect. 3.5.3.

  2. (ii)

    Although strong convergence of the densities is no longer available at this stage, we can still show that

    $$\displaystyle{ \overline{b(\varrho )G(\vartheta )} = \overline{b(\varrho )}\;\overline{G(\vartheta )}, }$$
    (3.199)

    where bC([0, )) ∩ L ((0, )), and G is the same as in the previous step. In order to prove this identity, we use the properties of renormalized solutions to the approximate continuity equation (cf. Sect. 11.19 in Appendix). Very roughly indeed, we can say that possible oscillations in the sequence of approximate densities and temperatures take place in orthogonal directions of the space-time.

  3. (iii)

    The simple monotonicity argument used in formula (3.150) has to be replaced by a more sophisticated tool. Here, the desired relation

    $$\displaystyle{s_{M}(\varrho _{\varepsilon },t,x)(G(\vartheta _{\varepsilon }) -\overline{G(\vartheta )}) \rightarrow 0}$$

    is shown to follow directly from (3.199) by means of a general argument borrowed from the theory of parameterized (Young) measures. An elementary alternative proof of this step involving a compactness argument based on the renormalized continuity equation (more precisely on Theorem 11.37 in Appendix) is shown in Sect. 3.7.3.

In the remaining part of this section, we develop the ideas delineated in the above program in a more specific way.

Uniform Estimates Seeing that the sequence {ϱ ɛ , u ɛ , ϑ ɛ } ɛ > 0 admits the bounds obtained in (3.188), we infer that {ϑ ɛ } ɛ > 0 satisfies the estimates stated in (3.121)–(3.127), namely

$$\displaystyle{ \begin{array}{c} \{\vartheta _{\varepsilon }\}_{\varepsilon>0},\ \{\vartheta _{\varepsilon }^{\Gamma /2}\}_{\varepsilon>0}\ \mbox{ are bounded in }L^{2}(0,T;W^{1,2}(\Omega )),\\ \\ \{\nabla (\vartheta _{\varepsilon }^{-1/2})\}_{\varepsilon>0}\;\mbox{ is bounded in }L^{2}((0,T) \times \Omega; \mathbb{R}^{3}),\\ \\ \{\vartheta _{\varepsilon }^{-1}\}_{\varepsilon>0}\;\mbox{ is bounded in }L^{3}((0,T) \times \Omega ),\\ \\ \{\log \vartheta _{\varepsilon }\}_{\varepsilon>0}\;\mbox{ is bounded in }L^{2}(0,T;W^{1,2}(\Omega )) \cap L^{\Gamma }(0,T;L^{3\Gamma }(\Omega )).\end{array} }$$
(3.200)

Moreover, relations (3.128), (3.129) imply that

$$\displaystyle{ \{\sqrt{\varepsilon }\varrho _{\varepsilon }\}_{\varepsilon>)},\,\{\sqrt{\varepsilon }\varrho _{\varepsilon }^{\frac{\Gamma } {2} }\}_{\varepsilon>0}\quad \mbox{ are bounded in }L^{2}(0,T;W^{1,2}(\Omega )). }$$
(3.201)

Application of Div-Curl Lemma Now we rewrite the approximate entropy balance (3.183) in the form

$$\displaystyle{\partial _{t}\Big(\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\Big) + \mathrm{div}_{x}\Big(\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\mathbf{u}_{\varepsilon } + \frac{\kappa _{\delta }(\vartheta _{\varepsilon })\nabla _{x}\vartheta _{\varepsilon }} {\vartheta _{\varepsilon }} + \varepsilon \mathbf{r}_{\varepsilon }\Big)}$$
$$\displaystyle{= \Sigma _{\varepsilon,\delta } + \frac{\varrho _{\varepsilon }} {\vartheta _{\varepsilon }}\mathcal{Q}_{\delta }-\varepsilon \vartheta _{\varepsilon }^{4}}$$

to be understood in the weak sense specified in Sects. 1.2 and 3.5.6.

Similarly to Sect. 3.5.3, we intend to apply Div-Curl lemma (Proposition 3.3) to the four-component vector fields

$$\displaystyle{ \mathbf{U}_{\varepsilon }:= \left [\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon }),\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\mathbf{u}_{\varepsilon } + \frac{\kappa _{\delta }(\vartheta _{\varepsilon })\nabla _{x}\vartheta _{\varepsilon }} {\vartheta } + \varepsilon \mathbf{r}_{\varepsilon }\right ], }$$
(3.202)
$$\displaystyle{ \mathbf{V}_{\varepsilon }:= \left [G(\vartheta _{\varepsilon }),0,0,0\right ], }$$
(3.203)

where G is a bounded globally Lipschitz function on [0, ).

First observe that the families

$$\displaystyle{\mathrm{div}_{t,x}\mathbf{U}_{\varepsilon } = \Sigma _{\varepsilon,\delta }+\frac{\varrho _{\varepsilon }} {\vartheta _{\varepsilon }}\mathcal{Q}_{\delta }-\varepsilon \vartheta _{\varepsilon }^{4},\quad \mathrm{curl}_{ t,x}\mathbf{V}_{\varepsilon } = G'(\vartheta _{\varepsilon })\left (\begin{array}{cc} 0 &\nabla \vartheta _{\varepsilon } \\ \nabla ^{T}\vartheta _{\varepsilon }& \mathbf{0}\\ \end{array} \right )}$$

are relatively compact in \(W^{-1,s}((0,T) \times \Omega ))\), \(W^{-1,s}((0,T) \times \Omega; \mathbb{R}^{4\times 4})\) for \(s \in [1, \frac{4} {3})\), respectively. Indeed, it is enough to use estimates (3.187), (3.191), (3.193), (3.200), and compactness of the embeddings \(\mathcal{M}^{+}([0,T] \times \overline{\Omega })\hookrightarrow W^{-1,s}((0,T) \times \Omega ))\), \(L^{1}((0,T) \times \Omega ))\hookrightarrow W^{-1,s}((0,T) \times \Omega ))\). Notice that we have, in particular,

$$\displaystyle{ \varepsilon \vartheta _{\varepsilon }^{4} \rightarrow 0\ \mbox{ in}\ L^{1}((0,T) \times \Omega ) }$$
(3.204)

as a direct consequence of (3.193).

As the sequence {G(ϑ ɛ )} ɛ > 0 is bounded in \(L^{\infty }((0,T) \times \Omega )\), it is enough to show boundedness of the family {U ɛ } ɛ > 0 in \(L^{\,p}((0,T) \times \Omega; \mathbb{R}^{4})\) for a certain 1 < p < . Combining the arguments already used in (3.144), (3.145) with the bounds (3.191), (3.200), we infer that

$$\displaystyle{ \{\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\}_{\varepsilon>0}\ \mbox{ is bounded in}\ L^{\,p}((0,T) \times \Omega )\ \mbox{ for a certain}\ p> 2, }$$
(3.205)

while

$$\displaystyle{ \{\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\mathbf{u}_{\varepsilon }\}_{\varepsilon>0},\ \left \{\frac{\kappa _{\delta }(\vartheta _{\varepsilon })} {\vartheta _{\varepsilon }} \nabla \vartheta _{\varepsilon }\right \}_{\varepsilon>0}\ \mbox{ are bounded in}\ L^{q}((0,T) \times \Omega; \mathbb{R}^{3}) }$$
(3.206)
$$\displaystyle{\mbox{ for a certain}\ q> 1\ \mbox{ provided}\ \Gamma> 6.}$$

Finally, following the reasoning of (3.146)–(3.148), we use (3.200) and (3.201) to obtain

$$\displaystyle{ \varepsilon \mathbf{r}_{\varepsilon } \rightarrow 0\ \mbox{ in}\ L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3}))\ \mbox{ for a certain}\ p> 1. }$$
(3.207)

Having verified all hypotheses of Proposition 3.3 we conclude that

$$\displaystyle{ \overline{\varrho s_{\delta }(\varrho,\vartheta )G(\vartheta )} = \overline{\varrho s_{\delta }(\varrho,\vartheta )}\ \overline{G(\vartheta )} }$$
(3.208)

for any bounded and continuous function G.

Monotonicity of the Entropy and Strong Convergence of the Approximate Temperatures: Application of the Theory of Parametrized (Young) Measures Similarly to Sect. 3.5.3, relation (3.208) can be used to show strong (pointwise) convergence of {ϑ ɛ } ɛ > 0. Decomposing

$$\displaystyle{\varrho s_{\delta }(\varrho,\vartheta ) = \varrho s_{M}(\varrho,\vartheta ) +\delta \varrho \log (\vartheta ) + \frac{4} {3}a\vartheta ^{3},}$$

we have to show that

$$\displaystyle{ \begin{array}{c} \overline{\varrho s_{M}(\varrho,\vartheta )G(\vartheta )} \geq \overline{\varrho s_{M}(\varrho,\vartheta )}\ \overline{G(\vartheta )},\ \overline{\varrho \log (\vartheta )G(\vartheta )} \geq \overline{\varrho \log (\vartheta )}\ \overline{G(\vartheta )},\\ \\ \overline{\vartheta ^{3}G(\vartheta )} \geq \overline{\vartheta ^{3}}\ \overline{G(\vartheta )} \end{array} }$$
(3.209)

for any continuous and increasing G chosen in such a way that all the weak limits exist et least in L 1. Indeed, relations (3.209) combined with (3.208) imply

$$\displaystyle{ \overline{\vartheta ^{3}G(\vartheta )} = \overline{\vartheta ^{3}}\ \overline{G(\vartheta )};\ \mbox{ whence}\ \overline{\vartheta ^{4}} = \overline{\vartheta ^{3}}\vartheta }$$
(3.210)

yielding, up to a subsequence, the desired conclusion

$$\displaystyle{ \vartheta _{\varepsilon } \rightarrow \vartheta \ \mbox{ a.a. in}\ (0,T) \times \Omega. }$$
(3.211)

In order to see (3.209), write

$$\displaystyle{0 \leq \left (\varrho _{\varepsilon }s_{M}\Big(\varrho _{\varepsilon },G^{-1}(G(\vartheta _{\varepsilon }))\Big) -\varrho _{\varepsilon }s_{ M}\Big(\varrho _{\varepsilon },G^{-1}(\overline{G(\vartheta )})\Big)\right )\Big(G(\vartheta _{\varepsilon }) -\overline{G(\vartheta )}\Big)}$$
$$\displaystyle{= \varrho _{\varepsilon }s_{M}(\varrho _{\varepsilon },\vartheta _{\varepsilon })\Big(G(\vartheta _{\varepsilon }) -\overline{G(\vartheta )})\Big) -\varrho _{\varepsilon }s_{M}\Big(\varrho _{\varepsilon },G^{-1}(\overline{G(\vartheta )})\Big)\Big(G(\vartheta _{\varepsilon }) -\overline{G(\vartheta )}\Big).}$$

Consequently, the first inequality in (3.209) follows as soon as we can show that

$$\displaystyle{ \varrho _{\varepsilon }s_{M}\Big(\varrho _{\varepsilon },G^{-1}(\overline{G(\vartheta )})\Big)\Big(G(\vartheta _{\varepsilon }) -\overline{G(\vartheta )}\Big) \rightarrow 0\ \mbox{ weakly in}\ L^{1}((0,T) \times \Omega ). }$$
(3.212)

The quantity

$$\displaystyle{\varrho _{\varepsilon }s_{M}\Big(\varrho _{\varepsilon },\Big[G^{-1}(\overline{G(\vartheta )})\Big](t,x)\Big) =\psi (t,x,\varrho _{\varepsilon })}$$

may be regarded as a superposition of a Carathéodory function with a weakly convergent sequence. In such a situation, a general argument of the theory of parameterized (Young) measures asserts that (3.212) follows as soon as we show that

$$\displaystyle{ \overline{b(\varrho )G(\vartheta )} = \overline{b(\varrho )}\ \overline{G(\vartheta )} }$$
(3.213)

for arbitrary smooth and bounded functions b and G (see Theorem 13).

Indeed, if ν (t, x) (ϱ, ϑ), ν (t, x) ϱ and ν (t, x) ϑ are families of parametrized Young measures associated to sequences {(ϱ ɛ , ϑ ɛ )} ɛ > 0, {ϱ ɛ } ɛ > 0 and{ϑ ɛ } ɛ > 0, respectively, then (3.213) implies

$$\displaystyle{\int _{\mathbb{R}^{2}}b(\lambda )G(\mu )\,\mathrm{d}\nu _{(t,x)}^{(\varrho,\vartheta )}(\lambda,\mu ) =\int _{ \mathbb{R}}b(\lambda )\,\mathrm{d}\nu _{(t,x)}^{\varrho }(\lambda ) \times \int _{ \mathbb{R}}G(\mu )\,\mathrm{d}\nu _{(t,x)}^{\vartheta }(\mu ).}$$

This evidently yields a decomposition

$$\displaystyle{\nu _{(t,x)}^{(\varrho,\vartheta )}(A \times B) =\nu _{ (t,x)}^{\varrho }(A)\nu _{ (t,x)}^{\vartheta }(B),}$$

where A, B are open subsets in \(\mathbb{R}\). Consequently, for any Carathéodory function ψ(t, x, λ) and a continuous function G(ϑ), such that sequences ψ(⋅ , ⋅ , ϱ n )G(ϑ n ) and ψ(⋅ , ⋅ , ϱ n ), G(ϑ n ) are weakly convergent in \(L^{1}((0,T) \times \Omega; \mathbb{R}^{2})\) and \(L^{1}((0,T) \times \Omega )\), respectively, we have

$$\displaystyle{[\overline{\psi (\cdot,\cdot,\varrho )G(\vartheta )}](t,x) =\int _{\mathbb{R}^{2}}\psi (t,x,\lambda )G(\mu )\,\mathrm{d}\nu _{(t,x)}^{(\varrho,\vartheta )}(\lambda,\mu ) =}$$
$$\displaystyle{\int _{\mathbb{R}^{2}}\psi (t,x,\lambda )G(\mu )\,\mathrm{d}\nu _{(t,x)}^{\varrho }(\lambda )\mathrm{d}\nu _{ (t,x)}^{\vartheta }(\mu ) = [\overline{\psi (\cdot,\cdot,\varrho )}\;\overline{G(\vartheta )}](t,x)}$$

which is nothing other than (3.212).

In order to verify (3.213), multiply the approximate continuity equation (3.171) by b′(ϱ)φ, \(\varphi \in C_{c}^{\infty }(\Omega )\), and integrate over \(\Omega\) to obtain

$$\displaystyle{ \begin{array}{c} \frac{\mathrm{d}} {\mathrm{d}t}\int _{\Omega }b(\varrho )\varphi \mathrm{d}x -\int _{\Omega }b(\varrho )\mathbf{u} \cdot \nabla _{x}\varphi \mathrm{d}x\\ \\ + \varepsilon \int _{\Omega }b''(\varrho )\vert \nabla _{x}\varrho \vert ^{2}\varphi \mathrm{d}x + \varepsilon \int _{\Omega }b'(\varrho )\nabla _{x}\varrho \cdot \nabla _{x}\varphi \mathrm{d}x\\ \\ +\int _{\Omega }(\varrho b'(\varrho ) - b(\varrho ))\mathrm{div}_{x}\mathbf{u}\varphi \mathrm{d}x = 0. \end{array} }$$
(3.214)

Consequently, the sequence \(\{t\mapsto \int _{\Omega }b(\varrho _{\varepsilon })\varphi \}_{\varepsilon>0}\) is uniformly bounded and equi-continuous in C([0, T]); whence

$$\displaystyle{ b(\varrho _{\varepsilon }) \rightarrow \overline{b(\varrho )}\;\mbox{ in }C_{\mathrm{weak}}([0,T];L^{\Gamma }(\Omega )) }$$
(3.215)

at least for any smooth function b with bounded second derivative.

Now, we use compactness of the embedding \(L^{\Gamma }(\Omega )\hookrightarrow W^{-1,2}(\Omega )\) to deduce

$$\displaystyle{ b(\varrho _{\varepsilon }) \rightarrow \overline{b(\varrho )}\ \mbox{ in}\ C([0,T];W^{-1,2}(\Omega )). }$$
(3.216)

On the other hand, in accordance with the uniform bounds established in (3.200),

$$\displaystyle{ G(\vartheta _{\varepsilon }) \rightarrow \overline{G(\vartheta )}\;\mbox{ weakly in }L^{2}(0,T;W^{1,2}(\Omega )); }$$
(3.217)

whence (3.213) follows from (3.216), (3.217).

In addition to (3.211), the limit temperature field ϑ is positive a.a. on the set \((0,T) \times \Omega\), more precisely, we have

$$\displaystyle{ \vartheta ^{-3} \in L^{1}((0,T) \times \Omega ). }$$
(3.218)

Indeed, (3.218) follows from the uniform bounds (3.200), the pointwise convergence of {ϑ ɛ } ɛ > 0 established in (3.211), and the property of weak lower semi-continuity of convex functionals (see Theorem 11.27 in Appendix).

Asymptotic Limit in the Entropy Balance At this stage, we are ready to let ɛ → 0 in the approximated entropy equality (3.183).

