Sbornik: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
License agreement
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Mat. Sb.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Sbornik: Mathematics, 2023, Volume 214, Issue 4, Pages 444–478
DOI: https://doi.org/10.4213/sm9800e
(Mi sm9800)
 

This article is cited in 1 scientific paper (total in 1 paper)

Properties and errors of second-order parabolic and hyperbolic perturbations of a first-order symmetric hyperbolic system

A. A. Zlotnikab, B. N. Chetverushkinb

a Higher School of Economics University, Moscow, Russia
b Keldysh Institute of Applied Mathematics of Russian Academy of Sciences, Moscow, Russia
References:
Abstract: The Cauchy problems are studied for a first-order multidimensional symmetric linear hyperbolic system of equations with variable coefficients and its singular perturbations that are second-order strongly parabolic and hyperbolic systems of equations with a small parameter $\tau>0$ multiplying the second derivatives with respect to $x$ and $t$. The existence and uniqueness of weak solutions of all three systems and $\tau$-uniform estimates for solutions of systems with perturbations are established. Estimates for the difference of solutions of the original system and the systems with perturbations are given, including ones of order $O(\tau^{\alpha/2})$ in the norm of $C(0,T;L^2(\mathbb{R}^n))$, for an initial function $\mathbf w_0$ in the Sobolev space $H^\alpha(\mathbb{R}^n)$, $\alpha=1,2$, or the Nikolskii space $H_2^{\alpha}(\mathbb{R}^n)$, $0<\alpha<2$, $\alpha\neq 1$, and under appropriate assumptions on the free term of the first-order system. For ${\alpha=1/2}$ a wide class of discontinuous functions $\mathbf w_0$ is covered. Estimates for derivatives of any order with respect to $x$ for solutions and of order $O(\tau^{\alpha/2})$ for their differences are also deduced. Applications of the results to the first-order system of gas dynamic equations linearized at a constant solution and to its perturbations, namely, the linearized second-order parabolic and hyperbolic quasi-gasdynamic systems of equations, are presented.
Bibliography: 34 titles.
Keywords: linear systems of partial differential equations, small parameter, linearized system of gas dynamic equations, quasi-gasdynamic systems of equations, estimates of the difference of solutions.
Funding agency Grant number
Russian Science Foundation 22-11-00126
This research was supported by the Russian Science Foundation under grant no. 22-11-00126, https://rscf.ru/en/project/22-11-00126/.
Received: 07.06.2022 and 07.01.2023
Russian version:
Matematicheskii Sbornik, 2023, Volume 214, Number 4, Pages 3–37
DOI: https://doi.org/10.4213/sm9800
Bibliographic databases:
Document Type: Article
MSC: Primary 35L40; Secondary 35L51, 35K40
Language: English
Original paper language: Russian

§ 1. Introduction

We study the Cauchy problems for a first-order $n$-dimensional symmetric linear hyperbolic system of equations with variable coefficients and for its singular perturbations that are second-order strongly parabolic and hyperbolic systems of equations with a small parameter $\tau>0$ multiplying the second derivatives with respect to $x$ and $t$. The perturbations involving second derivatives with respect to $x$ have a divergent form and contain matrices with variable coefficients. Perturbations of this type can be useful, in particular, when numerical methods to solve first-order hyperbolic systems are developed; a similar approach has already been implemented and used in practice for many years for the quasilinear system of gas dynamic equations (see [1]–[3]). The parabolic perturbation is a version of the well-known artificial viscosity method (see, for example, [4]). Hyperbolic perturbations in various settings were also considered in the literature, mainly when the original problem is posed for a first-order hyperbolic equation, rather than a system, including in the quasilinear case (see, in particular, [5]–[11]).

This paper begins with presenting auxiliary results on weak and strong solutions of the first-order system in § 2. Though much attention has been paid to systems of this type in the literature (see, for example, [4], [12]–[14]), these results improve slightly and supplement the known results regarding conditions on the free term $\mathbf f$ and the coefficients. This is achieved using an appropriate combination of different known approaches. In § 3 the properties, including $\tau$-uniform estimates, of weak solutions of second-order parabolic and hyperbolic systems are deduced. For the second system the structural condition that the composite matrix of its leading coefficients must dominate the composite matrix of leading coefficients of the first-order system of type known in the literature is imposed.

Estimates for the differences $\mathbf r_\tau$ of solutions of the original system and systems with perturbations, including estimates in the norm of $C(0,T;L^2(\mathbb{R}^n))$, under appropriate conditions on the initial data $\mathbf w_0$, the free term $\mathbf f$ and the coefficients of the systems are proved in § 4 on the basis of results from § 2 and § 3. They include estimates of order $O(\tau^{\alpha/2})$ for $\mathbf w_0$ in the Sobolev spaces $H^\alpha(\mathbb{R}^n)$ for $\alpha=1,2$ and in the Nikolskii spaces $H_2^{\alpha}(\mathbb{R}^n)$ (see [15]) for $0<\alpha<2$, $\alpha\neq 1$. For $\alpha=1/2$ we cover $\mathbf w_0\in \mathrm{BV}(\mathbb{R}^n)\cap L^\infty(\mathbb{R}^n)$, where $\mathrm{BV}(\mathbb{R}^n)$ is the space of functions of bounded variation on $\mathbb{R}^n$, thus covering a wide class of discontinuous functions $\mathbf w_0$ which is of importance for applications. For the second-order hyperbolic system, to have these estimates for $\alpha<2$ it is essential to use averaging (for example, in the sense of Steklov) with step size $\tau$ for $\mathbf w_0$. The corresponding conditions of smoothness of order $\alpha$ with respect to $x$, and also of order $\alpha/2$ with respect to $t$ in the case of a second-order hyperbolic perturbation, in the sense of the Sobolev or Nikolskii spaces constructed on the basis of anisotropic Lebesgue space $L^{2,1}(\mathbb{R}^n\times (0,T))$, are imposed on $\mathbf f$. For the second-order hyperbolic perturbation it is sufficient to set the initial value of the derivative of the solution with respect to $t$ equal to zero.

In § 5, estimates for derivatives, of any order with respect to $x$ for solutions of all systems under consideration and of order $O(\tau^{\alpha/2})$, $0<\alpha\leqslant 2$, for $\mathbf r_\tau$, are derived from the previous results. For greater transparency the coefficients of the systems are assumed to be independent of $x$. The dependence on $T$ is traceable in these estimates. Estimates for $\mathbf r_\tau$ in the uniform norm are indicated.

The final section, § 6, is devoted to applications of our results to the linearized (at a constant solution) first-order system of gas dynamic equations and its perturbations that are linearized second-order parabolic and hyperbolic quasi-gasdynamic systems of equations with constant coefficients (see [16] and [17]). We do this in a uniform manner for both systems, by deducing preliminarily the appropriate properties of the viscous terms (containing the second derivatives with respect to $x$) in these systems. The results of this paper were briefly presented in [18].

Note that second-order hyperbolic perturbations of parabolic problems have also been intensively studied in the literature, including estimates for the difference of their solutions obtained in [19]–[22].

§ 2. Notation. First-order symmetric hyperbolic system of equations and its properties

We introduce the Hilbert spaces $L^2(\mathbb{R}^n)=H^0(\mathbb{R}^n)$ (the Lebesgue space) and $H^l(\mathbb{R}^n)$, $l=1,2$ (the Sobolev spaces), with the inner products

$$ \begin{equation*} (v,w)_{L^2(\mathbb{R}^n)}=\int_{\mathbb{R}^n}vw\,dx, \quad (v,w)_{H^l(\mathbb{R}^n)}=\sum_{0\leqslant k\leqslant l}\int_{\mathbb{R}^n}\nabla^kv\cdot\nabla^kw\,dx, \qquad n\geqslant 1 \end{equation*} \notag $$
(all spaces are assumed to be real). For brevity let $\|\cdot\|_{\mathbb{R}^n}=\|\cdot\|_{L^2(\mathbb{R}^n)}$ in the proofs. Throughout, $\nabla=(\partial_1,\dots,\partial_n)$, $\nabla^2=\{\partial_i\partial_j\}_{i,j=1}^n$ is the matrix of second-order partial derivatives with respect to $x$, and the symbol $\cdot$ denotes the scalar product of vectors or matrices (unless otherwise specified).

We consider a layer $\Pi_T:=\mathbb{R}^n\times (0,T)$ and will use the Sobolev space $H^1(\Pi_T)$. Let $L^{2,q}(\Pi_T)$ and $W_{2,q}^{l,0}(\Pi_T)$, $W_{2,q}^{l,1}(\Pi_T)$, where $1\leqslant q\leqslant\infty$, $l=1,2$, denote the anisotropic Lebesgue and Sobolev spaces with the norms

$$ \begin{equation*} \begin{gathered} \, \|v\|_{L^{2,q}(\Pi_T)}=\bigl\|\|v(x,t)\|_{L^2(\mathbb{R}^n)}\bigr\|_{L^q(0,T)}, \\ \|v\|_{W_{2,q}^{1,0}(\Pi_T)}=\|\{v,\nabla v\}\|_{L^{2,q}(\Pi_T)} \equiv\|v\|_{L^{2,q}(\Pi_T)}+\|\nabla v\|_{L^{2,q}(\Pi_T)}, \\ \|v\|_{W_{2,q}^{2,0}(\Pi_T)} =\|\{v,\nabla v,\nabla^2v\}\|_{L^{2,q}(\Pi_T)} \end{gathered} \end{equation*} \notag $$
and
$$ \begin{equation*} \|v\|_{W_{2,q}^{l,1}(\Pi_T)}=\|v\|_{W_{2,q}^{l,0}(\Pi_T)} +\|\partial_t v\|_{L^{2,q}(\Pi_T)}. \end{equation*} \notag $$
Let $V_2(\Pi_T)$ be the space of functions $v\in L^{2,\infty}(\Pi_T)$ such that $\nabla v \in L^2(\Pi_T)$ (see [23] and [24]) and $C(0,T;L^2(\mathbb{R}^n))$ be the space of continuous functions $v$: ${[0,T]\to L^2(\mathbb{R}^n)}$ with the norm $\|v\|_{C(0,T;L^2(\mathbb{R}^n))}=\max_{0\leqslant t\leqslant T}\|v(\,\cdot\,,t)\|_{L^2(\mathbb{R}^n)}$.

We introduce the products

$$ \begin{equation*} (\mathbf v,\mathbf w)_{\mathbb{R}^n}=\int_{\mathbb{R}^n}\mathbf v\cdot\mathbf w\,dx \quad\text{and}\quad (\mathbf v,\mathbf w)_{\Pi_T}=\int_{\Pi_T}\mathbf v\cdot\mathbf w\,dx\,dt \end{equation*} \notag $$
of vector functions $\mathbf v$ and $\mathbf w$ such that $\mathbf v\cdot\mathbf w\in L^1(\mathbb{R}^n)$ or $L^1(\Pi_T)$, respectively. Below, all vectors and vector functions are assumed to be columns, and the operators $\nabla^l$, $\partial_t^l$, $l=1,2$, are applied elementwise to vector- and matrix-valued functions. Unless otherwise specified, the Lebesgue norms of vector functions $\mathbf v$ and $\nabla\mathbf v$ and square matrix functions $A$ are understood as the norms in these spaces of the Euclidean norms $|\mathbf v|$ and $(|\partial_1\mathbf v|^2+\dots+|\partial_n\mathbf v|^2)^{1/2}$ and the spectral norm $\|A\|$. Furthermore, the Lebesgue norms of $\nabla^2\mathbf v$, $\nabla A$ and $\nabla^2 A$ are understood as the sums of the norms of the corresponding partial derivatives of $\mathbf v$ and $A$. For composite rectangular matrix functions and their derivatives, the suitable norms of the square matrices forming them are added.

We introduce the Cauchy problem

$$ \begin{equation} \mathcal{H}\mathbf w:=\partial_t\mathbf w+B_i\partial_i\mathbf w+C\mathbf w=\mathbf f \quad \text{in } \Pi_T, \qquad \mathbf w|_{t=0}=\mathbf w_0 \quad \text{in } \mathbb{R}^n \end{equation} \tag{2.1} $$
for a first-order symmetric hyperbolic system. Here $\mathbf w(x,t),\mathbf f(x,t)\colon\Pi_T\to\mathbb{R}^m$, $\mathbf w_0(x)\colon\mathbb{R}^n\to\mathbb{R}^m$ are the sought-for and given vector functions, respectively, and the coefficients $B_i(x,t)=B_i^T(x,t)$, $i=1,\dots,n$, and $C(x,t)$ are matrix functions of order $m$. Throughout, summation from 1 to $n$ over the repeated indices $i$ and $j$ (and only over such indices) is assumed; $\delta^{(ij)}$ is the Kronecker symbol.

Assume that

$$ \begin{equation} \mathbf B,\operatorname{div}\mathbf B, C\in L^\infty(\Pi_T), \qquad \mathbf f\in L^{2,1}(\Pi_T)\quad\text{and} \quad \mathbf w_0\in L^2(\mathbb{R}^n), \end{equation} \tag{2.2} $$
where $\mathbf B=(B_1,\dots,B_n)$ is the composite matrix of leading coefficients of size $m\times mn$ and $\operatorname{div}\mathbf B:=\partial_iB_i$. Here and in what follows the membership of a matrix- or vector-valued functions in some space means that all its components belong to this space. To simplify the presentation, conditions on the coefficients and their derivatives are stated mainly in terms of the spaces $L^\infty(\Pi_T)$, rather than $L^{p,q}(\Pi_T)$.

A weak solution of the Cauchy problem (2.1) is a function $\mathbf w\in L^{2,\infty}(\Pi_T)$ satisfying the integral identity

$$ \begin{equation} (\mathbf w,\mathcal{H}^*\boldsymbol\varphi)_{\Pi_T}=\ell(\mathbf w_0,\mathbf f;\boldsymbol\varphi):=(\mathbf w_0,\boldsymbol\varphi_0)_{\mathbb{R}^n} +(\mathbf f,\boldsymbol\varphi)_{\Pi_T} \quad \forall\,\boldsymbol\varphi\in W_{2,1}^{1,1}(\Pi_T), \ \ \boldsymbol\varphi|_{t=T}=0. \end{equation} \tag{2.3} $$
Throughout, $\mathcal{H}^*\boldsymbol\varphi=-\partial_t\boldsymbol\varphi-B_i\partial_i \boldsymbol\varphi-(\operatorname{div}\mathbf B-C^T)\boldsymbol\varphi$ is the Lagrange-conjugate operator of $\mathcal{H}$, $\boldsymbol\varphi(x,t)$: $\Pi_T\to\mathbb{R}^m$ and $\boldsymbol\varphi_0:=\boldsymbol\varphi|_{t=0}$.

A strong solution of this Cauchy problem is a function $\mathbf w\in L^{2,\infty}(\Pi_T)$ with $\nabla\mathbf w\in L^{2,\infty}(\Pi_T)$ and $\partial_t\mathbf w\in L^{2,1}(\Pi_T)$ that satisfies the equation in (2.1) in the space $L^{2,1}(\Pi_T)$ and the initial condition $\mathbf w|_{t=0}=\mathbf w_0$ in the space $C(0,T;L^2(\mathbb{R}^n))$. A strong solution is also a weak solution, while a weak solution such that $\nabla\mathbf w\in L^{2,\infty}(\Pi_T)$ and $\partial_t\mathbf w\in L^{2,1}(\Pi_T)$ is a strong solution.

Below we use conditions like $\|\mathbf B\|_{L^\infty(\Pi_T)}\leqslant N$, where $N\geqslant 1$ is a parameter; for brevity we always assume automatically that $\mathbf B\in L^\infty(\Pi_T)$ in these conditions. Nonnegative constants $C(N,T)\geqslant 0$, $C_1(N,T)\geqslant 0$, $\dots$, which are nondecreasing with respect to $N$ and $T$ arise in the text; the same notation can be used for different constants. Let $I_m$ be the identity matrix of order $m$. Let $\Delta_\xi C(x,t):=C(x+\xi,t)-C(x,t)$.

We begin with the properties of weak and strong solutions of the Cauchy problem (2.1).

Theorem 1. 1. a) Let conditions (2.2) hold and $\frac{1}{2}\operatorname{div}\mathbf B-C\leqslant c_0I_m$ almost everywhere (a.e.) in $\Pi_T$ for a constant $c_0>0$ (we can assume that $c_0\leqslant\|\frac{1}{2}\operatorname{div}\mathbf B-C\|_{L^\infty(\Pi_T)}$). Then there exists a weak solution $\mathbf w$ of the Cauchy problem (2.1) such that

$$ \begin{equation} \|\mathbf w\|_{L^{2,\infty}(\Pi_T)}\leqslant e^{c_0T}\bigl(\|\mathbf w_0\|_{L^2(\mathbb{R}^n)}+2\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{2.4} $$

b) In addition, if $\nabla\mathbf B\in L^\infty(\Pi_T)$ and $\displaystyle \lim_{\xi\to 0}\int_0^T\|\Delta_\xi C\|_{L^\infty(\mathbb{R}^n)}\,dt=0$, then the weak solution is unique. It also has the property $\mathbf w\in C(0,T;L^2(\mathbb{R}^n))$; therefore, the above estimate takes the form

$$ \begin{equation} \|\mathbf w\|_{C(0,T;L^2(\mathbb{R}^n))}\leqslant e^{c_0T}\bigl(\|\mathbf w_0\|_{L^2(\mathbb{R}^n)}+2\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{2.5} $$

2. Let $\|\{\mathbf B,C,\nabla\mathbf B,\nabla C\}\|_{L^\infty(\Pi_T)}\leqslant N$, $\mathbf f,\nabla\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_0\in H^1(\mathbb{R}^n)$. Then the weak solution $\mathbf w$ is a strong solution, which is unique and satisfies the estimate

$$ \begin{equation} \|\nabla\mathbf w\|_{L^{2,\infty}(\Pi_T)}+\|\partial_t\mathbf w\|_{L^{2,1}(\Pi_T)} \leqslant C_1(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{1,0}(\Pi_T)}\bigr). \end{equation} \tag{2.6} $$

If also $\mathbf f\in L^{2,q}(\Pi_T)$ for some $1\leqslant q\leqslant\infty$, then $\partial_t\mathbf w\in L^{2,q}(\Pi_T)$ and

$$ \begin{equation} \|\partial_t\mathbf w\|_{L^{2,q}(\Pi_T)}\leqslant C_2(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)} +\|\mathbf f\|_{L^{2,q}(\Pi_T)}+\|\nabla\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{2.7} $$

3. a) Let, in addition, $\|\{\nabla^2\mathbf B,\nabla^2C\}\|_{L^\infty(\Pi_T)}\leqslant N$, $\nabla^2\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_0\in H^2(\mathbb{R}^n)$. Then the strong solution $\mathbf w$ has $\nabla^2\mathbf w \in L^{2,\infty}(\Pi_T)$, $\partial_t\nabla\mathbf w \in L^{2,1}(\Pi_T)$, and

$$ \begin{equation} \|\nabla^2\mathbf w\|_{L^{2,\infty}(\Pi_T)}+\|\partial_t\nabla\mathbf w\|_{L^{2,1}(\Pi_T)} \leqslant C_3(N,T)\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr). \end{equation} \tag{2.8} $$

b) If also $\|\{\partial_t\mathbf B,\partial_tC\}\|_{L^\infty(\Pi_T)}\leqslant N$ and $\mathbf f,\nabla\mathbf f,\partial_t\mathbf f\in L^{2,q}(\Pi_T)$ for some $q$, ${1\leqslant q\leqslant\infty}$, then, in addition, there exists $\partial_t^2\mathbf w\in L^{2,q}(\Pi_T)$ and

$$ \begin{equation} \|\partial_t^2\mathbf w\|_{L^{2,q}(\Pi_T)} \leqslant C_4(N,T)\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)} +\|\{\mathbf f,\nabla\mathbf f,\partial_t\mathbf f\}\|_{L^{2,q}(\Pi_T)}\bigr). \end{equation} \tag{2.9} $$

Proof. а) We introduce averaging with respect to $x$ as follows:
$$ \begin{equation} (\sigma^{(h)}v)(x):=\int_{\Omega_h}e_h(\xi)v(x+\xi)\,d\xi =\int_{\mathbb{R}^n}e_h(x-\xi)v(\xi)\,d\xi \quad \forall\, v\in L_{\mathrm{loc}}^1(\mathbb{R}^n), \end{equation} \tag{2.10} $$
where $h>0$, $e_h(\xi)=h^{-n}\widehat{e}(\xi_1/h)\cdots \widehat{e}(\xi_n/h)$, $\xi=(\xi_1,\dots,\xi_n)$, $\widehat{e}(a)\,{=}\,\max\{1- |a|,0\}$ and $\Omega_h=(-h,h)^n$.

First of all, we recall that the property $B_i=B_i^T$, $i=1,\dots,n$, and integration by parts with respect to $x$ (that is, with respect to $x_1,\dots,x_n$) yield $(B_i\partial_i\mathbf v,\mathbf v)_{\mathbb{R}^n}=-(\mathbf v,(\operatorname{div}\mathbf B)\mathbf v+B_i\partial_i\mathbf v)_{\mathbb{R}^n}$, which implies that

$$ \begin{equation} (B_i\partial_i\mathbf v,\mathbf v)_{\mathbb{R}^n} =-\frac{1}{2}((\operatorname{div}\mathbf B)\mathbf v,\mathbf v)_{\mathbb{R}^n} \quad \forall\, \mathbf v\in H^1(\mathbb{R}^n), \end{equation} \tag{2.11} $$
where $B_i=B_i(\,\cdot\,,t)$ and $\operatorname{div}\mathbf B=\operatorname{div}\mathbf B(\,\cdot\,,t)$ for almost all $t\in (0,T)$. Here we suppose first that $\partial_kB_k\in L^\infty(\Pi_T)$, $k=1,\dots,n$. This assumption is eliminated by introducing the averagings $\sigma^{(h)}B_k$, using the estimates $\|\sigma^{(h)}B_k\|_{L^\infty(\Pi_T)}\leqslant\|B_k\|_{L^\infty(\Pi_T)}$, $\|\operatorname{div}\sigma^{(h)}\mathbf B\|_{L^\infty(\Pi_T)}\leqslant\|\operatorname{div}\mathbf B\|_{L^\infty(\Pi_T)}$, and taking the $*$-weak limits in $L^\infty(\mathbb{R}^n)$
$$ \begin{equation*} \sigma^{(h)}B_k(\,\cdot\,,t)\to B_k(\,\cdot\,,t)\quad\text{and} \quad \operatorname{div}\sigma^{(h)}\mathbf B(\,\cdot\,,t)=\sigma^{(h)}\operatorname{div}\mathbf B(\,\cdot\,,t)\to\operatorname{div}\mathbf B(\,\cdot\,,t) \end{equation*} \notag $$
for some sequence $h=h_l(t)\to 0$ as $l\to \infty$ and almost all $t\in (0,T)$.