Using relations (3.200)–(3.211) we obtain, in the same way as in (3.155), (3.156),

$$\displaystyle{\frac{\kappa _{\delta }(\vartheta _{\varepsilon })} {\vartheta _{\varepsilon }} \nabla _{x}\vartheta _{\varepsilon } \rightarrow \frac{\kappa _{\delta }(\vartheta )} {\vartheta } \nabla _{x}\vartheta }$$
$$\displaystyle{\mbox{ weakly in }L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3})\mbox{ for some }p> 1.}$$

Furthermore, in accordance with (3.191), (3.211), we get

$$\displaystyle{\frac{\varrho _{\varepsilon }} {\vartheta _{\varepsilon }}\mathcal{Q}_{\delta }\rightarrow \frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\,\mbox{ weakly in }L^{\,p}((0,T) \times \Omega )\mbox{ for some }p> 1.}$$

Applying Div-Curl Lemma (Proposition 3.3) to the sequence {U ɛ } ɛ > 0 defined in (3.202) and {V ɛ } ɛ > 0,

$$\displaystyle{\mathbf{V}_{\varepsilon } = [(u_{\varepsilon })_{i},0,0,0],\;i = 1,2,3,}$$

we deduce

$$\displaystyle{\varrho _{\varepsilon }s_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\mathbf{u}_{\varepsilon } \rightarrow \overline{\varrho s_{\delta }(\varrho,\vartheta )}\mathbf{u}\,\mbox{ weakly in }L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3})\mbox{ for a certain }p> 1.}$$

The terms \(\frac{1} {\vartheta _{\varepsilon }} S_{\delta }(\vartheta _{\varepsilon },\mathbf{u}_{\varepsilon }): \nabla \mathbf{u}_{\varepsilon }\), \(\frac{\kappa _{\delta }(\vartheta _{\varepsilon })} {\vartheta _{\varepsilon }} \vert \nabla \vartheta _{\varepsilon }\vert ^{2}\) appearing in σ ɛ, δ are weakly lower semi-continuous as we have already observed in (3.159)–(3.164), while the remaining ɛ-dependent quantities in σ ɛ, δ are non-negative. Finally, by virtue of (3.187), we may assume

$$\displaystyle{\Sigma _{\varepsilon,\delta } \rightarrow \sigma _{\delta }\in \;\mbox{ weakly-(*) in }\mathcal{M}([0,T] \times \overline{\Omega })\mbox{, where }\sigma _{\delta } \in \mathcal{M}^{+}([0,T] \times \overline{\Omega }).}$$

Recalling the limits (3.204) and (3.207), we let ɛ → 0 in (3.183) to obtain

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\overline{\varrho s_{\delta }(\varrho,\vartheta )}\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.219)
$$\displaystyle{+\int _{0}^{T}\int _{ \Omega }\frac{\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta } {\vartheta } \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t+ <\sigma _{\delta },\varphi> _{[C;\mathcal{M}]([0,T]\times \overline{\Omega })} =}$$
$$\displaystyle{-\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x -\int _{0}^{T}\int _{ \Omega }\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\varphi \ \mathrm{d}x\ \mathrm{d}t,\,\mbox{ for all }\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega }),}$$

where

$$\displaystyle{\sigma _{\delta } \geq \frac{1} {\vartheta } \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u} + \frac{\kappa _{\delta }(\vartheta )} {\vartheta } \vert \nabla \vartheta \vert ^{2}.}$$

Consequently, in order to perform the limit ɛ → 0 in the remaining equations of the approximate system (3.171)–(3.181), we have to show

  1. (i)

    uniform pressure estimates analogous to those established in Sect. 2.2.5 or, alternatively, in Sect. 2.2.6,

  2. (ii)

    strong (pointwise) convergence of the approximate densities.

3.6.3 Uniform Pressure Estimates

The pressure estimates are derived in the same way as in Sect. 2.2.5, namely we use the quantities

$$\displaystyle{ \varphi =\psi \phi,\ \psi \in C_{c}^{\infty }(0,T),\ \phi = \mathcal{B}[\varrho _{\varepsilon } -\overline{\varrho }] }$$
(3.220)

as test functions in the approximate momentum equation (3.174), where

$$\displaystyle{\overline{\varrho } = \frac{1} {\vert \Omega \vert }\int _{\Omega }\varrho _{\varepsilon }\ \mathrm{d}x,}$$

and \(\mathcal{B}\approx \mathrm{div}_{x}^{-1}\) is the Bogovskii operator introduced in Sect. 2.2.5 and investigated in Sect. 11.6 in Appendix.

Since ϱ ɛ satisfies the approximate continuity equation (3.171), we have

$$\displaystyle{ \partial _{t}\phi = -\mathcal{B}\left [\mathrm{div}_{x}(\varrho \mathbf{u} -\varepsilon \nabla _{x}\varrho )\right ]. }$$
(3.221)

Consequently, by virtue of the basic properties of the operator \(\mathcal{B}\) listed in Sect. 2.2.5,

$$\displaystyle{ \|\phi (t,\cdot )\|_{W^{1,p}(\Omega;\mathbb{R}^{3})} \leq c(\,p,\Omega )\|\varrho _{\varepsilon }(t,\cdot )\|_{L^{\,p}(\Omega )}\;\mbox{ for a.a. t } \in (0,T), }$$
(3.222)

and

$$\displaystyle{ \left \|\partial _{t}\phi (t,\cdot )\right \|_{L^{\,p}(\Omega;\mathbb{R}^{3})} \leq c(\,p,\Omega )\,\Big\|\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }(t,\cdot ) + \varepsilon \nabla _{x}\varrho _{\varepsilon }(t,\cdot )\Big\|_{L^{\,p}(\Omega;\mathbb{R}^{3})}\mbox{ for a.a. }t \in (0,T) }$$
(3.223)

for any 1 < p < .

The last two estimates, together with those previously established in (3.191)–(3.196), (3.200), and (3.201), render the test functions (3.220) admissible in (3.174) provided, say, \(\Gamma \geq 4\). Note that, unlike in Sect. 2.2.5, the argument of the operator \(\mathcal{B}\) is an affine function of ϱ ɛ , whereas the necessary uniform estimate on {ϱ ɛ } ɛ > 0 in \(L^{\infty }(0,T;L^{\Gamma }(\Omega ))\) is provided by the extra pressure term \(\delta \varrho ^{\Gamma }\).

In view of these arguments, we can write, similarly to (2.94),

$$\displaystyle{ \int _{0}^{T}\Big[\psi \int _{ \Omega }\Big(\,p(\varrho _{\varepsilon },\vartheta _{\varepsilon }) +\delta (\varrho _{\varepsilon }^{\Gamma } +\varrho _{ \varepsilon }^{2})\Big)\varrho _{\varepsilon }\ \mathrm{d}x\Big]\ \mathrm{d}t =\sum _{ j=1}^{7}I_{ j}, }$$
(3.224)

where

$$\displaystyle{I_{1} =\int _{ 0}^{T}\Big[\psi \overline{\varrho }\int _{ \Omega }\Big(\,p(\varrho _{\varepsilon },\vartheta _{\varepsilon } +\delta (\varrho _{\varepsilon }^{\Gamma } +\varrho _{ \varepsilon }^{2})\Big)\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$
$$\displaystyle{I_{2} = -\int _{0}^{T}\Big[\psi \int _{ \Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \cdot \partial _{t}\boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$
$$\displaystyle{I_{3} = -\int _{0}^{T}\Big[\psi \int _{ \Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \otimes \mathbf{u}_{\varepsilon }: \nabla _{x}\boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$
$$\displaystyle{I_{4} =\int _{ 0}^{T}\Big[\psi \int _{ \Omega }\mathbb{S}_{\delta }(\mathbf{u}_{\varepsilon },\vartheta _{\varepsilon }): \nabla _{x}\boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$
$$\displaystyle{I_{5} = -\int _{0}^{T}\Big[\psi \int _{ \Omega }\varrho _{\varepsilon }\mathbf{f}\cdot \boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$
$$\displaystyle{I_{6} = -\int _{0}^{T}\Big[\psi '\int _{ \Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }\cdot \boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t,}$$

and

$$\displaystyle{I_{7} =\int _{ 0}^{T}\psi \Big[\int _{ \Omega }\varepsilon \nabla _{x}\varrho _{\varepsilon }\nabla _{x}\mathbf{u}_{\varepsilon }\cdot \boldsymbol{\phi }\ \mathrm{d}x\Big]\ \mathrm{d}t.}$$

The simple form of I 7 conditioned by the specific form of the test function φ, where the argument of \(\mathcal{B}\) is an affine function of ϱ ɛ , is the only technical reason why the limit processes for ɛ → 0 and δ → 0 must be separated.

The integral identity (3.224) can be used to obtain uniform bounds on the pressure independent of ɛ. Exactly as in Sect. 2.2.5, we deduce that

$$\displaystyle{ \|\varrho _{\varepsilon }\|_{L^{\Gamma +1}((0,T)\times \Omega )} \leq c(\mathrm{data},\delta ), }$$
(3.225)

and

$$\displaystyle{ \|p_{M}(\varrho _{\varepsilon },\vartheta _{\varepsilon })\|_{L^{\,p}((0,T)\times \Omega )} \leq c(\mathrm{data},\delta )\quad \mbox{ for a certain }p> 1. }$$
(3.226)

Indeed, these bounds can be obtained by dominating the integrals I 1I 7 in the spirit of Sect. 2.2.5, specifically, by means of estimates (3.222), (3.223), (3.191)–(3.196), and (3.200), provided \(\Gamma \geq 4\). In particular, by virtue of (3.192), (3.194),

$$\displaystyle{ \varepsilon \nabla _{x}\varrho _{\varepsilon }\nabla _{x}\mathbf{u}_{\varepsilon } \rightarrow 0\ \mbox{ in}\ L^{1}((0,T) \times \Omega; \mathbb{R}^{3})) }$$
(3.227)

yielding boundedness of integral I 7.

3.6.4 Limit in the Approximate Momentum Equation and in the Energy Balance

In accordance with estimates (3.225), (3.226), together with (3.193), (3.200), and (3.211),

$$\displaystyle{ \begin{array}{c} p_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon }) \rightarrow \overline{p_{\delta }(\varrho,\vartheta )} = \overline{p_{M}(\varrho,\vartheta )} + \frac{a} {4} \vartheta ^{4} +\delta (\overline{\varrho ^{\Gamma }} + \overline{\varrho ^{2}})\\ \\ \mbox{ weakly in }L^{\,p}((0,T) \times \Omega )\mbox{ for a certain }p> 1, \end{array} }$$
(3.228)

where we have denoted

$$\displaystyle{ p_{\delta }(\varrho,\vartheta ) = p_{M}(\varrho,\vartheta ) + \frac{a} {4}\vartheta ^{4} +\delta (\varrho ^{\Gamma } +\varrho ^{2}). }$$
(3.229)

On the other hand, by virtue of (3.17), (3.23), (3.193), and (3.211),

$$\displaystyle{\mu (\vartheta _{\varepsilon }) \rightarrow \mu (\vartheta ),\,\eta (\vartheta _{\varepsilon }) \rightarrow \eta (\vartheta )\,\mbox{ (strongly) in }L^{\,p}((0,T) \times \Omega )\mbox{ for any }1 \leq p <4.}$$

Moreover, since \(\mathbb{S}_{\delta }\) takes the form specified in (3.53), we can use (3.192) in order to deduce

$$\displaystyle{ \mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }) \rightarrow \mathbb{S}_{\delta }(\vartheta,\mathbf{u})\,\mbox{ weakly in }L^{\,p}((0,T) \times \Omega )\mbox{, for a certain }p> 1. }$$
(3.230)

As the limits of the families ϱ ɛ f, ϱ ɛ u ɛ , and ɛϱ ɛ u ɛ have already been identified through (3.191), (3.196) and (3.227), we are left with the convective term ϱ ɛ u ɛ u ɛ . Following the arguments of Sect. 3.5.4 we observe that

$$\displaystyle{ \varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \rightarrow \varrho \mathbf{u}\,\mbox{ in }C_{\mathrm{weak}}([0,T];L^{ \frac{2\Gamma } {\Gamma +1} }(\Omega; \mathbb{R}^{3})). }$$
(3.231)

Consequently, because of compact embedding \(L^{s}(\Omega )\hookrightarrow W^{-1,2}(\Omega )\), \(s> \frac{6} {5}\),

$$\displaystyle{\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \rightarrow \varrho \mathbf{u}\,\mbox{ (strongly) in }L^{\,p}(0,T;W^{-1,2}(\Omega; \mathbb{R}^{3}))}$$

for any 1 ≤ p < . In accordance with (3.192),

$$\displaystyle{ \varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \otimes \mathbf{u}_{\varepsilon } \rightarrow \varrho \mathbf{u} \otimes \mathbf{u}\,\mbox{ weakly in }L^{\,p}((0,T) \times \Omega )\mbox{ for a certain }p> 1. }$$
(3.232)

Letting ɛ → 0 in the approximate momentum equation (3.174) we get

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \mathbf{u} \cdot \partial _{t}\boldsymbol{\varphi } +\varrho [\mathbf{u} \otimes \mathbf{u}]: \nabla _{x}\boldsymbol{\varphi } + \overline{p_{\delta }(\varrho,\vartheta )}\mathrm{div}_{x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.233)
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\Big(\mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}_{\delta }\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }(\varrho \mathbf{u})_{0}\cdot \boldsymbol{\varphi }\ \mathrm{d}x,}$$

for any test function \(\boldsymbol{\varphi }\in C_{c}^{\infty }([0,T) \times \overline{\Omega };R^{3}))\) such that either

$$\displaystyle{\boldsymbol{\varphi }\cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ in the case of the complete slip boundary conditions,}}$$

or

$$\displaystyle{\boldsymbol{\varphi }\vert _{\partial \Omega } = 0\ \mbox{ in the case of the no-slip boundary conditions.}}$$

Finally, as the sequence {ϱ ɛ e δ (ϱ ɛ , ϑ ɛ )} ɛ > 0 is bounded in \(L^{\,p}((0,T) \times \Omega )\) (see (3.30), (3.191)–(3.193), (3.200)), we are allowed to let ɛ → 0 in the approximate energy balance (3.181) to obtain

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + \overline{\varrho e_{\delta }(\varrho,\vartheta )} + \delta ( \frac{\overline{\varrho ^{\Gamma }}} {\Gamma - 1} + \overline{\varrho ^{2}})\Big)(\tau )\ \mathrm{d}x }$$
(3.234)
$$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0,\delta }\vert ^{2}} {\varrho _{0,\delta }} +\varrho _{0,\delta }e_{0,\delta } + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \mathcal{Q}_{\delta } +\delta \frac{1} {\vartheta ^{2}} -\varepsilon \vartheta ^{5}\Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for a.a.}\ \tau \in [0,T].}$$

3.6.5 Strong Convergence of the Densities

In order to show strong (pointwise) convergence of {ϱ ɛ } ɛ > 0, we adapt the method introduced in the context of barotropic fluids with constant viscosity coefficients by Lions [192], and further developed in [103] in order to accommodate the variable transport coefficients.

Similarly to Sect. 2.2.6, we use the quantities

$$\displaystyle{ \varphi (t,x) =\psi (t)\zeta (x)\boldsymbol{\phi },\ \boldsymbol{\phi }= (\nabla _{x}\Delta _{x}^{-1})[1_{ \Omega }\varrho _{\varepsilon }],\ \psi \in C_{c}^{\infty }((0,T)),\,\zeta \in C_{ c}^{\infty }(\Omega ), }$$
(3.235)

as test functions in the approximate momentum equation (3.174), where the symbol \(\Delta _{x}^{-1}\) stands for the inverse Laplace operator considered on the whole space \(\mathbb{R}^{3}\) introduced in (2.100). The operator \(\nabla _{x}\Delta _{x}^{-1}\) is investigated in Sect. 11.17 in Appendix.

Since ϱ ɛ u ɛ and ∇ϱ ɛ possess zero normal traces, the approximate continuity equation (3.171) can be extended to the whole \(\mathbb{R}^{3}\), specifically,

$$\displaystyle{ \partial _{t}(1_{\Omega }\varrho _{\varepsilon }) + \mathrm{div}_{x}(1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }) -\varepsilon \mathrm{div}_{x}(1_{\Omega }\nabla \varrho _{\varepsilon }) = 0\,\mbox{ a.e. in }(0,T) \times \mathbb{R}^{3}. }$$
(3.236)

Accordingly, we have

$$\displaystyle{ \partial _{t}\boldsymbol{\phi } = -(\nabla _{x}\Delta _{x}^{-1})\left [\mathrm{div}_{ x}(1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } -\varepsilon 1_{\Omega }\nabla _{x}\varrho )\right ], }$$
(3.237)

cf. Theorem 11.33 in Appendix.

Now, exactly as in Sect. 2.2.6, we can use the uniform estimates (3.191)–(3.196), (3.200), and (3.201), in order to observe that φ defined through (3.235) is admissible in the integral identity (3.174) as soon as \(\Gamma \geq 4\). Thus we get

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(p_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\varrho _{\varepsilon } - \mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\Big)\ \mathrm{d}x\ \mathrm{d}t =\sum _{ j=1}^{8}I_{ j,\varepsilon }, }$$
(3.238)

where

$$\displaystyle{I_{1,\varepsilon } =\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \cdot \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }] - (\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \otimes \mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\Big)\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{2,\varepsilon } = -\varepsilon \int _{0}^{T}\int _{ \Omega }\psi \zeta \ \varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \cdot \nabla _{x}\Delta ^{-1}[\mathrm{div}_{ x}(1_{\Omega }\nabla _{x}\varrho _{\varepsilon })]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{3,\varepsilon } = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \varrho _{\varepsilon }\mathbf{f}_{\delta } \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{4,\varepsilon } = -\int _{0}^{T}\int _{ \Omega }\psi p_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\nabla _{x}\zeta \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{5,\varepsilon } =\int _{ 0}^{T}\int _{ \Omega }\psi \mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{6,\varepsilon } = -\int _{0}^{T}\int _{ \Omega }\psi (\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \otimes \mathbf{u}_{\varepsilon }): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{7,\varepsilon } = -\int _{0}^{T}\int _{ \Omega }\partial _{t}\psi \ \zeta \varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$

and

$$\displaystyle{I_{8,\varepsilon } = \varepsilon \int _{0}^{T}\int _{ \Omega }\nabla _{x}\varrho _{\varepsilon }\nabla _{x}\mathbf{u}_{\varepsilon } \cdot (\nabla _{x}\Delta _{x}^{-1})[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t.}$$

Here, the symbol \(\mathcal{R}\) stands for the double Riesz transform, defined componentwise as \(\mathcal{R}_{i,j} = \partial _{x_{i}}\Delta _{x}^{-1}\partial _{x_{j}}\), introduced in (2.101).