To prove the existence of the weak solution, it is convenient to use the artificial viscosity method (see, for example, [4], Section 7.3) by introducing the following Cauchy problem for a strongly parabolic system of equations with the simplest principal part:

$$ \begin{equation} \partial_t\mathbf y-\tau\Delta\mathbf y+B_i\partial_i\mathbf y+C\mathbf y=\mathbf f \quad \text{in } \Pi_T, \qquad \mathbf y|_{t=0}=\mathbf w_0 \quad \text{in } \mathbb{R}^n, \end{equation} \tag{2.12} $$
with a small parameter $\tau>0$, where $\Delta$ is the Laplace operator. From Theorem 2 (see below), which is not related to the result we prove, it follows, in particular, that this problem has a weak solution $\mathbf y=\mathbf y_\tau\in V_2(\Pi_T)$ satisfying the integral identity
$$ \begin{equation} (-\mathbf y,\partial_t\boldsymbol\varphi)_{\Pi_T}+\tau(\nabla\mathbf y,\nabla\boldsymbol\varphi)_{\Pi_T} +(B_i\partial_i\mathbf y+C\mathbf y,\boldsymbol\varphi)_{\Pi_T} =\ell(\mathbf w_0,\mathbf f;\boldsymbol\varphi) \end{equation} \tag{2.13} $$
for all $\boldsymbol\varphi\in H^1(\Pi_T)$, $\boldsymbol\varphi|_{t=T}=0$. In addition, the following $\tau$-uniform estimate is true:
$$ \begin{equation} \max\bigl\{\|\mathbf y_\tau\|_{L^{2,\infty}(\Pi_T)},\sqrt{2\tau}\|\nabla\mathbf y_\tau\|_{L^2(\Pi_T)}\bigr\} \leqslant e^{c_0T}\bigl(\|\mathbf y_0\|_{\mathbb{R}^n}+2\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{2.14} $$

Due to this estimate there exists a function $\mathbf w\in L^{2,\infty}(\Pi_T)$ such that ${\mathbf y_{\tau_k}\to\mathbf w}$ $*$-weakly in $L^{2,\infty}(\Pi_T)$ and $\tau_k\nabla\mathbf y_{\tau_k}\to 0$ in $L^2(\Pi_T)$ for some sequence $\tau_k\to 0$. For $\mathbf w$, estimate (2.4) holds. Taking the limit in the integral identity (2.13) for $\mathbf y=\mathbf y_{\tau_k}$ (using that $(B_i\partial_i\mathbf y,\boldsymbol\varphi)_{\Pi_T}=-(\mathbf y,B_i\partial_i\boldsymbol\varphi+(\operatorname{div}\mathbf B)\boldsymbol\varphi)_{\Pi_T}$) leads to the integral identity (2.3), that is, the function $\mathbf w$ constructed is a weak solution of the Cauchy problem (2.1).

b) We establish the uniqueness of a weak solution under the assumptions in part 1, b). With this aim in view, we obtain estimate (2.4) for any weak solution $\mathbf w$. It suffices to assume that $\mathbf f=0$ and $\mathbf w_0=0$. We use the Friedrichs method (see [12], Ch. 6, § 2). We substitute the function $\boldsymbol\varphi=\sigma^{(h)}\boldsymbol\psi$ into the integral identity (2.3). Transferring $\sigma^{(h)}$ to the other factors (which is possible for averagings with kernel $e_h(\xi)$ that is even with respect to $\xi_1,\dots,\xi_n$) and integrating by parts with respect to $x$, we obtain

$$ \begin{equation*} -(\sigma^{(h)}\mathbf w,\partial_t\boldsymbol\psi)_{\Pi_T} +\bigl(\partial_i\sigma^{(h)}(B_i\mathbf w)-\sigma^{(h)}[(\operatorname{div}\mathbf B-C)\mathbf w],\boldsymbol\psi\bigr)_{\Pi_T}=0 \end{equation*} \notag $$
for any $\boldsymbol\psi\in W_{2,1}^{1,1}(\Pi_T)$, $\boldsymbol\psi|_{t=T}=0$. Note that $\sigma^{(h)}\mathbf w,\nabla\sigma^{(h)}\mathbf w\in L^{2,\infty}(\Pi_T)$; by virtue of the last identity it is true that
$$ \begin{equation*} \partial_t\sigma^{(h)}\mathbf w=-\partial_i\sigma^{(h)}(B_i\mathbf w)+\sigma^{(h)}[(\operatorname{div}\mathbf B-C)\mathbf w]\in L^{2,\infty}(\Pi_T) \end{equation*} \notag $$
and $\sigma^{(h)}\mathbf w\big|_{t=0}=0$. We rewrite these results as the following Cauchy problem with the strong solution $\sigma^{(h)}\mathbf w$:
$$ \begin{equation} \begin{gathered} \, \mathcal{H}\sigma^{(h)}\mathbf w =\mathbf f_1^{(h)}+\mathbf f_2^{(h)}, \qquad \sigma^{(h)}\mathbf w|_{t=0}=0, \\ \mathbf f_1^{(h)}:=B_i\partial_i\sigma^{(h)}\mathbf w-\partial_i\sigma^{(h)}(B_i\mathbf w)+\sigma^{(h)}[(\operatorname{div}\mathbf B)\mathbf w], \nonumber \\ \mathbf f_2^{(h)}:=C\sigma^{(h)}\mathbf w-\sigma^{(h)}(C\mathbf w). \nonumber \end{gathered} \end{equation} \tag{2.15} $$

An estimate similar to (2.4) holds for $\sigma^{(h)}\mathbf w$. It can be derived like (2.14), but for $\tau=0$; therefore, it is slightly simpler. More precisely, taking the inner product in $L^2(\mathbb{R}^n)$ of the equation in (2.15) and $\sigma^{(h)}\mathbf w$, in view of (2.11) we obtain

$$ \begin{equation*} \begin{aligned} \, \frac{1}{2}\partial_t\bigl(\|\sigma^{(h)}\mathbf w\|_{\mathbb{R}^n}^2\bigr) &=\Bigl(\Bigl(\frac{1}{2}\operatorname{div}\mathbf B-C\Bigr)\sigma^{(h)}\mathbf w,\sigma^{(h)}\mathbf w\Bigr)_{\mathbb{R}^n} +(\mathbf f_1^{(h)}+\mathbf f_2^{(h)},\sigma^{(h)}\mathbf w)_{\mathbb{R}^n} \\ &\leqslant c_0\|\sigma^{(h)}\mathbf w\|_{\mathbb{R}^n}^2 +\|\mathbf f_1^{(h)}+\mathbf f_2^{(h)}\|_{\mathbb{R}^n}\|\sigma^{(h)}\mathbf w\|_{\mathbb{R}^n} \end{aligned} \end{equation*} \notag $$
almost everywhere on $(0,T)$. This differential inequality implies the estimate
$$ \begin{equation} \begin{aligned} \, \nonumber T^{-1/2}\|\sigma^{(h)}\mathbf w\|_{L^2(\Pi_T)} &\leqslant \|\sigma^{(h)}\mathbf w\|_{C(0,T;L^2(\mathbb{R}^n))} \\ &\leqslant 2e^{c_0T}\bigl(\|\mathbf f_1^{(h)}\|_{L^{2,1}(\Pi_T)}+\|\mathbf f_2^{(h)}\|_{L^{2,1}(\Pi_T)}\bigr) \end{aligned} \end{equation} \tag{2.16} $$
in the standard way, where the added left-hand inequality is obvious (cf. [12], Proposition 6.1).

We bound the functions $\mathbf f_1^{(h)}$ and $ \mathbf f_2^{(h)}$. Supposing provisionally that $\partial_i\mathbf w\in L_{\mathrm{loc}}^1(\mathbb{R}^n)$, $i=1,\dots,n$, we obtain

$$ \begin{equation*} \begin{aligned} \, \mathbf f_1^{(h)} &=\bigl(B_i\sigma^{(h)}\partial_i\mathbf w-\sigma^{(h)}(B_i\partial_i\mathbf w)\bigr)(x,t) \\ &=-\int_{\Omega_h}e_h(\xi)[B_i(x+\xi,t)-B_i(x,t)] \partial_i\mathbf w(x+\xi,t)\,d\xi. \end{aligned} \end{equation*} \notag $$
Since $\partial_i\mathbf w(x+\xi,t)=\partial_{\xi_i}\mathbf w(x+\xi,t)=\partial_{\xi_i}[\mathbf w(x+\xi,t)-\mathbf w(x,t)]$, integration by parts with respect to $x$ results in the formula
$$ \begin{equation*} \mathbf f_1^{(h)}(x,t)=\int_{\Omega_h}\bigl[(\partial_ie_h)(\xi)\Delta_\xi B_i(x,t) +e_h(\xi)\partial_iB_i(x+\xi,t)\bigr]\Delta_\xi\mathbf w(x,t)\,d\xi. \end{equation*} \notag $$
Like the original formula, this one does not involve $\partial_i\mathbf w$. Hence, by replacing $\mathbf w$ by $\sigma^{(h)}\mathbf w$ in this formula and taking the limit as $h\to 0$ we verify it for any ${\mathbf w\in L^2\mkern-1mu(\Pi_T)}$. It is also clear that
$$ \begin{equation*} \mathbf f_2^{(h)}(x,t)=-\int_{\Omega_h}e_h(\xi)(\Delta_\xi C)(x,t)\mathbf w(x+\xi,t)\,d\xi. \end{equation*} \notag $$
By the generalized Minkowski inequality and the assumptions in part 1, b), we have
$$ \begin{equation} \begin{aligned} \, \nonumber \|\mathbf f_1^{(h)}\|_{L^{2,1}(\Pi_T)} & \leqslant\int_{\Omega_h}\biggl[e_h^{(i)}(\xi) \frac{|\xi|}{h}\|\partial_\xi B_i\|_{L^\infty(\Pi_T)} \\ \nonumber &\qquad\qquad +e_h(\xi)\|\operatorname{div}\mathbf B(\cdot+\xi)\|_{L^\infty(\Pi_T)}\biggr]\|\Delta_\xi\mathbf w\|_{L^{2,1}(\Pi_T)}\,d\xi \\ &\leqslant\bigl(c\|\nabla\mathbf B\|_{L^\infty(\Pi_T)}+\|\operatorname{div}\mathbf B\|_{L^\infty(\Pi_T)}\bigr) \sup_{\xi\in\Omega_h}\|\Delta_\xi\mathbf w\|_{L^{2,1}(\Pi_T)}\to 0 \end{aligned} \end{equation} \tag{2.17} $$
and
$$ \begin{equation} \begin{aligned} \, \nonumber \|\mathbf f_2^{(h)}\|_{L^{2,1}(\Pi_T)} &\leqslant\int_{\Omega_h}e_h(\xi)\|\Delta_\xi C\|_{L^{\infty,1}(\Pi_T)}\|\mathbf w(\cdot+\xi)\|_{L^{2,\infty}(\Pi_T)}\,d\xi \\ &\leqslant\biggl(\frac{1}{h^n}\int_{\Omega_h}\|\Delta_\xi C\|_{L^{\infty,1}(\Pi_T)}\,d\xi\biggr) \|\mathbf w\|_{L^{2,\infty}(\Pi_T)}\to 0 \end{aligned} \end{equation} \tag{2.18} $$
as $h\to 0$, where $e_h^{(i)}(\xi)$ is obtained from $e_h(\xi)$ by replacing the factor $\widehat{e}_h(\xi_i)$ by $h^{-1}$ and $\partial_\xi$ is the derivative in the direction $\xi$. Since also $\sigma^{(h)}\mathbf w\to\mathbf w$ in $L^2(\Pi_T)$, we have $\|\sigma^{(h)}\mathbf w\|_{L^2(\Pi_T)}\to\|\mathbf w\|_{L^2(\Pi_T)}$, and taking the limit on the left- and right-hand sides of (2.16) yields $\|\mathbf w\|_{L^2(\Pi_T)}=0$, that is, $\mathbf w=0$. The uniqueness of the weak solution is proved. The condition on $\nabla\mathbf B$ in (2.17), and therefore also in this proof, can be relaxed to $\nabla\mathbf B\in L^{\infty,q}(\Pi_T)$ for some $q>1$. We deduce the property $\mathbf w\in C(0,T;L^2(\mathbb{R}^n))$ below after we prove part 2.

c) We introduce the following difference quotient and shift operator with respect to the coordinate $x_j$:

$$ \begin{equation*} \delta_{jh}v(x)=\frac{1}{h}[v(x+h\check{\mathbf e}_j)-v(x)] \quad (h\neq 0)\quad\text{and}\quad s_{jh}v=v(x+h\check{\mathbf e}_j), \qquad j=1,\dots,n, \end{equation*} \notag $$
where $\check{\mathbf e}_1,\dots,\check{\mathbf e}_n$ is the canonical basis in $\mathbb{R}^n$. Recall that by the generalized Newton-Leibniz formula and Minkowski’s inequality we have the estimate
$$ \begin{equation} \|\delta_{jh}v\|_{L^p(\mathbb{R}^n)}\leqslant\|\partial_jv\|_{L^p(\mathbb{R}^n)} \quad \forall\, v,\partial_jv\in L^p(\mathbb{R}^n), \quad 1\leqslant p\leqslant\infty. \end{equation} \tag{2.19} $$
Applying the operator $\delta_{jh}$ formally to the Cauchy problem (2.12) for $\mathbf y=\mathbf y_\tau$ and using the formula $\delta_{jh}(vw)=v\delta_{jh}w+(\delta_{jh}v)s_{jh}w$, we obtain
$$ \begin{equation*} \begin{gathered} \, \begin{split} &\partial_t\delta_{jh}\mathbf y_\tau-\tau\Delta\delta_{jh}\mathbf y_\tau+B_i\partial_i\delta_{jh}\mathbf y_\tau+C\delta_{jh}\mathbf y_\tau \\ &\qquad =-(\delta_{jh}B_i)\partial_is_{jh}\mathbf y_\tau-(\delta_{jh}C)s_{jh}\mathbf y_\tau+\delta_{jh}\mathbf f \quad \text{in } \Pi_T, \end{split} \\ \delta_{jh}\mathbf y_\tau|_{t=0}=\delta_{jh}\mathbf w_0 \quad \text{in } \mathbb{R}^n. \end{gathered} \end{equation*} \notag $$
This Cauchy problem arises in a rigorous way if we substitute the function ${\boldsymbol\varphi=-\delta_{j(-h)}\boldsymbol\psi}$ into the integral identity (2.13) and use the elementary formula
$$ \begin{equation*} (\mathbf v,-\delta_{j(-h)}\boldsymbol\psi)_{\mathbb{R}^n}=(\delta_{jh}\mathbf v,\boldsymbol\psi)_{\mathbb{R}^n} \quad \forall\,\mathbf v\in L^2(\mathbb{R}^n). \end{equation*} \notag $$

We apply (2.14) to the weak solution $\delta_{jh}\mathbf y_\tau$ of this problem and then use (2.19) under the assumptions on $\mathbf B$ and $C$ from part 2. Therefore, we obtain

$$ \begin{equation*} \begin{aligned} \, &\max\bigl\{\|\delta_{jh}\mathbf y_\tau\|_{L^{2,\infty}(\Pi_T)},\sqrt{2\tau}\|\nabla\delta_{jh}\mathbf y_\tau\|_{L^2(\Pi_T)}\bigr\} \leqslant 2e^{c_0T}\bigl(\|\delta_{jh}\mathbf w_0\|_{\mathbb{R}^n} \\ &\quad\qquad +\|(\delta_{jh}B_i)\partial_is_{jh}\mathbf y_\tau\|_{L^{2,1}(\Pi_T)} +\|(\delta_{jh}C)s_{jh}\mathbf y_\tau\|_{L^{2,1}(\Pi_T)} +\|\delta_{jh}\mathbf f\|_{L^{2,1}(\Pi_T)} \bigr) \\ &\quad \leqslant 2e^{c_0T}\bigl(\|\partial_j\mathbf w_0\|_{\mathbb{R}^n} +\|\partial_jB_i\|_{L^\infty(\Pi_T)}\|\partial_i\mathbf y_\tau\|_{L^{2,1}(\Pi_T)} +\|\partial_jC\|_{L^\infty(\Pi_T)}\|\mathbf y_\tau\|_{L^{2,1}(\Pi_T)} \\ &\quad\qquad +\|\partial_j\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{aligned} \end{equation*} \notag $$
Thus, in view of (2.14) we have
$$ \begin{equation*} \begin{aligned} \, &\max_{j=1,\dots,n}\|\delta_{jh}\mathbf y_\tau\|_{L^{2,\infty}(\Pi_T)} \\ &\qquad \leqslant C(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\nabla\mathbf f\}\|_{L^{2,1}(\Pi_T)} +\|\nabla\mathbf y_\tau\|_{L^{2,1}(\Pi_T)}\bigr), \end{aligned} \end{equation*} \notag $$
where $C(N,T)$ is independent of $h$ and, we recall, nondecreasing with respect to $T$. Using the theorem on the relationship between the derivative and the difference quotient (see, for example, [15], § 4.8) we conclude first that $\nabla\mathbf y_\tau\in L^2(\Pi_T)$ and then that $\nabla\mathbf y_\tau\in L^{2,\infty}(\Pi_T)$ and, moreover, that
$$ \begin{equation*} \|\nabla\mathbf y_\tau\|_{L^{2,\infty}(\Pi_T)} \leqslant C(N,T)\biggl(\!\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\nabla\mathbf f\}\|_{L^{2,1}(\Pi_T)} +\int_0^T\!\|\nabla\mathbf y_\tau(\,\cdot\,,t)\|_{\mathbb{R}^n}\,dt\!\biggr). \end{equation*} \notag $$

It is possible to replace $T$ in this inequality by any $t$, $0<t\leqslant T$; therefore, it is also true that

$$ \begin{equation} \begin{aligned} \, &\operatorname*{ess\,sup}_{0<\theta<t}\|\nabla\mathbf y_\tau(\,\cdot\,,\theta)\|_{\mathbb{R}^n} \nonumber \\ & \qquad \leqslant C(N,T)\biggl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\nabla\mathbf f\}\|_{L^{2,1}(\Pi_T)} +\int_0^t\|\nabla\mathbf y_\tau(\,\cdot\,,\theta)\|_{\mathbb{R}^n}\,d\theta\biggr). \end{aligned} \end{equation} \tag{2.20} $$

We have

$$ \begin{equation*} y(t)=\lim_{h\to +0}\frac{1}{h}\int_{t-h}^ty(\theta)\,d\theta \leqslant\operatorname*{ess\,sup}_{0<\theta<t}y(\theta) \quad \text{a.e. on}\ (0,T) \quad \forall\, y\in L^\infty(0,T), \end{equation*} \notag $$
where we set $y(t):=0$ for $t<0$. The limit exists owing to a variant of the theorem on Lebesgue points of an integrable function (see [25], Theorem 1.34). Therefore, the left-hand side of (2.20) can be estimated from below by $\|\nabla\mathbf y_\tau(\,\cdot\,,t)\|_{\mathbb{R}^n}$ for almost all $t\in (0,T)$. Then from Grönwall’s lemma in an integral form (see, for example, [4], Appendix B.2k)) we obtain the $\tau$-uniform estimate
$$ \begin{equation*} \|\nabla\mathbf y_\tau\|_{L^{2,\infty}(\Pi_T)} \leqslant C_1(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\nabla\mathbf f\}\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation*} \notag $$
In the $*$-weak limit in $L^{2,\infty}(\Pi_T)$ as $\tau\to 0$ it yields the estimate
$$ \begin{equation} \|\nabla\mathbf w\|_{L^{2,\infty}(\Pi_T)} \leqslant C_1(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\nabla\mathbf f\}\|_{L^{2,1}(\Pi_T)}\bigr), \end{equation} \tag{2.21} $$
where $\mathbf w$ can be regarded as the weak solution constructed above.

Integrating by parts with respect to $x$ we write the integral identity (2.3) for $\mathbf w$ in the form

$$ \begin{equation*} -(\mathbf w,\partial_t\boldsymbol\varphi)_{\Pi_T} {=}\,(\mathbf w_0,\boldsymbol\varphi|_{t=0})_{\mathbb{R}^n} +(\mathbf f-B_i\partial_i\mathbf w-C\mathbf w,\boldsymbol\varphi)_{\Pi_T} \quad \forall\,\boldsymbol\varphi\,{\in}\, W_{2,1}^{1,1}(\Pi_T),\ \boldsymbol\varphi|_{t=T}\,{=}\,0. \end{equation*} \notag $$
It implies that $\partial_t\mathbf w=\mathbf f-B_i\partial_i\mathbf w-C\mathbf w\in L^{2,1}(\Pi_T)$ and, furthermore, $\mathbf w\in C(0,T; L^2(\mathbb{R}^n))$ and $\mathbf w|_{t=0}=\mathbf w_0$, that is, the weak solution is a strong solution.

If $\mathbf f\in L^{2,q}(\Pi_T)$ for some $1\leqslant q\leqslant\infty$, then by the equation in (2.1) we have

$$ \begin{equation*} \begin{aligned} \, \|\partial_t\mathbf w\|_{L^{2,q}(\Pi_T)} & \leqslant \|B_i\|_{L^\infty(\Pi_T)}\|\partial_i\mathbf w\|_{L^{2,q}(\Pi_T)} +\|C\|_{L^\infty(\Pi_T)}\|\mathbf w\|_{L^{2,q}(\Pi_T)} +\|\mathbf f\|_{L^{2,q}(\Pi_T)} \\ &\leqslant c(N)T^{1/q}\|\{\nabla\mathbf w,\mathbf w\}\|_{L^{2,\infty}(\Pi_T)}+\|\mathbf f\|_{L^{2,q}(\Pi_T)}. \end{aligned} \end{equation*} \notag $$
The last estimate, (2.4) and (2.21) imply (2.6) and (2.7).

The assumption on $C$ in part 1, b) is much weaker than $\nabla C\in L^\infty(\Pi_T)$; therefore, the strong solution is unique.

d) Now, under the assumptions in part 1, b) we deduce the property $\mathbf w\in C(0,T; L^2(\mathbb{R}^n))$ of the weak solution, which will complete the proof of part 1. Let $C_k=\sigma^{(1/k)}C$, $k=1,2,\dots$; then $\nabla C_k\in L^\infty(\Pi_T)$ and

$$ \begin{equation*} \begin{aligned} \, \|C-C_k\|_{L^{\infty,1}(\Pi_T)} & =\biggl\|\int_{\Omega_{1/k}}e_{1/k}(\xi)\Delta_\xi C\,d\xi\biggr\|_{L^{\infty,1}(\Pi_T)} \\ &\leqslant k^n\int_{\Omega_{1/k}}\|\Delta_\xi C\|_{L^{\infty,1}(\Pi_T)}\,d\xi\to 0 \end{aligned} \end{equation*} \notag $$
as $k\to\infty$ similarly to (2.18). For $\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_0\in L^2(\mathbb{R}^n)$ we construct sequences of functions $\mathbf f_k=\sigma^{(1/k)}\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_{0k}=\sigma^{(1/k)}\mathbf w_0\in H^1(\mathbb{R}^n)$ such that $\nabla\mathbf f_k\in L^{2,1}(\Pi_T)$, $\mathbf f_k\to\mathbf f$ in $L^{2,1}(\Pi_T)$ and $\mathbf w_{0k}\to\mathbf w_0$ in $L^2(\mathbb{R}^n)$ as $k\to\infty$.