Repeating the same procedure we use the quantities

$$\displaystyle{\boldsymbol{\varphi }(t,x) =\psi (t)\zeta (x)(\nabla _{x}\Delta _{x}^{-1})[1_{ \Omega }\varrho ],\ \psi \in C_{c}^{\infty }(0,T),\ \zeta \in C_{ c}^{\infty }(\Omega ),}$$

as test functions in the limit momentum equation (3.233) in order to obtain

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(\overline{p_{\delta }(\varrho,\vartheta )}\varrho - \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\Big)\ \mathrm{d}x\ \mathrm{d}t =\sum _{ j=1}^{6}I_{ j}, }$$
(3.239)

where

$$\displaystyle{I_{1} =\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho \mathbf{u} \cdot \mathcal{R}[1_{\Omega }\varrho \mathbf{u}] - (\varrho \mathbf{u} \otimes \mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\Big)\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{2} = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \varrho \mathbf{f}_{\delta } \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{3} = -\int _{0}^{T}\int _{ \Omega }\psi \overline{p_{\delta }(\varrho,\vartheta )}\nabla _{x}\zeta \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho ]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{4} =\int _{ 0}^{T}\int _{ \Omega }\psi \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho ]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{5} = -\int _{0}^{T}\int _{ \Omega }\psi (\varrho \mathbf{u} \otimes \mathbf{u}): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho ]\ \mathrm{d}x\ \mathrm{d}t,}$$

and

$$\displaystyle{I_{6} = -\int _{0}^{T}\int _{ \Omega }\partial _{t}\psi \ \zeta \varrho \mathbf{u} \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\varrho ]\ \mathrm{d}x\ \mathrm{d}t.}$$

Combining (3.191) with (3.215) we get

$$\displaystyle{\varrho _{\varepsilon } \rightarrow \varrho \ \mbox{ in}\ C_{\mathrm{weak}}([0,T];L^{\Gamma }(\Omega )).}$$

In accordance with the standard theory of elliptic problems, the pseudodifferential operator \((\nabla _{x}\Delta _{x}^{-1})\) “gains” one spatial derivative, in particular, by virtue of the embedding \(W^{1,\Gamma }(\Omega )\hookrightarrow C(\overline{\Omega })\), we get

$$\displaystyle{(\nabla _{x}\Delta ^{-1})[1_{ \Omega }\varrho _{\varepsilon }] \rightarrow (\nabla _{x}\Delta ^{-1})[1_{ \Omega }\varrho ]\ \mbox{ in}\ C([0,T] \times \overline{\Omega }; \mathbb{R}^{3})}$$

provided \(\Gamma> 3\) (see Theorem 11.33 in Appendix). Consequently, we can use relations (3.191), (3.196), (3.228)–(3.232) in order to see that (i) I 2,ɛ , I 8,ɛ → 0, while (ii) the integrals I j, ɛ , j = 3, , 7, converge for ɛ → 0 to their counterparts in (3.239). We infer that

$$\displaystyle{ \lim _{\varepsilon \rightarrow 0}\int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(p_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\varrho _{\varepsilon } - \mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.240)
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\overline{p_{\delta }(\varrho,\vartheta )}\varrho - \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{+\lim _{\varepsilon \rightarrow 0}\int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \cdot \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }] - (\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \otimes \mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{-\int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho \mathbf{u} \cdot \mathcal{R}[1_{\Omega }\varrho \mathbf{u}] - (\varrho \mathbf{u} \otimes \mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\Big)\ \mathrm{d}x\ \mathrm{d}t.}$$

Now, the crucial observation is that the difference of the two most right quantities in (3.240) vanishes. In order to see this, we need the following assertion (Theorem 11.34 in Appendix) that can be viewed as a straightforward consequence of the Div-Curl lemma.

Lemma 3.5

Let

$$\displaystyle{\begin{array}{c} \mathbf{U}_{\varepsilon } \rightarrow \mathbf{U}\ \mathit{\mbox{ weakly in}}\ L^{\,p}(\mathbb{R}^{3}; \mathbb{R}^{3}),\\ \\ \mathbf{V}_{\varepsilon } \rightarrow \mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{q}(\mathbb{R}^{3}; \mathbb{R}^{3}),\end{array} }$$

where

$$\displaystyle{\frac{1} {p} + \frac{1} {q} = \frac{1} {r} <1.}$$

Then

$$\displaystyle{\mathbf{U}_{\varepsilon } \cdot \mathcal{R}[\mathbf{V}_{\varepsilon }] -\mathcal{R}[\mathbf{U}_{\varepsilon }] \cdot \mathbf{V}_{\varepsilon } \rightarrow \mathbf{U} \cdot \mathcal{R}[\mathbf{V}] -\mathcal{R}[\mathbf{U}] \cdot \mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{r}(\mathbb{R}^{3}).}$$

This statement provides the following corollary:

Corollary 3.3

Let

$$\displaystyle{\mathbf{V}_{\varepsilon } \rightarrow \mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{\,p}(\mathbb{R}^{3}; \mathbb{R}^{3}),}$$
$$\displaystyle{r_{\varepsilon } \rightarrow r\ \mathit{\mbox{ weakly in}}\ L^{q}(\mathbb{R}^{3}),}$$

where

$$\displaystyle{\frac{1} {p} + \frac{1} {q} = \frac{1} {s} <1.}$$

Then

$$\displaystyle{r_{\varepsilon }\mathcal{R}[\mathbf{V}_{\varepsilon }] -\mathcal{R}[r_{\varepsilon }]\mathbf{V}_{\varepsilon } \rightarrow r\mathcal{R}[\mathbf{V}] -\mathcal{R}[r]\mathbf{V}\ \mathit{\mbox{ weakly in}}\ L^{s}(\mathbb{R}^{3}; \mathbb{R}^{3}).}$$

Hereafter, we shall use Corollary 3.3 to show that

$$\displaystyle{ \lim _{\varepsilon \rightarrow 0}\int _{0}^{T}\int _{ \Omega }\psi \zeta \mathbf{u}_{\varepsilon } \cdot \Big (\varrho _{\varepsilon }\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }] -\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.241)
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \mathbf{u} \cdot \Big (\varrho \mathcal{R}[1_{\Omega }\varrho \mathbf{u}] -\mathcal{R}[1_{\Omega }\varrho ]\varrho \mathbf{u}\Big)\ \mathrm{d}x\ \mathrm{d}t,}$$

where, recall, \(\mathcal{R}[\mathbf{v}]\) is a vector field with i-th component \(\sum _{j=1}^{3}\mathcal{R}_{i,j}[v_{j}]\) while \(\mathcal{R}[a]\mathbf{v}\) is a vector field with i-th component \(\sum _{j=1}^{3}\mathcal{R}_{i,j}[a]v_{j}\).

As shown in (3.195), (3.231),

$$\displaystyle{\left \{\begin{array}{c} \varrho _{\varepsilon }(t,\cdot ) \rightarrow \varrho (t,\cdot )\,\mbox{ weakly in }L^{\Gamma }(\Omega ),\\ \\ (\varrho _{\varepsilon }\mathbf{u}_{\varepsilon })(t,\cdot ) \rightarrow (\varrho \mathbf{u})(t,\cdot )\,\mbox{ weakly in }L^{ \frac{2\Gamma } {\Gamma +1} }(\Omega; \mathbb{R}^{3}) \end{array} \right \}\ \mbox{ for all}\ t \in [0,T].}$$

Applying Corollary 3.3 to r ɛ = ϱ ɛ (t, ⋅ ), U ɛ = ϱ ɛ u ɛ (t, ⋅ ) (extended by 0 outside \(\Omega\)), we obtain

$$\displaystyle{\begin{array}{c} \left (\varrho _{\varepsilon }\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }] -\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }\right )(t,\cdot ) \rightarrow \left (\varrho \mathcal{R}[1_{\Omega }\varrho \mathbf{u}] -\mathcal{R}[1_{\Omega }\varrho ]\varrho \mathbf{u}\right )(t,\cdot ))\\ \\ \mbox{ weakly in}\,L^{ \frac{2\Gamma } {\Gamma +3} }(\Omega ),\,\mbox{ provided }\Gamma> \frac{9} {2} \end{array} }$$

for all t ∈ [0, T].

As the embedding \(L^{ \frac{2\Gamma } {\Gamma +3} }(\Omega )\hookrightarrow W^{-1,2}(\Omega )\) is compact for \(\Gamma> 9/2\), we conclude that

$$\displaystyle{ \begin{array}{c} \varrho _{\varepsilon }\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }] -\mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\varrho _{\varepsilon }\mathbf{u}_{\varepsilon } \rightarrow \varrho \mathcal{R}[1_{\Omega }\varrho \mathbf{u}] -\mathcal{R}[1_{\Omega }\varrho ]\varrho \mathbf{u}\\ \\ \mbox{ in}\ L^{q}(0,T;W^{-1,2}(\Omega; \mathbb{R}^{3}))\ \mbox{ for any}\ q \geq 1, \end{array} }$$
(3.242)

which, together with (3.192), yields (3.241). Consequently, (3.240) reduces to

$$\displaystyle{ \lim _{\varepsilon \rightarrow 0}\int _{0}^{T}\int _{ \Omega }\psi \zeta \Big(p_{\delta }(\varrho _{\varepsilon },\vartheta _{\varepsilon })\varrho _{\varepsilon } - \mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.243)
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\overline{p_{\delta }(\varrho,\vartheta )}\varrho - \mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\Big)\ \mathrm{d}x\ \mathrm{d}t.}$$

Our next goal is to replace in (3.243) the quantity \(\mathbb{S}_{\delta }(\vartheta _{\varepsilon },\nabla _{x}\mathbf{u}_{\varepsilon }): \mathcal{R}[1_{\Omega }\varrho _{\varepsilon }]\) by \(\varrho _{\varepsilon }\Big(\frac{4} {3}\mu _{\delta }(\vartheta _{\varepsilon })\) \(+\eta (\vartheta _{\varepsilon })\Big)\mathrm{div}_{x}\mathbf{u}_{\varepsilon }\), and, similarly, \(\mathbb{S}_{\delta }(\vartheta,\nabla _{x}\mathbf{u}): \mathcal{R}[1_{\Omega }\varrho ]\) by the expression \(\varrho \Big(\frac{4} {3}\mu _{\delta }(\vartheta )\) \(+\eta (\vartheta )\Big)\mathrm{div}_{x}\mathbf{u}\) in (3.243), where μ δ (ϑ) = μ(ϑ) + δϑ.

To this end write

$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\psi \zeta \mu _{\delta }(\vartheta _{\varepsilon })\Big(\nabla _{x}\mathbf{u}_{\varepsilon } + \nabla _{x}^{T}\mathbf{u}_{\varepsilon }\Big): \mathcal{R}[1_{ \Omega }\varrho _{\varepsilon }]\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \mathcal{R}:\Big [\zeta \mu _{\delta }(\vartheta _{\varepsilon })\Big(\nabla _{x}\mathbf{u}_{\varepsilon } + \nabla _{x}^{T}\mathbf{u}_{\varepsilon }\Big)\Big]\varrho _{\varepsilon }\ \mathrm{d}x\ \mathrm{d}t,}$$

and

$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\psi \zeta \mu _{\delta }(\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big): \mathcal{R}[1_{ \Omega }\varrho ]\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \mathcal{R}:\Big [\zeta \mu _{\delta }(\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big)\Big]\varrho \ \mathrm{d}x\ \mathrm{d}t,}$$

where we have used the evident properties of the double Riesz transform recalled in Sect. 11.17 in Appendix. Furthermore,

$$\displaystyle{\mathcal{R}:\Big [\zeta \mu _{\delta }(\vartheta _{\varepsilon })\Big(\nabla _{x}\mathbf{u}_{\varepsilon } + \nabla _{x}^{T}\mathbf{u}_{\varepsilon }\Big)\Big] = 2\zeta \mu _{\delta }(\vartheta _{\varepsilon })\mathrm{div}_{ x}\mathbf{u}_{\varepsilon } +\omega (\vartheta _{\varepsilon },\mathbf{u}_{\varepsilon }),}$$

and

$$\displaystyle{\mathcal{R}:\Big [\zeta \mu _{\delta }(\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big)\Big] = 2\zeta \mu _{\delta }(\vartheta )\mathrm{div}_{ x}\mathbf{u} +\omega (\vartheta,\mathbf{u}),}$$

with the commutator

$$\displaystyle{\omega (\vartheta,\mathbf{u}) = \mathcal{R}:\Big [\zeta \mu _{\delta }(\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big)\Big] -\zeta \mu _{\delta }(\vartheta )\mathcal{R}:\Big [\nabla _{ x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big].}$$

In order to proceed, we report the following result in the spirit of Coifman and Meyer [62] proved as Theorem 11.35 in Appendix.

      Commutator Lemma:

Lemma 3.6

Let \(w \in W^{1,2}(\mathbb{R}^{3})\) and \(\mathbf{Z} \in L^{\,p}(\mathbb{R}^{3}; \mathbb{R}^{3})\) be given, where \(\frac{6} {5} <p <\infty\) .

Then for any \(1 <s <\frac{6p} {6+p}\) ,

$$\displaystyle{\Big\|\mathcal{R}[w\mathbf{Z}] - w\mathcal{R}[\mathbf{Z}]\Big\|_{W^{\beta,s}(\mathbb{R}^{3};\mathbb{R}^{3})} \leq c\|w\|_{W^{1,2}(\mathbb{R}^{3})}\|\mathbf{Z}\|_{L^{\,p}(\mathbb{R}^{3};\mathbb{R}^{3})},}$$

where \(0 <\beta = \frac{3} {s} -\frac{6+p} {6p} <1\) and c = c( p, s) is a positive constant.

Applying Lemma 3.6 to w = w ɛ = ζ(μ(ϑ ɛ ) + δϑ ɛ ), Z = Z ɛ = [Z ɛ, 1, Z ɛ, 2, Z ɛ, 3], with \(Z_{\varepsilon,i} = \partial _{x_{i}}u_{\varepsilon,j} + \partial _{x_{j}}u_{\varepsilon,i}\), j = 1, 2, 3, where {w ɛ } ɛ > 0, {Z ɛ } ɛ > 0 are bounded in \(L^{2}(0,T;W^{1,2}(\Omega ))\) and \(L^{2}((0,T) \times \Omega; \mathbb{R}^{3})\), respectively, cf. (3.192), (3.200), (3.17)–(3.18), we deduce that

$$\displaystyle{ \{\omega (\vartheta _{\varepsilon },\mathbf{u}_{\varepsilon })\}_{\varepsilon>0}\ \mbox{ is bounded in}\ L^{1}(0,T;W^{\beta,s}(\Omega )) }$$
(3.244)

for certain \(1 <s <\frac{3} {2}\), \(0 <\beta = \frac{3-2s} {s} <1\).

Next we claim that

$$\displaystyle{ \omega (\vartheta _{\varepsilon },\mathbf{u}_{\varepsilon })\varrho _{\varepsilon } \rightarrow \overline{\omega (\vartheta,\mathbf{u})}\varrho \ \mbox{ weakly in}\ L^{1}((0,T) \times \Omega ), }$$
(3.245)

where, in accordance with relations (3.17), (3.23), (3.192), (3.193), and (3.211),

$$\displaystyle{ \overline{\omega (\vartheta,\mathbf{u})} =\omega (\vartheta,\mathbf{u}). }$$
(3.246)

In order to show (3.245), we apply the Div-Curl Lemma (see Proposition 3.3) to the four-component vector fields

$$\displaystyle{\mathbf{U}_{\varepsilon } = [\varrho _{\varepsilon },\varrho _{\varepsilon }\mathbf{u}_{\varepsilon }],\ \mathbf{V}_{\varepsilon } = [\omega (\vartheta _{\varepsilon },\mathbf{u}_{\varepsilon }),0,0,0].}$$

In view of relations (3.171), (3.194), (3.244) yielding the sequences {div t, x U ɛ } ɛ > 0 and {curl t, x V ɛ } ɛ > 0 compact in \(W^{-1,s}((0,T) \times \Omega )\) and \(W^{-1,s}((0,T) \times \Omega; \mathbb{R}^{3\times 3})\) for a certain s > 1, it is enough to observe that

$$\displaystyle{\left.\begin{array}{r} \{\mathbf{U}_{\varepsilon }\}_{\varepsilon>0}\\ \\ \{\mathbf{V}_{\varepsilon }\}_{\varepsilon>0} \end{array} \right \}\mbox{ are bounded in}\left \{\begin{array}{l} L^{q}((0,T) \times \Omega; \mathbb{R}^{4})),\\ \\ L^{r}((0,T) \times \Omega; \mathbb{R}^{4})), \end{array} \right.\ \mbox{ respectively,}}$$

with 1∕r + 1∕q < 1. This is certainly true provided \(\Gamma\) is large enough.

Relations (3.243), (3.245), (3.246) give rise to a remarkable identity

     Weak Compactness Identity for Effective Pressure (Level ɛ):

$$\displaystyle{ \overline{p_{\delta }(\varrho,\vartheta )\varrho } -\Big (\frac{4} {3}\mu (\vartheta ) + \frac{4} {3}\delta \vartheta +\eta (\vartheta )\Big)\overline{\varrho \mathrm{div}_{x}\mathbf{u}} = }$$
(3.247)
$$\displaystyle{\overline{p_{\delta }(\varrho,\vartheta )}\varrho -\Big (\frac{4} {3}\mu (\vartheta ) + \frac{4} {3}\delta \vartheta +\eta (\vartheta )\Big)\varrho \mathrm{div}_{x}\mathbf{u},}$$

where the quantity \(p - (\frac{4} {3}\mu +\eta )\mathrm{div}_{x}\mathbf{u}\) is usually termed effective viscous flux or effective pressure. As we will see below, the quantity

$$\displaystyle{\overline{\varrho \mathrm{div}_{x}\mathbf{u}} -\varrho \mathrm{div}_{x}\mathbf{u}}$$

plays a role of a “defect” measure of the density oscillations described through the (renormalized) equation of continuity. Relation (3.247) enables us to relate these oscillations to the changes in the pressure.

In order to exploit (3.247), we multiply the approximate continuity equation (3.171) on G′(ϱ ɛ ), where G is a smooth convex function, integrate by parts, and let ɛ → 0 to obtain

$$\displaystyle{ \int _{\Omega }\overline{G(\varrho )}(\tau )\ \mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\overline{\Big(G'(\varrho )\varrho - G(\varrho )\Big)\mathrm{div}_{x}\mathbf{u}}\ \mathrm{d}x\ \mathrm{d}t \leq \int _{\Omega }G(\varrho _{0,\delta })\ \mathrm{d}x }$$
(3.248)

from which we easily deduce that

$$\displaystyle{ \int _{\Omega }\overline{\varrho \log (\varrho )}(\tau )\ \mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\overline{\varrho \mathrm{div}_{x}\mathbf{u}}\ \mathrm{d}x\ \mathrm{d}t = \int _{\Omega }\varrho _{0,\delta }\log (\varrho _{0,\delta })\ \mathrm{d}x }$$
(3.249)

for a.a. τ ∈ (0, T).

To derive a relation similar to (3.249) for the limit functions ϱ, u, we need the renormalized continuity equation introduced in (1.20). Note that we have already shown that the quantities ϱ, u solve the continuity equation (3.197) in \((0,T) \times \mathbb{R}^{3}\). On the other hand, the general theory of transport equations developed by DiPerna-Lions asserts that any solution of (3.197) is automatically a renormalized one as soon as, roughly speaking, the quantity ϱdiv x u is integrable.

More precisely, we report the following result proved in Sect. 11.19 in Appendix.

Lemma 3.7

Assume that \(\varrho \in L^{2}((0,T) \times \mathbb{R}^{3})\) , \(\mathbf{u} \in L^{2}(0,T;W^{1,2}(\mathbb{R}^{3}))\) solve the equation of continuity ( 3.197 ) in \(\mathcal{D}'((0,T) \times \mathbb{R}^{3}))\) .

Then ϱ, u represent a renormalized solution in the sense specified in ( 2.2 ).

As a consequence of Lemma 3.7 (see also Theorem 11.36 and Lemma 11.13 for more details), we deduce

$$\displaystyle{ \int _{\Omega }\varrho \log (\varrho )(\tau )\ \mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\varrho \mathrm{div}_{x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t \leq \int _{\Omega }\varrho _{0,\delta }\log (\varrho _{0,\delta })\ \mathrm{d}x. }$$
(3.250)

Since the pressure p δ is non-decreasing with respect to ϱ and we already know that ϑ ɛ ϑ strongly in \(L^{1}((0,T) \times \Omega )\), we have

$$\displaystyle{\overline{p_{\delta }(\varrho,\vartheta )\varrho } \geq \overline{p_{\delta }(\varrho,\vartheta )}\varrho.}$$

Indeed,

$$\displaystyle{\lim _{n\rightarrow \infty }\int _{B}\Big(\,p_{\delta }(\varrho _{n},\vartheta _{n})\varrho _{n} - p_{\delta }(\varrho _{n},\vartheta _{n})\varrho \Big)\mathrm{d}x\ \mathrm{d}t =}$$
$$\displaystyle{\lim _{n\rightarrow \infty }\int _{B}\Big(\,p_{\delta }(\varrho _{n},\vartheta _{n}) - p_{\delta }(\varrho,\vartheta _{n})\Big)(\varrho _{n}-\varrho )\mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{+\lim _{n\rightarrow \infty }\int _{B}p_{\delta }(\varrho,\vartheta _{n})(\varrho _{n}-\varrho )\mathrm{d}x\ \mathrm{d}t,}$$

where the first term is non-negative, and the second term tends to zero by virtue of the asymptotic limits established in (3.191), (3.211), the bounds (3.193), (3.200), (3.225), (3.226), and the structural properties of p δ stated in (3.229).

Consequently, relation (3.247) yields

$$\displaystyle{\overline{\varrho \mathrm{div}_{x}\mathbf{u}} \geq \varrho \ \mathrm{div}_{x}\mathbf{u};}$$

whence (3.249) together with (3.250) imply the desired conclusion

$$\displaystyle{\overline{\varrho \log (\varrho )} =\varrho \log (\varrho ).}$$

As zzlog(z) is a strictly convex function, we may infer that

$$\displaystyle{ \varrho _{\varepsilon } \rightarrow \varrho \ \mbox{ a.a. in}\ (0,T) \times \Omega, }$$
(3.251)

in agreement with Theorem 11.27 in Appendix.

3.6.6 Artificial Diffusion Asymptotic Limit

Strong convergence of the sequence of approximate densities established in (3.251) completes the second step in the proof of Theorem 3.1 eliminating completely the ɛ-dependent terms in the approximate system. For any δ > 0, we have constructed a trio {ϱ, u, ϑ} solving the following problem:

  1. (i)

    Renormalized continuity equation:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varrho B(\varrho )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
    (3.252)
    $$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }b(\varrho )\mathrm{div}_{x}\mathbf{u}\varphi \ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }\varrho _{0,\delta }B(\varrho _{0,\delta })\varphi (0,\cdot )\ \mathrm{d}x}$$

    for any

    $$\displaystyle{b \in L^{\infty }\cap C[0,\infty ),\ B(\varrho ) = B(1) +\int _{ 1}^{\varrho }\frac{b(z)} {z^{2}} \ \mathrm{d}z,}$$

    and any test function

    $$\displaystyle{\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega }).}$$
  2. (ii)

    Approximate balance of momentum:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \mathbf{u} \cdot \partial _{t}\boldsymbol{\varphi } +\varrho [\mathbf{u} \otimes \mathbf{u}]: \nabla _{x}\boldsymbol{\varphi } +\Big (\,p +\delta (\varrho ^{\Gamma } +\varrho ^{2})\Big)\mathrm{div}_{ x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
    (3.253)
    $$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\Big(\mathbb{S}_{\delta }: \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}_{\delta }\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }(\varrho \mathbf{u})_{0}\cdot \boldsymbol{\varphi }\ \mathrm{d}x,}$$

    for any test function \(\boldsymbol{\varphi }\in C_{c}^{\infty }([0,T) \times \overline{\Omega }; \mathbb{R}^{3})\), where either

    $$\displaystyle{ \boldsymbol{\varphi }\cdot \mathbf{n}\vert _{\partial \Omega } = 0\ \mbox{ in the case of the complete slip boundary conditions,} }$$
    (3.254)

    or

    $$\displaystyle{ \boldsymbol{\varphi }\vert _{\partial \Omega } = 0\ \mbox{ in the case of the no-slip boundary conditions.} }$$
    (3.255)

    Furthermore,

    $$\displaystyle{ \mathbb{S}_{\delta } = (\mu (\vartheta )+\delta \vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{\perp }\mathbf{u} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}\Big) +\eta (\vartheta )\mathrm{div}_{x}\mathbf{u}\ \mathbb{I}. }$$
    (3.256)
  3. (iii)

    Approximate entropy balance:

    $$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\varrho s_{\delta }(\varrho,\vartheta )\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
    (3.257)
    $$\displaystyle{+\int _{0}^{T}\int _{ \Omega }\frac{\kappa _{\delta }(\vartheta )\nabla _{x}\vartheta } {\vartheta } \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t+ <\sigma _{\delta };\varphi> _{[\mathcal{M},C]([0,T]\times \overline{\Omega })} =}$$
    $$\displaystyle{-\int _{\Omega }(\varrho s)_{0,\delta }\varphi (0,\cdot )\ \mathrm{d}x -\int _{0}^{T}\int _{ \Omega }\frac{\varrho } {\vartheta }\mathcal{Q}_{\delta }\varphi \ \mathrm{d}x\ \mathrm{d}t}$$

    for all \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\), where \(\sigma _{\delta } \in \mathcal{M}^{+}([0,T] \times \overline{\Omega })\) satisfies

    $$\displaystyle{ \sigma _{\delta } \geq \frac{1} {\vartheta } \Big[\mathbb{S}_{\delta }: \nabla _{x}\mathbf{u} +\Big (\frac{\kappa (\vartheta )} {\vartheta } + \frac{\delta } {2}(\vartheta ^{\Gamma -1} + \frac{1} {\vartheta ^{2}} )\Big)\vert \nabla _{x}\vartheta \vert ^{2} +\delta \frac{1} {\vartheta ^{2}} \Big], }$$
    (3.258)

    and where we have set

    $$\displaystyle{ s_{\delta }(\varrho,\vartheta ) = s(\varrho,\vartheta ) +\delta \log (\vartheta ),\ \kappa _{\delta }(\vartheta ) =\kappa (\vartheta ) +\delta \Big (\vartheta ^{\Gamma } + \frac{1} {\vartheta } \Big). }$$
    (3.259)
  4. (iv)

    Approximate energy balance:

    $$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} +\varrho e(\varrho,\vartheta ) + \delta ( \frac{\varrho ^{\Gamma }} {\Gamma - 1} +\varrho ^{2})\Big)(\tau )\ \mathrm{d}x }$$
    (3.260)
    $$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0,\delta }} +\varrho _{0,\delta }e_{0,\delta } + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
    $$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f}_{\delta } \cdot \mathbf{u} +\varrho \mathcal{Q}_{\delta } +\delta \frac{1} {\vartheta ^{2}} \Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for a.a.}\ \tau \in [0,T].}$$

3.7 Vanishing Artificial Pressure

The last and probably the most illuminative step in the proof of Theorem 3.1 is to let δ → 0 in the approximate system (3.252)–(3.260). Although many arguments are almost identical or mimick closely those discussed in the previous text, there are still some new ingredients coming into play. Notably, we introduce a concept of oscillation defect measure in order to control the density oscillations beyond the theory of DiPerna and Lions. Moreover, weighted estimates of this quantity are used in order to accommodate the physically realistic growth restrictions on the transport coefficients imposed through hypotheses (3.17), (3.23).

3.7.1 Uniform Estimates

From now on, let {ϱ δ , u δ , ϑ δ } δ > 0 be a family of approximate solutions satisfying (3.252)–(3.260). To begin, we recall that the total mass is a constant of motion, specifically,

$$\displaystyle{ \int _{\Omega }\varrho _{\delta }(t,\cdot )\ \mathrm{d}x = \int _{\Omega }\varrho _{0,\delta }\ \mathrm{d}x\ \mbox{ for any}\ t \in [0,T]. }$$
(3.261)

Since we assume that

$$\displaystyle{ \varrho _{0,\delta } \rightarrow \varrho _{0}\ \mbox{ in}\ L^{1}(\Omega ), }$$
(3.262)

the bound (3.261) is uniform for δ → 0.

The next step is the dissipation balance

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho _{\delta }\vert \mathbf{u}_{\delta }\vert ^{2}(\tau ) + H_{\overline{\vartheta }}(\varrho _{\delta },\vartheta _{\delta })(\tau ) +\delta ( \frac{1} {\Gamma - 1}\varrho _{\delta }^{\Gamma } + \varrho _{\delta }^{2})(\tau )\Big)\ \mathrm{d}x }$$
(3.263)
$$\displaystyle{+\overline{\vartheta }\ \sigma _{\delta }\Big[[0,\tau ] \times \overline{\Omega }\Big]}$$
$$\displaystyle{= \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0,\delta }} + H_{\overline{\vartheta }}(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x}$$
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho _{\delta }\mathbf{f}_{\delta } \cdot \mathbf{u}_{\delta } + \varrho _{\delta }(1 -\frac{\overline{\vartheta }} {\vartheta _{\delta }})\mathcal{Q}_{\delta } +\delta \frac{1} {\vartheta _{\delta }^{2}}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$

satisfied for a.a. τ ∈ [0, T], which can be deduced from (3.257), (3.260), with the Helmholtz function \(H_{\overline{\vartheta }}\) introduced in (2.48). Accordingly, in order to get uniform estimates, we have to take

$$\displaystyle{ \begin{array}{c} \{\mathbf{f}_{\delta }\}_{\delta>0}\ \mbox{ bounded in}\ L^{\infty }((0,T) \times \Omega; \mathbb{R}^{3}),\\ \\ \mathcal{Q}_{\delta }\geq 0,\ \{\mathcal{Q}_{\delta }\}_{\delta>0}\ \mbox{ bounded in}\ L^{\infty }((0,T) \times \Omega ) \end{array} }$$
(3.264)

as well as

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0,\delta }} + H_{\overline{\vartheta }}(\varrho _{0,\delta },\vartheta _{0,\delta }) + \delta ( \frac{\varrho _{0,\delta }^{\Gamma }} {\Gamma - 1} +\varrho _{ 0,\delta }^{2})\Big)\ \mathrm{d}x \leq c }$$
(3.265)

uniformly for δ → 0.

As the term δϑ δ 2 is “absorbed” by its counterpart in the entropy production σ δ satisfying (3.258), the dissipation balance (3.263) gives rise, exactly as in Sect. 3.6.1, to the following uniform estimates:

$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\|\sqrt{\varrho _{\delta }}\mathbf{u}_{\delta }(t)\|_{L^{2}(\Omega;\mathbb{R}^{3})} \leq c, }$$
(3.266)
$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\|\varrho _{\delta }(t)\|_{ L^{\frac{5} {3} }(\Omega )} \leq c, }$$
(3.267)
$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\|\varrho _{\delta }(t)\|_{L^{\Gamma }(\Omega )} \leq \delta ^{-\frac{1} {\Gamma } }c, }$$
(3.268)

and

$$\displaystyle{ \mathrm{ess}\sup _{t\in (0,T)}\|\vartheta _{\delta }(t)\|_{L^{4}(\Omega )} \leq c. }$$
(3.269)

In addition, we have

$$\displaystyle{ \sigma _{\delta }\Big[[0,T] \times \overline{\Omega }\Big] \leq c, }$$
(3.270)

and, as a consequence of (3.258),

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\vert \nabla _{x}\log (\vartheta _{\delta })\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.271)
$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\vert \nabla _{x}\vartheta _{\delta }^{\frac{3} {2} }\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.272)

and

$$\displaystyle{ \delta \int _{0}^{T}\int _{ \Omega }\frac{1} {\vartheta _{\delta }^{3}}\ \mathrm{d}x\ \mathrm{d}t \leq c, }$$
(3.273)
$$\displaystyle{ \delta \int _{0}^{T}\int _{ \Omega }\Big(\vartheta _{\delta }^{\Gamma -2} + \frac{1} {\vartheta _{\delta }^{3}}\Big)\vert \nabla _{x}\vartheta _{\delta }\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c. }$$
(3.274)

Finally, making use of Korn’s inequality established in Proposition 2.1 we deduce, exactly as in (2.65), (2.66), that

$$\displaystyle{ \|\ \mathbf{u}_{\delta }\ \|_{L^{2}(0,T;W^{1,p}(\Omega;\mathbb{R}^{3}))} \leq c\ \mbox{ for}\ p = \frac{8} {5-\alpha }, }$$
(3.275)

and

$$\displaystyle{ \|\ \mathbf{u}_{\delta }\ \|_{L^{q}(0,T;W^{1,s}(\Omega;\mathbb{R}^{3}))} \leq c\ \mbox{ for}\ q = \frac{6} {4-\alpha },\ s = \frac{18} {10-\alpha }, }$$
(3.276)

where α was introduced in hypotheses (3.17)–(3.23). Moreover,

$$\displaystyle{ \delta \int _{0}^{T}\int _{ \Omega }\Big\vert \nabla _{x}\mathbf{u}_{\delta } + \nabla _{x}^{T}\mathbf{u}_{\delta } -\frac{2} {3}\mathbb{I}\Big\vert ^{2}\ \mathrm{d}x\ \mathrm{d}t \leq c. }$$
(3.277)

Note that estimates (3.269)–(3.272) yield

$$\displaystyle{ \{\vartheta _{\delta }^{\beta }\}_{\delta>0}\ \mbox{ bounded in}\ L^{2}(0,T;W^{1,2}(\Omega ))\ \mbox{ for any}\ 1 \leq \beta \leq \frac{3} {2}, }$$
(3.278)

while (3.275), (3.276), together with hypotheses (3.17), (3.19), and (3.23), imply that

$$\displaystyle{ \{\mathbb{S}_{\delta }\}_{\delta>0}\ \mbox{ is bounded in}\ L^{q}((0,T) \times \Omega; \mathbb{R}^{3\times 3}))\ \mbox{ for a certain}\ q> 1, }$$
(3.279)

(cf. estimate (2.68)).

Now, positivity of the absolute temperature can be shown by help of Proposition 2.2 and Lemma 2.1, exactly as in Sect. 2.2.4. In particular, estimate (3.271) can be strengthened to

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\vert \log \vartheta _{\delta }\vert ^{2} + \vert \nabla _{ x}\log \vartheta _{\delta }\vert ^{2}\Big)\ \mathrm{d}x\ \mathrm{d}t \leq c. }$$
(3.280)

In order to complete our list of uniform bounds, we evoke the pressure estimates obtained in Sect. 2.2.5. In the present context, relation (2.95) reads

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\delta \varrho _{\delta }^{\Gamma } + p_{\delta }(\varrho _{\delta },\vartheta _{\delta })\Big)\varrho _{\delta }^{\nu }\ \mathrm{d}x\ \mathrm{d}t \leq c(\mathrm{data}), }$$
(3.281)

where ν > 0 is a constant exponent.

3.7.2 Asymptotic Limit for Vanishing Artificial Pressure

The piece of information provided by the uniform bounds established in the previous section is sufficient for taking δ → 0 in the approximate system of equations (3.252)–(3.260).

Due to the structural properties of the molecular pressure p M derived in (3.32), and because of (3.229), estimates (3.267), (3.269), and (3.275), (3.276) imply that

$$\displaystyle{ \varrho _{\delta } \rightarrow \varrho \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{\frac{5} {3} }(\Omega )), }$$
(3.282)
$$\displaystyle{ \vartheta _{\delta } \rightarrow \vartheta \ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{4}(\Omega )), }$$
(3.283)

and

$$\displaystyle{ \mathbf{u}_{\delta } \rightarrow \mathbf{u}\left \{\begin{array}{l} \ \mbox{ weakly in}\ L^{2}(0,T;W^{1,p}(\Omega; \mathbb{R}^{3})),\ p = \frac{8} {5-\alpha },\\ \\ \ \mbox{ weakly in}\ L^{q}(0,T;W^{1,s}(\Omega; \mathbb{R}^{3})),\ q = \frac{6} {4-\alpha },\ s = \frac{18} {10-\alpha }, \end{array} \right \} }$$
(3.284)

at least for suitable subsequences.