Let $\mathbf w_k$ be the strong solution of the Cauchy problem corresponding to $C=C_k$, $\mathbf f=\mathbf f_k$ and $\mathbf w_0=\mathbf w_{0k}$. Then it satisfies an estimate of type (2.4) or, more precisely, of type (2.5):

$$ \begin{equation} \begin{aligned} \, \|\mathbf w_k\|_{C(0,T;L^2(\mathbb{R}^n))} &\leqslant e^{\widetilde{c}_0T}\bigl(\|\mathbf w_{0k}\|_{L^2(\mathbb{R}^n)}+2\|\mathbf f_k\|_{L^{2,1}(\Pi_T)}\bigr) \nonumber \\ &\leqslant e^{\widetilde{c}_0T}\bigl(\|\mathbf w_{0}\|_{L^2(\mathbb{R}^n)}+2\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr), \end{aligned} \end{equation} \tag{2.22} $$
where $\widetilde{c}_0=\frac{1}{2}\|\operatorname{div}\mathbf B\|_{L^\infty(\Pi_T)}+\|C\|_{L^\infty(\Pi_T)}$. In addition, the difference $\mathbf w_k-\mathbf w_l$ is the strong solution of the Cauchy problem
$$ \begin{equation*} \partial_t(\mathbf w_k-\mathbf w_l)+B_i\partial_i(\mathbf w_k-\mathbf w_l)+C_k(\mathbf w_k-\mathbf w_l) =\mathbf f_k-\mathbf f_l+(C_l-C_k)\mathbf w_l \quad \text{in } \Pi_T \end{equation*} \notag $$
and $(\mathbf w_k-\mathbf w_l)|_{t=0}=\mathbf w_{0k}-\mathbf w_{0l}$. Therefore, it obeys an estimate similar to the left-hand estimate in (2.22), namely,
$$ \begin{equation*} \begin{aligned} \, &\|\mathbf w_k-\mathbf w_l\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant 2e^{\widetilde{c}_0T}\bigl(\|\mathbf w_{0k}-\mathbf w_{0l}\|_{L^2(\mathbb{R}^n)} \\ &\qquad\qquad +\|\mathbf f_k-\mathbf f_l\|_{L^{2,1}(\Pi_T)} +\|C_k-C_l\|_{L^{\infty,1}(\Pi_T)}\|\mathbf w_l\|_{L^{2,\infty}(\Pi_T)}\bigr). \end{aligned} \end{equation*} \notag $$
Now the $k$-uniform estimate (2.22) and the fact that $C_k$, $\mathbf f_k$ and $\mathbf w_{0k}$ are Cauchy sequences in $L^{\infty,1}(\Pi_T)$, $L^{2,1}(\Pi_T)$ and $L^2(\mathbb{R}^n)$, respectively, imply that $\mathbf w_k$ is a Cauchy sequence in $C(0,T;L^2(\mathbb{R}^n))$. Thus, $\mathbf w_k\to\mathbf w$ in $C(0,T;L^2(\mathbb{R}^n))$. Passing to the limit in identity (2.3) written for $\mathbf w=\mathbf w_k$, $C=C_k$, $\mathbf f=\mathbf f_k$ and $\mathbf w_0=\mathbf w_{0k}$ we conclude that the function $\mathbf w$ constructed is a (unique) weak solution of the Cauchy problem (2.1).

e) Assume that the assumptions in part 3, a) of the theorem hold. For the strong solution $\mathbf w$, we introduce the functions $\mathbf v_k=\partial_k\mathbf w\in L^{2,\infty}(\Pi_T)$, $k=1,\dots,n$, and obtain an extended system of equations for them (see [13], § 12). We choose $\boldsymbol\varphi:=-\partial_k\boldsymbol\psi$ in the integral identity (2.3), use the formula

$$ \begin{equation*} \mathcal{H}^*\partial_k\boldsymbol\psi=\partial_k\mathcal{H}^*\boldsymbol\psi+\partial_j[(\partial_kB_j)\boldsymbol\psi]-(\partial_kC^T)\boldsymbol\psi, \end{equation*} \notag $$
integrate by parts with respect to $x_k$ and $x_j$, and obtain the integral identity
$$ \begin{equation*} (\mathbf v_k,\mathcal{H}^*\boldsymbol\psi)_{\Pi_T}+((\partial_kB_j)\mathbf v_j,\boldsymbol\psi)_{\Pi_T} =(\partial_k\mathbf w_0,\boldsymbol\psi|_{t=0})_{\mathbb{R}^n} +(\partial_k\mathbf f-(\partial_kC)\mathbf v_k,\boldsymbol\psi)_{\Pi_T} \end{equation*} \notag $$
for all $\boldsymbol\psi$ such that $\boldsymbol\psi,\nabla\boldsymbol\psi\in W_{2,1}^{1,1}(\Pi_T)$ and $\boldsymbol\psi|_{t=T}=0$. Functions of this type are dense in the space of functions $\boldsymbol\psi\in W_{2,1}^{1,1}(\Pi_T)$ such that $\boldsymbol\psi|_{t=T}=0$. This means that the composite vector function $\nabla\mathbf w=(\mathbf v_1,\dots,\mathbf v_n)$: $\Pi_T\to\mathbb{R}^{mn}$ is a weak solution of the following Cauchy problem for the extended hyperbolic system of equations with $n$ sought-for vector functions:
$$ \begin{equation*} \begin{gathered} \, \partial_t\mathbf v_k+B_i\partial_i\mathbf v_k+C\mathbf v_k+(\partial_kB_j)\mathbf v_j= \mathbf f_k:= \partial_k\mathbf f-(\partial_kC)\mathbf w \quad \text{in } \Pi_T, \\ \mathbf v_k|_{t=0}=\partial_k\mathbf w_0 \quad \text{in } \mathbb{R}^n, \end{gathered} \end{equation*} \notag $$
where $k=1,\dots,n$ and the functions $\mathbf f_k$ are considered as known.

This Cauchy problem is similar to the original one. Under the assumptions in part 3, a), we can apply parts 1 and 2 to it. Consequently, its weak solution $\nabla\mathbf w$ is unique, is a strong solution, and satisfies the estimate

$$ \begin{equation*} \begin{aligned} \, &\|\nabla(\nabla\mathbf w)\|_{L^{2,\infty}(\Pi_T)}+\|\partial_t(\nabla\mathbf w)\|_{L^{2,1}(\Pi_T)} \\ &\qquad \leqslant C(N,T)\bigl(\|\nabla(\nabla\mathbf w_0)\|_{\mathbb{R}^n} +\|\{\nabla\mathbf f-(\nabla C)\mathbf w,\nabla[\nabla\mathbf f-(\nabla C)\mathbf w]\}\|_{L^{2,1}(\Pi_T)}\bigr) \\ &\qquad \leqslant C_1(N,T)\bigl(\|\nabla^2\mathbf w_0\|_{\mathbb{R}^n} +\|\{\nabla\mathbf f,\nabla^2\mathbf f\}\|_{L^{2,1}(\Pi_T)} +\|\{\mathbf w,\nabla\mathbf w\}\|_{L^{2,1}(\Pi_T)}\bigr), \end{aligned} \end{equation*} \notag $$
where $(\nabla C)\mathbf w=((\partial_1C)\mathbf w,\dots,(\partial_nC)\mathbf w)$ and, in addition to parts 1 and 2, the inequality $\|\{\nabla^2\mathbf B,\nabla^2 C\}\|_{L^\infty(\Pi_T)}\leqslant N$ is taken into account. This inequality and the above estimates (2.4) and (2.6) yield (2.8).

If also $\partial_t\mathbf B,\partial_tC\in L^\infty(\Pi_T)$ and $\mathbf f,\nabla\mathbf f,\partial_t\mathbf f\in L^{2,q}(\Pi_T)$ for some $1\leqslant q\leqslant\infty$, then we can apply $\partial_t$ to the equation in (2.1), then use it again and derive the formula

$$ \begin{equation*} \begin{aligned} \, \partial_t^2\mathbf w &=-B_i\partial_t\partial_i\mathbf w\mathbf -(\partial_tB_i)\partial_i\mathbf w\mathbf -C\partial_t\mathbf w-(\partial_tC)\mathbf w+\partial_t\mathbf f \\ &=B_i\partial_i(B_j\partial_j\mathbf w+C\mathbf w-\mathbf f)-(\partial_tB_i)\partial_i\mathbf w\mathbf -C\partial_t\mathbf w-(\partial_tC)\mathbf w+\partial_t\mathbf f. \end{aligned} \end{equation*} \notag $$
From the conditions imposed in part 3 on the norms of $\mathbf B$ and $C$ it follows that
$$ \begin{equation*} \begin{aligned} \, &\|\partial_t^2\mathbf w\|_{L^{2,q}(\Pi_T)} \leqslant c(N)\bigl(T^{1/q}\|\{\mathbf w,\nabla\mathbf w,\nabla^2\mathbf w\}\|_{L^{2,\infty}(\Pi_T)} \\ &\qquad\qquad +\|\partial_t\mathbf w\|_{L^{2,q}(\Pi_T)} +\|\nabla\mathbf f\|_{L^{2,q}(\Pi_T)}\bigr) +\|\partial_t\mathbf f\|_{L^{2,q}(\Pi_T)}. \end{aligned} \end{equation*} \notag $$
By virtue of (2.4)(2.8) estimate (2.9) is also true. Theorem 1 is proved.

§ 3. Second-order strongly parabolic and hyperbolic systems of equations with a small parameter $\tau$

We now introduce the Cauchy problem for a second-order parabolic system of equations of the following form that is a perturbation of the system in (2.1) with a small parameter $0<\tau\leqslant\overline{\tau}$ multiplying the second derivatives with respect to $x$:

$$ \begin{equation} \mathcal{P}_\tau\mathbf y:=\mathcal{H}\mathbf y-\tau\partial_i(A_{ij}\partial_j\mathbf y)=\mathbf f_\tau+\tau\partial_i\mathbf g_{i\tau} \quad \text{in } \Pi_T, \qquad \mathbf y|_{t=0}=\mathbf y_{0\tau} \quad \text{in } \mathbb{R}^n. \end{equation} \tag{3.1} $$
Here $\mathbf y(x,t),\mathbf f_\tau(x,t),\mathbf g_{i\tau}(x,t)$: $\Pi_T\to\mathbb{R}^m$ and $\mathbf y_{0\tau}(x)$: $\mathbb{R}^n\to\mathbb{R}^m$ are the sought-for and given vector functions, and the coefficients $A_{ij}\in L^\infty(\Pi_T)$ are matrix functions of order $m$, $i,j=1,\dots,n$.

Following [24], Ch. V, § 1, we assume that

$$ \begin{equation} \nu\|\nabla\mathbf v\|_{L^2(\mathbb{R}^n)}^2-\mu\|\mathbf v\|_{L^2(\mathbb{R}^n)}^2 \leqslant (A_{ij}\partial_j\mathbf v,\partial_i\mathbf v)_{\mathbb{R}^n} \quad \forall\,\mathbf v\in H^1(\mathbb{R}^n) \quad \text{a.e. in } (0,T) \end{equation} \tag{3.2} $$
for some constants $\nu>0$ and $\mu\geqslant 0$. For $\mu=0$ this condition is obviously implied by the cruder algebraic inequality
$$ \begin{equation} \nu\bigl(|\mathbf w_1|^2+\dots+|\mathbf w_n|^2\bigr) \leqslant (A_{ij}(x,t)\mathbf w_j)\cdot \mathbf w_i \quad \forall\,\mathbf w_1,\dots,\mathbf w_n\in \mathbb{R}^m \quad \text{a.e. in } \Pi_T \end{equation} \tag{3.3} $$
for the same $\nu>0$. If we introduce the block matrix $A=\{A_{ij}\}_{i,j=1}^n$ of order $mn$ and the block column vector $\mathbf w=(\mathbf w_1,\dots,\mathbf w_n)$ of length $mn$, then the last inequality takes the compact form
$$ \begin{equation*} \nu|\mathbf w|^2\leqslant (A(x,t)\mathbf w)\cdot\mathbf w \quad \forall\,\mathbf w\in \mathbb{R}^{mn} \quad \text{a.e. in } \Pi_T. \end{equation*} \notag $$
Assume that
$$ \begin{equation} B_i,C,A_{ij}\in L^\infty(\Pi_T), \quad i,j=1,\dots,n, \qquad \mathbf f_\tau\in L^{2,1}(\Pi_T), \qquad \mathbf y_{0\tau}\in L^2(\mathbb{R}^n), \end{equation} \tag{3.4} $$
and $\mathbf g_\tau:=(\mathbf g_{1\tau},\dots,\mathbf g_{n\tau})\in L^2(\Pi_T)$. A weak solution of the Cauchy problem (3.1) is a function $\mathbf y\in V_2(\Pi_T)$ satisfying the integral identity
$$ \begin{equation} \begin{aligned} \, &-(\mathbf y,\partial_t\boldsymbol\varphi)_{\Pi_T}+(B_i\partial_i\mathbf y+C\mathbf y,\boldsymbol\varphi)_{\Pi_T} +\tau(A_{ij}\partial_j\mathbf y,\partial_i\boldsymbol\varphi)_{\Pi_T} \nonumber \\ &\qquad =\ell(\mathbf y_{0\tau},\mathbf f_{\tau};\boldsymbol\varphi)-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\Pi_T} \end{aligned} \end{equation} \tag{3.5} $$
for any $\boldsymbol\varphi\in H^1(\Pi_T)$, $\boldsymbol\varphi|_{t=T}=0$.

Theorem 2. Assume that conditions (3.2) and (3.4) hold, $\operatorname{div}\mathbf B\in L^\infty(\Pi_T)$, $\frac{1}{2}\operatorname{div}\mathbf B-C\leqslant c_0I_m$ almost everywhere in $\Pi_T$ and $\mathbf g_\tau\in L^2(\Pi_T)$. Then there exists a unique weak solution $\mathbf y=\mathbf y_\tau$ of the Cauchy problem (3.1); it belongs to $C(0,T;L^2(\mathbb{R}^n))$ and satisfies the energy estimate

$$ \begin{equation} \begin{aligned} \, &\max\bigl\{\|\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \sqrt{\nu\tau}\|\nabla\mathbf y_\tau\|_{L^2(\Pi_T)}\bigr\} \nonumber \\ &\qquad \leqslant e^{\overline{c}_0T}\bigl(\|\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +2\|\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}+\sqrt{\nu^{-1}\tau}\|\mathbf g_\tau\|_{L^2(\Pi_T)}\bigr) \textit{ with } \overline{c}_0:=c_0+\overline{\tau}\mu. \end{aligned} \end{equation} \tag{3.6} $$

Proof. 1. First we deduce (3.6) as an a priori estimate under the assumption that $\partial_t\mathbf y\in L^2(\Pi_T)$. Then the integral identity (3.5) yields the integral identity
$$ \begin{equation*} (\partial_t\mathbf y,\boldsymbol\varphi)_{\mathbb{R}^n}+(B_i\partial_i\mathbf y+C\mathbf y,\boldsymbol\varphi)_{\mathbb{R}^n} +\tau(A_{ij}\partial_j\mathbf y,\partial_i\boldsymbol\varphi)_{\mathbb{R}^n} =(\mathbf f_{\tau},\boldsymbol\varphi)_{\mathbb{R}^n}-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\mathbb{R}^n} \end{equation*} \notag $$
for any $\boldsymbol\varphi\in H^1(\mathbb{R}^n)$ almost everywhere on $(0,T)$ and the initial condition $\mathbf y|_{t=0}=\mathbf y_{0\tau}$ in $C(0,T;L^2(\mathbb{R}^n))$. For $\boldsymbol\varphi=\mathbf y$, using (2.11), we derive the following energy relations from it:
$$ \begin{equation*} \begin{aligned} \, &\frac{1}{2}\partial_t\bigl(\|\mathbf y\|_{\mathbb{R}^n}^2\bigr) +\tau\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n} \\ &\qquad =\Bigl(\Bigl(\frac{1}{2}\operatorname{div}\mathbf B-C\Bigr)\mathbf y,\mathbf y\Bigr)_{\mathbb{R}^n}+(\mathbf f_\tau,\mathbf y)_{\mathbb{R}^n} -\tau(\mathbf g_{i\tau},\partial_i\mathbf y)_{\mathbb{R}^n} \\ &\qquad \leqslant c_0\|\mathbf y\|_{\mathbb{R}^n}^2+\|\mathbf f_\tau\|_{\mathbb{R}^n}\|\mathbf y\|_{\mathbb{R}^n} +\tau\|\mathbf g_\tau\|_{\mathbb{R}^n}\|\nabla\mathbf y\|_{\mathbb{R}^n} \quad \text{a.e. on } (0,T). \end{aligned} \end{equation*} \notag $$
By (3.2), the inequality $\frac{1}{2}\operatorname{div}\mathbf B-C\leqslant c_0I_m$ and Cauchy’s inequality with $\varepsilon$ we have
$$ \begin{equation*} \frac{1}{2}\partial_t\bigl(\|\mathbf y\|_{\mathbb{R}^n}^2\bigr) +\frac{1}{2}\nu\tau\|\nabla\mathbf y\|_{\mathbb{R}^n}^2 \leqslant (c_0+\overline{\tau}\mu)\|\mathbf y\|_{\mathbb{R}^n}^2+\|\mathbf f_\tau\|_{\mathbb{R}^n}\|\mathbf y\|_{\mathbb{R}^n} +\frac{1}{2}\nu^{-1}\tau\|\mathbf g_\tau\|_{\mathbb{R}^n}^2 \end{equation*} \notag $$
almost everywhere on $(0,T)$. This differential inequality implies the a priori estimate (3.6) in a standard way.

2. To establish the existence of a solution, we introduce the related initial-boundary value problem

$$ \begin{equation*} \begin{gathered} \, \mathcal{P}_\tau\mathbf y^{(k)}=\mathbf f_\tau+\tau\partial_i\mathbf g_{i\tau} \quad \text{in } \Pi_{kT}:=\Omega_k\times (0,T), \\ \mathbf y^{(k)}|_{\Gamma_{kT}}=0, \qquad \mathbf y^{(k)}|_{t=0}=\mathbf y_{0\tau} \quad \text{in } \Omega_k \end{gathered} \end{equation*} \notag $$
with parameter $k\geqslant 1$, where $\Omega_k:=(-k,k)^n$, $\Gamma_{kT}=\partial\Omega_k\times (0,T)$, and $\partial\Omega_k$ is the boundary of $\Omega_k$. By definition, its weak solution $\mathbf y^{(k)}\in V_2(\Pi_{kT})$ satisfies the integral identity
$$ \begin{equation} \begin{aligned} \, &-(\mathbf y^{(k)},\partial_t\boldsymbol\varphi)_{\Pi_{kT}}+(B_i\partial_i\mathbf y^{(k)}+C\mathbf y^{(k)},\boldsymbol\varphi)_{\Pi_{kT}} +\tau(A_{ij}\partial_j\mathbf y^{(k)},\partial_i\boldsymbol\varphi)_{\Pi_{kT}} \nonumber \\ &\qquad =(\mathbf y_{0\tau},\boldsymbol\varphi|_{t=0})_{\Omega_k} +(\mathbf f_\tau,\boldsymbol\varphi)_{\Pi_{kT}}-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\Pi_{kT}} \end{aligned} \end{equation} \tag{3.7} $$
for any $\boldsymbol\varphi\in H^1(\Pi_{kT})$ such that $\boldsymbol\varphi|_{\Gamma_{kT}}=0$, $\boldsymbol\varphi|_{t=T}=0$, and also satisfies the boundary condition $\mathbf y^{(k)}|_{\Gamma_{kT}}=0$. Here $V_2(\Pi_{kT})$, $(\,\cdot\,{,}\,\cdot\,)_{\Pi_{kT}}$, and $(\,\cdot\,{,}\,\cdot\,)_{\Omega_k}$ are introduced similarly to $V_2(\Pi_T)$, $(\,\cdot\,{,}\,\cdot\,)_{\Pi_T}$, and $(\,\cdot\,{,}\,\cdot\,)_{\mathbb{R}^n}$.

The existence of a weak solution $\mathbf y^{(k)}$ is proved in a standard way using the Galerkin method (see [23], Ch. III). First for the Galerkin approximations and then for $\mathbf y^{(k)}$ itself, after taking the limit, we prove quite similarly to (3.6) that

$$ \begin{equation} \begin{aligned} \, &\max\bigl\{\|\mathbf y^{(k)}\|_{L^{2,\infty}(\Pi_{kT})}, \sqrt{\nu\tau}\|\nabla\mathbf y_\tau\|_{L^2(\Pi_{kT})}\bigr\} \nonumber \\ &\qquad \leqslant e^{\overline{c}_0T}\bigl(\|\mathbf y_{0\tau}\|_{L^2(\Omega_k)} +2\|\mathbf f_\tau\|_{L^{2,1}(\Pi_{kT})}+\sqrt{\nu^{-1}\tau}\|\mathbf g_\tau\|_{L^2(\Pi_{kT})}\bigr). \end{aligned} \end{equation} \tag{3.8} $$

We set additionally $\mathbf y^{(k)}=0$ on $\Pi_T\setminus\Pi_{kT}$. Then $\mathbf y^{(k)}\in V_2(\Pi_T)$ and inequality (3.8) yields a $k$-uniform estimate similar to (3.6) with $\mathbf y^{(k)}$ in place of $\mathbf y_\tau$ and $L^{2,\infty}(\Pi_T)$ in place of $C(0,T;L^2(\mathbb{R}^n))$. By virtue of this estimate, there exists a function $\mathbf y\in V_2(\Pi_T)$ such that $\mathbf y^{(k)}\to\mathbf y$ $*$-weakly in $L^{2,\infty}(\Pi_T)$ and $\nabla\mathbf y^{(k)}\to \nabla\mathbf y$ weakly in $L^2(\Pi_T)$ for some subsequence $k=k_l\to\infty$. Estimate (3.6) holds for $\mathbf y=\mathbf y_\tau$ with the norm in $L^{2,\infty}(\Pi_T)$ instead of $C(0,T;L^2(\mathbb{R}^n))$.

We rewrite the integral identity (3.7) as (3.5) with $\mathbf y^{(k)}$ playing the role of $\mathbf y$ as follows:

$$ \begin{equation*} \begin{aligned} \, &-(\mathbf y^{(k)},\partial_t\boldsymbol\varphi)_{\Pi_T}+(B_i\partial_i\mathbf y^{(k)}+C\mathbf y^{(k)},\boldsymbol\varphi)_{\Pi_T} +\tau(A_{ij}\partial_j\mathbf y^{(k)},\partial_i\boldsymbol\varphi)_{\Pi_T} \\ &\qquad =\ell(\mathbf y_{0\tau},\mathbf f_{\tau};\boldsymbol\varphi)-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\Pi_T}, \end{aligned} \end{equation*} \notag $$
where $\boldsymbol\varphi$ is the same as in (3.7) and $\boldsymbol\varphi=0$ on $\Pi_T\setminus\Pi_{kT}$. Taking the limit as $k=k_l\to\infty$ on the left-hand side leads to the integral identity (3.5), where, in addition, $\boldsymbol\varphi$ has a compact support with respect to $x$. This property does not restrict the generality; therefore, the function $\mathbf y$ constructed is a weak solution of the Cauchy problem (3.1).

3. We establish that $\mathbf y\in C(0,T;L^2(\mathbb{R}^n))$ and also the uniqueness of the weak solution. To prove the first result we assume provisionally that $\mathbf f_\tau\in L^2(\Pi_T)$. Then for any weak solution $\mathbf y$ we can rewrite (3.5) in the form

$$ \begin{equation} -(\mathbf y,\partial_t\boldsymbol\varphi)_{\Pi_T}=(\mathbf b,\boldsymbol\varphi)_{\Pi_T}-(\mathbf b_i,\partial_i\boldsymbol\varphi)_{\Pi_T} \quad \forall\,\boldsymbol\varphi\in H^1(\Pi_T), \quad \boldsymbol\varphi|_{t=0,T}=0, \end{equation} \tag{3.9} $$
where $\mathbf b=\mathbf f_\tau-(B_i\partial_i\mathbf y+C\mathbf y),\mathbf b_i=\tau(A_{ij}\partial_j\mathbf y+\mathbf g_{i\tau})\in L^2(\Pi_T)$. Hence there exists a derivative with respect to $t$ in the sense of distributions $\partial_t\mathbf y\in L^2(0,T; H^{-1}(\mathbb{R}^n))$, where $H^{-1}(\mathbb{R}^n)=[H^1(\mathbb{R}^n)]^*$; since we also have $\mathbf y\in L^2(0,T;H^1(\mathbb{R}^n))$, it follows that $\mathbf y\in C(0,T;L^2(\mathbb{R}^n))$ (see, for example, [26], Ch. IV, Theorem 1.17). In addition, it follows from (3.9) that
$$ \begin{equation*} \begin{aligned} \, &\int_0^T\langle\partial_t\mathbf y(\,\cdot\,,t),\boldsymbol\varphi(\,\cdot\,,t)\rangle_{\mathbb{R}^n}\,dt+(B_i\partial_i\mathbf y+C\mathbf y,\boldsymbol\varphi)_{\Pi_T} +\tau(A_{ij}\partial_j\mathbf y,\partial_i\boldsymbol\varphi)_{\Pi_T} \\ &\qquad =(\mathbf f_{\tau},\boldsymbol\varphi)_{\Pi_T}-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\Pi_T} \end{aligned} \end{equation*} \notag $$
for any $\boldsymbol\varphi\in L^2(\Pi_T)$ such that $\nabla\boldsymbol\varphi\in L^2(\Pi_T)$, where $\langle\,\cdot\,{,}\,\cdot\,\rangle_{\mathbb{R}^n}$ is the duality relation on $H^{-1}(\mathbb{R}^n)\times H^1(\mathbb{R}^n)$. It is also true that $\mathbf y(\,\cdot\,,0)=\mathbf y_{0\tau}(\,\cdot\,)$.

For the homogeneous Cauchy problem (that is, the problem with $\mathbf f_\tau=0$, $\mathbf g_\tau=0$, and $\mathbf y_{0\tau}=0$) we choose $\boldsymbol\varphi=\mathbf y$ in $\Pi_t$ and $\boldsymbol\varphi=0$ in $\Pi_T\setminus\Pi_t$, $0<t\leqslant T$ in the last identity. By the formula

$$ \begin{equation*} 2\int_0^t\langle\partial_t\mathbf y(\,\cdot\,,\theta),\mathbf y(\,\cdot\,,\theta)\rangle_{\mathbb{R}^n}\,d\theta =\|\mathbf y(\,\cdot\,,t)\|_{\mathbb{R}^n}^2-\|\mathbf y(\,\cdot\,,0)\|_{\mathbb{R}^n}^2 \quad \text{on } [0,T] \end{equation*} \notag $$
(see [26], Ch. IV, Theorem 1.17 and Remark 1.22), like in part 1 of the proof, we have
$$ \begin{equation*} \frac{1}{2}\|\mathbf y(\,\cdot\,,t)\|_{\mathbb{R}^n}^2\leqslant (c_0+\overline{\tau}\mu)\int_0^t\|\mathbf y(\,\cdot\,,\theta)\|_{\mathbb{R}^n}^2\,d\theta \quad \text{on } [0,T]. \end{equation*} \notag $$
Thus, $\mathbf y=0$, so that the weak solution is unique under the assumptions of the theorem. (In a more general version, independently of the previous part of the proof, we can similarly prove that (3.6) is valid for any weak solution if $\mathbf f_\tau\in L^2(\Pi_T)$.)

It remains to show that $\mathbf y\in C(0,T;L^2(\mathbb{R}^n))$ for $\mathbf f_\tau\in L^{2,1}(\Pi_T)$. With this aim in view (similarly to part d) of the proof of Theorem 1) it suffices to take functions $\mathbf f_{\tau k}\in L^2(\Pi_T)$, $k=1,2,\dots$, such that $\mathbf f_{\tau k}\to\mathbf f_\tau$ in $L^{2,1}(\Pi_T)$ as $k\to\infty$ instead of $\mathbf f_\tau$ and consider the corresponding weak solutions $\mathbf y_k$ of the Cauchy problem. By what we proved above, we have

$$ \begin{equation*} \max\bigl\{\|\mathbf y_k-\mathbf y_l\|_{C(0,T;L^2(\mathbb{R}^n))}, \sqrt{\nu\tau}\, \|\nabla(\mathbf y_k-\mathbf y_l)\|_{L^2(\Pi_T)}\bigr\} \leqslant e^{\overline{c}_0T}2\|\mathbf f_{\tau k}-\mathbf f_{\tau l}\|_{L^{2,1}(\Pi_T)} \end{equation*} \notag $$
for all $k,l=1,2,\dots$ . Therefore, $\mathbf y_k\to\mathbf z$ in $C(0,T;L^2(\mathbb{R}^n))$ and $\nabla\mathbf y_k\to\nabla\mathbf z$ in $L^2(\Pi_T)$ as $k\to\infty$. Passing to the limit in identity (3.5) for $\mathbf y=\mathbf y_k$ and $\mathbf f_\tau=\mathbf f_{\tau k}$ shows that $\mathbf z=\mathbf y$ is a weak solution of the Cauchy problem.

Theorem 2 is proved.

Now we introduce the Cauchy problem for a second-order hyperbolic system of equations that is a perturbation of the original system in (2.1) with parameter $\tau$, $0<\tau\leqslant\overline{\tau}$, in front of the leading derivatives:

$$ \begin{equation} \mathcal{H}_\tau\mathbf y:=\tau\partial_t^2\mathbf y+\mathcal{H}\mathbf y-\tau\partial_i(A_{ij}\partial_j\mathbf y)=\mathbf f_\tau \quad \text{in } \Pi_T, \end{equation} \tag{3.10} $$
$$ \begin{equation} \mathbf y|_{t=0}=\mathbf y_{0\tau}, \quad \partial_t\mathbf y|_{t=0}=\mathbf y_{1\tau} \quad \text{in } \mathbb{R}^n. \end{equation} \tag{3.11} $$
Here $\mathbf y(x,t),\mathbf f_\tau(x,t)$: $\Pi_T\to\mathbb{R}^m$ and $\mathbf y_{0\tau}(x),\mathbf y_{1\tau}(x)$: $\mathbb{R}^n\to\mathbb{R}^m$ are the sought-for and given vector functions.

Assume that conditions (3.4) are satisfied and $\mathbf y_{1\tau}\in L^2(\mathbb{R}^n)$. A weak solution of the Cauchy problem (3.10), (3.11) is a function $\mathbf y\in W_{2,\infty}^{1,1}(\Pi_T)$ satisfying the integral identity

$$ \begin{equation} -\tau(\partial_t\mathbf y,\partial_t\boldsymbol\varphi)_{\Pi_T} +(\mathcal{H}\mathbf y,\boldsymbol\varphi)_{\Pi_T} +\tau(A_{ij}\partial_j\mathbf y,\partial_i\boldsymbol\varphi)_{\Pi_T} =\ell(\mathbf y_{1\tau},\mathbf f_{\tau};\boldsymbol\varphi) \end{equation} \tag{3.12} $$
for any $\boldsymbol\varphi\in W_{2,1}^{1,1}(\Pi_T)$ such that $\boldsymbol\varphi|_{t=T}=0$ and satisfying the initial condition $\mathbf y|_{t=0}=\mathbf y_{0\tau}$ in $C(0,T;L^2(\mathbb{R}^n))$.

Below we need the following condition, which means that the matrix $A$ of leading coefficients of the second-order system dominates the matrix $\mathbf B$ of leading coefficients of the first-order system:

$$ \begin{equation} \|\mathbf B\cdot\nabla\mathbf v\|_{L^2(\mathbb{R}^n)}^2 \leqslant (1-\delta)^2\bigl[(A_{ij}\partial_j\mathbf v,\partial_i\mathbf v)_{\mathbb{R}^n}+\mu_1\|\mathbf v\|_{L^2(\mathbb{R}^n)}^2\bigr] \quad \forall\,\mathbf v\in H^1(\mathbb{R}^n) \end{equation} \tag{3.13} $$
for some $\delta$, $0\leqslant\delta<1$, and $\mu_1\geqslant 0$, where $\mathbf B\cdot\nabla\mathbf v:=B_i\partial_i\mathbf v$. Its left-hand side is $(B_j\partial_j\mathbf v,B_i\partial_i\mathbf v)_{\mathbb{R}^n}=(B_i B_j\partial_j\mathbf v,\partial_i\mathbf v)_{\mathbb{R}^n}$; therefore, it is equivalent to the property
$$ \begin{equation*} \begin{aligned} \, -(1-\delta)^2\mu_1\|\mathbf v\|_{L^2(\mathbb{R}^n)}^2 &\leqslant \bigl([(1-\delta)^2A_{ij}-B_iB_j]\partial_j\mathbf v,\partial_i\mathbf v\bigr)_{\mathbb{R}^n} \\ &=\bigl([(1-\delta)^2A-B]\nabla\mathbf v,\nabla\mathbf v\bigr)_{\mathbb{R}^n} \end{aligned} \end{equation*} \notag $$
for all $\mathbf v\in H^1(\mathbb{R}^n)$, where $B=\mathbf B^T\mathbf B=\{B_iB_j\}_{i,j=1}^n$ is a block matrix of order $mn$ and $\nabla\mathbf v$ is understood as the column vector $(\partial_1\mathbf v,\dots,\partial_n\mathbf v)$ of length $mn$.

Property (3.13) for $\mu_1=0$ is implied by the cruder algebraic inequality

$$ \begin{equation} |\mathbf B(x,t)\cdot\mathbf w|^2 \leqslant (1-\delta)^2(A_{ij}(x,t)\mathbf w_j)\cdot \mathbf w_i \quad \forall\,\mathbf w_1,\dots,\mathbf w_n\in \mathbb{R}^m \quad \text{a.e. in } \Pi_T \end{equation} \tag{3.14} $$
with the same $\delta$ and $\mathbf B\cdot\mathbf w:=B_i\mathbf w_i$, where $\mathbf w=(\mathbf w_1,\dots,\mathbf w_n)$ is a column vector of length $mn$. To be more precise, it is obtained by integration over $\mathbb{R}^n$ for $\mathbf w_i=\partial_i\mathbf v$, $i=1,\dots,n$. The left-hand side of (3.14) equals $(B_j(x,t)\mathbf w_j)\cdot B_i(x,t)\mathbf w_i$; thus, it is equivalent to the nonnegative definiteness of the matrix $(1-\delta)^2A-B$:
$$ \begin{equation*} 0\,{\leqslant}\, \{[(1-\delta)^2A_{ij}-B_iB_j](x,t)\mathbf w_j\}\cdot \mathbf w_i \,{=}\, [(1-\delta)^2A-B]\mathbf w\cdot\mathbf w \quad \forall\,\mathbf w\,{\in}\, \mathbb{R}^{mn} \quad \text{a.e. in } \Pi_T. \end{equation*} \notag $$
For similar conditions, see [5]–[9] and [11], where $0<\delta<1$, and [10], [16], [17] and [27], where $\delta=0$.

We also introduce the condition $A\leqslant c_AI_{mn}$ almost everywhere in $\Pi_T$, that is,

$$ \begin{equation} (A(x,t)\mathbf w)\cdot\mathbf w\leqslant c_A|\mathbf w|^2 \quad \forall\,\mathbf w\in \mathbb{R}^{mn} \quad \text{a.e. in } \Pi_T \end{equation} \tag{3.15} $$
for some $c_A>0$, for example, $c_A=n\|A\|_{L^\infty(\Pi_T)}$, where
$$ \begin{equation*} \|A\|_{L^\infty(\Pi_T)}:=\max_{i,j=1,\dots,n}\|A_{ij}\|_{L^\infty(\Pi_T)}. \end{equation*} \notag $$
We indicate some properties of a weak solution of the Cauchy problem (3.10), (3.11).

Theorem 3. Assume that conditions (3.2), (3.4), (3.13), and (3.15) hold and that $\operatorname{div}\mathbf B$, $\partial_tA_{ij}\in L^\infty(\Pi_T)$, $A_{ij}=A_{ji}^T$, $i,j=1,\dots,n$, $\frac{1}{2}\operatorname{div}\mathbf B-C\leqslant c_0I_m$ almost everywhere in $\Pi_T$, $\mathbf y_{0\tau}\in H^1(\mathbb{R}^n)$, $\mathbf y_{1\tau}\in L^2(\mathbb{R}^n)$, and $8\overline{\tau}^2\mu\leqslant 1$. Then there exists a unique weak solution $\mathbf y=\mathbf y_\tau$ of the Cauchy problem (3.10), (3.11), and the following energy estimate holds:

$$ \begin{equation} \begin{aligned} \, &\nu_0\max\Bigl\{\|\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \tau\|\{\nabla\mathbf y_\tau,\partial_t\mathbf y_\tau\}\|_{L^{2,\infty}(\Pi_T)}, \sqrt{\delta\tau}\|\{\partial_t\mathbf y_\tau,\nabla\mathbf y_\tau\}\|_{L^2(\Pi_T)} \Bigr\} \nonumber \\ &\quad \leqslant e^{\overline{c}_1T}\Bigl(\|\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +\tau\sqrt{2c_A} \|\nabla\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +2\tau\|\mathbf y_{1\tau}\|_{L^2(\mathbb{R}^n)} +9\|\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}\Bigr), \end{aligned} \end{equation} \tag{3.16} $$
where $\nu_0\!=\!\min\{\sqrt{\nu/2},\sqrt{1/6}\}$, $\overline{c}_1\!=\!\max\bigl\{4\bigl[c_0 +\overline{\tau}(\delta\mu+(1-\delta)\mu_1) +\sqrt{3/2}\|C\|_{L^\infty(\Pi_T)}\bigr], (2\nu)^{-1}c_{A1}\bigr\}$, and $c_{A1}\geqslant 0$ is such that $\partial_tA\leqslant c_{A1}I_{mn}$ almost everywhere in $\Pi_T$ (for example, $c_{A1}=n\|\partial_tA\|_{L^\infty(\Pi_T)}$, cf. (3.15)).

Proof. 1. Since $8\overline{\tau}^2\mu\leqslant 1$, we can introduce the following norm of a pair of functions $\mathbf y_0\in H^1(\mathbb{R}^n)$ and $\mathbf y_1\in L^2(\mathbb{R}^n)$:
$$ \begin{equation*} \begin{aligned} \, &\|\{\mathbf y_0,\mathbf y_1\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)}(t) \\ &\qquad =\bigl\{\|\tau\mathbf y_1\|_{L^2(\mathbb{R}^n)}^2 +\|\tau\mathbf y_1+\mathbf y_0\|_{L^2(\mathbb{R}^n)}^2 +2\tau^2\bigl(A_{ij}(\,\cdot\,,t)\partial_j\mathbf y_0(\,\cdot\,),\partial_i\mathbf y_0(\,\cdot\,)\bigr)_{\mathbb{R}^n} \bigr\}^{1/2}, \end{aligned} \end{equation*} \notag $$
with the parameter $0\leqslant t\leqslant T$. By (3.2) and (3.15) this norm satisfies the two-sided estimate
$$ \begin{equation} \begin{aligned} \, & \max\biggl\{\frac12\|\mathbf y_0\|_{L^2(\mathbb{R}^n)},\tau\sqrt{2\nu}\|\nabla\mathbf y_0\|_{L^2(\mathbb{R}^n)}, \sqrt{\frac23}\tau\|\mathbf y_1\|_{L^2(\mathbb{R}^n)}\biggr\} \nonumber \\ &\qquad \leqslant\|\{\mathbf y_0,\mathbf y_1\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)} \leqslant\|\mathbf y_0\|_{L^2(\mathbb{R}^n)}+\tau\sqrt{2c_A} \|\nabla\mathbf y_0\|_{L^2(\mathbb{R}^n)} +2\tau\|\mathbf y_1\|_{L^2(\mathbb{R}^n)}. \end{aligned} \end{equation} \tag{3.17} $$
For $\mu_1=0$ the constant $1/2$ can be improved to $1/\sqrt{2}$, while $\sqrt{2/3}$ can be improved to 1, which makes it possible to improve also the constants in (3.16) (see Theorem 8, part 3, below). First we derive (3.16) as an a priori estimate under the assumption $\partial_t^2\mathbf y,\partial_t\nabla\mathbf y\in L^2(\Pi_T)$. Then the integral identity (3.12) implies that
$$ \begin{equation*} \tau(\partial_t^2\mathbf y,\boldsymbol\varphi)_{\mathbb{R}^n} +(\mathcal{H}\mathbf y,\boldsymbol\varphi)_{\mathbb{R}^n} +\tau(A_{ij}\partial_j\mathbf y,\partial_i\boldsymbol\varphi)_{\mathbb{R}^n} =(\mathbf f_{\tau},\boldsymbol\varphi)_{\mathbb{R}^n} \quad \forall\,\boldsymbol\varphi\in H^1(\mathbb{R}^n) \end{equation*} \notag $$
almost everywhere on $(0,T)$, and $\partial_t\mathbf y|_{t=0}=\mathbf y_{1\tau}$ in $C(0,T;L^2(\mathbb{R}^n))$. By analogy with Proposition 10 in [16] we substitute $\boldsymbol\varphi=2\tau\partial_t\mathbf y+\mathbf y=\tau\partial_t\mathbf y+(\tau\partial_t\mathbf y+\mathbf y)$ into the last identity. The following formulae hold:
$$ \begin{equation*} \begin{gathered} \, (\tau\partial_t^2\mathbf y+\partial_t\mathbf y,\tau\partial_t\mathbf y+\tau\partial_t\mathbf y+\mathbf y)_{\mathbb{R}^n} = \frac{1}{2}\partial_t\bigl[\|\tau\partial_t\mathbf y\|_{\mathbb{R}^n}^2 +\|\tau\partial_t\mathbf y+\mathbf y\|_{\mathbb{R}^n}^2\bigr]+\tau\|\partial_t\mathbf y\|_{\mathbb{R}^n}^2, \\ \begin{aligned} \, (\tau A_{ij}\partial_j\mathbf y,2\tau\partial_t\partial_i\mathbf y+\partial_i\mathbf y)_{\mathbb{R}^n} &=\tau^2\partial_t\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n} \\ &\qquad -\tau^2\bigl((\partial_tA_{ij})\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n} +\tau\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n}, \end{aligned} \end{gathered} \end{equation*} \notag $$
where we use the property that $A_{ij}=A_{ji}^T$. This implies the energy equality
$$ \begin{equation*} \begin{aligned} \, &\frac{1}{2}\partial_t\bigl[\|\tau\partial_t\mathbf y\|_{\mathbb{R}^n}^2 +\|\tau\partial_t\mathbf y+\mathbf y\|_{\mathbb{R}^n}^2 +2\tau^2\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n}\bigr] \\ &\qquad\qquad +\tau\bigl[\|\partial_t\mathbf y\|_{\mathbb{R}^n}^2 +2(B_i\partial_i\mathbf y,\partial_t\mathbf y)_{\mathbb{R}^n} +\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n}\bigr] \\ &\qquad\qquad -\tau^2\bigl((\partial_tA_{ij})\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n} +(B_i\partial_i\mathbf y,\mathbf y)_{\mathbb{R}^n} +(C\mathbf y,2\tau\partial_t\mathbf y+\mathbf y)_{\mathbb{R}^n} \\ &\qquad =(\mathbf f_\tau,2\tau\partial_t\mathbf y+\mathbf y)_{\mathbb{R}^n}. \end{aligned} \end{equation*} \notag $$

According to the dominance condition (3.13), we have

$$ \begin{equation*} \begin{aligned} \, 2|(B_i\partial_i\mathbf y,\partial_t\mathbf y)_{\mathbb{R}^n}| &\leqslant 2\|\mathbf B\cdot\nabla\mathbf y\|_{\mathbb{R}^n}\|\partial_t\mathbf y\|_{\mathbb{R}^n} \\ &\leqslant (1-\delta)\bigl[\|\partial_t\mathbf y\|_{\mathbb{R}^n}^2 +\bigl(A_{ij}\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n}+\mu_1\|\mathbf y\|_{\mathbb{R}^n}^2\bigr]. \end{aligned} \end{equation*} \notag $$
Owing to (3.2) and (2.11), and then (3.17) we obtain
$$ \begin{equation*} \begin{aligned} \, &\frac{1}{2}\partial_t\bigl(\|\{\mathbf y,\partial_t\mathbf y\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)}^2\bigr) +\delta\tau\bigl(\|\partial_t\mathbf y\|_{\mathbb{R}^n}^2 +\nu\|\nabla\mathbf y\|_{\mathbb{R}^n}^2\bigr) \\ &\qquad \leqslant \tau(\delta\mu+(1-\delta)\mu_1)\|\mathbf y\|_{\mathbb{R}^n}^2 +\tau^2\bigl((\partial_tA_{ij})\partial_j\mathbf y,\partial_i\mathbf y\bigr)_{\mathbb{R}^n} \\ &\qquad\qquad +((\frac{1}{2}\operatorname{div}\mathbf B-C)\mathbf y,\mathbf y)_{\mathbb{R}^n} -(C\mathbf y,2\tau\partial_t\mathbf y)_{\mathbb{R}^n} +(\mathbf f_\tau,2\tau\partial_t\mathbf y+\mathbf y)_{\mathbb{R}^n} \\ &\qquad \leqslant c_{A1}\tau^2\|\nabla\mathbf y\|_{\mathbb{R}^n}^2 +[\overline{\tau}(\delta\mu+(1-\delta)\mu_1)+c_0]\|\mathbf y\|_{\mathbb{R}^n}^2 \\ &\qquad\qquad +2\|C\|_{L^\infty(\Pi_T)}\|\mathbf y\|_{\mathbb{R}^n}\|\tau\partial_t\mathbf y\|_{\mathbb{R}^n} +\|\mathbf f_\tau\|_{\mathbb{R}^n}(2+\sqrt{6})\|\{\mathbf y,\partial_t\mathbf y\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)} \\ &\qquad \leqslant \overline{c}_1\|\{\mathbf y,\partial_t\mathbf y\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)}^2 +4.5\|\mathbf f_\tau\|_{\mathbb{R}^n}\|\{\mathbf y,\partial_t\mathbf y\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)} \end{aligned} \end{equation*} \notag $$
with the constant $\overline{c}_1$ indicated in the theorem. This differential inequality yields the estimate
$$ \begin{equation*} \begin{aligned} \, &\max\Bigl\{\max_{0\leqslant t\leqslant T}\|\{\mathbf y,\partial_t\mathbf y\}(\,\cdot\,,t)\|_{\mathcal{E}_\tau(\mathbb{R}^n)}(t), 2\sqrt{\delta\tau}\bigl(\|\partial_t\mathbf y\|_{L^2(\Pi_T)}^2 +\nu\|\nabla\mathbf y\|_{L^2(\Pi_T)}^2\bigr)^{1/2} \Bigr\} \\ &\qquad\qquad \leqslant e^{\overline{c}_1T}\bigl(\|\{\mathbf y_{0\tau},\mathbf y_{1\tau}\}\|_{\mathcal{E}_\tau(\mathbb{R}^n)}(0) +9\|\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}\bigr) \end{aligned} \end{equation*} \notag $$
in the standard way. In view of (3.17) the a priori estimate (3.16) is also true.