Taking b ≡ 0 in the renormalized Eq. (3.252), we deduce, in view of the previous estimates, that

$$\displaystyle{ \varrho _{\delta } \rightarrow \varrho \ \mbox{ in}\ C_{\mathrm{weak}}([0,T];L^{\frac{5} {3} }(\Omega )). }$$
(3.285)

On the other hand, as the Lebesgue space \(L^{\frac{5} {3} }(\Omega )\) is compactly embedded into the dual \(W^{-1,p'}(\Omega )\), p′ = 8∕(3 + α) as soon as α ∈ (2∕5, 1], we conclude, taking (3.282) together with (3.266), (3.267) into account, that

$$\displaystyle{ \varrho _{\delta }\mathbf{u}_{\delta } \rightarrow \varrho \mathbf{u}\ \mbox{ weakly-(*) in}\ L^{\infty }(0,T;L^{\frac{5} {4} }(\Omega; \mathbb{R}^{3})). }$$
(3.286)

A similar argument in the case when the time derivative of the momentum ϱ δ u δ is expressed via the approximate momentum equation (3.253) gives rise to

$$\displaystyle{ \varrho _{\delta }\mathbf{u}_{\delta } \rightarrow \varrho \mathbf{u}\ \mbox{ in}\ C_{\mathrm{weak}}([0,T];L^{\frac{5} {4} }(\Omega; \mathbb{R}^{3})). }$$
(3.287)

Since

$$\displaystyle{ W^{1,s}(\Omega )\ \mbox{ is compactly embedded into}\ L^{5}(\Omega )\ \mbox{ for}\ s = \frac{18} {10-\alpha }, }$$
(3.288)

we can use (3.284) to conclude that

$$\displaystyle{ \varrho _{\delta }\mathbf{u}_{\delta } \otimes \mathbf{u}_{\delta } \rightarrow \varrho \mathbf{u} \otimes \mathbf{u}\ \mbox{ weakly in}\ L^{q}(0,T;L^{q}(\Omega; \mathbb{R}^{3\times 3}))\ \mbox{ for a certain}\ q> 1. }$$
(3.289)

In order to handle the approximate pressure in the momentum equation (3.253), we first observe that, as a direct consequence of (3.281),

$$\displaystyle{ \delta \varrho _{\delta } \rightarrow 0\ \mbox{ in}\ L^{1}((0,T) \times \Omega ). }$$
(3.290)

Moreover, writing

$$\displaystyle{p(\varrho _{\delta },\vartheta _{\delta }) = p_{M}(\varrho _{\delta },\vartheta _{\delta }) + \frac{a} {3}\vartheta _{\delta }^{4},}$$

and interpolating estimates (3.269), (3.278), we have

$$\displaystyle{ \vartheta _{\delta }^{4} \rightarrow \overline{\vartheta ^{4}}\ \mbox{ weakly in}\ L^{q}((0,T) \times \Omega )\ \mbox{ for a certain}\ q> 1. }$$
(3.291)

In accordance with hypotheses (3.15), (3.16), the asymptotic structure of p M derived in (3.32), and in agreement with (3.281), (3.291),

$$\displaystyle{ p(\varrho _{\delta },\vartheta _{\delta }) = p_{M}(\varrho _{\delta },\vartheta _{\delta }) + \frac{a} {3}\vartheta _{\delta }^{4} \rightarrow \overline{p_{ M}(\varrho,\vartheta )} + \frac{a} {3}\overline{\vartheta ^{4}}\ \mbox{ weakly in}\ L^{1}((0,T) \times \Omega ). }$$
(3.292)

At this stage, it is possible to let δ → 0 in Eqs. (3.252), (3.253) to obtain

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\overline{\varrho B(\varrho )}\partial _{t}\varphi + \overline{\varrho B(\varrho )}\mathbf{u} \cdot \nabla _{x}\varphi -\overline{b(\varrho )\mathrm{div}_{x}\mathbf{u}}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.293)
$$\displaystyle{= -\int _{\Omega }\varrho _{0}B(\varrho _{0})\varphi (0,\cdot )\ \mathrm{d}x}$$

for any test function \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\) and any

$$\displaystyle{b \in L^{\infty }\cap C[0,\infty ),\ B(\varrho ) = B(1) +\int _{ 1}^{\varrho }\frac{b(z)} {z} \ \mathrm{d}z.}$$

Similarly, we get

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho \mathbf{u} \cdot \partial _{t}\boldsymbol{\varphi } +\varrho \mathbf{u} \otimes \mathbf{u}: \nabla _{x}\boldsymbol{\varphi } + (\overline{p_{M}(\varrho,\vartheta )} + \frac{a} {3}\overline{\vartheta ^{4}})\mathrm{div}_{x}\boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.294)
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\Big(\overline{\mathbb{S}(\vartheta,\nabla _{x}\mathbf{u})}: \nabla _{x}\boldsymbol{\varphi } -\varrho \mathbf{f}\cdot \boldsymbol{\varphi }\Big)\ \mathrm{d}x\ \mathrm{d}t -\int _{\Omega }(\varrho \mathbf{u})_{0} \cdot \boldsymbol{\varphi } (0,\cdot )\ \mathrm{d}x}$$

for any test function \(\boldsymbol{\varphi }\in C_{c}^{\infty }([0,T) \times \overline{\Omega }; \mathbb{R}^{3})\) satisfying \(\boldsymbol{\varphi }\cdot \mathbf{n}\vert _{\partial \Omega } = 0\), or, in addition, \(\boldsymbol{\varphi }\vert _{\partial \Omega } = 0\) in the case of the no-slip boundary conditions. Here we have set

$$\displaystyle{ \overline{\mathbb{S}(\vartheta,\nabla _{x}\mathbf{u})} = \overline{\mu (\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{\perp }\mathbf{u} -\frac{2} {3}\mathrm{div}_{x}\mathbf{u}\mathbb{I}\Big)} + \overline{\eta (\vartheta )\mathrm{div}_{x}\mathbf{u}\mathbb{I}}. }$$
(3.295)

Finally, letting δ → 0 in the approximate total energy balance (3.260) we conclude

$$\displaystyle{ \int _{\Omega }\Big(\frac{1} {2}\varrho \vert \mathbf{u}\vert ^{2} + \overline{\varrho e(\varrho,\vartheta )}\Big)(\tau )\ \mathrm{d}x = \int _{ \Omega }\Big(\frac{1} {2} \frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0}} +\varrho _{0}e(\varrho _{0},\vartheta _{0})\Big)\ \mathrm{d}x }$$
(3.296)
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\Big(\varrho \mathbf{f} \cdot \mathbf{u} +\varrho \mathcal{Q}\Big)\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for a.a.}\ \tau \in (0,T),}$$

where we have used estimate (3.273) in order to eliminate the singular term δϑ δ 2. Moreover, we have assumed strong convergence (a.a.) of the approximate data f δ , ϱ 0,δ , ϑ 0,δ , and \(\mathcal{Q}_{\delta }\).

3.7.3 Entropy Balance and Pointwise Convergence of the Temperature

Similarly to the preceding parts, specifically Sect. 3.6.2, our aim is to use Div-Curl lemma (Proposition 3.3), together with the monotonicity of the entropy, in order to show

$$\displaystyle{ \vartheta _{\delta } \rightarrow \vartheta \ \mbox{ a.a. on}\ (0,T) \times \Omega. }$$
(3.297)

Uniform Estimates We have to show that all terms appearing on the left-hand side of the approximate entropy balance (3.257) are either non-negative or belong to an L p-space, with p > 1.

To this end, we use the structural properties of the specific entropy s stated in (3.34), (3.39), together with the uniform estimates (3.267), (3.269), (3.280), to deduce that

$$\displaystyle{ \varrho _{\delta }s(\varrho _{\delta },\vartheta _{\delta }) \rightarrow \overline{\varrho s(\varrho,\vartheta )}\ \mbox{ weakly in}\ L^{\,p}((0,T) \times \Omega )\ \mbox{ for a certain}\ p> 1. }$$
(3.298)

Similarly, we have

$$\displaystyle{\vert \varrho _{\delta }s(\varrho _{\delta },\vartheta _{\delta })\mathbf{u}_{\delta }\vert \leq c\Big(\vert \vartheta _{\delta }\vert ^{3}\vert \mathbf{u}_{\delta }\vert + \varrho _{\delta }\vert \log (\varrho _{\delta })\vert \vert \mathbf{u}_{\delta }\vert + \vert \mathbf{u}_{\delta }\vert + \varrho _{\delta }\vert \log (\vartheta _{\delta })\vert \vert \mathbf{u}_{\delta }\vert \Big);}$$

whence, by virtue of (3.288), combined with estimates (3.282)–(3.284), there is p > 1 such that

$$\displaystyle{ \Big\{\vert \vartheta _{\delta }\vert ^{3}\vert \mathbf{u}_{\delta }\vert + \varrho _{\delta }\vert \log (\varrho _{\delta })\vert \vert \mathbf{u}_{\delta }\vert + \vert \mathbf{u}_{\delta }\vert \Big\}_{\delta>0}\ \mbox{ is bounded in}\ L^{\,p}((0,T) \times \Omega ). }$$
(3.299)

In addition, relations (3.277), (3.290) give rise to

$$\displaystyle{ \{\varrho _{\delta }\log (\vartheta _{\delta })\mathbf{u}_{\delta }\}_{\delta>0}\ \mbox{ bounded in}\ L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3})\ \mbox{ for a certain}\ p> 1. }$$
(3.300)

The entropy flux can be handled by means of the uniform estimates established in (3.269), (3.278). Indeed, writing

$$\displaystyle{\frac{\kappa (\vartheta _{\delta })} {\vartheta _{\delta }} \vert \nabla _{x}\vartheta _{\delta }\vert \leq c\Big(\vert \nabla _{x}\log (\vartheta _{\delta })\vert + \vartheta _{\delta }^{\frac{3} {2} }\vert \nabla _{x}\vartheta _{\delta }^{\frac{3} {2} }\vert \Big)}$$

we observe easily that

$$\displaystyle{ \Big\{\frac{\kappa (\vartheta _{\delta })} {\vartheta _{\delta }} \nabla _{x}\vartheta _{\delta }\Big\}_{\delta>0}\ \mbox{ is bounded in}\ L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3}) }$$
(3.301)

for a suitable p > 1.

Finally, relations (3.269), (3.274), (3.280) can be used to obtain

$$\displaystyle{ \left \{\begin{array}{c} \delta \int _{0}^{T}\|\vartheta _{\delta }^{\frac{\Gamma } {2} }(t,\cdot )\|_{W^{1,2}(\Omega )}^{2}\ \mathrm{d}t \leq c,\\ \\ \delta \int _{0}^{T}\|\vartheta _{\delta }^{-\frac{1} {2} }(t,\cdot )\|_{W^{1,2}(\Omega )}^{2}\ \mathrm{d}t \leq c,\end{array} \right \} }$$
(3.302)

uniformly for δ → 0. Consequently, seeing that

$$\displaystyle{\delta \vartheta _{\delta }^{\Gamma -1}\nabla _{ x}\vartheta _{\delta } =\delta \frac{\Gamma } {2} \vartheta _{\delta }^{\frac{\Gamma } {2} }\nabla _{x}\vartheta _{\delta }^{\frac{\Gamma } {2} } =\delta \frac{\varGamma } {2}\vartheta _{\delta }^{\frac{1} {4} }\vartheta _{\delta }^{ \frac{\Gamma } {2} -\frac{1} {4} }\nabla _{x}\vartheta _{\delta }^{\frac{\Gamma } {2} },}$$

we can use (3.283), (3.302), together with Hölder’s inequality and the embedding relation \(W^{1,2}(\Omega )\hookrightarrow L^{6}(\Omega )\), in order to conclude that

$$\displaystyle{ \delta \vartheta _{\delta }^{\Gamma -1}\nabla _{ x}\vartheta _{\delta } \rightarrow 0\ \mbox{ in}\ L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3}))\ \mbox{ for}\ \delta \rightarrow 0\ \mbox{ and a certain}\ p> 1. }$$
(3.303)

Similarly, by the same token,

$$\displaystyle{ \frac{\delta } {\vartheta _{\delta }^{2}}\nabla _{x}\vartheta _{\delta } \rightarrow 0\ \mbox{ in}\ L^{\,p}((0,T) \times \Omega; \mathbb{R}^{3})),\ \mbox{ where}\ p> 1. }$$
(3.304)

Strong Convergence of Temperature via the Young Measures Having established all necessary estimates we can proceed as in Sect. 3.6.2.

By virtue of (3.280),

$$\displaystyle{ \delta \log (\vartheta _{\delta })\,G(\vartheta _{\delta }) \rightarrow 0\;\mbox{ in }L^{1}((0,T) \times \Omega ). }$$
(3.305)

We can apply Div-Curl lemma (Proposition 3.3) in order to obtain identity

$$\displaystyle{ \overline{\varrho s(\varrho,\vartheta )G(\vartheta )} = \overline{\varrho s(\varrho,\vartheta )}\;\overline{G(\vartheta )}. }$$
(3.306)

Consequently, employing Theorem 11.37, we show identity (3.213). Now we apply Theorem 13 in the same way as in Sect. 3.6.2 and conclude that

$$\displaystyle{ \overline{\varrho s_{M}(\varrho,\vartheta )G(\vartheta )} \geq \overline{\varrho s_{M}(\varrho,\vartheta )}\;\overline{G(\vartheta )}. }$$
(3.307)

We also observe that, according to Theorem 11.26,

$$\displaystyle{ \overline{\vartheta ^{3}G(\vartheta )} \geq \overline{\vartheta ^{3}}\;\overline{G(\vartheta )}. }$$
(3.308)

The symbol G in the last four formulas denotes an arbitrary nondecreasing and continuous function on [0, ), chosen in such a way that all the L 1-weak limits in the above formulas exist.

Relations (3.307)–(3.308) combined with identity (3.306) yield (3.210). The latter identity implies the pointwise convergence (3.297).

Strong Convergence of Temperature: An Alternative Proof The departure point is formula (3.306) with G(ϑ) = T k (ϑ), where the truncation functions T k is defined by formula (3.316) below. The goal is to show the inequality (3.307) by using more elementary arguments than in the previous section. Once this is done, (3.306) and Theorem 11.26 yield

$$\displaystyle{\overline{\vartheta ^{3}T_{k}(\vartheta )} = \overline{\vartheta ^{3}}\;\overline{T_{k}(\vartheta )}.}$$

Since the sequence ϑ δ is bounded in \(L^{\infty }(0,T;L^{4}(\Omega )) \cap L^{2}(0,T;L^{6}(\Omega ))\), the last inequality and Corollary 11.2 in Appendix, imply

$$\displaystyle{\overline{\vartheta ^{4}} = \overline{\vartheta ^{3}}\;\vartheta }$$

which proves (3.297).

Accordingly, it is enough to show

$$\displaystyle{ \overline{\varrho s_{M}(\varrho,\vartheta )T_{k}(\vartheta )} \geq \overline{\varrho s_{M}(\varrho,\vartheta )}\;\overline{T_{k}(\vartheta )}. }$$
(3.309)

Due to Corollary 11.2 and property (3.39), we have

$$\displaystyle{\sup _{\varepsilon>0}\|\varrho _{\delta }s_{M}(\varrho _{\delta },\vartheta _{\delta })T_{k}(\vartheta _{\delta }) -\varrho _{\delta }s_{M}(\varrho _{\delta },T_{k}(\vartheta _{\delta }))T_{k}(\vartheta _{\delta })\|_{L^{1}((0,T)\times \Omega )} \rightarrow 0}$$

and

$$\displaystyle{\sup _{\varepsilon>0}\|\varrho _{\delta }s_{M}(\varrho _{\delta },\vartheta _{\delta })\overline{T_{k}(\vartheta )} -\varrho _{\delta }s_{M}(\varrho _{\delta },T_{k}(\vartheta _{\delta }))\overline{T_{k}(\vartheta )}\|_{L^{1}((0,T)\times \Omega )} \rightarrow 0}$$

as k. It is therefore sufficient to prove

$$\displaystyle{ \overline{\varrho s_{M}(\varrho,T_{k}(\vartheta ))T_{k}(\vartheta )} \geq \overline{\varrho s_{M}(\varrho,T_{k}(\vartheta ))}\;\overline{T_{k}(\vartheta )}. }$$
(3.310)

Due to the monotonicity of function ϑs M (ϱ, ϑ), we have

$$\displaystyle{\Big(\varrho _{\delta }s_{M}(\varrho _{\delta },T_{k}(\vartheta _{\delta })) -\varrho _{\delta }s_{M}(\varrho _{\delta },\overline{T_{k}(\vartheta )})\Big)\Big(T_{k}(\vartheta _{\delta }) -\overline{T_{k}(\vartheta )}\Big) \geq 0.}$$

Therefore, (3.310) will be verified if we show that

$$\displaystyle{ \int _{B}\varrho _{\delta }s_{M}(\varrho _{\delta },\overline{T_{k}(\vartheta )})\Big(T_{k}(\vartheta _{\delta }) -\overline{T_{k}(\vartheta )}\Big)\,\mathrm{d}x\mathrm{d}t \rightarrow 0\;\mbox{ as }\varepsilon \rightarrow 0+, }$$
(3.311)

where B is an arbitrary ball in \((0,T) \times \Omega\).

Since log is a concave function, we have \(\overline{\log (T_{k}(\vartheta ))} \leq \log (\overline{T_{k}(\vartheta )})\). Moreover, the sequence {log(ϑ δ )} δ > 0 is bounded in \(L^{2}(0,T;W^{1,2}(\Omega ))\) and the same holds for {log(T k (ϑ δ ))} δ > 0. Consequently,

$$\displaystyle{\log (\overline{T_{k}(\vartheta )})\;1_{\{\overline{T_{ k}(\vartheta )}\leq 1\}} \in L^{2}(0,T;L^{6}(\Omega )),}$$
$$\displaystyle{0 <\log (\overline{T_{k}(\vartheta )})\;1_{\{\overline{T_{ k}(\vartheta )}>1\}} \leq \overline{T_{k}(\vartheta )} \in L^{2}(0,T;L^{6}(\Omega )),}$$

therefore \(\log (\overline{T_{k}(\vartheta )})\) belongs to the space \(L^{2}(0,T;L^{6}(\Omega ))\). In particular, there exists \(z_{\epsilon } \in C^{1}([0,T] \times \overline{\Omega })\) such that

$$\displaystyle{\|z_{\epsilon } -\log (\overline{T_{k}(\vartheta )})\|_{L^{2}(0,T;L^{6}(\Omega ))} <\varepsilon }$$

where ɛ > 0 is a parameter that can be taken arbitrarily small. Setting \(\Theta =\exp (z_{\varepsilon })\) we have

$$\displaystyle{\Theta \in C^{1}([0,T] \times \overline{\Omega }),\;\min _{ (t,x)\in [0,T]\times \overline{\Omega }}\Theta (t,x)> 0.}$$

Now, we write

$$\displaystyle{\int _{B}\varrho _{\delta }s_{M}(\varrho _{\delta },\overline{T_{k}(\vartheta )})\Big(T_{k}(\vartheta _{\delta }) -\overline{T_{k}(\vartheta )}\Big)\,\mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{ =\int _{B}\Big(\varrho _{\delta }s_{M}(\varrho _{\delta },\overline{T_{k}(\vartheta )}) -\varrho _{\delta }s_{M}(\varrho _{\delta },\Theta )\Big)\Big(T_{k}(\vartheta _{\delta }) -\overline{T_{k}(\vartheta )}\Big)\,\mathrm{d}x\ \mathrm{d}t }$$
(3.312)
$$\displaystyle{+\int _{B}\varrho _{\delta }s_{M}(\varrho _{\delta },\Theta )\Big(T_{k}(\vartheta _{\delta }) -\overline{T_{k}(\vartheta )}\Big)\,\mathrm{d}x\ \mathrm{d}t.}$$

We may use (3.11), (3.34) to verify that

$$\displaystyle{\Big\vert \varrho _{\delta }s_{M}(\varrho _{\delta },\overline{T_{k}(\vartheta )}) -\varrho _{\delta }s_{M}(\varrho _{\delta },\Theta )\Big\vert =}$$
$$\displaystyle{\varrho _{\delta }\Big\vert \int _{\overline{T_{ k}(\vartheta )}}^{\Theta }\frac{1} {r} \frac{\partial e_{M}} {\partial \vartheta } (\varrho _{\delta },r)\,\mathrm{d}r\Big\vert \leq c\varrho _{\delta }\Big\vert \log (\overline{T_{k}(\vartheta )}) -\log (\Theta )\Big\vert.}$$

Since ϱ δ is bounded in \(L^{\infty }(0,T;L^{\frac{5} {3} }(\Omega ))\), we infer that

$$\displaystyle{\sup _{\delta>0}\Big\|\varrho _{\delta }\Big(\log (\overline{T_{k}(\vartheta )}) -\log (\Theta )\Big)\Big\|_{ L^{2}(0,T;L^{\frac{30} {23} }(\Omega ))} \leq c\varepsilon;}$$

whence the first integral on the right-hand side of (3.311) tends to 0 as ɛ → 0+.

As a consequence of (3.39), the sequence \(B(t,x,\varrho _{\delta }) =\varrho _{\delta }s_{M}(\varrho _{\delta },\Theta (t,x))\) satisfies hypothesis (11.131) of Theorem 11.37 in Appendix. We can therefore conclude that

$$\displaystyle{\{\varrho _{\delta }s_{M}(\varrho _{\delta },\Theta )\}_{\delta>0}\ \mbox{ is precompact in}\ L^{2}(0,T;W^{-1,2}(\Omega )),}$$

which, together with the fact that \(T_{k}(\vartheta _{\delta }) \rightarrow \overline{T_{k}(\vartheta )}\) weakly in \(L^{2}(0,T;W^{1,2}(\Omega ))\), concludes the proof of inequality (3.309).