2. Like in the previous proof, to verify the existence of a solution we introduce the auxiliary initial-boundary value problem

$$ \begin{equation} \begin{gathered} \, \mathcal{H}_\tau\mathbf y^{(k)}=\mathbf f_\tau \quad \text{in } \Pi_{kT}, \qquad \mathbf y^{(k)}|_{\Gamma_{kT}}=0, \qquad \mathbf y^{(k)}|_{t=0}=\mathbf y_{0\tau k}, \\ \partial_t\mathbf y^{(k)}|_{t=0}=\mathbf y_{1\tau}\quad \text{in } \Omega_k \end{gathered} \end{equation} \tag{3.18} $$
with the parameter $k\geqslant 1$, where $\mathbf y_{0\tau k}\in H^1(\Omega_k)$ and $\mathbf y_{0\tau k}|_{\partial\Omega_k}=0$. Let $\mathbf y_{0\tau k}(x)=\mathbf y_{0\tau}(x)\alpha_k(x_1)\dotsb\alpha_k(x_n)$, where $\alpha_k(\xi)=1$ for $|\xi|\leqslant k-1$, $\alpha_k(\xi)=0$ for $|\xi|\geqslant k$, and $\alpha_k(\xi)$ is linear on $[-k,-k-1]$ and $[k-1,k]$. Then $\mathbf y_{0\tau k}\to\mathbf y_{0\tau}$ in $H^1(\mathbb{R}^n)$ as ${k\to\infty}$ and also $\|\mathbf y_{0\tau k}\|_{L^2(\Omega_k)}\leqslant\|\mathbf y_{0\tau}\|_{\mathbb{R}^n}$ and
$$ \begin{equation*} \|\nabla\mathbf y_{0\tau k}\|_{L^2(\Omega_k)}\leqslant\|\nabla\mathbf y_{0\tau}\|_{\mathbb{R}^n} +\|\mathbf y_{0\tau}\|_{L^2(\Omega_k\setminus\Omega_{k-1})}=\|\nabla\mathbf y_{0\tau}\|_{\mathbb{R}^n}+o(1). \end{equation*} \notag $$
By definition, a weak solution $\mathbf y^{(k)}\in W_{2,\infty}^{1,1}(\Pi_{kT})$ of (3.18) satisfies the integral identity
$$ \begin{equation} \begin{aligned} \, &-\tau(\partial_t\mathbf y^{(k)},\partial_t\boldsymbol\varphi)_{\Pi_{kT}} +(\mathcal{H}\mathbf y^{(k)},\boldsymbol\varphi)_{\Pi_{kT}} +\tau(A_{ij}\partial_j\mathbf y^{(k)},\partial_i\boldsymbol\varphi)_{\Pi_{kT}} \nonumber \\ &\qquad =(\mathbf y_{1\tau},\boldsymbol\varphi|_{t=0})_{\Omega_k} +(\mathbf f_{\tau},\boldsymbol\varphi)_{\Pi_{kT}} \end{aligned} \end{equation} \tag{3.19} $$
for any $\boldsymbol\varphi\in W_{2,1}^{1,1}(\Pi_{kT})$ such that $\boldsymbol\varphi|_{\Gamma_{kT}}=0$, $\boldsymbol\varphi|_{t=T}=0$, and also satisfies the boundary and initial conditions $\mathbf y^{(k)}|_{\Gamma_{kT}}=0$ and $\mathbf y^{(k)}|_{t=0}\,{=}\,\mathbf y_{0\tau k}$ in $C(0,T;L^2(\Omega_k))$. Here the spaces $W_{2,q}^{1,1}(\Pi_{kT})$ are introduced similarly to $W_{2,q}^{1,1}(\Pi_T)$.

The existence of the weak solution $\mathbf y^{(k)}$ is deduced in a standard way using the Galerkin method (see [24], Ch. IV, Theorem 3.2; the $*$-weak convergence of the Galerkin approximations and their first derivatives in $L^{2,\infty}(\Pi_{kT})$ must be used instead of their weak convergence in $L^2(\Pi_{kT})$). First for the Galerkin approximations and then for $\mathbf y^{(k)}$ itself, after taking the limit, we prove quite similarly to (3.16) that

$$ \begin{equation} \begin{aligned} \, &\nu_0\max\bigl\{\|\mathbf y^{(k)}\|_{C(0,T;L^2(\Omega_k))}, \tau\|\{\nabla\mathbf y^{(k)},\partial_t\mathbf y^{(k)}\}\|_{L^{2,\infty}(\Pi_{kT})}, \nonumber \\ &\qquad\qquad \sqrt{\delta\tau}\, \|\{\partial_t\mathbf y^{(k)},\nabla\mathbf y^{(k)}\}\|_{L^2(\Pi_{kT})}\bigr\} \nonumber \\ &\quad\leqslant e^{\overline{c}_1T}\bigl(\|\mathbf y_{0\tau k}\|_{L^2(\Omega_k)} +\tau\sqrt{2c_A}\, \|\nabla\mathbf y_{0\tau k}\|_{L^2(\Omega_k)} +2\tau\|\mathbf y_{1\tau}\|_{L^2(\Omega_k)} \,{+}\,9\|\mathbf f_\tau\|_{L^{2,1}(\Pi_{kT})}\bigr) \nonumber \\ &\quad\leqslant e^{\overline{c}_1T}\bigl(\|\mathbf y_{0\tau}\|_{\mathbb{R}^n}+o(1) +\tau\sqrt{2c_A}\, \|\nabla\mathbf y_{0\tau}\|_{\mathbb{R}^n} +2\tau\|\mathbf y_{1\tau}\|_{\mathbb{R}^n} +9\|\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}\bigr). \end{aligned} \end{equation} \tag{3.20} $$

We set additionally $\mathbf y^{(k)}=0$ on $\Pi_T\setminus\Pi_{kT}$. Then $\mathbf y^{(k)}\in W_{2,\infty}^{1,1}(\Pi_{kT})$, and we can replace $\Omega_k$ by $\mathbb{R}^n$ and $\Pi_{kT}$ by $\Pi_T$ on the left-hand side of (3.20). By this estimate there is a function $\mathbf y\in W_{2,\infty}^{1,1}(\Pi_T)$ such that $\mathbf y^{(k)}\to\mathbf y$, $\nabla\mathbf y^{(k)}\to \nabla\mathbf y$, and $\partial_t\mathbf y^{(k)}\to\partial_t\mathbf y$ $*$-weakly in $L^{2,\infty}(\Pi_T)$ for some subsequence $k=k_l\to\infty$. Thus, estimate (3.16) is true for $\mathbf y=\mathbf y_\tau$.

We rewrite the integral identity (3.19) in the form of (3.12) with $\mathbf y^{(k)}$ in the role of $\mathbf y$:

$$ \begin{equation*} \begin{aligned} \, &-\tau(\partial_t\mathbf y^{(k)},\partial_t\boldsymbol\varphi)_{\Pi_T} +(\partial_t\mathbf y^{(k)}+B_i\partial_i\mathbf y^{(k)}+C\mathbf y^{(k)},\boldsymbol\varphi)_{\Pi_T} +\tau(A_{ij}\partial_j\mathbf y^{(k)},\partial_i\boldsymbol\varphi)_{\Pi_T} \\ &\qquad =\ell(\mathbf y_{1\tau},\mathbf f_{\tau};\boldsymbol\varphi), \end{aligned} \end{equation*} \notag $$
where $\boldsymbol\varphi$ is the same as in (3.19) and $\boldsymbol\varphi=0$ on $\Pi_T\setminus\Pi_{kT}$. Taking the limit as $k=k_l\to\infty$ on its left-hand side leads to the integral identity (3.12), where, in addition, $\boldsymbol\varphi$ has a compact support with respect to $x$, which does not restrict the generality. Passing to the limit as $k=k_l\to\infty$ in the integral identity
$$ \begin{equation*} -(\partial_t\mathbf y^{(k)},\boldsymbol\psi)_{\Pi_T}=(\mathbf y^{(k)},\partial_t\boldsymbol\psi)_{\Pi_T}+(\mathbf y_{0\tau k},\boldsymbol\psi_0)_{\mathbb{R}^n}, \end{equation*} \notag $$
where $\boldsymbol\psi,\partial_t\boldsymbol\psi\in L^{2,1}(\Pi_T)$ and $\boldsymbol\psi|_{t=T}=0$, yields additionally $-(\partial_t\mathbf y,\boldsymbol\psi)_{\Pi_T}=(\mathbf y,\partial_t\boldsymbol\psi)_{\Pi_T}+(\mathbf y_{0\tau},\boldsymbol\psi_0)_{\mathbb{R}^n},$ which implies that $\mathbf y|_{t=0}=\mathbf y_{0\tau}$. As a result, the function $\mathbf y$ constructed is a weak solution of the Cauchy problem (3.10), (3.11). In fact, the limiting procedure could be avoided completely if we have instead verified, among other things, that for a weak solution of this Cauchy problem the velocity of propagation of perturbations is finite (see [24], Ch. IV, § 2). However, this would not reduce the length of the proof.

3. The uniqueness of a weak solution is deduced using the known method (see [24], Ch. IV, Theorem 3.1) that consists in taking $\displaystyle\boldsymbol\varphi(x,t)=\int_t^\theta\mathbf y(x,t')\,dx\,dt'$ in $\Pi_\theta$, $\boldsymbol\varphi(x,t)=0$ in $\Pi_T\setminus\Pi_\theta$, in the integral identity (3.12), where $0<\theta\leqslant T$ is a parameter. This makes it possible to prove that $\mathbf y=0$ for $\mathbf y_{0\tau}=\mathbf y_{1\tau}=0$ and $\mathbf f_\tau=0$. As the arguments are standard, we omit the details. Theorem 3 is proved.

§ 4. Estimates for the difference of solutions of the Cauchy problems for the first-order hyperbolic system of equations and its perturbations with a small parameter $\tau$

First we obtain estimates for the difference of solutions of the Cauchy problems for a first-order hyperbolic system of equations and for its parabolic perturbation.

Theorem 4. 1. Assume that the assumptions of Theorem 1, parts 1 and 2 (for $q= 1$), and Theorem 2 hold and that $\|A\|_{L^\infty(\Pi_T)}\leqslant N$. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of solutions of the Cauchy problems (2.1) and (3.1) satisfies the estimate

$$ \begin{equation} \begin{aligned} \, &\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} +\sqrt{\tau}\, \|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)} \leqslant C(N,T)\bigl[\|\mathbf w_0-\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} \nonumber \\ &\qquad +\|\mathbf f-\mathbf f_\tau\|_{L^{2,1}(\Pi_T)} +\sqrt{\tau}\, \bigl(\|\mathbf g_\tau\|_{L^2(\Pi_T)}+\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{1,0}(\Pi_T)}\bigr)\bigr]. \end{aligned} \end{equation} \tag{4.1} $$
In this and the subsequent estimates $C(N,T)$ is independent of $\tau$.

In particular, for $\mathbf f_\tau=\mathbf f$, $\mathbf g_\tau=0$ and $\mathbf y_{0\tau}=\mathbf w_0$ the following simplified estimate is true:

$$ \begin{equation} \begin{aligned} \, &\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\sqrt{\tau}\, \|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)} \nonumber \\ &\qquad \leqslant C(N,T)\sqrt{\tau}\, \bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{1,0}(\Pi_T)}\bigr). \end{aligned} \end{equation} \tag{4.2} $$

2. Let the assumptions of Theorem 1, part 3, a), also hold, let $\|\operatorname{div} A_{\cdot j}\|_{L^\infty(\Pi_T)}\leqslant N$, $j=1,\dots,n$, where $\operatorname{div} A_{\cdot j}:=\partial_iA_{ij}$ and, again, let $\mathbf f_\tau=\mathbf f$, $\mathbf g_\tau=0$ and $\mathbf y_{0\tau}=\mathbf w_0$. Then $\mathbf r_\tau$ also satisfies the estimate

$$ \begin{equation} \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\sqrt{\tau}\, \|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)} \leqslant C(N,T)\tau\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr). \end{equation} \tag{4.3} $$

Proof. 1. According to (2.1) and (3.1) (more precisely, see the integral identity (3.5)), the function $\mathbf r_\tau$ is the weak solution of the Cauchy problem
$$ \begin{equation*} \begin{gathered} \, \mathcal{P}_\tau\mathbf r_\tau=\mathbf f-\mathbf f_\tau-\tau\partial_i\mathbf g_{i\tau}+(\mathcal{P}_\tau-\mathcal{H})\mathbf w =\mathbf f-\mathbf f_\tau-\tau\partial_i(\mathbf g_{i\tau}+A_{ij}\partial_j\mathbf w) \quad \text{in } \Pi_T, \\ \mathbf r_\tau|_{t=0}=\mathbf w_0-\mathbf y_{0\tau} \quad \text{in } \mathbb{R}^n. \end{gathered} \end{equation*} \notag $$
Therefore, by virtue of estimate (3.6) as applied to the solution of the last problem, we have
$$ \begin{equation} \begin{aligned} \, & \max\{\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))},\sqrt{\nu\tau}\,\|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)}\} \nonumber \\ &\quad \leqslant e^{\overline{c}_0T}\bigl[\|\mathbf w_0-\mathbf y_{0\tau}\|_{\mathbb{R}^n} +2\|\mathbf f-\mathbf f_\tau\|_{L^{2,1}(\Pi_T)} \nonumber \\ &\qquad\quad +\sqrt{\nu^{-1}\tau}\, \|\mathbf g_\tau\|_{L^2(\Pi_T)} +\sqrt{\nu^{-1}\tau}\, \delta^{(ii)}\sqrt{T}\, \|A_{ij}\|_{L^\infty(\Pi_T)}\|\partial_j\mathbf w\|_{L^{2,\infty}(\Pi_T)}\bigr]. \end{aligned} \end{equation} \tag{4.4} $$
Using estimate (2.6) for $\mathbf w$ we obtain (4.1).

2. Under the assumptions of part 2 of the theorem the last term on the right-hand side of (4.4) can be replaced by $2\tau\|\partial_i(A_{ij}\partial_j\mathbf w)\|_{L^{2,1}(\Pi_T)}$ (where the sum over $i,j=1,\dots,n$ is assumed under the norm sign). Since, using estimates (2.6) and (2.8) for $\mathbf w$, we have

$$ \begin{equation} \begin{aligned} \, \|\partial_i(A_{ij}\partial_j\mathbf w)\|_{L^{2,1}(\Pi_T)} &\leqslant\|A_{ij}\|_{L^\infty(\Pi_T)}T\|\partial_i\partial_j\mathbf w\|_{L^{2,\infty}(\Pi_T)} \nonumber \\ &\qquad +\|\operatorname{div} A_{\cdot j}\|_{L^\infty(\Pi_T)}T\|\partial_j\mathbf w\|_{L^{2,\infty}(\Pi_T)} \nonumber \\ & \leqslant C(N,T)\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr), \end{aligned} \end{equation} \tag{4.5} $$
the above modification of (4.4) yields (4.3). The theorem is proved.

Note that (4.3) implies not only the estimate $\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}=O(\tau)$ but also the estimate $\|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)}=O(\sqrt{\tau})$.

We deduce additional estimates of fractional order for $\mathbf r_\tau$ in the case when $\mathbf w_0$ and $\mathbf f$ are less smooth. We introduce the following Banach spaces constructed using the $K_{\alpha,\infty}$-method of real interpolation of Banach spaces, $0<\alpha<1$ (see, for example, [28], Ch. 3):

$$ \begin{equation*} \begin{gathered} \, \mathcal{H}^\alpha:=\bigl(L^2(\mathbb{R}^n),H^1(\mathbb{R}^n)\bigr)_{\alpha,\infty}, \quad \mathcal{H}^1=H^1(\mathbb{R}^n), \\ \mathcal{H}^{1+\alpha}:=\bigl(H^1(\mathbb{R}^n),H^2(\mathbb{R}^n)\bigr)_{\alpha,\infty}, \\ \mathcal{W}_{2,1}^{\alpha,0}:=\bigl(L^{2,1}(\Pi_T),W_{2,1}^{1,0}(\Pi_T)\bigr)_{\alpha,\infty}, \quad \mathcal{W}_{2,1}^{1+\alpha,0}:=\bigl(W_{2,1}^{1,0}(\Pi_T),W_{2,1}^{2,0}(\Pi_T)\bigr)_{\alpha,\infty}. \end{gathered} \end{equation*} \notag $$

Theorem 5. For simplicity let $\mathbf f_\tau=\mathbf f$, $\mathbf g_\tau=0$ and $\mathbf y_{0\tau}=\mathbf w_0$.

1. Let the assumptions on matrix coefficients in Theorem 4, part 1, hold. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of the solutions of the Cauchy problems (2.1) and (3.1) satisfies the estimate

$$ \begin{equation} \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant C(N,T)\tau^{\alpha/2} \bigl(\|\mathbf w_0\|_{\mathcal{H}^{\alpha}}+\|\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,0}}\bigr), \qquad 0<\alpha<1. \end{equation} \tag{4.6} $$

2. Let all assumptions on matrix coefficients of Theorem 4 hold. Then for ${1<\alpha<2}$ $\mathbf r_\tau$ also satisfies the estimate

$$ \begin{equation} \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\sqrt{\tau}\|\nabla\mathbf r_\tau\|_{L^2(\Pi_T)} \leqslant C(N,T)\tau^{\alpha/2} \bigl(\|\mathbf w_0\|_{\mathcal{H}^{\alpha}}+\|\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,0}}\bigr). \end{equation} \tag{4.7} $$

In these estimates and similar estimates below it is automatically assumed that $\mathbf w_0$ and $\mathbf f$ belong to the spaces whose norms are applied.

Proof. Estimates (2.5) for $\mathbf w$ and (3.6) for $\mathbf y_\tau$ yield
$$ \begin{equation} \begin{aligned} \, \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} & \leqslant \|\mathbf w\|_{C(0,T;L^2(\mathbb{R}^n))}+\|\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \nonumber \\ & \leqslant C(N,T)\bigl(\|\mathbf w_0\|_{\mathbb{R}^n}+\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{aligned} \end{equation} \tag{4.8} $$
Applying the interpolation theorem for linear operators (see [28], Theorem 3.1.2) to estimates (4.8) and (4.2) (without the second term on the left) for $\mathbf r_\tau$ or, more precisely, to the corresponding operator
$$ \begin{equation} \mathbf r_\tau=R(\mathbf w_0,\mathbf f)\colon X_p\to C(0,T;L^2(\mathbb{R}^n)),\qquad p=0,1, \end{equation} \tag{4.9} $$
where $X_0=L^2(\mathbb{R}^n)\times L^{2,1}(\Pi_T)$, $X_1=H^1(\mathbb{R}^n)\times W_{2,1}^{1,0}(\Pi_T)$, leads to (4.6). Similarly, applying this theorem to (4.2) and (4.3) leads to (4.7). The theorem is proved.

We now pass to estimating the difference of solutions of the Cauchy problems for the first-order hyperbolic system of equations and its second-order hyperbolic perturbation.

Theorem 6. Let all assumptions of Theorem 1 for $q=1$ and of Theorem 3 hold, and let $\|A\|_{L^\infty(\Pi_T)}\leqslant N$ and $\|\operatorname{div} A_{\cdot j}\|_{L^\infty(\Pi_T)}\leqslant N$, $j=1,\dots,n$. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of solutions of the Cauchy problems (2.1) and (3.10), (3.11) satisfies the estimate

$$ \begin{equation} \begin{aligned} \, & \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\tau\|\{\nabla\mathbf r_\tau,\partial_t\mathbf r_\tau\}\|_{L^{2,\infty}(\Pi_T)} +\sqrt{\delta\tau}\, \|\{\nabla\mathbf r_\tau,\partial_t\mathbf r_\tau\}\|_{L^2(\Pi_T)} \nonumber \\ &\qquad\leqslant C(N,T)\bigl[\|\mathbf w_0-\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +\|\mathbf f-\mathbf f_\tau\|_{L^{2,1}(\Pi_T)} \nonumber \\ &\qquad\qquad +\tau\bigl(\|\nabla\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +\|\mathbf y_{1\tau}\|_{L^2(\mathbb{R}^n)} +\|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr)\bigr]. \end{aligned} \end{equation} \tag{4.10} $$

In particular, if $\mathbf f_\tau=\mathbf f$, $\mathbf y_{0\tau}=\mathbf w_0$ and $\|\mathbf y_{1\tau}\|_{L^2(\mathbb{R}^n)} \leqslant C_1(N,T)\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr)$ (for example, if $\mathbf y_{1\tau}=0$ or $\mathbf y_{1\tau}=(\partial_t\mathbf w)_0\equiv(\partial_t\mathbf w)|_{t=0}$), then the following simplified estimate is true:

$$ \begin{equation} \begin{aligned} \, & \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\tau\|\{\nabla\mathbf r_\tau,\partial_t\mathbf r_\tau\}\|_{L^{2,\infty}(\Pi_T)} +\sqrt{\delta\tau}\, \|\{\nabla\mathbf r_\tau,\partial_t\mathbf r_\tau\}\|_{L^2(\Pi_T)} \nonumber \\ &\qquad \leqslant C_2(N,T)\tau\bigl( \|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr). \end{aligned} \end{equation} \tag{4.11} $$

Proof. Under the assumptions in the theorem, according to (2.1) and (3.10), (3.11) (see the integral identity (3.12)), the function $\mathbf r_\tau$ is the weak solution of the Cauchy problem
$$ \begin{equation*} \begin{gathered} \, \mathcal{H}_\tau\mathbf r_\tau=\mathbf f-\mathbf f_\tau+(\mathcal{H}_\tau-\mathcal{H})\mathbf w =\mathbf f-\mathbf f_\tau+\tau[\partial_t^2\mathbf w-\partial_i(A_{ij}\partial_j\mathbf w)] \quad \text{in } \Pi_T, \\ \mathbf r_\tau|_{t=0}=\mathbf w_0-\mathbf y_{0\tau}, \quad \partial_t\mathbf r_\tau|_{t=0}=(\partial_t\mathbf w)_0-\mathbf y_{1\tau} \quad \text{in } \mathbb{R}^n. \end{gathered} \end{equation*} \notag $$
Estimate (3.16), as applied to the solution of this problem, yields
$$ \begin{equation} \begin{aligned} \, &\nu_0 \max\bigl\{\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \tau\|\{\nabla\mathbf r_\tau,\partial_t\mathbf r_\tau\}\|_{L^{2,\infty}(\Pi_T)}, \sqrt{\delta\tau}\, \|\{\partial_t\mathbf r_\tau,\nabla\mathbf r_\tau\}\|_{L^2(\Pi_T)}\bigr\} \nonumber \\ &\qquad\leqslant e^{\overline{c}_1T}\bigl\{\|\mathbf w_0-\mathbf y_{0\tau}\|_{\mathbb{R}^n} +\tau\sqrt{2c_A}\, \|\nabla(\mathbf w_0-\mathbf y_{0\tau})\|_{\mathbb{R}^n} +2\tau\|(\partial_t\mathbf w)_0-\mathbf y_{1\tau}\|_{\mathbb{R}^n} \nonumber \\ &\qquad\qquad +9\bigl[\|\mathbf f-\mathbf f_\tau\|_{L^{2,1}(\Pi_T)} +\tau\|\partial_t^2\mathbf w-\partial_i(A_{ij}\partial_j\mathbf w)\|_{L^{2,1}(\Pi_T)}\bigr]\bigr\}. \end{aligned} \end{equation} \tag{4.12} $$
It is obvious that
$$ \begin{equation*} \begin{aligned} \, & \tau\|\nabla(\mathbf w_0-\mathbf y_{0\tau})\|_{\mathbb{R}^n} +2\tau\|(\partial_t\mathbf w)_0-\mathbf y_{1\tau}\|_{\mathbb{R}^n} \\ &\qquad\leqslant \tau\bigl(\|\mathbf y_{0\tau}\|_{\mathbb{R}^n} +2\|\mathbf y_{1\tau}\|_{\mathbb{R}^n} +\|\nabla\mathbf w_0\|_{\mathbb{R}^n}+2\|(\partial_t\mathbf w)_0\|_{\mathbb{R}^n} \bigr). \end{aligned} \end{equation*} \notag $$
Using estimates (2.9) for $q=1$ and (4.5) for $\mathbf w$ we also obtain
$$ \begin{equation*} \|\partial_t^2\mathbf w-\partial_i(A_{ij}\partial_j\mathbf w)\|_{L^{2,1}(\Pi_T)} \leqslant C(N,T)\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr). \end{equation*} \notag $$
Substituting the last two estimates into (4.12) and using the fact that
$$ \begin{equation} \begin{aligned} \, \|(\partial_t\mathbf w)_0\|_{\mathbb{R}^n} &=\|-(B_i)_0\partial_i\mathbf w_0-C_0\mathbf w_0+\mathbf f_0\|_{\mathbb{R}^n} \nonumber \\ &\leqslant C(N,T)\bigl(\|\mathbf w_0\|_{H^1(\mathbb{R}^n)}+\|\{\mathbf f,\partial_t\mathbf f\}\|_{L^{2,1}(\Pi_T)}\bigr), \end{aligned} \end{equation} \tag{4.13} $$
we derive the required estimate (4.10). Estimate (4.13) also justifies the estimate for $\mathbf y_{1\tau}$ in the assumptions of the theorem in the case when $\mathbf y_{1\tau}=(\partial_t\mathbf w)_0$. The theorem is proved.

Estimate (4.10) for $\mathbf y_{0\tau}\neq\mathbf w_0$ is used below in Theorem 7. In addition, the case when $\mathbf f_\tau\neq\mathbf f$ has some applications (see [29]).

Estimate (4.11) implies not only the estimate $\|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}=O(\tau)$ but also the estimate $\sqrt{\delta}\, \|\{\partial_t\mathbf r_\tau,\nabla\mathbf r_\tau\}\|_{L^2(\Pi_T)}=O(\sqrt{\tau})$.

The term containing the factor $\tau$ on the left-hand sides of the above estimates is also useful. For example, when $n=1$, by the inequality $\|v\|_{L^\infty(\mathbb{R})}\leqslant c\|v\|_{L^2(\mathbb{R})}^{1/2} \|v\|_{H^1(\mathbb{R})}^{1/2}$ and the property that $\mathbf w,\mathbf y_\tau\in C(\overline{\Pi}_T)$ estimate (4.11) yields the following estimate in the uniform norm:

$$ \begin{equation*} \|\mathbf r_\tau\|_{C_b(\overline{\Pi}_T)}:=\sup_{\overline{\Pi}_T}|\mathbf r_\tau(x,t)| \leqslant C(N,T)\bigl(1+\sqrt{\overline{\tau}}\bigr)\sqrt{\tau}\, \bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr). \end{equation*} \notag $$

We derive additional estimates of fractional order in the case when $\mathbf w_0$ and $\mathbf f$ are less smooth. We introduce the simplest Steklov averaging with step size $\tau>0$ with respect to $x_1,\dots,x_n$ as follows:

$$ \begin{equation*} (\overline{\sigma}^{(\tau)}v)(x):=\frac{1}{\tau^n}\int_{(-\tau/2,\tau/2)^n} v(x+\xi)\,d\xi \quad \forall\, v\in L_{\mathrm{loc}}^1(\mathbb{R}^n) \end{equation*} \notag $$
(other averagings, including $\sigma^{(\tau)}$, can also be used) and also introduce the spaces $\mathcal{W}_{2,1}^{2\alpha,\alpha}:=\bigl(L^{2,1}(\Pi_T),W_{2,1}^{2,1}(\Pi_T)\bigr)_{\alpha,\infty}$ constructed using the $K_{\alpha,\infty}$-method of real interpolation, $0<\alpha\leqslant 1$.

Theorem 7. Let all assumptions on matrix coefficients of Theorem 6 hold, let $\mathbf y_{0\tau}=\overline{\sigma}^{(\tau)}\mathbf w_0$, and let $\mathbf f_\tau=\mathbf f$ and $\mathbf y_{1\tau}=0$ for simplicity. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of solutions of the Cauchy problems (2.1) and (3.10), (3.11) satisfies the estimate

$$ \begin{equation} \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant C(N,T)\tau^{\alpha/2} \bigl(\|\mathbf w_0\|_{\mathcal{H}^\alpha}+\|\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,\alpha/2}}\bigr), \qquad 0<\alpha\leqslant 2. \end{equation} \tag{4.14} $$

Proof. It is known that the averagings obey the estimates
$$ \begin{equation*} \|\overline{\sigma}^{(\tau)} v\|_{\mathbb{R}^n}+\tau\|\nabla \overline{\sigma}^{(\tau)} v\|_{\mathbb{R}^n} \leqslant c\|v\|_{\mathbb{R}^n} \end{equation*} \notag $$
and
$$ \begin{equation} \|v-\overline{\sigma}^{(\tau)} v\|_{\mathbb{R}^n}\leqslant\tau c\|\nabla v\|_{\mathbb{R}^n}, \quad \|\nabla \overline{\sigma}^{(\tau)} v\|_{\mathbb{R}^n}\leqslant c\|\nabla v\|_{\mathbb{R}^n} \quad \forall\, v\in H^1(\mathbb{R}^n). \end{equation} \tag{4.15} $$
Using the first of them, by virtue of Theorem 3, for $\mathbf y_{1\tau}=0$ and $\mathbf f_\tau=\mathbf f$ we obtain
$$ \begin{equation*} \|\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant C(N,T)\bigl(\|\mathbf w_0\|_{\mathbb{R}^n}+\|\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation*} \notag $$
Therefore, (4.8) remains valid here. Relying on the properties of the averagings (4.15), we derive from (4.10) under the assumptions of this theorem that
$$ \begin{equation*} \|\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant C(N,T)\tau\bigl(\|\mathbf w_0\|_{H^2(\mathbb{R}^n)}+\|\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr). \end{equation*} \notag $$
Applying the interpolation theorem for linear operators to (4.8) and the last estimate of $\mathbf r_\tau$, in view of the embedding of $H^1(\mathbb{R}^n)$ in $\bigl(L^2(\mathbb{R}^n),H^2(\mathbb{R}^n)\bigr)_{1/2,\infty}$ (see [28], § 6.2), leads to the required estimate (4.14) for $\mathbf r_\tau$. The theorem is proved.

In particular, according to (4.14), for $\mathbf w_0\in H^1(\mathbb{R}^n)$ and $\mathbf f\in \mathcal{W}_{2,1}^{1,1/2}$ we have the estimate $O(\sqrt{\tau})$; cf. (4.1). It is impossible to deduce it in the framework of the above proof of Theorem 6.

The estimates in Theorems 6 and 7 are similar to those obtained in [10] under other conditions and by another method.

For $0<\alpha<2$, $\alpha\neq 1$, the space $\mathcal{H}^{\alpha}$ coincides with the Nikolskii space $H_2^{\alpha}(\mathbb{R}^n)$ (up to equivalence of norms); see [28], § 6.2. Approximating functions in the Nikolskii spaces by their averagings, we can verify that

$$ \begin{equation*} H_{2,1}^{\alpha,0}(\Pi_T)\subset \mathcal{W}_{2,1}^{\alpha,0}\quad\text{and} \quad H_{2,1}^{\alpha,\alpha/2}(\Pi_T)\subset \mathcal{W}_{2,1}^{\alpha,\alpha/2}, \qquad 0<\alpha<2, \quad \alpha\neq 1; \end{equation*} \notag $$
in a similar way, $WH_{2,1}^{1,1/2}(\Pi_T):=\{v\in H_{2,1}^{0,1/2}(\Pi_T),\nabla v\in L^{2,1}(\Pi_T)\}\subset \mathcal{W}_{2,1}^{1,1/2}$ (the case $\alpha=1$); cf. [15], § 6.10, and [30], Ch. 1, § 2.

In addition, the space $H_2^{1/2}(\mathbb{R}^n)$ contains $\mathrm{BV}(\mathbb{R}^n)\cap L^\infty(\mathbb{R}^n)$, where $\mathrm{BV}(\mathbb{R}^n)$ is the space of functions of bounded variation on $\mathbb{R}^n$ (see, for example, [31], Chs. 37 and 38). In a similar way, the space $H_{2,1}^{1/2,1/4}(\Pi_T)$ obviously contains the space $\mathrm{BV}(\Pi_T)\cap L^\infty(\Pi_T)$, which follows from the elementary estimate

$$ \begin{equation*} |\xi|^{-1/2}\|\Delta_\xi f\|_{L^{2,1}(\Pi_T)} \leqslant \bigl(T|\xi|^{-1}\|\Delta_\xi f\|_{L^1(\Pi_T)}\bigr)^{1/2}(2\|f\|_{L^\infty(\Pi_T)})^{1/2}, \qquad \xi\neq 0, \end{equation*} \notag $$
and a similar estimate for the difference $f(\,\cdot\,,t+\theta)-f(\,\cdot\,,t)$. This makes it possible to cover a wide class of discontinuous functions $\mathbf w_0$ and $\mathbf f$ which is important for applications, and this ensures the estimate $O(\tau^{1/4})$ for them in (4.6) and (4.14) for $\alpha=1/2$. To justify the last result it has been essential to use just the $K_{\alpha,\theta}$-method of real interpolation for $\theta=\infty$ and the spaces $H_2^{\alpha}(\mathbb{R}^n)$ as the widest spaces in the family of Besov spaces $B_{2,\theta}^{\alpha}(\mathbb{R}^n)$, which correspond to $\theta=\infty$.

§ 5. Estimates for derivatives of any order with respect to $x$ of solutions of the systems under consideration and their differences

In this section we limit ourselves to the case when the matrices $\mathbf B,C,A\in L^\infty(0,T)$ are independent of $x$ and $-C\leqslant c_0I_m$ almost everywhere on $(0,T)$. This case is essential in the next section and makes it possible to clarify and significantly simplify the statements of the results and their derivation from the previous theorems, although this restriction is not fundamental. We introduce the Sobolev derivative with respect to $x$ of any order:

$$ \begin{equation*} \partial^{\mathbf k}=\partial_1^{k_1}\dotsb \partial_n^{k_n},\qquad \mathbf k=(k_1,\dots,k_n),\qquad |\mathbf k|_1=k_1+\dots+k_n. \end{equation*} \notag $$
Below we assume that $\mathbf k$ and $q$ satisfying $|\mathbf k|_1\geqslant 1$ and $1\leqslant q\leqslant\infty$ are arbitrary.

Theorem 8. 1. a) Let $\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_0\in L^2(\mathbb{R}^n)$. Then the weak solution $\mathbf w$ of the Cauchy problem (2.1) satisfies the estimate

$$ \begin{equation} \|\partial^{\mathbf k}\mathbf w\|_{C(0,T;L^2(\mathbb{R}^n))} \leqslant e^{c_0T}\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{L^2(\mathbb{R}^n)}+2\|\partial^{\mathbf k}\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{5.1} $$
More specifically, estimate (5.1) means that if $\partial^{\mathbf k}\mathbf w_0\in L^2(\mathbb{R}^n)$ and $\partial^{\mathbf k}\mathbf f\in L^{2,1}(\Pi_T)$, then $\partial^{\mathbf k}\mathbf w\in C(0,T;L^2(\mathbb{R}^n))$ and the above estimate holds. For more compact statements all estimates in this theorem and Theorems 9 and 10 below are understood similarly.

b) Let $p=1,2$. If $\mathbf f,\nabla^p\mathbf f\in L^{2,1}(\Pi_T)$ and $\mathbf w_0\in H^p(\mathbb{R}^n)$, then the strong solution $\mathbf w$ of the Cauchy problem (2.1) satisfies the estimates

$$ \begin{equation} \|\partial^{\mathbf k}\nabla^p\mathbf w\|_{L^{2,\infty}(\Pi_T)} \leqslant ce^{c_0T}\bigl(\|\partial^{\mathbf k}\nabla^p\mathbf w_0\|_{L^2(\mathbb{R}^n)} +\|\partial^{\mathbf k}\nabla^p\mathbf f\|_{L^{2,1}(\Pi_T)}\bigr). \end{equation} \tag{5.2} $$
If also $\|\{\mathbf B,C\}\|_{L^\infty(0,T)}\leqslant N$, then for $p=1,2$ we have the estimates
$$ \begin{equation} \begin{aligned} \, &\|\partial_t\partial^{\mathbf k}\mathbf w\|_{L^{2,q}(\Pi_T)} \nonumber \\ &\qquad\leqslant cNe^{c_0T}T^{1/q}\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^1(\mathbb{R}^n)} +\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{1,0}(\Pi_T)}\bigr) +\|\partial^{\mathbf k}\mathbf f\|_{L^{2,q}(\Pi_T)} \end{aligned} \end{equation} \tag{5.3} $$
and
$$ \begin{equation} \begin{aligned} \, &\|\partial_t\nabla\partial^{\mathbf k}\mathbf w\|_{L^{2,q}(\Pi_T)} \nonumber \\ &\qquad\leqslant cNe^{c_0T}T^{1/q}\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr)+\|\partial^{\mathbf k}\nabla\mathbf f\|_{L^{2,q}(\Pi_T)} \end{aligned} \end{equation} \tag{5.4} $$
respectively.

If $p=2$ and, in addition, $\|\{\partial_t\mathbf B,\partial_tC\}\|_{L^\infty(0,T)}\leqslant N$ and $\partial_t\mathbf f\in L^{2,1}(\Pi_T)$, then

$$ \begin{equation} \begin{aligned} \, \|\partial_t^2\partial^{\mathbf k}\mathbf w\|_{L^{2,q}(\Pi_T)} &\leqslant cN^2e^{c_0T}T^{1/q}\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr) \nonumber \\ &\qquad +cN\|\partial^{\mathbf k}\{\mathbf f,\nabla\mathbf f,\partial_t\mathbf f\}\|_{L^{2,q}(\Pi_T)}. \end{aligned} \end{equation} \tag{5.5} $$

2. Assume that conditions (3.2) hold for $\mu=0$, and let $\mathbf f_\tau\in L^{2,1}(\Pi_T)$, $\mathbf g_\tau\in L^2(\Pi_T)$ and $\mathbf y_{0\tau}\in L^2(\mathbb{R}^n)$. Then the weak solution $\mathbf y=\mathbf y_\tau$ of the Cauchy problem (3.1) satisfies the estimate

$$ \begin{equation} \begin{aligned} \, &\max\bigl\{\|\partial^{\mathbf k}\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \sqrt{\nu\tau}\, \|\nabla\partial^{\mathbf k}\mathbf y_\tau\|_{L^2(\Pi_T)}\bigr\} \nonumber \\ &\qquad \leqslant e^{c_0T}\bigl(\|\partial^{\mathbf k}\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +2\|\partial^{\mathbf k}\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}+\sqrt{\nu^{-1}\tau}\, \|\partial^{\mathbf k}\mathbf g_\tau\|_{L^2(\Pi_T)}\bigr). \end{aligned} \end{equation} \tag{5.6} $$

3. Assume that conditions (3.2) for $\mu=0$, (3.13) for $\mu_1=0$ and (3.15) hold, let $\partial_tA_{ij}\in L^\infty(0,T)$, $A_{ij}=A_{ji}^T$, $i,j=1,\dots,n$, and also let $\mathbf f_\tau\in L^{2,1}(\Pi_T)$, $\mathbf y_{0\tau}\in H^1(\mathbb{R}^n)$ and $\mathbf y_{1\tau}\in L^2(\mathbb{R}^n)$. Then the weak solution $\mathbf y=\mathbf y_\tau$ of the Cauchy problem (3.10), (3.11) satisfies the estimate

$$ \begin{equation} \begin{aligned} \, &\nu_1\max\bigl\{\|\partial^{\mathbf k}\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \tau\|\partial^{\mathbf k}\{\nabla\mathbf y_\tau,\partial_t\mathbf y_\tau\}\|_{L^{2,\infty}(\Pi_T)}, \nonumber \\ &\qquad\qquad\sqrt{\delta\tau}\, \|\partial^{\mathbf k}\{\partial_t\mathbf y_\tau,\nabla\mathbf y_\tau\}\|_{L^2(\Pi_T)}\bigr\} \nonumber \\ &\leqslant e^{c_1T}\bigl(\|\partial^{\mathbf k}\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} \nonumber \\ &\qquad +\tau\sqrt{2c_A}\, \|\nabla\partial^{\mathbf k}\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +2\tau\|\partial^{\mathbf k}\mathbf y_{1\tau}\|_{L^2(\mathbb{R}^n)} +2\sqrt{2}\, \|\partial^{\mathbf k}\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}\bigr), \end{aligned} \end{equation} \tag{5.7} $$
where $\nu_1=\min\{\sqrt{2\nu},1/\sqrt{2}\}$, $c_1=\max\bigl\{2\bigl(c_0+\sqrt{2}\, \|C\|_{ L^\infty(0,T)}\bigr),(2\nu)^{-1}c_{A1}\bigr\}$, and $c_{A1}\geqslant 0$ is as before.

Proof. By assumption $\mu=\mu_1=0$; therefore, $\overline{c}_0=c_0$ in (3.6), while the constants in (3.16) can be specified as follows:
$$ \begin{equation} \begin{aligned} \, &\nu_1\max\bigl\{\|\mathbf y_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}, \tau\|\{\nabla\mathbf y_\tau,\partial_t\mathbf y_\tau\}\|_{L^{2,\infty}(\Pi_T)}, \sqrt{\delta\tau}\, \|\{\partial_t\mathbf y_\tau,\nabla\mathbf y_\tau\}\|_{L^2(\Pi_T)} \bigr\} \nonumber \\ &\leqslant e^{c_1T}\bigl(\|\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +\tau\sqrt{2c_A} \|\nabla\mathbf y_{0\tau}\|_{L^2(\mathbb{R}^n)} +2\tau\|\mathbf y_{1\tau}\|_{L^2(\mathbb{R}^n)} +2\sqrt{2}\|\mathbf f_\tau\|_{L^{2,1}(\Pi_T)}\bigr) , \end{aligned} \end{equation} \tag{5.8} $$
where $\nu_1$ and $c_1$ are defined above and the condition $8\overline{\tau}^2\mu\leqslant 1$ disappears.

Now, in averaging (2.10) we take a smooth kernel $\widehat{e}\in C^\infty(\mathbb{R})$ such that

$$ \begin{equation*} \begin{gathered} \, \widehat{e}(a)>0 \quad \text{for } |a|<1, \qquad \widehat{e}(a)=0 \quad \text{for } |a|\geqslant 1, \\ \widehat{e}(a)=\widehat{e}(-a), \qquad \int_{\mathbb{R}}\widehat{e}(a)\,da=1. \end{gathered} \end{equation*} \notag $$
Under the assumptions in parts 1, 2 and 3, the functions $\sigma^{(h)}\mathbf w(\,\cdot\,,t)$ and $\sigma^{(h)}\mathbf y(\,\cdot\,,t)$ belong to the space $C^\infty(\mathbb{R}^n)$ for all $t$, $0\leqslant t\leqslant T$ and are the strong solutions of the corresponding Cauchy problems
$$ \begin{equation} \mathcal{H}\sigma^{(h)}\mathbf w=\sigma^{(h)}\mathbf f \quad \text{in } \Pi_T, \qquad \sigma^{(h)}\mathbf w\big|_{t=0}=\sigma^{(h)}\mathbf w_0 \quad \text{in } \mathbb{R}^n, \end{equation} \tag{5.9} $$
$$ \begin{equation} \mathcal{P}_\tau\sigma^{(h)}\mathbf y=\sigma^{(h)}\mathbf f_\tau+\tau\partial_i\sigma^{(h)}\mathbf g_{i\tau} \quad \text{in } \Pi_T, \qquad \sigma^{(h)}\mathbf y\big|_{t=0}=\sigma^{(h)}\mathbf y_{0\tau} \quad \text{in } \mathbb{R}^n, \end{equation} \tag{5.10} $$
and
$$ \begin{equation} \begin{gathered} \, \mathcal{H}_\tau\sigma^{(h)}\mathbf y=\sigma^{(h)}\mathbf f_\tau \quad \text{in } \Pi_T, \\ \sigma^{(h)}\mathbf y\big|_{t=0}=\sigma^{(h)}\mathbf y_{0\tau}, \quad \sigma^{(h)}\partial_t\mathbf y\big|_{t=0}=\sigma^{(h)}\mathbf y_{1\tau} \quad \text{in } \mathbb{R}^n. \end{gathered} \end{equation} \tag{5.11} $$
For the first problem in part 1, a), this follows from part b) of the proof of Theorem 1 (while in part 1, b), this is obvious); for the other two problems this is deduced similarly, including the properties that $\partial_t\sigma^{(h)}\mathbf y\in L^{2,1}(\Pi_T)$ and $\partial_t^2\sigma^{(h)}\mathbf y\in L^{2,1}(\Pi_T)$, respectively. By definition the strong solutions of the last two problems satisfy the equations in $L^{2,1}(\Pi_T)$ and the initial conditions in $C(0,T;L^2(\mathbb{R}^n))$. We also have $\partial_t^l\sigma^{(h)}\mathbf y(\,\cdot\,,t)\in C^\infty(\mathbb{R}^n)$ almost everywhere on $(0,T)$, where $l=1$ for problem (5.10) and $l=1,2$ for problem (5.11).

Under the assumptions in parts 1, 2 and 3, applying the operator $\partial^{\mathbf k}$ to the Cauchy problems (5.9), (5.10) and (5.11) yields, respectively,

$$ \begin{equation} \mathcal{H}\partial^{\mathbf k}\sigma^{(h)}\mathbf w=\sigma^{(h)}\partial^{\mathbf k}\mathbf f \quad \text{in } \Pi_T, \qquad \partial^{\mathbf k}\sigma^{(h)}\mathbf w\big|_{t=0} =\sigma^{(h)}\partial^{\mathbf k}\mathbf w_0 \quad \text{in } \mathbb{R}^n, \end{equation} \tag{5.12} $$
$$ \begin{equation} \begin{gathered} \, \mathcal{P}_\tau\partial^{\mathbf k}\sigma^{(h)}\mathbf y =\sigma^{(h)}\partial^{\mathbf k}\mathbf f_\tau+\tau\partial_i\sigma^{(h)}\partial^{\mathbf k}\mathbf g_{i\tau} \quad \text{in } \Pi_T, \\ \partial^{\mathbf k}\sigma^{(h)}\mathbf y\big|_{t=0} =\sigma^{(h)}\partial^{\mathbf k}\mathbf y_{0\tau} \quad \text{in } \mathbb{R}^n, \end{gathered} \end{equation} \tag{5.13} $$
and
$$ \begin{equation} \mathcal{H}_\tau\partial^{\mathbf k}\sigma^{(h)}\mathbf y=\sigma^{(h)}\partial^{\mathbf k}\mathbf f_\tau \quad \text{in } \Pi_T, \end{equation} \tag{5.14} $$
$$ \begin{equation} \partial^{\mathbf k}\sigma^{(h)}\mathbf y\big|_{t=0} =\sigma^{(h)}\partial^{\mathbf k}\mathbf y_{0\tau}, \quad \partial^{\mathbf k}\sigma^{(h)}\partial_t\mathbf y\big|_{t=0} =\sigma^{(h)}\partial^{\mathbf k}\mathbf y_{1\tau} \quad \text{in } \mathbb{R}^n. \end{equation} \tag{5.15} $$
In addition, $\sigma^{(h)}\partial^{\mathbf k}\mathbf f\to\partial^{\mathbf k}\mathbf f, \sigma^{(h)}\partial^{\mathbf k}\mathbf f_\tau\to\partial^{\mathbf k}\mathbf f_\tau$ in $L^{2,1}(\Pi_T)$, $\sigma^{(h)}\partial^{\mathbf k}\mathbf g_\tau\to\partial^{\mathbf k}\mathbf g_\tau$ in $L^2(\Pi_T)$, $\sigma^{(h)}\partial^{\mathbf k}\mathbf y_{0\tau}\to \partial^{\mathbf k}\mathbf y_{0\tau}$ in $H^{l-1}(\mathbb{R}^n)$ and $\sigma^{(h)}\partial^{\mathbf k}\mathbf y_{1\tau}\to \partial^{\mathbf k}\mathbf y_{1\tau}$ in $L^2(\mathbb{R}^n)$ as $h\to 0$. By (2.5), (3.6) and (5.8), for solutions of the Cauchy problems we have $\partial^{\mathbf k}\sigma^{(h)}\mathbf w\to\mathbf w^{(\mathbf k)}$ in $C(0,T;L^2(\mathbb{R}^n))$ and $\partial^{\mathbf k}\sigma^{(h)}\mathbf y\to\mathbf y^{(\mathbf k)}$ in $C(0,T;L^2(\mathbb{R}^n))\cap V_2(\Pi_T)$ (for $l=1$) or in $W_{2,\infty}^{1,1}(\Pi_T)$ (for $l=2$) as $h\to 0$. The limit functions $\mathbf w^{(\mathbf k)}$ as $\partial^{\mathbf k}\mathbf w$ and $\mathbf y^{(\mathbf k)}$ as $\partial^{\mathbf k}\mathbf y$ satisfy estimates (5.1), (5.6) (for $l=1$) and (5.7) (for ${l=2}$).