Asymptotic Limit in the Entropy Balance Using weak lower semicontinuity of convex functionals, we can let δ → 0 in the approximate entropy balance (3.257) to conclude that

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\overline{\varrho s(\varrho,\vartheta )}\Big(\partial _{t}\varphi + \mathbf{u} \cdot \nabla _{x}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{ \Omega }\frac{\mathbf{q}} {\vartheta } \cdot \nabla _{x}\varphi \ \mathrm{d}x\ \mathrm{d}t }$$
(3.313)
$$\displaystyle{+ <\sigma;\varphi> _{[\mathcal{M};C]([0,T]\times \overline{\Omega })} = -\int _{\Omega }(\varrho s)_{0}\varphi (0,\cdot )\ \mathrm{d}x -\int _{0}^{T}\int _{ \Omega }\frac{\varrho } {\vartheta }\mathcal{Q}\varphi \ \mathrm{d}x\ \mathrm{d}t,}$$

for any \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\). In this equation

$$\displaystyle{ \mathbf{q} = -\kappa (\vartheta )\nabla _{x}\vartheta, }$$
(3.314)

and \(\sigma \in \mathcal{M}^{+}([0,T] \times \overline{\Omega })\) is a weak-(*) limit in \(\mathcal{M}([0,T] \times \overline{\Omega })\) of the sequence σ δ that exists at least for a chosen subsequence due to estimate (3.270). Employing (3.258), (3.270), (3.284), (3.297) and lower weak semicontinuity of convex functionals, using the fact that all δ-dependent quantities in the entropy production rate at the right hand side of (3.258) are non negative, we show that

$$\displaystyle{ \sigma \geq \frac{1} {\vartheta } \Big(\mathbb{S}(\vartheta,\nabla _{x}\mathbf{u}): \nabla _{x}\mathbf{u} + \frac{\kappa (\vartheta )} {\vartheta } \vert \nabla \vartheta \vert ^{2}\Big). }$$
(3.315)

For more details see the similar reasoning between formulas (3.158)–(3.160) in Sect. 3.5.3.

Consequently, in order to complete the proof of Theorem 3.1, we have to show pointwise convergence of the densities. This will be done in the next section.

3.7.4 Pointwise Convergence of the Densities

We follow the same strategy as in Sect. 3.6.5, however, some essential steps have to be considerably modified due to lower L p-integrability available for {ϱ δ } δ > 0, {u δ } δ > 0.

To begin, we introduce a family of cut-off functions

$$\displaystyle{ T_{k}(z) = kT\Big(\frac{z} {k}\Big),\ z \geq 0,\ k \geq 1, }$$
(3.316)

where TC [0, ),

$$\displaystyle{ T(z) = \left \{\begin{array}{l} z\ \mbox{ for}\ 0 \leq z \leq 1,\\ \\ \mbox{ concave on}\ [0,\infty ),\\ \\ 2\ \mbox{ for}\ z \geq 3. \end{array} \right. }$$
(3.317)

Similarly to Sects. 2.2.63.6.5, we use the quantities

$$\displaystyle{\boldsymbol{\varphi }(t,x) =\psi (t)\zeta (x)(\nabla _{x}\Delta _{x}^{-1})[1_{ \Omega }T_{k}(\varrho _{\delta })],\ \psi \in C_{c}^{\infty }(0,T),\ \zeta \in C_{ c}^{\infty }(\Omega ),}$$

with the operators \((\nabla _{x}\Delta _{x}^{-1})\) introduced in (2.100), as test functions in the approximate momentum equation (3.253) to deduce

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\psi \zeta \Big[\Big(p(\varrho _{\delta },\vartheta _{\delta }) +\delta (\varrho _{\delta }^{\Gamma } + \varrho _{\delta }^{2})\Big)T_{ k}(\varrho _{\delta }) - \mathbb{S}_{\delta }: \mathcal{R}[1_{\Omega }T_{k}(\varrho _{\delta })]\Big]\ \mathrm{d}x\ \mathrm{d}t =\sum _{ j=1}^{7}I_{ j,\delta }, }$$
(3.318)

where \(\mathbb{S}_{\delta }:= \mathbb{S}_{\delta }(\vartheta _{\delta },\nabla _{x}\mathbf{u}_{\delta })\) and

$$\displaystyle{I_{1,\delta } =\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho _{\delta }\mathbf{u}_{\delta } \cdot \mathcal{R}[1_{\Omega }T_{k}(\varrho _{\delta })\mathbf{u}_{\delta }] - (\varrho _{\delta }\mathbf{u}_{\delta } \otimes \mathbf{u}_{\delta }): \mathcal{R}[1_{\Omega }T_{k}(\varrho _{\delta })]\Big)\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{2,\delta } = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \ \varrho _{\delta }\mathbf{u}_{\delta } \cdot \nabla _{x}\Delta _{x}^{-1}\Big[1_{ \Omega }(T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta })\mathrm{div}_{ x}\mathbf{u}_{\delta }\Big]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{3,\delta } = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \varrho _{\delta }\mathbf{f}_{\delta } \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }T_{k}(\varrho _{\delta })]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{4,\delta } = -\int _{0}^{T}\int _{ \Omega }\psi \Big(\,p(\varrho _{\delta },\vartheta _{\delta }) +\delta (\varrho _{\delta }^{\Gamma } + \varrho _{\delta }^{2})\Big)\nabla _{ x}\zeta \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }T_{k}(\varrho _{\delta })]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{5,\delta } =\int _{ 0}^{T}\int _{ \Omega }\psi \mathbb{S}_{\delta }: \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }T_{k}(\varrho _{\delta })]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{6,\delta } = -\int _{0}^{T}\int _{ \Omega }\psi (\varrho _{\delta }\mathbf{u}_{\delta } \otimes \mathbf{u}_{\delta }): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }T_{k}(\varrho _{\delta })]\ \mathrm{d}x\ \mathrm{d}t,}$$

and

$$\displaystyle{I_{7,\delta } = -\int _{0}^{T}\int _{ \Omega }\partial _{t}\psi \ \zeta \varrho _{\delta }\mathbf{u}_{\delta } \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }T_{k}(\varrho _{\delta })]\ \mathrm{d}x\ \mathrm{d}t.}$$

Now, mimicking the strategy of Sect. 3.6.5, we use

$$\displaystyle{\boldsymbol{\varphi }(t,x) =\psi (t)\zeta (x)(\nabla _{x}\Delta _{x}^{-1})[1_{ \Omega }\overline{T_{k}(\varrho )}],\ \psi \in C_{c}^{\infty }(0,T),\ \zeta \in C_{ c}^{\infty }(\Omega )}$$

as test functions in the limit momentum balance (3.294) to obtain

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\psi \zeta \Big[\Big(\overline{p_{M}(\varrho,\vartheta )} + \frac{a} {4}\vartheta ^{4})\Big)\overline{T_{ k}(\varrho )} - \mathbb{S}: \mathcal{R}[1_{\Omega }\overline{T_{k}(\varrho )}]\Big]\ \mathrm{d}x\ \mathrm{d}t =\sum _{ j=1}^{7}I_{ j}, }$$
(3.319)

where

$$\displaystyle{I_{1} =\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \Big(\varrho \mathbf{u} \cdot \mathcal{R}[1_{\Omega }\overline{T_{k}(\varrho )}\mathbf{u}] - (\varrho \mathbf{u} \otimes \mathbf{u}): \mathcal{R}[1_{\Omega }\overline{T_{k}(\varrho )}]\Big)\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{2} = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \ \varrho \mathbf{u} \cdot \nabla _{x}\Delta _{x}^{-1}\Big[1_{ \Omega }\overline{(T_{k}(\varrho ) - T_{k}^{{\prime}}(\varrho )\varrho )\mathrm{div}_{ x}\mathbf{u}}\Big]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{3} = -\int _{0}^{T}\int _{ \Omega }\psi \zeta \varrho \mathbf{f} \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\overline{T_{k}(\varrho )}]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{4} = -\int _{0}^{T}\int _{ \Omega }\psi \overline{p(\varrho,\vartheta )}\nabla _{x}\zeta \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\overline{T_{k}(\varrho )}]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{5} =\int _{ 0}^{T}\int _{ \Omega }\psi \mathbb{S}: \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\overline{T_{k}(\varrho )}]\ \mathrm{d}x\ \mathrm{d}t,}$$
$$\displaystyle{I_{6} = -\int _{0}^{T}\int _{ \Omega }\psi (\varrho \mathbf{u} \otimes \mathbf{u}): \nabla _{x}\zeta \otimes \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\overline{T_{k}(\varrho )}]\ \mathrm{d}x\ \mathrm{d}t,}$$

and

$$\displaystyle{I_{7} = -\int _{0}^{T}\int _{ \Omega }\partial _{t}\psi \ \zeta \varrho \mathbf{u} \cdot \nabla _{x}\Delta _{x}^{-1}[1_{ \Omega }\overline{T_{k}(\varrho )}]\ \mathrm{d}x\ \mathrm{d}t.}$$

We recall that \(\mathcal{R} = \mathcal{R}_{i,j}\) is the double Riesz transform introduced in Sect. 5.

To get formula (3.319) we have used (3.284), (3.297) to identify \(\overline{\vartheta ^{4}}\) with ϑ 4 and \(\overline{\mathbb{S}(\vartheta,\nabla _{x}\mathbf{u})}\) with \(\mathbb{S}:= \mathbb{S}(\vartheta,\nabla _{x}\mathbf{u})\). We also recall that \(\mathcal{R} = \mathcal{R}_{i,j}\) is the double Riesz transform introduced in Sect. 5.

Now, letting δ → 0+ in (3.318), we get

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\psi \zeta \left [\overline{p_{M}(\varrho,\vartheta )\,T_{k}(\varrho )} + a\vartheta ^{4}\overline{T_{ k}(\varrho )} -\overline{\mathbb{S}: \mathcal{R}[1_{\Omega }\,T_{k}(\varrho )]}\right ]\,\mathrm{d}x\mathrm{d}t =\sum _{ j=1}^{7}I_{ j}, }$$
(3.320)

where the right hand side is the same as the right hand side in (3.319). Here, we have used the commutator lemma in form of Corollary 3.3 with \(r_{\delta } = 1_{\Omega }\,T_{k}(\varrho _{\delta })\) and \(\mathbf{V}_{\delta } = 1_{\Omega }\varrho _{\delta }\mathbf{u}_{\delta }\) to show that

$$\displaystyle{I_{1,\delta } \rightarrow I_{1}\ \mbox{ as }\delta \rightarrow 0+,}$$

exactly in the same way as explained in detail in Sect. 3.6.5. We have also employed the pointwise convergence (3.297) to verify that \(\overline{\vartheta ^{4}} =\vartheta ^{4}\) and that \(\overline{\vartheta ^{4}\,T_{k}(\varrho )} =\) \(\vartheta ^{4}\overline{T_{k}(\varrho )}\).

Combining (3.319) and (3.320), we get identity

$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\psi \zeta \left [\overline{p_{M}(\varrho,\vartheta )\,T_{k}(\varrho )} -\overline{p_{M}(\varrho,\vartheta )}\;\overline{T_{k}(\varrho )}\right ]\,\mathrm{d}x\mathrm{d}t}$$
$$\displaystyle{=\int _{ 0}^{T}\int _{ \Omega }\psi \zeta \left [\overline{\mathbb{S}: \mathcal{R}[1_{\Omega }T_{k}(\varrho )]} - \mathbb{S}: \mathcal{R}[1_{\Omega }\overline{T_{k}(\varrho )}]\right ]\,\mathrm{d}x\mathrm{d}t.}$$

We again follow the great lines of Sect. 3.6.5. Employing the evident properties of the Riesz transform evoked in formulas (11.107), we may write

$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\psi \zeta \overline{\mathbb{S}: \mathcal{R}[1_{\Omega }\,T_{k}(\varrho )]}\,\mathrm{d}x\mathrm{d}t =\lim _{\delta \rightarrow 0+}\int _{0}^{T}\int _{ \Omega }\psi \omega (\vartheta _{\delta },\mathbf{u}_{\delta })\,T_{k}(\varrho _{\delta })\,\mathrm{d}x\mathrm{d}t}$$
$$\displaystyle{+\lim _{\delta \rightarrow 0+}\int _{0}^{T}\int _{ \Omega }\psi \zeta (\frac{4} {3}\mu (\vartheta _{\delta }) +\eta (\vartheta _{\delta }))\mathrm{div}_{x}\mathbf{u}\,T_{k}(\varrho _{\delta })\,\mathrm{d}x\mathrm{d}t}$$

and

$$\displaystyle{\int _{0}^{T}\int _{ \Omega }\psi \zeta \mathbb{S}: \mathcal{R}[1_{\Omega }\overline{T_{k}(\varrho )}]\,\mathrm{d}x\mathrm{d}t =\int _{ 0}^{T}\int _{ \Omega }\psi \omega (\vartheta,\mathbf{u})\,\overline{T_{k}(\varrho )}\,\mathrm{d}x\mathrm{d}t}$$
$$\displaystyle{+\int _{0}^{T}\int _{ \Omega }\psi \zeta (\frac{4} {3}\mu (\vartheta ) +\eta (\vartheta ))\mathrm{div}_{x}\mathbf{u}\,\overline{T_{k}(\varrho )}\,\mathrm{d}x\mathrm{d}t,}$$

where

$$\displaystyle{\omega (\vartheta,\mathbf{u}) = \mathcal{R}:\Big [\zeta \mu (\vartheta )\Big(\nabla _{x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big)\Big] -\zeta \mu (\vartheta )\mathcal{R}:\Big [\nabla _{ x}\mathbf{u} + \nabla _{x}^{T}\mathbf{u}\Big].}$$

Applying Lemma 3.6 to w = ζμ(ϑ δ ), \(\mathbf{Z} = [\partial _{x_{i}}u_{\varepsilon,j} + \partial _{x_{j}}u_{\varepsilon,i}]_{i=1}^{3}\), j ∈ {1, 2, 3} fixed, where, according to (3.17)–(3.18), (3.278), (3.275), the sequences w, Z are bounded in \(L^{2}(0,T;W^{1,2}(\Omega ))\) and \(L^{8/(5-\alpha )}((0,T) \times \Omega )\), respectively, we deduce that

$$\displaystyle{ \{\omega (\vartheta _{\delta },\mathbf{u}_{\delta })\}_{\delta>0}\ \mbox{ is bounded in}\ L^{1}(0,T;W^{\beta,s}(\Omega ))\;\mbox{ for certain }\beta \in (0,1),s> 1. }$$
(3.321)

Now, we consider four-dimensional vector fields

$$\displaystyle{\mathbf{U}_{\delta } = [T_{k}(\varrho _{\delta }),T_{k}(\varrho _{\delta })\mathbf{u}_{\delta }],\;\mathbf{V}_{\delta } = [\omega (\varrho _{\delta },\vartheta _{\delta }),0,0,0]}$$

and take advantage of relations (3.252), (3.266), (3.267), (3.269), (3.278), (3.275) and (3.321) in order to show that U δ , V δ verify all hypotheses of the Div-Curl lemma stated in Proposition 3.3. Using this proposition, we may conclude that

$$\displaystyle{ \omega (\vartheta _{\delta },\mathbf{u}_{\delta })T_{k}(\varrho _{\delta }) \rightarrow \overline{\omega (\vartheta,\mathbf{u})}\;\overline{T_{k}(\varrho )} =\omega (\vartheta,\mathbf{u})\;\overline{T_{k}(\varrho )}\ \mbox{ weakly in}\ L^{1}((0,T) \times \Omega ), }$$
(3.322)

where we have used (3.284), (3.297) to identify \(\overline{\omega (\vartheta,\mathbf{u})}\) with ω(ϑ, u).

We thus discover on this level of approximations again the weak compactness identity for the effective pressure

      Weak Compactness Identity for Effective Pressure (Level δ):

$$\displaystyle{ \overline{p_{M}(\varrho,\vartheta )T_{k}(\varrho )} -\Big (\frac{4} {3}\mu (\vartheta ) +\eta (\vartheta )\Big)\;\overline{T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u}} }$$
(3.323)
$$\displaystyle{= \overline{p_{M}(\varrho,\vartheta )}\;\overline{T_{k}(\varrho )} -\Big (\frac{4} {3}\mu (\vartheta ) +\eta (\vartheta )\Big)\overline{T_{k}(\varrho )}\;\mathrm{div}_{x}\mathbf{u}.}$$

Thus our ultimate goal is to use relation (3.323) in order to show pointwise convergence of the family of approximate densities {ϱ δ } δ > 0. To this end, we revoke the “renormalized” limit Eq. (3.293) yielding

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\overline{\varrho L_{k}(\varrho )}\partial _{t}\varphi + \overline{\varrho L_{k}(\varrho )}\mathbf{u} \cdot \nabla _{x}\varphi -\overline{T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u}}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.324)
$$\displaystyle{= -\int _{\Omega }\varrho _{0}L_{k}(\varrho _{0})\varphi (0,\cdot )\ \mathrm{d}x}$$

for any test function \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\), where we have set

$$\displaystyle{L_{k}(\varrho ) =\int _{ 1}^{\varrho }\frac{T_{k}(z)} {z^{2}} \ \mathrm{d}z.}$$

Assume, for a moment, that the limit functions ϱ, u also satisfy the equation of continuity in the sense of renormalized solutions, in particular,

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\Big(\varrho L_{k}(\varrho )\partial _{t}\varphi + \varrho L_{k}(\varrho )\mathbf{u} \cdot \nabla _{x}\varphi - T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u}\varphi \Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.325)
$$\displaystyle{= -\int _{\Omega }\varrho _{0}L_{k}(\varrho _{0})\varphi (0,\cdot )\ \mathrm{d}x}$$

for any test function \(\varphi \in C_{c}^{\infty }([0,T) \times \overline{\Omega })\).