Taking the limit in the integral identities that are the definitions of weak solutions of the Cauchy problems (5.12), (5.13) and (5.14), (5.15) shows that $\mathbf w^{(\mathbf k)}$ and $\mathbf y^{(\mathbf k)}$ are weak solutions of the corresponding Cauchy problems

$$ \begin{equation} \mathcal{H}\mathbf w^{(\mathbf k)}=\partial^{\mathbf k}\mathbf f \quad \text{in } \Pi_T, \qquad \mathbf w^{(\mathbf k)}\big|_{t=0}=\partial^{\mathbf k}\mathbf w_0 \quad \text{in } \mathbb{R}^n, \end{equation} \tag{5.16} $$
$$ \begin{equation} \mathcal{P}_\tau\mathbf y^{(\mathbf k)}=\partial^{\mathbf k}\mathbf f_\tau+\tau\partial_i\partial^{\mathbf k}\mathbf g_{i\tau} \quad \text{in } \Pi_T, \qquad \mathbf y^{(\mathbf k)}\big|_{t=0}=\partial^{\mathbf k}\mathbf y_{0\tau} \quad \text{in } \mathbb{R}^n, \end{equation} \tag{5.17} $$
and
$$ \begin{equation} \mathcal{H}_\tau\mathbf y^{(\mathbf k)}=\partial^{\mathbf k}\mathbf f_\tau \quad \text{in } \Pi_T, \qquad \mathbf y^{(\mathbf k)}\big|_{t=0}=\partial^{\mathbf k}\mathbf y_{0\tau}, \quad \partial_t\mathbf y^{(\mathbf k)}\big|_{t=0}=\partial^{\mathbf k}\mathbf y_{1\tau} \quad \text{in } \mathbb{R}^n. \end{equation} \tag{5.18} $$
Taking the limit in the definition of the Sobolev derivative
$$ \begin{equation*} (\sigma^{(h)}\mathbf w,\partial^{\mathbf k}\boldsymbol\varphi)_{\Pi_T}=(-1)^{|\mathbf k|_1}(\partial^{\mathbf k}\sigma^{(h)}\mathbf w,\boldsymbol\varphi)_{\Pi_T}, \end{equation*} \notag $$
where $\boldsymbol\varphi\in C^\infty(\Pi_T)$, $\boldsymbol\varphi$ has a compact support in $\Pi_T$, yields $ (\mathbf w,\partial^{\mathbf k}\boldsymbol\varphi)_{\Pi_T}=(-1)^{|\mathbf k|_1}(\mathbf w^{(\mathbf k)},\boldsymbol\varphi)_{\Pi_T}, $ that is, the derivative $\partial^{\mathbf k}\mathbf w=\mathbf w^{(\mathbf k)}$ exists too. In a similar way, the derivative $\partial^{\mathbf k}\mathbf y=\mathbf y^{(\mathbf k)}$ exists. This completes the derivation of estimates (5.1), (5.6), and (5.7).

Now we deduce the estimates in part 1, b). Estimate (5.2) follows from (5.1) after we replace $\partial^{\mathbf k}$ by $\partial_i\partial^{\mathbf k}$ and $\partial_i\partial_j\partial^{\mathbf k}$ and sum over $i,j=1,\dots,n$. By virtue of the equation in (5.16) for $\mathbf w^{(\mathbf k)}=\partial^{\mathbf k}\mathbf w$, the derivatives $\partial_t^l\partial^{\mathbf k}\mathbf w$, $l=1,2$, can be expressed successively in terms of the $\nabla^p\partial^{\mathbf k}\mathbf w$, $p=0,l$, and the derivatives of $\mathbf f$ (like in parts c) and e) of the proof of Theorem 1); therefore, (5.3) and (5.5) follow from (5.1) and (5.2). Finally, for $p=2$, by virtue of the equation in (5.16) again, $\nabla\partial_t\mathbf w^{(\mathbf k)}=-B_i\nabla\partial_i\mathbf w^{(\mathbf k)}-C\nabla\mathbf w^{(\mathbf k)}+\nabla\partial^{\mathbf k}\mathbf f\in L^{2,1}(\Pi_T)$, and estimate (5.4) follows from (5.2). Theorem 8 is proved.

The estimates in the last theorem are also valid for $|\mathbf k|_1=0$ when $\partial^{\mathbf k}v=v$.

We switch to estimates for derivatives of the differences of solutions of the systems under consideration. For brevity we restrict ourselves to the simplest choice of $\mathbf f_\tau$, $\mathbf g_\tau$, $\mathbf y_{0\tau}$ and $\mathbf y_{1\tau}$. For $|\mathbf k|_1=0$ we will have the same estimates as above but with refined constants.

Theorem 9. Assume that the assumptions of Theorem 8, parts 1 and 2, hold except for the conditions on $\partial_t\mathbf f$ and $\partial_t\mathbf B,\partial_tC$, and that $\|A\|_{L^\infty(0,T)}\leqslant N$. Let $\mathbf f_\tau=\mathbf f$, $\mathbf g_\tau=0$, $\mathbf y_{0\tau}=\mathbf w_0$ and $|\mathbf k|_1\geqslant 0$. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of solutions of the Cauchy problems (2.1) and (3.1) satisfies the estimates

$$ \begin{equation*} \begin{aligned} \, &\|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} +\sqrt{\tau}\, \|\nabla\partial^{\mathbf k}\mathbf r_\tau\|_{L^2(\Pi_T)} \\ &\qquad\qquad \leqslant cNe^{2c_0T}\sqrt{T\tau}\, \bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^1(\mathbb{R}^n)} +\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{1,0}(\Pi_T)}\bigr), \\ & \|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\sqrt{\tau}\, \|\nabla\partial^{\mathbf k}\mathbf r_\tau\|_{L^2(\Pi_T)} \\ &\qquad\qquad \leqslant cNe^{2c_0T}T\tau\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^2(\mathbb{R}^n)} +\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{2,0}(\Pi_T)}\bigr), \\ &\|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \\ &\qquad\qquad \leqslant c N^\alpha e^{(1+\alpha)c_0T}(T\tau)^{\alpha/2} \bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{\mathcal{H}^{\alpha}}+\|\partial^{\mathbf k}\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,0}}\bigr), \qquad 0<\alpha<1, \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, & \|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))}+\sqrt{\tau}\, \|\nabla\partial^{\mathbf k}\mathbf r_\tau\|_{L^2(\Pi_T)} \\ &\qquad\qquad \leqslant cNe^{2c_0T}(T\tau)^{\alpha/2} \bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{\mathcal{H}^{\alpha}}+\|\partial^{\mathbf k}\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,0}}\bigr), \qquad 1<\alpha<2. \end{aligned} \end{equation*} \notag $$

Proof. By assumption and according to Theorem 8, part 1 for $q=1$ and part 2, and equations (5.16) and (5.17), the functions $\mathbf w^{(\mathbf k)}=\partial^{\mathbf k}\mathbf w$ and $\mathbf y_\tau^{(\mathbf k)}=\partial^{\mathbf k}\mathbf y_\tau$ are the weak solutions of the Cauchy problems
$$ \begin{equation*} \mathcal{H}\mathbf w^{(\mathbf k)}=\mathbf f^{(\mathbf k)} \quad \text{in } \Pi_T, \qquad \mathbf w^{(\mathbf k)}\big|_{t=0}=\mathbf w_0^{(\mathbf k)} \quad \text{in } \mathbb{R}^n \end{equation*} \notag $$
and
$$ \begin{equation*} \mathcal{P}_\tau\mathbf y_\tau^{(\mathbf k)}=\mathbf f^{(\mathbf k)} \quad \text{in } \Pi_T, \qquad \mathbf y_\tau^{(\mathbf k)}\big|_{t=0}=\mathbf w_0^{(\mathbf k)} \quad \text{in } \mathbb{R}^n, \end{equation*} \notag $$
where $\mathbf f^{(\mathbf k)}:=\partial^{\mathbf k}\mathbf f$ and $\mathbf w_0^{(\mathbf k)}:=\partial^{\mathbf k}\mathbf w_0$. Hence their difference $\mathbf r_\tau^{(\mathbf k)}:=\mathbf w^{(\mathbf k)}-\mathbf y_\tau^{(\mathbf k)}=\partial^{\mathbf k}\mathbf r_\tau$ in the role of $\mathbf r_\tau$ merely satisfies estimates (4.2), (4.3), (4.6) and (4.7) with $\mathbf f^{(\mathbf k)}$ as $\mathbf f$ and $\mathbf w_0^{(\mathbf k)}$ as $\mathbf w_0$. In addition, owing to (5.1), (5.2) and (5.6) in Theorem 8, it is straightforward to infer from the proofs of the estimates listed above that under the assumptions of this theorem the constants in these estimates can be refined in the way described above. The theorem is proved.

Theorem 10. Let all assumptions of Theorem 8, part 1 for $q=1$ and part 3, hold and let $\|A\|_{L^\infty(0,T)}\leqslant N$. Let $\mathbf f_\tau=\mathbf f$, $\mathbf y_{0\tau}=\mathbf w_0$, $\mathbf y_{1\tau}=0$ and $|\mathbf k|_1\geqslant 0$. Then the difference $\mathbf r_\tau=\mathbf w-\mathbf y_\tau$ of solutions of the Cauchy problems (2.1) and (3.10), (3.11) satisfies the estimate

$$ \begin{equation*} \begin{aligned} \, &\|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} {+}\tau\|\{\nabla\partial^{\mathbf k}\mathbf r_\tau,\partial_t\partial^{\mathbf k}\mathbf r_\tau\}\|_{L^{2,\infty}(\Pi_T)}{+}\sqrt{\delta\tau}\, \|\{\nabla\partial^{\mathbf k}\mathbf r_\tau,\partial_t\partial^{\mathbf k}\mathbf r_\tau\}\|_{L^2(\Pi_T)} \\ &\qquad \leqslant cN^2e^{(c_0+c_1)T}(T+1)\tau\bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{H^2(\mathbb{R}^n)}+\|\partial^{\mathbf k}\mathbf f\|_{W_{2,1}^{2,1}(\Pi_T)}\bigr). \end{aligned} \end{equation*} \notag $$

For $\mathbf y_{0\tau}=\overline{\sigma}^{(\tau)}\mathbf w_0$ (instead of $\mathbf y_{0\tau}=\mathbf w_0$) and $0<\alpha\leqslant 2$, it is also true that

$$ \begin{equation*} \begin{aligned} \, & \|\partial^{\mathbf k}\mathbf r_\tau\|_{C(0,T;L^2(\mathbb{R}^n))} \\ &\qquad \leqslant cN^\alpha \exp\biggl(\biggl(\frac{\alpha c_0}2+c_1\biggr)T\biggr)[(T+1)\tau]^{\alpha/2} \bigl(\|\partial^{\mathbf k}\mathbf w_0\|_{\mathcal{H}^\alpha}+\|\partial^{\mathbf k}\mathbf f\|_{\mathcal{W}_{2,1}^{\alpha,\alpha/2}}\bigr). \end{aligned} \end{equation*} \notag $$

We can prove this theorem similarly to the previous proof using Theorem 8, part 1 for $q=1$ and part 3, for the Cauchy problem (5.18), as well as using Theorems 6 and 7 for the Cauchy problem (3.10), (3.11). The formula $\mathbf y_{0\tau}=\overline{\sigma}^{(\tau)}\mathbf w_0$ implies that $\mathbf y_\tau^{(\mathbf k)}\big|_{t=0}=\overline{\sigma}^{(\tau)}\mathbf w_0^{(\mathbf k)}$. The refined constants in the estimates are derived from the proofs of Theorems 6 and 7 in view of (5.1), (5.2), (5.5) and (5.7).

It is important that for $c_0=0$ — that is, for $C\geqslant 0$ almost everywhere on $(0,T)$ — and $c_1=0$ — that is, for $C=0$ and $\partial_tA\geqslant 0$ almost everywhere on $(0,T)$ (the case considered in the next section) — the constants in the corresponding estimates in Theorems 810 are bounded or have a power growth with respect to $T$, rather than an exponential one.

Since each estimate obtained in Theorems 810 can be summed over all $\mathbf k$ with $|\mathbf k|_1=p$, similar estimates are true if we replace $\partial^{\mathbf k}$ on the left- and right-hand sides by $\nabla^p$ for any $p\geqslant 1$. In particular, the estimates for $\mathbf r_\tau$ modified in this way for $p>n/2$ (and the original estimates for $\mathbf k=0$), in view of Sobolev’s embedding theorem (see, for example, [32], Theorem 1.2.1) imply estimates of all the orders mentioned in Theorems 9 and 10 in the uniform norm $\|\mathbf r_\tau\|_{C_b(\overline{\Pi}_T)}$.

The above estimates for the derivatives $\partial^{\mathbf k}\mathbf v$, where $\mathbf v=\mathbf w,\mathbf y_\tau,\mathbf r_\tau$, can be supplemented with estimates for the derivatives $\partial_t^s\partial^{\mathbf k}\mathbf v$ for $s\geqslant 1$ and $|\mathbf k|_1\geqslant 0$; however, we do not dwell on them.

§ 6. Linearized quasi-gasdynamic systems of equations and the linearized system of gas dynamic equations

We write linearized (at a constant solution) quasi-gasdynamic systems of equations with respect to the normalized (dimensionless) vector of small perturbations $\widetilde{\mathbf z}(x,t):=(\widetilde{\rho},\widetilde{\mathbf u},\widetilde{\varepsilon})(x,t)$ of the gas density, velocity and specific internal energy, where $\widetilde{\mathbf u}=(\widetilde{u}_1,\dots,\widetilde{u}_n)$, $n=1,2,3$. These are a system of differential equations of order $(\ell+1)$ with respect to $t$ and of the second order with respect to $x$ with constant coefficients (see [16] and [17]), where $\ell=0$ for the parabolic quasi-gasdynamic system and $\ell=1$ for the hyperbolic one:

$$ \begin{equation} \ell\tau\partial_t^2\widetilde{\rho} +\partial_t\widetilde{\rho}+c_*\biggl(\mathbf M\nabla\widetilde{\rho}+\frac{1}{\sqrt{\gamma}}\operatorname{div}\widetilde{\mathbf u}\biggr) \nonumber \end{equation} \notag $$
$$ \begin{equation} \qquad =\tau c_*^2\biggl[\frac{1}{\gamma}\Delta\widetilde{\rho}+(\mathbf M\nabla)\mathbf M\nabla\widetilde{\rho} +\frac{2}{\sqrt{\gamma}}(\mathbf M\nabla)\operatorname{div}\widetilde{\mathbf u} +\frac{1}{\sqrt{\gamma\gamma_*}}\Delta\widetilde{\varepsilon}\biggr]+f_{0\tau}, \end{equation} \tag{6.1} $$
$$ \begin{equation} \ell\tau\partial_t^2\widetilde{\mathbf u} +\partial_t\widetilde{\mathbf u}+c_*\biggl(\frac{1}{\sqrt{\gamma}}\nabla\widetilde{\rho}+(\mathbf M\nabla)\widetilde{\mathbf u} +\frac{1}{\sqrt{\gamma_*}}\nabla\widetilde{\varepsilon}\biggr) =\tau c_*^2\biggl[\frac{2}{\sqrt{\gamma}}(\mathbf M\nabla)\nabla\widetilde{\rho} \nonumber \end{equation} \notag $$
$$ \begin{equation} \qquad\qquad +\widehat{\alpha}_s\Delta\widetilde{\mathbf u}+(\widehat{a}_0+1)\nabla\operatorname{div}\widetilde{\mathbf u}+(\mathbf M\nabla)(\mathbf M\nabla)\widetilde{\mathbf u} +\frac{2}{\sqrt{\gamma_*}}(\mathbf M\nabla)\nabla\widetilde{\varepsilon}\biggr]+\widetilde{\mathbf f}_\tau, \end{equation} \tag{6.2} $$
$$ \begin{equation} \ell\tau\partial_t^2\widetilde{\varepsilon}+\partial_t\widetilde{\varepsilon}+c_*\biggl(\frac{1}{\sqrt{\gamma_*}}\operatorname{div}\widetilde{\mathbf u}+\mathbf M\nabla\widetilde{\varepsilon}\biggr) =\tau c_*^2\biggl[\frac{1}{\sqrt{\gamma\gamma_*}}\Delta\widetilde{\rho} \nonumber \end{equation} \notag $$
$$ \begin{equation} \qquad\qquad +\frac{2}{\sqrt{\gamma_*}}(\mathbf M\nabla)\operatorname{div}\widetilde{\mathbf u} +\biggl(\widehat{\alpha}_P+\frac{1}{\gamma_*}\biggr)\Delta\widetilde{\varepsilon} +(\mathbf M\nabla)\mathbf M\nabla\widetilde{\varepsilon}\biggr]+f_{(n+1)\tau} \end{equation} \tag{6.3} $$
in $\Pi_T$ (a typo in the coefficient of the first term on the right-hand side of (6.2) is corrected here). The free terms $f_{0\tau}(x,t)$, $\widetilde{\mathbf f}_\tau(x,t)$ and $f_{(n+1)\tau}(x,t)$ are added formally instead of zeros for a greater generality of the subsequent analysis. Here $c_*>0$ is the background speed of sound, $\mathbf M=(M_1,\dots,M_n)$ is the background velocity normalized by $c_*$, $M=|\mathbf M|$ is thus the background Mach number and, furthermore,
$$ \begin{equation*} \mathbf M\nabla=\mathbf M\cdot\nabla, \qquad \gamma_*=\frac{\gamma}{\gamma-1}\quad\text{and} \quad \widehat{a}_0=\frac{1}{3}\widehat{\alpha}_s+\widehat{\alpha}_{1s}\geqslant 0, \end{equation*} \notag $$
where $\gamma>1$ is the heat capacity ratio in the gas equation of state; $\widehat{\alpha}_s\geqslant 0$, $\widehat{\alpha}_{1s}\geqslant 0$ and $\widehat{\alpha}_P\geqslant 0$ are constant quasi-gasdynamic parameters in the artificial coefficients of dynamic and volumetric viscosity and thermal conductivity; $\tau>0$ is the relaxation parameter.

We can rewrite this system of equations in a symmetrized matrix form; then the Cauchy problem for it takes the form

$$ \begin{equation} \ell\tau\partial_t^2\widetilde{\mathbf z}+\partial_t\widetilde{\mathbf z} +c_*B^{(i)}\partial_i\widetilde{\mathbf z} -\tau c_*^2\bigl(A^{(ii)}\partial_i^2\widetilde{\mathbf z}+(1-\delta^{(ij)})\widehat{A}^{(ij)}\partial_i\partial_j\widetilde{\mathbf z}\bigr) =\mathbf f_\tau \quad \text{in } \Pi_T, \end{equation} \tag{6.4} $$
$$ \begin{equation} \widetilde{\mathbf z}|_{t=0}=\widetilde{\mathbf z}_{0\tau}, \quad \partial_t\widetilde{\mathbf z}|_{t=0}=\widetilde{\mathbf z}_{1\tau} \quad (\text{for } \ell=1) \quad \text{in } \mathbb{R}^n, \end{equation} \tag{6.5} $$
where $B^{(i)}$ and $A^{(ii)}$, $\widehat{A}^{(ij)}$ are constant matrices of convective and viscous terms of order $m=n+2$ (cf. [16] and [17]) and $\mathbf f_\tau=(f_{0\tau},\widetilde{\mathbf f}_\tau,f_{(n+1)\tau})$.

For $\tau=0$ this turns to the Cauchy problem for the first-order linearized system of gas dynamic equations

$$ \begin{equation} \partial_t\mathbf w+c_*B^{(i)}\partial_i\mathbf w=\mathbf f \quad \text{in } \Pi_T, \qquad \mathbf w|_{t=0}=\mathbf w_0 \quad \text{in } \mathbb{R}^n. \end{equation} \tag{6.6} $$

We write out the arising matrices. Let $\mathbf e_0,\dots,\mathbf e_{n+1}$ be vectors of the canonical basis in $\mathbb{R}^{n+2}$. We introduce the matrices $E^{(k,l)}:=\mathbf e_k\mathbf e_l^T+\mathbf e_l\mathbf e_k^T$; then

$$ \begin{equation*} \begin{aligned} \, B^{(k)} &=M_kI_{n+2} +\frac{1}{\sqrt\gamma}E^{(0,k)}+\frac{1}{\sqrt{\gamma_*}}E^{(k,n+1)}, \\ A^{(kk)} &=D_\gamma+M_k^2I_{n+2} +\frac{2}{\sqrt\gamma}M_kE^{(0,k)}+\frac{2}{\sqrt{\gamma_*}}M_kE^{(k,n+1)} \\ &\qquad\qquad +(\widehat{a}_0+1)\mathbf e_k\mathbf e_k^T+\frac{1}{\sqrt{\gamma\gamma_*}}E^{(0,n+1)}, \\ D_\gamma &:= \operatorname{diag}\biggl\{\frac{1}{\gamma},\widehat{\alpha}_s, \dots,\widehat{\alpha}_s,\widehat{\alpha}_P+\frac{1}{\gamma_*}\biggr\} \end{aligned} \end{equation*} \notag $$
and
$$ \begin{equation*} \begin{aligned} \, \widehat{A}^{(kl)} &=M_kM_lI_{n+2} +\frac{1}{\sqrt\gamma}\bigl(M_kE^{(0,l)}+M_lE^{(0,k)}\bigr) \\ &\qquad\qquad +\frac{1}{\sqrt{\gamma_*}}\bigl(M_kE^{(l,n+1)}+M_lE^{(k,n+1)}\bigr) +\frac12(\widehat{a}_0+1)E^{(k,l)} \end{aligned} \end{equation*} \notag $$
for $k$ and $l$ from $1$ to $n$ (like in [33] and [17]). Throughout, $\operatorname{diag}\{d_1,\dots,d_{n+2}\}$ is the diagonal matrix with the indicated diagonal entries. The matrices $E^{(k,l)}$, $B^{(k)}$, $A^{(kk)}$ and $\widehat{A}^{(kl)}$ are symmetric and $\widehat{A}^{(kl)}=\widehat{A}^{(lk)}$.

For completeness we also present the matrices $B^{(k)}$, $A^{(kk)}$, $\widehat{A}^{(kl)}$ ($k\neq l$) in the $ 3\times 3 $ block form (see [17])

$$ \begin{equation*} \begin{gathered} \, B^{(k)} =\begin{pmatrix} M_k&\dfrac{1}{\sqrt\gamma}\widetilde{\mathbf e}_k^T& 0 \\ \dfrac{1}{\sqrt\gamma}\check{\mathbf e}_k &M_kI_n &\dfrac{1}{\sqrt\gamma_*}\check{\mathbf e}_k \\ 0&\dfrac{1}{\sqrt\gamma_*}\check{\mathbf e}_k^T& M_k \end{pmatrix}, \\ A^{(kk)} =\begin{pmatrix} M_k^2+\dfrac{1}{\gamma}& \dfrac{2}{\sqrt\gamma}M_k\check{\mathbf e}_k^T &\dfrac{1}{\sqrt{\gamma\gamma_*}} \\ \dfrac{2}{\sqrt\gamma}M_k\check{\mathbf e}_k& (M_k^2+\widehat{\alpha}_s)I_n+(\widehat{a}_0+1)\check{\mathbf e}_k\check{\mathbf e}_k^T &\dfrac{2}{\sqrt\gamma_*}M_k\check{\mathbf e}_k \\ \dfrac{1}{\sqrt{\gamma\gamma_*}}&\dfrac{2}{\sqrt\gamma_*}M_k\check{\mathbf e}_k^T & M_k^2+\widehat{\alpha}_P+\dfrac{1}{\gamma_*} \end{pmatrix}, \end{gathered} \end{equation*} \notag $$
where typos in the blocks $(1,1)$ and $(3,3)$ are corrected,
$$ \begin{equation*} \widehat{A}^{(kl)} =\begin{pmatrix} M_kM_l&\dfrac{1}{\sqrt\gamma}(M_k\check{\mathbf e}_l+M_l\check{\mathbf e}_k)^T&0 \\ \dfrac{1}{\sqrt\gamma}(M_k\check{\mathbf e}_l+M_l\check{\mathbf e}_k) & \widehat{A}^{(kl)}_{22} &\dfrac{1}{\sqrt\gamma_*}(M_k\check{\mathbf e}_l+M_l\check{\mathbf e}_k) \\ 0&\dfrac{1}{\sqrt\gamma_*}(M_k\check{\mathbf e}_l+M_l\check{\mathbf e}_k)^T& M_kM_l \end{pmatrix}, \end{equation*} \notag $$
where $\widehat{A}^{(kl)}_{22} =M_kM_lI_n+{\frac12(\widehat{a}_0+1)}(\check{\mathbf e}_k\check{\mathbf e}_l^T+\check{\mathbf e}_l\check{\mathbf e}_k^T)$, and we recall that $\check{\mathbf e}_1,\dots,\check{\mathbf e}_n$ are the vectors of the canonical basis in $\mathbb{R}^n$.