Now, relations (3.324), (3.325) give rise to

$$\displaystyle{ \int _{\Omega }\Big(\overline{\varrho L_{k}(\varrho )} -\varrho L_{k}(\varrho )\Big)(\tau )\ \mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\Big(\overline{T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u}} -\overline{T_{k}(\varrho )}\mathrm{div}_{x}\mathbf{u}\Big)\ \mathrm{d}x\ \mathrm{d}t }$$
(3.326)
$$\displaystyle{=\int _{ 0}^{\tau }\int _{ \Omega }\Big(T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u} -\overline{T_{k}(\varrho )}\mathrm{div}_{x}\mathbf{u}\ \mathrm{d}x\ \mathrm{d}t\ \mbox{ for any}\ \tau \in [0,T].}$$

As {ϑ δ } δ > 0 converges strongly in L 1 and p M is a non-decreasing function of ϱ, we can use relation (3.323) to obtain

$$\displaystyle{\overline{T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u}} -\overline{T_{k}(\varrho )}\mathrm{div}_{x}\mathbf{u} \geq 0.}$$

Letting k in (3.326) we obtain

$$\displaystyle{ \overline{\varrho \log (\varrho )} =\varrho \log (\varrho )\ \mbox{ a.a. on}\ (0,T) \times \Omega, }$$
(3.327)

as soon as we are able to show that

$$\displaystyle{ \int _{0}^{\tau }\int _{ \Omega }\Big(T_{k}(\varrho )\mathrm{div}_{x}\mathbf{u} -\overline{T_{k}(\varrho )}\mathrm{div}_{x}\mathbf{u}\Big)\ \mathrm{d}x\ \mathrm{d}t \rightarrow 0\ \mbox{ for}\ k \rightarrow \infty. }$$
(3.328)

Relation (3.327) yields

$$\displaystyle{ \varrho _{\delta } \rightarrow \varrho \ \mbox{ in}\ L^{1}((0,T) \times \Omega ), }$$
(3.329)

see Theorem 11.27 in Appendix. This completes the proof of Theorem 3.1.

Note, however, that two fundamental issues have been left open in the preceding discussion, namely

  • the validity of the renormalized Eq. (3.325),

  • relation (3.328).

These two intimately related questions will be addressed in the following section.

3.7.5 Oscillations Defect Measure

The oscillations defect measure introduced in [117] represents a basic tool for studying density oscillations. Given a family {ϱ δ } δ > 0, a set Q, and q ≥ 1, we introduce:

      Oscillations Defect Measure:

$$\displaystyle{ \mathbf{osc}_{q}[\varrho _{\delta } \rightarrow \varrho ](Q) =\sup _{k\geq 1}\Big(\limsup _{\delta \rightarrow 0+}\int _{Q}\Big\vert T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\Big\vert ^{q}\ \mathrm{d}x\ \mathrm{d}t\Big), }$$
(3.330)

where T k are the cut-off functions introduced in (3.316).

Assume that

$$\displaystyle{ \mathrm{div}_{x}\mathbf{u} \in L^{r}((0,T) \times \Omega ),\ \mathbf{osc}_{ q}[\varrho _{\delta } \rightarrow \varrho ]((0,T) \times \Omega ) <\infty,\ \mbox{ with}\ \frac{1} {r} + \frac{1} {q} <1. }$$
(3.331)

Seeing that

$$\displaystyle{T_{k}(\varrho ) \rightarrow \varrho,\ \overline{T_{k}(\varrho )} \rightarrow \varrho \ \mbox{ in}\ L^{1}((0,T) \times \Omega )\ \mbox{ for}\ k \rightarrow \infty,}$$

we conclude easily that (3.331) implies (3.328).

A less obvious statement is the following assertion.

Lemma 3.8

Let \(Q \subset \mathbb{R}^{4}\) be an open set. Suppose that

$$\displaystyle{\varrho _{\delta } \rightarrow \varrho \ \mathit{\mbox{ weakly in}}\ L^{1}(Q),}$$
$$\displaystyle{ \mathbf{u}_{\delta } \rightarrow \mathbf{u}\ \mathit{\mbox{ weakly in}}\ L^{r}(Q; \mathbb{R}^{3}), }$$
(3.332)
$$\displaystyle{ \nabla _{x}\mathbf{u}_{\delta } \rightarrow \nabla _{x}\mathbf{u}\ \mathit{\mbox{ weakly in}}\ L^{r}(Q; \mathbb{R}^{3\times 3}),\ r> 1, }$$
(3.333)

and

$$\displaystyle{ \mathbf{osc}_{q}[\varrho _{\delta } \rightarrow \varrho ](Q) <\infty \ \mathit{\mbox{ for}}\ \frac{1} {q} + \frac{1} {r} <1, }$$
(3.334)

where ϱ δ , u δ solve the renormalized Eq. ( 2.2 ) in \(\mathcal{D}'(Q)\) .

Then the limit functions ϱ, u solve the renormalized Eq. ( 2.2 ) in \(\mathcal{D}'(Q)\) .

Proof

Clearly, it is enough to show the result on the set J × K, where J is a bounded time interval and K is a ball such that \(\overline{J \times K} \subset Q\). Since ϱ δ is a renormalized solution of (2.2), we get

$$\displaystyle{T_{k}(\varrho _{\delta }) \rightarrow \overline{T_{k}(\varrho )}\ \mbox{ in}\ C_{\mathrm{weak}}(\overline{J};L^{\beta }(\Omega ))\ \mbox{ for any}\ 1 \leq \beta <\infty;}$$

whence, by virtue of hypotheses (3.332), (3.333),

$$\displaystyle{T_{k}(\varrho _{\delta })\mathbf{u}_{\delta } \rightarrow \overline{T_{k}(\varrho )}\mathbf{u}\ \mbox{ weakly in}\ L^{r}(J \times K; \mathbb{R}^{3}).}$$

Consequently, we deduce

$$\displaystyle{ \partial _{t}\overline{T_{k}(\varrho )} + \mathrm{div}_{x}\Big(\overline{T_{k}(\varrho )}\mathbf{u}\Big) + \overline{\Big(T_{k}^{{\prime}}(\varrho )\varrho - T_{k}(\varrho )\Big)\mathrm{div}_{x}\mathbf{u}\Big)} = 0\ \mbox{ in}\ \mathcal{D}'(J \times \Omega ). }$$
(3.335)

Since \(\overline{T_{k}(\varrho )}\) are bounded, we can apply the regularization technique introduced by DiPerna and Lions [85] (Theorem 11.36), already used in Lemma 3.7, in order to deduce

$$\displaystyle{\partial _{t}h(\overline{T_{k}(\varrho )}) + \mathrm{div}_{x}\Big(h(\overline{T_{k}(\varrho )})\mathbf{u}\Big) +\Big (h'(\overline{T_{k}(\varrho )})\ \overline{T_{k}(\varrho )} -\overline{T_{k}(\varrho )}\Big)\mathrm{div}_{x}\mathbf{u}}$$
$$\displaystyle{= h'(\overline{T_{k}(\varrho )})\overline{\Big(T_{k}(\varrho ) - T_{k}^{{\prime}}(\varrho )\varrho \Big)\mathrm{div}_{x}\mathbf{u}}\ \mbox{ in}\ \mathcal{D}'(J \times K),}$$

where h is a continuously differentiable function such that h′(z) = 0 for all z large enough, say, zM.

Consequently, it is enough to show

$$\displaystyle{ h'(\overline{T_{k}(\varrho )})\overline{\Big(T_{k}(\varrho ) - T_{k}^{{\prime}}(\varrho )\varrho \Big)\mathrm{div}_{x}\mathbf{u}} \rightarrow 0\ \mbox{ in}\ L^{1}(J \times K)\ \mbox{ for}\ k \rightarrow \infty. }$$
(3.336)

To this end, denote

$$\displaystyle{Q_{k,M} =\{ (t,x) \in J \times K\ \vert \ \vert \overline{T_{k}(\varrho )}\vert \leq M\}.}$$

Consequently,

$$\displaystyle{ \Big\|h'(\overline{T_{k}(\varrho )})\overline{\Big(T_{k}(\varrho ) - T_{k}^{{\prime}}(\varrho )\varrho \Big)\mathrm{div}_{x}\mathbf{u}}\Big\|_{L^{1}(J\times K)} }$$
(3.337)
$$\displaystyle{\leq \Big (\sup _{0\leq z\leq M}\vert h'(z)\vert \Big)\Big(sup_{\delta>0}\|\mathrm{div}_{x}\mathbf{u}_{\delta }\|_{L^{r}(J\times K)}\Big)\liminf _{\delta \rightarrow 0}\|T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{r'}(Q_{k,M})},}$$

where 1∕r + 1∕r′ = 1.

Furthermore, a simple interpolation argument yields

$$\displaystyle{ \|T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{r'}(Q_{k,M})} }$$
(3.338)
$$\displaystyle{\leq \| T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{1}(J\times K)}^{\beta }\|T_{ k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{q}(Q_{k,M})}^{1-\beta },}$$

with β ∈ (0, 1).

As the family {ϱ δ } δ > 0 is equi-integrable, we deduce

$$\displaystyle{ \sup _{\delta>0}\Big\{\|T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{1}(J\times K)}\Big\} \rightarrow 0\ \mbox{ for}\ k \rightarrow \infty. }$$
(3.339)

Finally, seeing that | T k (ϱ δ )ϱ δ | ≤ T k (ϱ δ ), we conclude

$$\displaystyle{\|T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{q}(Q_{k,M})}}$$
$$\displaystyle{\leq 2\Big(\|T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\|_{L^{q}(J\times K)} +\| T_{k}(\varrho ) -\overline{T_{k}(\varrho )}\|_{L^{q}(J\times K)} +\| \overline{T_{k}(\varrho )}\|_{L^{q}(Q_{k,M})}\Big)}$$
$$\displaystyle{\leq 2\Big(\|T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\|_{L^{q}(J\times K)} + \mathbf{osc}_{q}[\varrho _{\delta } \rightarrow \varrho ](J \times K + \vert J \times K\vert ^{\frac{1} {q} };}$$

whence

$$\displaystyle{ \limsup _{\delta \rightarrow 0}\|T_{k}(\varrho _{\delta }) - T_{k}^{{\prime}}(\varrho _{\delta })\varrho _{\delta }\|_{ L^{q}(Q_{k,M})} \leq 4\mathbf{osc}_{q}[\varrho _{\delta } \rightarrow \varrho ](J \times K) + 2M\vert J \times K\vert ^{\frac{1} {q} }. }$$
(3.340)

Clearly, relation (3.336) follows from (3.337) to (3.340).

In order to apply Lemma 3.8, we need to establish suitable bounds on osc q [ϱ δ ϱ]. To this end, revoking (3.42)–(3.44) we write

$$\displaystyle{ p_{M}(\varrho,\vartheta ) = d\varrho ^{\frac{5} {3} } + p_{m}(\varrho,\vartheta ) + p_{b}(\varrho,\vartheta ),\ d> 0, }$$
(3.341)

where

$$\displaystyle{ \frac{\partial p_{m}(\varrho,\vartheta )} {\partial \varrho } \geq 0, }$$
(3.342)

and

$$\displaystyle{ \vert p_{b}(\varrho,\vartheta )\vert \leq c(1 + \vartheta ^{\frac{5} {2} }) }$$
(3.343)

for all ϱ, ϑ > 0.

Consequently,

$$\displaystyle{d\limsup _{\delta \rightarrow 0+}\int _{0}^{T}\int _{ \Omega }\varphi \vert T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{\frac{8} {3} }\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{\leq d\int _{0}^{T}\int _{ \Omega }\varphi \Big(\overline{\varrho ^{\frac{5} {3} }T_{k}(\varrho )} -\overline{\varrho ^{\frac{5} {3} }}\ \overline{T_{k}(\varrho )}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{+d\int _{0}^{T}\int _{ \Omega }\varphi \Big(\varrho ^{\frac{5} {3} } -\overline{\varrho ^{\frac{5} {3} }}\Big)\Big(T_{k}(\varrho ) -\overline{T_{k}(\varrho )}\Big)\ \mathrm{d}x\ \mathrm{d}t}$$
$$\displaystyle{\leq \int _{0}^{T}\int _{ \Omega }\varphi \Big(\overline{p_{M}(\varrho,\vartheta )T_{k}(\varrho )} -\overline{p_{M}(\varrho,\vartheta )}\ \overline{T_{k}(\varrho )}\Big)\ \mathrm{d}x}$$
$$\displaystyle{+\left \vert \int _{0}^{T}\int _{ \Omega }\varphi \Big(\overline{p_{b}(\varrho,\vartheta )T_{k}(\varrho )} -\overline{p_{b}(\varrho,\vartheta )}\ \overline{T_{k}(\varrho )}\Big)\ \mathrm{d}x\ \mathrm{d}t\right \vert }$$

for any \(\varphi \in C_{c}^{\infty }((0,T) \times \Omega )\), φ ≥ 0, where we have used (3.342), convexity of \(\varrho \mapsto \varrho ^{\frac{5} {3} }\), and concavity of T k on [0, ).

In accordance with the uniform bound (3.269) and (3.343), we have

$$\displaystyle{ \left \vert \int _{0}^{T}\int _{ \Omega }\varphi \Big(\overline{p_{b}(\varrho,\vartheta )T_{k}(\varrho )} -\overline{p_{b}(\varrho,\vartheta )}\ \overline{T_{k}(\varrho )}\Big)\ \mathrm{d}x\ \mathrm{d}t\right \vert }$$
(3.344)
$$\displaystyle{\leq c_{1}\Big(1 +\sup _{\delta>0}\|\vartheta _{\delta }^{\frac{5} {2} }\|_{ L^{\frac{8} {5} }((0,T)\times \Omega )}\Big)\left (\int _{0}^{T}\int _{ \Omega }\varphi \vert T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{\frac{8} {3} }\ \mathrm{d}x\ \mathrm{d}t\right )^{\frac{3} {8} }}$$
$$\displaystyle{\leq c_{2}\limsup _{\delta \rightarrow 0}\left (\int _{0}^{T}\int _{ \Omega }\varphi \vert T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{\frac{8} {3} }\ \mathrm{d}x\ \mathrm{d}t\right )^{\frac{3} {8} }.}$$

Furthermore, introducing a Carathéodory function

$$\displaystyle{G_{k}(t,x,z) = \vert T_{k}(z) - T_{k}(\varrho (t,x))\vert ^{\frac{8} {3} }}$$

we get, in accordance with (3.344),

$$\displaystyle{\overline{G_{k}(\cdot,\cdot,\varrho )} \leq c\Big(1 + \overline{p_{M}(\varrho,\vartheta )T_{k}(\varrho )} -\overline{p_{M}(\varrho,\vartheta )}\ \overline{T_{k}(\varrho )}\Big),\ \mbox{ with}\ c\ \mbox{ independent of}\ k \geq 1.}$$

Thus, evoking once more (3.323) we infer that

$$\displaystyle{ \overline{G_{k}(\cdot,\cdot,\varrho )} \leq c\Big(1 + (\frac{4} {3}\mu (\vartheta ) +\eta (\vartheta ))(\overline{\mathrm{div}_{x}\mathbf{u}\ T_{k}(\varrho )} -\mathrm{div}_{x}\mathbf{u}\ \overline{T_{k}(\varrho )})\Big)\ \mbox{ for all}\ k \geq 1. }$$
(3.345)

On the other hand, by virtue of hypothesis (3.17) and estimate (3.275), we get

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }(1+\vartheta )^{-\alpha }\overline{G_{ k}(t,x,\varrho )}\ \mathrm{d}x\ \mathrm{d}t }$$
(3.346)
$$\displaystyle{\leq c\Big(1 +\sup _{\delta>0}\|\mathrm{div}_{x}\mathbf{u}_{\delta }\|_{ L^{ \frac{8} {5-\alpha }}((0,T)\times \Omega )}\limsup _{\delta \rightarrow 0+}\|T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\|_{ L^{ \frac{8} {3+\alpha } }((0,T)\times \Omega )}\Big)}$$
$$\displaystyle{\leq c\Big(1 +\limsup _{\delta \rightarrow 0+}\|T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\|_{ L^{ \frac{8} {3+\alpha } }((0,T)\times \Omega )}\Big).}$$

Taking

$$\displaystyle{ \frac{8} {3+\alpha } <q <\frac{8} {3},\ \beta = \frac{3q\alpha } {8} }$$

and using Hölder’s inequality, we obtain

$$\displaystyle{ \int _{0}^{T}\int _{ \Omega }\vert T_{k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{q}\ \mathrm{d}x\ \mathrm{d}t =\int _{ 0}^{T}\int _{ \Omega }(1+\vartheta )^{-\beta }(1+\vartheta )^{\beta }\vert T_{ k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{q}\ \mathrm{d}x\ \mathrm{d}t }$$
(3.347)
$$\displaystyle{\leq c\Big(\int _{0}^{T}\int _{ \Omega }(1+\vartheta )^{-\alpha }\vert T_{ k}(\varrho _{\delta }) - T_{k}(\varrho )\vert ^{\frac{8} {3} }\ \mathrm{d}x\ \mathrm{d}t +\int _{ 0}^{T}\int _{\Omega }(1+\vartheta )^{ \frac{3\alpha q} {8-3q} }\ \mathrm{d}x\ \mathrm{d}t\Big).}$$

Finally, choosing q such that

$$\displaystyle{ \frac{8} {3+\alpha } <q \leq \frac{32} {3\alpha + 12},\ \mbox{ meaning}\ \frac{3\alpha q} {8 - 3q} \leq 4,}$$

we can combine relations (3.346), (3.347), together with estimate (3.269), in order to conclude that

$$\displaystyle{ \mathbf{osc}_{q}[\varrho _{\delta } \rightarrow \varrho ]((0,T) \times \Omega ) <\infty \ \mbox{ for a certain}\ q> \frac{8} {3+\alpha }. }$$
(3.348)

Relation (3.348) together with (3.275) allow us to apply Lemma 3.8 in order to conclude that

  • the limit functions ϱ u satisfy the renormalized Eq. (3.325),

  • relation (3.328) holds.

Thus we have rigorously justified the strong convergence of {ϱ δ } δ > 0 claimed in (3.329). The proof of Theorem 3.1 is complete.

3.8 Regularity Properties of the Weak Solutions

The reader will have noticed that the weak solutions constructed in the course of the proof of Theorem 3.1 enjoy slightly better regularity and integrability properties than those required in Sect. 2.1. As a matter of fact, the uniform bounds obtained above can be verified for any weak solution of the Navier-Stokes-Fourier system in the sense of Sect. 2.1 and not only for the specific one resulting from our approximation procedure. Similarly, the restrictions on the geometry of the spatial domain can be considerably relaxed and other types of domains, for instance, the periodic slab, can be handled.

       Regularity of the Weak Solutions:

Theorem 3.2

Let \(\Omega \subset \mathbb{R}^{3}\) be a bounded Lipschitz domain. Assume the data ϱ 0 , (ϱ u)0 , E 0 , (ϱs)0 , the source terms f , \(\mathcal{Q}\) , the thermodynamic functions p, e, s, and the transport coefficients μ, η, κ satisfy the structural hypotheses ( 3.1 )–( 3.23 ) listed in Sect.  3.1 . Let {ϱ, u, ϑ} be a weak solution to the Navier-Stokes-Fourier system on \((0,T) \times \Omega\) in the sense specified in Sect.  2.1 .

Then, in addition to the minimal integrability and regularity properties required in ( 2.5 )–(2.6), (2.13)–(2.15), (2.30)–(2.31), there holds:

  1. (i)
    $$\displaystyle{ \begin{array}{c} \varrho \in C_{\mathrm{weak}}([0,T];L^{\frac{5} {3} }(\Omega )) \cap C([0,T];L^{1}(\Omega )),\\ \\ \varrho \mathbf{u} \in C_{\mathrm{weak}}([0,T];L^{\frac{5} {4} }(\Omega )),\end{array} }$$
    (3.349)
    $$\displaystyle{ \begin{array}{c} \vartheta \in L^{2}(0,T;W^{1,2}(\Omega )) \cap L^{\infty }(0,T;L^{4}(\Omega )),\\ \\ \log \vartheta \in L^{2}(0,T;W^{1,2}(\Omega )),\end{array} }$$
    (3.350)
    $$\displaystyle{ \left \{\begin{array}{c} \mathbb{S}(\vartheta,\nabla _{x}\mathbf{u}) \in L^{q}((0,T) \times \Omega; \mathbb{R}^{3\times 3})\quad \mathit{\mbox{ for a certain }}q> 1,\\ \\ \mathbf{u} \in L^{q}(0,T;W^{1,p}(\Omega; \mathbb{R}^{3}))\ \mathit{\mbox{ for}}\ q = \frac{6} {4-\alpha },p = \frac{18} {10-\alpha },\end{array} \right \} }$$
    (3.351)
    $$\displaystyle{ \left \{\begin{array}{c} \varrho \in L^{q}((0,T) \times \Omega )\quad \mathit{\mbox{ for a certain }}q> \frac{5} {3},\\ \\ p(\varrho,\vartheta ) \in L^{q}((0,T) \times \Omega )\quad \mathit{\mbox{ for a certain }}q> 1.\end{array} \right \} }$$
    (3.352)
  2. (ii)

    The total kinetic energy \(\int _{\Omega }\frac{\vert \varrho \mathbf{u}\vert ^{2}} {\varrho } 1_{\{\varrho>0\}}\,\mathrm{d}x\) is lower semicontinuous on (0, T), left lower semicontinuous at T and right lower semicontinuous at 0; in particular

    $$\displaystyle{ \liminf _{t\rightarrow 0+}\Big(\int _{\Omega }\frac{\vert \varrho \mathbf{u}\vert ^{2}} {\varrho } 1_{\{\varrho>0\}}\,\mathrm{d}x\Big)(t) \geq \int _{\Omega }\frac{\vert (\varrho \mathbf{u})_{0}\vert ^{2}} {\varrho _{0}} 1_{\{\varrho _{0}>0\}}\,\mathrm{d}x. }$$
    (3.353)
  3. (iii)

    The entropy satisfies

    $$\displaystyle{ \left \{\begin{array}{c} \mathrm{ess}\lim _{t\rightarrow 0+}\int _{\Omega }\varrho s(\varrho,\vartheta )(t,\cdot )\varphi \,\mathrm{d}x\, \geq \int _{\Omega }\varrho _{0}s(\varrho _{0},\vartheta _{0})\varphi \,\mathrm{d}x\\ \\ \mathit{\mbox{ for any}}\ \varphi \in C_{c}^{\infty }(\overline{\Omega }),\ \varphi \geq 0.\end{array} \right \} }$$
    (3.354)

    If, in addition, \(\vartheta _{0} \in W^{1,\infty }(\Omega )\) , then

    $$\displaystyle{ \mathrm{ess}\lim _{t\rightarrow 0+}\int _{\Omega }\varrho s(\varrho,\vartheta )(t,\cdot )\varphi \,\mathrm{d}x =\int _{\Omega }\varrho _{0}s(\varrho _{0},\vartheta _{0})\varphi \,\mathrm{d}x,\ \mathit{\mbox{ for all}}\ \varphi \in C_{c}^{\infty }(\overline{\Omega }). }$$
    (3.355)

Proof

Step 1: Unlike the proof of existence based on the classical theory of parabolic equations requiring \(\Omega\) to be a regular domain, the integrability properties (3.349)–(3.352) of the weak solutions follow directly from the total dissipation balance (2.52) and the space-time pressure estimates obtained by means of the operator \(\mathcal{B}\approx \mathrm{div}_{x}^{-1}\) introduced in Sect. 2.2.5; for more details, see estimates (2.40), (2.46), (2.66), (2.68), (2.73), (2.96) and (2.98). In particular, it is enough to assume \(\Omega \subset \mathbb{R}^{3}\) to be a bounded Lipschitz domain.

Step 2: Strong continuity in time of the density claimed in (3.349) is a general property of the renormalized solutions that follows from the DiPerna and Lions transport theory [85], see Lemma 11.14 in Appendix. Once \(\varrho \in C([0,T];L^{1}(\Omega ))) \cap C_{\mathrm{weak}}([0,T];L^{\frac{5} {3} }(\Omega ))\), we deduce from the momentum equation (2.9) and estimates (3.350)–(3.352) that one may take a representative of \(\mathbf{u} \in L^{q}(0,T;W^{1,p}(\Omega ))\) such that \(\mathbf{m}:=\varrho \mathbf{u} \in C_{\mathrm{weak}}([0,T];L^{\frac{5} {4} }(\Omega; \mathbb{R}^{3}))\). In addition, we may infer from the inequality

$$\displaystyle{\|\mathbf{m}(t)\|_{ L^{\frac{5} {4} }(\Omega )}^{2} \leq \|\varrho (t)\|_{ L^{\frac{5} {3} }(\Omega )}\|\varrho (t)\vert \mathbf{u}(t)\vert ^{2}\|_{ L^{\infty }(0,T;L^{1}(\Omega ))},\;t \in [0,T]}$$

that m(t) vanishes almost anywhere on the set \(\{x \in \Omega \,\vert \,\varrho (t) = 0\}\). The expression \(\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t)} 1_{\{\varrho (t)>0\}}\) is therefore defined for all t ∈ [0, T] and is equal to ϱ | u |2(t) a.a. on (0, T).

Since \(\int _{\Omega }\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t)+\varepsilon } \,\mathrm{d}x \leq \|\varrho \mathbf{u}\|_{L^{\infty }(0,T;L^{1}(\Omega ))}\) uniformly with ɛ → 0+, we deduce by the Beppo-Lévi monotone convergence theorem that

$$\displaystyle{\int _{\Omega }\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t) + \varepsilon } \,\mathrm{d}x \rightarrow \int _{\Omega }\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t)} 1_{\{\varrho (t)>0\}}\,\mathrm{d}x <\infty \;\mbox{ for all }t \in [0,T].}$$

This information together with (3.349) guarantees \(\mathbf{m}(t)/\sqrt{\varrho (t) +\varepsilon }\)C weak([0, T]; \(L^{2}(\Omega ))\). Therefore, for any α > 0 and sufficiently small 0 < ɛ < ɛ(α), and for any τ ∈ [0, T),

$$\displaystyle{\int _{\Omega }\frac{\vert \mathbf{m}(\tau )\vert ^{2}} {\varrho (\tau )} 1_{\{\varrho (\tau )>0\}}\,\mathrm{d}x-\alpha \leq \int _{\Omega }\frac{\vert \mathbf{m}(\tau )\vert ^{2}} {\varrho (\tau ) + \varepsilon } \,\mathrm{d}x}$$
$$\displaystyle{\leq \liminf _{t\rightarrow \tau +}\int _{\Omega }\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t)} 1_{\{\varrho (t)>\varepsilon \}}\,\mathrm{d}x \leq \liminf _{t\rightarrow \tau +}\int _{\Omega }\frac{\vert \mathbf{m}(t)\vert ^{2}} {\varrho (t)} 1_{\{\varrho (t)>0\}}\,\mathrm{d}x,}$$

where, to justify the inequality in the middle, we have used (3.349) and the lower weak semicontinuity of convex functionals discussed in Theorem 11.27 in Appendix. We have completed the proof of lower semicontinuity in time of the total kinetic energy, and, in particular, formula (3.353).

Step 3: In agreement with formulas (1.11)–(1.12), we deduce from the entropy balance (2.27) that

$$\displaystyle{[\varrho s(\varrho,\vartheta )](\tau +) \in \mathcal{M}^{+}(\overline{\Omega }),\;\tau \in [0,T),\quad [\varrho s(\varrho,\vartheta )](\tau -) \in \mathcal{M}^{+}(\overline{\Omega }),\;\tau \in (0,T],}$$
$$\displaystyle{[\varrho s(\varrho,\vartheta )](\tau +) \geq [\varrho s(\varrho,\vartheta )](\tau -),\;\tau \in (0,T),}$$

where the measures [ϱs(ϱ, ϑ)](τ+), τ ∈ [0, T) and [ϱs(ϱ, ϑ)](τ−), τ ∈ (0, T] are defined in the following way

$$\displaystyle{<[\varrho s(\varrho,\vartheta )](\tau \pm );\zeta> _{[\mathcal{M};C](\overline{\Omega })}:=\lim _{\delta \rightarrow 0+}\int _{I_{\tau,\delta }^{\pm }}\int _{\Omega }[\varrho s(\varrho,\vartheta )](t)[\psi _{\delta }^{\tau,\pm }]'(t)\zeta \,\mathrm{d}x\,\mathrm{d}t}$$
$$\displaystyle{ =\int _{\Omega }\varrho _{0}s(\varrho _{0},\vartheta _{0})\zeta \,\mathrm{d}x +\lim _{\delta \rightarrow 0+} <\sigma;\psi _{\delta }^{(\tau,\pm )}\zeta> _{ [\mathcal{M};C]([0,T]\times \overline{\Omega })} }$$
(3.356)
$$\displaystyle{+\int _{\Omega }\varrho _{0}s(\varrho _{0},\vartheta _{0})\zeta \,\mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\left (\varrho s(\varrho,\vartheta )\mathbf{u} + \frac{\mathbf{q}} {\vartheta } \right ) \cdot \nabla _{x}\zeta \,\mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\frac{\mathcal{Q}} {\vartheta } \zeta \,\mathrm{d}x.}$$

In this formula, \(\zeta \in C(\overline{\Omega })\), I τ, δ + = (τ, τ + δ), I τ, δ = (τδ, τ) and \(\psi _{\delta }^{(\tau,\pm )} \in C^{1}(\mathbb{R})\) are non increasing functions such that

$$\displaystyle{\psi _{\delta }^{(\tau,+)}(t) = \left \{\begin{array}{c} 1\;\mbox{ if }t \in (-\infty,\tau ),\\ \\ 0\;\mbox{ if }t \in [\tau +\delta,\infty ), \end{array} \right \},\;\psi _{\delta }^{(\tau,-)}(t) = \left \{\begin{array}{c} 1\;\mbox{ if }t \in (-\infty,\tau -\delta ),\\ \\ 0\;\mbox{ if }t \in [\tau,\infty ), \end{array} \right \}.}$$

According to the theorem about the Lebesgue points applied to function ϱs(ϱ, ϑ) (belonging to \(L^{\infty }(0,T;L^{1}(\Omega )\)), we may infer

$$\displaystyle{ <[\varrho s(\varrho,\vartheta )](\tau -);\zeta> _{[\mathcal{M};C](\overline{\Omega })} =<[\varrho s(\varrho,\vartheta )](\tau +);\zeta> _{[\mathcal{M};C](\overline{\Omega })} }$$
(3.357)
$$\displaystyle{=\int _{\Omega }[\varrho s(\varrho,\vartheta )](\tau )\zeta \,\mathrm{d}x,\;\zeta \in C_{c}^{\infty }(\overline{\Omega }),\;\zeta \geq 0\;\mbox{ for a.a. }\tau \in (0,T).}$$

Letting δ → 0+ in (3.356), we obtain

$$\displaystyle{ \int _{\Omega }[\varrho s(\varrho,\vartheta )](\tau +)\zeta \,\mathrm{d}x- <\sigma;\zeta> _{[\mathcal{M};C]([0,\tau ]\times \overline{\Omega })} }$$
(3.358)
$$\displaystyle{=\int _{\Omega }\varrho _{0}s(\varrho _{0},\vartheta _{0})\zeta \,\mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\left (\varrho \frac{\mathcal{Q}} {\vartheta } \zeta + (\varrho s(\varrho,\vartheta )\mathbf{u} + \frac{\mathbf{q}} {\vartheta } ) \cdot \nabla _{x}\zeta \right )\,\mathrm{d}x.}$$

In the remaining part of the proof, we shall show that

$$\displaystyle{ \mathrm{ess}\,\lim _{\tau \rightarrow 0+} <\sigma;\zeta> _{[\mathcal{M};C]([0,\tau ]\times \overline{\Omega })} = 0. }$$
(3.359)

Step 4: To this end we employ in the entropy balance (2.27) the test function φ(t, x) = ψ δ (τ, +)(t)ϑ 0(x), τ ∈ (0, T), which is admissible provided \(\vartheta _{0} \in W^{1,\infty }(\Omega )\). Using additionally (3.357), we get

$$\displaystyle{ \int _{\Omega }\left ([\varrho s(\varrho,\vartheta )](\tau ) -\varrho _{0}s(\varrho _{0},\vartheta _{0})\right )\vartheta _{0}\,\mathrm{d}x =<\sigma;\vartheta _{0}> _{[\mathcal{M};C]([0,\tau ]\times \overline{\Omega })} }$$
(3.360)
$$\displaystyle{+\int _{0}^{\tau }\int _{ \Omega }\left (\varrho s(\varrho,\vartheta )\mathbf{u} + \frac{\mathbf{q}} {\vartheta } \right ) \cdot \nabla _{x}\vartheta _{0}\,\mathrm{d}x +\int _{ 0}^{\tau }\int _{ \Omega }\frac{\mathcal{Q}} {\vartheta } \vartheta _{0}\,\mathrm{d}x}$$

for a.a. τ ∈ (0, T). On the other hand, the total energy balance (2.22) with the test function ψ = ψ δ (τ, +) yields

$$\displaystyle{ \int _{\Omega }\left ([\frac{1} {2\varrho }\vert \varrho \mathbf{u}\vert ^{2} +\varrho e(\varrho,\vartheta )](\tau ) - [ \frac{1} {2\varrho _{0}}\vert \varrho _{0}\mathbf{u}_{0}\vert ^{2} +\varrho _{ 0}e(\varrho _{0},\vartheta _{0})]\right )\,\mathrm{d}x }$$
(3.361)
$$\displaystyle{=\int _{ 0}^{\tau }\int _{ \Omega }\left (\varrho f\mathbf{u} +\varrho \mathcal{Q}\right )\,\mathrm{d}x\,\mathrm{d}t}$$

for a.a. τ ∈ (0, T). Now, we introduce the Helmholtz function

$$\displaystyle{H_{\vartheta _{0}}(\varrho,\vartheta ) =\varrho e(\varrho,\vartheta ) -\vartheta _{0}\varrho s(\varrho,\vartheta )}$$

and combine (3.360)–(3.361) to get

$$\displaystyle{\int _{\Omega }\left ([\frac{1} {2\varrho }\vert \varrho \mathbf{u}\vert ^{2}](\tau ) - \frac{1} {2\varrho _{0}}\vert \varrho _{0}\mathbf{u}_{0}\vert ^{2}\right )\,\mathrm{d}x +\int _{ \Omega }[H_{\vartheta _{0}}(\varrho,\vartheta ) - H_{\vartheta _{0}}(\varrho,\vartheta _{0})](\tau )\,\mathrm{d}x}$$
$$\displaystyle{ +\int _{\Omega }\left (H_{\vartheta _{0}}(\varrho (\tau ),\vartheta ) - H_{\vartheta _{0}}(\varrho _{0},\vartheta _{0}) - (\varrho (\tau ) -\varrho _{0})\frac{\partial H_{\vartheta _{0}}} {\partial \varrho } (\varrho _{0},\vartheta _{0})\right )\,\mathrm{d}x }$$
(3.362)
$$\displaystyle{+\int _{\Omega }(\varrho (\tau ) -\varrho _{0})\frac{\partial H_{\vartheta _{0}}} {\partial \varrho } (\varrho _{0},\vartheta _{0})\,\mathrm{d}x+ <\sigma;\vartheta _{0}> _{[\mathcal{M};C]([0,\tau ]\times \overline{\Omega })}}$$
$$\displaystyle{=\int _{ 0}^{\tau }\int _{ \Omega }\left (\varrho f\mathbf{u} +\varrho \mathcal{Q}\left (1 -\frac{\vartheta _{0}} {\vartheta } \right ) -\left (\varrho s(\varrho,\vartheta )\mathbf{u} + \frac{\mathbf{q}} {\vartheta } \right ) \cdot \nabla \vartheta _{0}\right )\,\mathrm{d}x\,\mathrm{d}t}$$

for a.a. τ in (0, T).

It follows from the thermodynamic stability hypothesis (1.44) that \(\varrho \mapsto H_{\vartheta _{0}}(\varrho,\vartheta _{0})\) is strictly convex for any fixed ϑ 0 and that \(\vartheta \mapsto H_{\vartheta _{0}}(\varrho,\vartheta )\) attains its global minimum at ϑ 0, see Sect. 2.2.3 for more details. Consequently,

$$\displaystyle{H_{\vartheta _{0}}(\varrho,\vartheta ) - H_{\vartheta _{0}}(\varrho,\vartheta _{0}) \geq 0,\;H_{\vartheta _{0}}(\varrho,\vartheta ) - H_{\vartheta _{0}}(\varrho _{0},\vartheta _{0}) - (\varrho -\varrho _{0})\frac{\partial H_{\vartheta _{0}}} {\partial \varrho } (\varrho _{0},\vartheta _{0}) \geq 0.}$$

Moreover, due to the strong continuity of density with respect to time stated in (3.349), we show

$$\displaystyle{\mathrm{lim}_{\tau \rightarrow 0+}\int _{\Omega }(\varrho (\tau ) -\varrho _{0})\frac{\partial H_{\vartheta _{0}}} {\partial \varrho } (\varrho _{0},\vartheta _{0})\,\mathrm{d}x = 0,}$$

while the last integral at the right hand side of (3.362) tends to 0 as τ → 0+ since the integrand belongs to \(L^{1}((0,T) \times \Omega )\). Thus, relation (3.362) reduces in the limit τ → 0+ to

$$\displaystyle{\mathrm{ess}\,\lim _{\tau \rightarrow 0+} <\sigma;\vartheta _{0}> _{[\mathcal{M};C]([0,\tau ]\times \overline{\Omega })},}$$

whence \(\mathrm{ess}\,\mathrm{lim}_{\tau \rightarrow 0+}\sigma \left [[0,\tau ] \times \overline{\Omega }\right ] = 0\) and (3.359) follows. Having in mind identity (3.357), statement (3.355) now follows by letting τ → 0+ in (3.358) (evidently, the right hand side in (3.358) tends to zero as the integrand belongs to \(L^{1}((0,T) \times \Omega )\)).

Theorem 3.2 is proved.