For $\widetilde{\mathbf z}=(\widetilde{\rho},\widetilde{\mathbf u},\widetilde{\varepsilon})$, $\mathbf z=(\rho,\mathbf u,\varepsilon)\in H^1(\mathbb{R}^n)$ we introduce the bilinear form

$$ \begin{equation*} \begin{aligned} \, &\mathcal{A}_{\mathbb{R}^n}(\widetilde{\mathbf z},\mathbf z) =\frac{1}{\gamma}(\nabla\widetilde{\rho},\nabla\rho)_{\mathbb{R}^n} +(\mathbf M\nabla\widetilde{\rho},\mathbf M\nabla\rho)_{\mathbb{R}^n} +\frac{2}{\sqrt{\gamma}}((\mathbf M\nabla)\widetilde{\mathbf u},\nabla\rho)_{\mathbb{R}^n} \\ &\qquad+\frac{1}{\sqrt{\gamma\gamma_*}}(\nabla\widetilde{\varepsilon},\nabla\rho)_{\mathbb{R}^n} +\frac{2}{\sqrt{\gamma}}(\mathbf M\nabla\widetilde{\rho},\operatorname{div}\mathbf u)_{\mathbb{R}^n} +\widehat{\alpha}_s(\nabla\widetilde{\mathbf u},\nabla\mathbf u)_{\mathbb{R}^n} \\ &\qquad +(\widehat{a}_0+1)(\operatorname{div}\widetilde{\mathbf u},\operatorname{div}\mathbf u)_{\mathbb{R}^n} +((\mathbf M\nabla)\widetilde{\mathbf u},(\mathbf M\nabla)\mathbf u)_{\mathbb{R}^n} \\ &\qquad +\frac{2}{\sqrt{\gamma_*}}(\mathbf M\nabla\widetilde{\varepsilon},\operatorname{div}\mathbf u)_{\mathbb{R}^n} +\frac{1}{\sqrt{\gamma\gamma_*}}(\nabla\widetilde{\rho},\nabla\varepsilon)_{\mathbb{R}^n} +\frac{2}{\sqrt{\gamma_*}}((\mathbf M\nabla)\widetilde{\mathbf u},\nabla\varepsilon)_{\mathbb{R}^n} \\ &\qquad +\biggl(\widehat{\alpha}_P+\frac{1}{\gamma_*}\biggr)(\nabla\widetilde{\varepsilon}, \nabla\varepsilon)_{\mathbb{R}^n} +(\mathbf M\nabla\widetilde{\varepsilon},\mathbf M\nabla\varepsilon)_{\mathbb{R}^n}. \end{aligned} \end{equation*} \notag $$
This bilinear form with minus sign is obtained formally by multiplying the right-hand sides of (6.1)(6.3) without the factors $\tau c_*^2$ and the free terms $f_{0\tau}$, $\widetilde{\mathbf f}_\tau$ and $f_{(n+1)\tau}$ by $\rho$, $\mathbf u$ and $\varepsilon$, respectively, integrating over $\mathbb{R}^n$ and then integrating by parts.

It is associated with the quadratic form

$$ \begin{equation*} \begin{aligned} \, &\mathcal{A}_{\mathbb{R}^n}(\mathbf z,\mathbf z) =\frac{1}{\gamma}\|\nabla\rho\|_{\mathbb{R}^n}^2 +\|\mathbf M\nabla\rho\|_{\mathbb{R}^n}^2 +\widehat{\alpha}_s\|\nabla\mathbf u\|_{\mathbb{R}^n}^2 +(\widehat{a}_0+1)\|\operatorname{div}\mathbf u\|_{\mathbb{R}^n}^2 \\ &\qquad+\|(\mathbf M\nabla)\mathbf u\|_{\mathbb{R}^n}^2 +\biggl(\widehat{\alpha}_P+\frac{1}{\gamma_*}\biggr)\|\nabla\varepsilon\|_{\mathbb{R}^n}^2 +\|\mathbf M\nabla\varepsilon\|_{\mathbb{R}^n}^2 +\frac{2}{\sqrt{\gamma\gamma_*}}(\nabla\rho,\nabla\varepsilon)_{\mathbb{R}^n} \\ &\qquad +\frac{2}{\sqrt{\gamma}}\bigl[(\nabla\rho,(\mathbf M\nabla)\mathbf u)_{\mathbb{R}^n}+(\mathbf M\nabla\rho,\operatorname{div}\mathbf u)_{\mathbb{R}^n}\bigr] \\ &\qquad+\frac{2}{\sqrt{\gamma_*}}\bigl[(\nabla\varepsilon,(\mathbf M\nabla)\mathbf u)_{\mathbb{R}^n}+(\mathbf M\nabla\varepsilon,\operatorname{div}\mathbf u)_{\mathbb{R}^n}\bigr]. \end{aligned} \end{equation*} \notag $$

The following lemmas are similar to those deduced recently in [17] (the difference case) and [34] (the differential case).

Lemma 1. The symmetry and nonnegative definiteness properties

$$ \begin{equation} \mathcal{A}_{\mathbb{R}^n}(\widetilde{\mathbf z},\mathbf z) =\mathcal{A}_{\mathbb{R}^n}(\mathbf z,\widetilde{\mathbf z})\qquad \forall\,\widetilde{\mathbf z},\mathbf z\in H^1(\mathbb{R}^n), \end{equation} \tag{6.7} $$
$$ \begin{equation} \mathcal{A}_{\mathbb{R}^n}(\mathbf z,\mathbf z) =\widehat{\alpha}_s\|\nabla\mathbf u\|_{\mathbb{R}^n}^2+\widehat{a}_0\|\operatorname{div}\mathbf u\|_{\mathbb{R}^n}^2 +\widehat{\alpha}_P\|\nabla\varepsilon\|_{\mathbb{R}^n}^2 +\biggl\|\mathbf M\nabla\rho+\frac{1}{\sqrt{\gamma}}\operatorname{div}\mathbf u\biggr\|_{\mathbb{R}^n}^2 \nonumber \end{equation} \notag $$
$$ \begin{equation} \qquad\qquad+\biggl\|\frac{1}{\sqrt{\gamma}}\nabla\rho+(\mathbf M\nabla)\mathbf u+\frac{1}{\sqrt{\gamma_*}}\nabla\varepsilon\biggr\|_{\mathbb{R}^n}^2 +\biggl\|\mathbf M\nabla\varepsilon+\frac{1}{\sqrt{\gamma_*}}\operatorname{div}\mathbf u\biggr\|_{\mathbb{R}^n}^2 \nonumber \end{equation} \notag $$
$$ \begin{equation} \qquad =\widehat{\alpha}_s\|\nabla\mathbf u\|_{\mathbb{R}^n}^2+\widehat{a}_0\|\operatorname{div}\mathbf u\|_{\mathbb{R}^n}^2 +\widehat{\alpha}_P\|\nabla\varepsilon\|_{\mathbb{R}^n}^2 +\|\mathbf B\cdot\nabla\mathbf z\|_{\mathbb{R}^n}^2 \geqslant\|\mathbf B\cdot\nabla\mathbf z\|_{\mathbb{R}^n}^2 \end{equation} \tag{6.8} $$
and
$$ \begin{equation} \mathcal{A}_{\mathbb{R}^n}(\mathbf z,\mathbf z) \geqslant\max\bigl\{\delta_0\|\nabla\rho\|_{\mathbb{R}^n}^2, \widehat{\alpha}_s\|\nabla\mathbf u\|_{\mathbb{R}^n}^2+\widehat{a}_0\|\operatorname{div}\mathbf u\|_{\mathbb{R}^n}^2 +\widehat{\alpha}_P\|\nabla\varepsilon\|_{\mathbb{R}^n}^2\bigr\} \geqslant\delta_1\|\nabla\mathbf z\|_{\mathbb{R}^n}^2 \end{equation} \tag{6.9} $$
hold for any $\mathbf z\in H^1(\mathbb{R}^n)$, with the composite matrix $\mathbf B:=(B^{(1)},\dots,B^{(n)})$ and
$$ \begin{equation*} \mathbf B\cdot\nabla\mathbf z=B^{(i)}\partial_i\mathbf z, \qquad\! \delta_0:=\frac{1}{3\gamma}\min\biggl\{1,\frac{\widehat{\alpha}_s}{M^2}, \gamma_*\widehat{\alpha}_P\biggr\}\quad\textit{and} \quad \delta_1:=\frac12\min\{\delta_0,\widehat{\alpha}_s,\widehat{\alpha}_P\}. \end{equation*} \notag $$

Proof. Integrating by parts twice yields
$$ \begin{equation*} \begin{gathered} \, ((\mathbf M\nabla)\mathbf u,\nabla v)_{\mathbb{R}^n} =-(\mathbf u,(\mathbf M\nabla)\nabla v)_{\mathbb{R}^n} =(\operatorname{div}\mathbf u,\mathbf M\nabla v)_{\mathbb{R}^n} \\ \forall\,\mathbf u\in H^1(\mathbb{R}^n),\quad v\in H^2(\mathbb{R}^n). \end{gathered} \end{equation*} \notag $$
Without the intermediate equality, this formula is also valid for $v\in H^1(\mathbb{R}^n)$ because $H^2(\mathbb{R}^n)$ is dense in $H^1(\mathbb{R}^n)$. It implies the symmetry property (6.7).

It is straightforward to see in view of the equality $1/\gamma+1/\gamma_*=1$ that the following pointwise formula is true:

$$ \begin{equation*} \begin{aligned} \, &\biggl|\frac{1}{\sqrt{\gamma}}\nabla\rho+(\mathbf M\nabla)\mathbf u+\frac{1}{\sqrt{\gamma_*}}\nabla\varepsilon\biggr|^2 +\biggl(\mathbf M\nabla\rho+\frac{1}{\sqrt{\gamma}}\operatorname{div}\mathbf u\biggr)^2 +\biggl(\mathbf M\nabla\varepsilon+\frac{1}{\sqrt{\gamma_*}}\operatorname{div}\mathbf u\biggr)^2 \\ &\qquad =\frac{1}{\gamma}|\nabla\rho|^2+(\mathbf M\nabla\rho)^2 +|(\mathbf M\nabla)\mathbf u|^2+(\operatorname{div}\mathbf u)^2 +\frac{1}{\gamma_*}|\nabla\varepsilon|^2+(\mathbf M\nabla\varepsilon)^2 \\ &\qquad\qquad +\frac{2}{\sqrt{\gamma\gamma_*}}(\nabla\rho)\cdot\nabla\varepsilon +\frac{2}{\sqrt{\gamma}}[(\mathbf M\nabla\rho)\operatorname{div}\mathbf u+(\nabla\rho)\cdot(\mathbf M\nabla)\mathbf u] \\ &\qquad\qquad +\frac{2}{\sqrt{\gamma_*}}[(\mathbf M\nabla\varepsilon)\operatorname{div}\mathbf u+(\nabla\rho)\cdot(\mathbf M\nabla)\mathbf u]. \end{aligned} \end{equation*} \notag $$
It yields the left-hand equality in (6.8). The other relations in (6.8) are obvious.

The left-hand inequality in (6.9) is ensured by (6.8), where the following estimate integrated over $\mathbb{R}^n$ is taken into account:

$$ \begin{equation*} \begin{aligned} \, \frac{1}{\gamma}|\nabla\rho|^2 &\leqslant 3\biggl(\biggl|\frac{1}{\sqrt{\gamma}}\nabla\rho+(\mathbf M\nabla)\mathbf u+\frac{1}{\sqrt{\gamma_*}}\nabla\varepsilon\biggr|^2 +(|\mathbf M||\nabla\mathbf u|)^2+\frac{1}{\gamma_*}|\nabla\varepsilon|^2\biggr) \\ &\leqslant 3\max\biggl\{1,\frac{M^2}{\widehat{\alpha}_s},\frac{1}{\gamma_*\widehat{\alpha}_P}\biggr\} \\ &\qquad\times \biggl(\biggl|\frac{1}{\sqrt{\gamma}}\nabla\rho+(\mathbf M\nabla)\mathbf u+\frac{1}{\sqrt{\gamma_*}}\nabla\varepsilon\biggr|^2 +\widehat{\alpha}_s|\nabla\mathbf u|^2+\widehat{\alpha}_P|\nabla\varepsilon|^2\biggr). \end{aligned} \end{equation*} \notag $$

The right-hand inequality in (6.9) is obvious. The lemma is proved.

We present another, more algebraic, derivation of the important relations (6.8), which shows how property (3.13) for $\mu_1=0$ can also hold when (3.14) fails.

Lemma 2. The following relations hold:

$$ \begin{equation*} \mathcal{A}_{\mathbb{R}^n}(\widetilde{\mathbf z},\mathbf z) =\bigl(A^{(ii)}\partial_i\widetilde{\mathbf z},\partial_i\mathbf z\bigr)_{\mathbb{R}^n} +(1-\delta^{(ij)})\bigl(\widehat{A}^{(ij)}\partial_j\widetilde{\mathbf z},\partial_i\mathbf z\bigr)_{\mathbb{R}^n} \quad \forall\,\widetilde{\mathbf z},\mathbf z\in H^1(\mathbb{R}^n) \end{equation*} \notag $$
and
$$ \begin{equation} \begin{aligned} \, &\bigl(A^{(ii)}\partial_i\mathbf z,\partial_i\mathbf z\bigr)_{\mathbb{R}^n} +(1-\delta^{(ij)})\bigl(\widehat{A}^{(ij)}\partial_j\mathbf z,\partial_i\mathbf z\bigr)_{\mathbb{R}^n} \nonumber \\ &\qquad =\bigl(D\partial_i\mathbf z,\partial_i\mathbf z)_{\mathbb{R}^n} +\widehat{a}_0\|\operatorname{div}\mathbf u\|_{L^2(\mathbb{R}^n)}^2 +\|\mathbf B\cdot\nabla\mathbf z\|_{L^2(\mathbb{R}^n)}^2 \nonumber \\ &\qquad \geqslant\|\mathbf B\cdot\nabla\mathbf z\|_{L^2(\mathbb{R}^n)}^2 \quad \forall\, \mathbf z=(\rho,\mathbf u,\varepsilon)\in H^1(\mathbb{R}^n), \end{aligned} \end{equation} \tag{6.10} $$
where $D:=\operatorname{diag}\bigl\{0,\widehat{\alpha}_s,\dots,\widehat{\alpha}_s,\widehat{\alpha}_P\bigr\}$ is a matrix of order $n+2$.

Proof. The first formula is simply derived from the above transition from equalities (6.1)(6.3) to their matrix form (6.4), the obvious equality
$$ \begin{equation*} \begin{aligned} \, & -\bigl(A^{(ii)}\partial_i^2\widetilde{\mathbf z} +(1-\delta^{(ij)})\widehat{A}^{(ij)}\partial_i\partial_j\widetilde{\mathbf z},\mathbf z\bigr)_{\mathbb{R}^n} \\ &\qquad =\bigl(A^{(ii)}\partial_i\widetilde{\mathbf z},\partial_i\mathbf z\bigr)_{\mathbb{R}^n} +(1-\delta^{(ij)})\bigl(\widehat{A}^{(ij)}\partial_j\widetilde{\mathbf z},\partial_i\mathbf z\bigr)_{\mathbb{R}^n} \end{aligned} \end{equation*} \notag $$
for $\widetilde{\mathbf z}\in H^2(\mathbb{R}^n)$ and $\mathbf z\in H^1(\mathbb{R}^n)$, and the density of $H^2(\mathbb{R}^n)$ in $H^1(\mathbb{R}^n)$.

We introduce the matrices

$$ \begin{equation*} \begin{gathered} \, \widetilde{A}^{(kk)}:=\widehat{A}^{(kk)} +\frac{1}{\gamma}\mathbf e_0\mathbf e_0^T+\frac{1}{\gamma_*}\mathbf e_{n+1}\mathbf e_{n+1}^T+\frac{1}{\sqrt{\gamma\gamma_*}}E^{(0,n+1)} =A^{(kk)}-D, \\ k=1,\dots,n. \end{gathered} \end{equation*} \notag $$
It was established in the proof of Theorem 1 in [17] that
$$ \begin{equation*} \begin{aligned} \, &\bigl(\widetilde{A}^{(ii)}\mathbf v_i\bigr)\cdot\mathbf v_i +\bigl(1-\delta^{(ij)}\bigr)(\widehat{A}^{(ij)}\mathbf v_j\bigr)\cdot\mathbf v_i -\biggl|\sum_{1\leqslant k\leqslant n}B^{(k)}\mathbf v_k\biggr|^2 \\ &\qquad =\frac12(\widehat{a}_0-1)\biggl(\sum_{1\leqslant k\leqslant n}v_{kk}\biggr)^2 +\frac12(\widehat{a}_0+1)v_{ij}v_{ji} \\ &\qquad \forall\, \mathbf v_i=(v_{i0},\dots,v_{i(n+1)})\in\mathbb{R}^{n+2}, \end{aligned} \end{equation*} \notag $$
where $i=1,\dots,n$. Note that the right-hand side of this formula does not contain the vector $\mathbf M$ or the components $v_{i0}$ and $v_{i(n+1)}$.

This formula does not imply an inequality of type (3.14) for $\delta=0$. Nevertheless, inequality (6.10) of type (3.13) for $\delta=0$ is valid. To derive it, like in the difference case (see [17]), for $\mathbf z=(\rho,\mathbf u,\varepsilon)\in H^1(\mathbb{R}^n)$ and $\mathbf u=(u_1,\dots,u_n)$ we set $\mathbf v_i=\partial_i\mathbf z$ and integrate the last formula over $\mathbb{R}^n$. For all $k\neq l$, integrating twice by parts we obtain $(\partial_ku_l,\partial_lu_k)_{\mathbb{R}^n} =(\partial_lu_l,\partial_ku_k)_{\mathbb{R}^n}$ (the arising additional condition $\partial_k\partial_lu_k\in L^2(\mathbb{R}^n)$ is eliminated because $H^2(\mathbb{R}^n)$ is dense in $H^1(\mathbb{R}^n)$). Owing to this formula, $(\partial_iu_j,\partial_ju_i)_{\mathbb{R}^n}=\|\operatorname{div}\mathbf u\|_{\mathbb{R}^n}^2$, which leads to (6.10). The lemma is proved.

According to the above definitions, a weak solution of the Cauchy problem for the parabolic quasi-gasdynamic system (6.4), (6.5) for $\ell=0$ is a function $\widetilde{\mathbf z}\in V_2(\Pi_T)$ satisfying the integral identity

$$ \begin{equation*} -(\widetilde{\mathbf z},\partial_t\boldsymbol\varphi)_{\Pi_T}+c_*(B^{(i)}\partial_i\widetilde{\mathbf z},\boldsymbol\varphi)_{\Pi_T} +\tau c_*^2\mathcal{A}_{\Pi_T}(\widetilde{\mathbf z},\boldsymbol\varphi) =\ell(\widetilde{\mathbf z}_{0\tau},\mathbf f_{\tau};\boldsymbol\varphi)-\tau(\mathbf g_{i\tau},\partial_i\boldsymbol\varphi)_{\Pi_T} \end{equation*} \notag $$
for any $\boldsymbol\varphi\in H^1(\Pi_T)$ such that $\boldsymbol\varphi|_{t=T}=0$. Here the bilinear form $\mathcal{A}_{\Pi_T}(\,\cdot\,{,}\,\cdot\,)$ is obtained from the bilinear form $\mathcal{A}_{\mathbb{R}^n}(\,\cdot\,{,}\,\cdot\,)$ introduced above by replacing $\mathbb{R}^n$ by $\Pi_T$ in the integrals, while the free term $\mathbf f_\tau$ is taken in the generalized form $\mathbf f_\tau+\tau\partial_i\mathbf g_{i\tau}$.

The weak solution of the Cauchy problem for the hyperbolic quasi-gasdynamic system (6.4), (6.5) for $\ell=1$ is a function $\widetilde{\mathbf z}\in W_{2,\infty}^{1,1}(\Pi_T)$ satisfying the integral identity

$$ \begin{equation*} -\tau(\partial_t\widetilde{\mathbf z},\partial_t\boldsymbol\varphi)_{\Pi_T} +(\partial_t\widetilde{\mathbf z}+c_*B^{(i)}\partial_i\widetilde{\mathbf z},\boldsymbol\varphi)_{\Pi_T} +\tau c_*^2\mathcal{A}_{\Pi_T}(\widetilde{\mathbf z},\boldsymbol\varphi) =\ell(\widetilde{\mathbf z}_{1\tau},\mathbf f_{\tau};\boldsymbol\varphi) \end{equation*} \notag $$
for any $\boldsymbol\varphi\in W_{2,1}^{1,1}(\Pi_T)$ such that $\boldsymbol\varphi|_{t=T}=0$ and satisfying the initial condition $\widetilde{\mathbf z}|_{t=0}=\widetilde{\mathbf z}_{0\tau}$ in $C(0,T;L^2(\mathbb{R}^n))$.

All the above theorems apply to the Cauchy problems for the linearized system of gas dynamic equations (6.6) and the linearized quasi-gasdynamic system (6.4), (6.5) (in view of Lemma 1). Therefore, the final result holds.

Theorem 11. 1. Theorem 1 and Theorem 8, part 1, for $c_0=0$ hold for the Cauchy problem for the linearized system of gas dynamic equations (6.6).

2. Let $\widehat{\alpha}_s>0$ and $\widehat{\alpha}_P>0$. Then for the Cauchy problems for the linearized quasi-gasdynamic systems (6.4), (6.5), property (3.2) holds for $\nu=c_*^2\delta_1$ and $\mu=0$ and property (3.13) holds for $\delta=\mu_1=0$. Hence Theorem 2 and Theorem 8, part 2, for $\overline{c}_0=c_0=0$ hold in the case $\ell=0$, while Theorem 3 and Theorem 8, part 3, for $\overline{c}_1=c_1=0$ hold in the case $\ell=1$.

The difference $\mathbf r_\tau=\mathbf w-\widetilde{\mathbf z}$ of solutions of the Cauchy problems for the linearized system of gas dynamic equations (6.6) and the linearized quasi-gasdynamic systems (6.4), (6.5) satisfies the estimates in Theorems 4 and 9 for $c_0=0$ in the case $\ell=0$ and in Theorems 6 and 10 for $c_0=c_1=0$ in the case $\ell=1$.

§ 7. Conclusions

This paper has studied the Cauchy problems for a first-order symmetric linear hyperbolic system of equations and its singular perturbations that are second-order strongly parabolic and hyperbolic systems of equations with a small parameter $\tau>0$ multiplying second derivatives with respect to $x$ and $t$, with matrix coefficients depending on $x$ and $t$. The existence and uniqueness results have been proved for weak solutions of these problems with $\tau$-uniform estimates of solutions, and regularity results for solutions of the first-order system of equations have been established.

As a basic result, estimates for the differences $\mathbf r_\tau$ of solutions of the first-order system of equations and its perturbations have been deduced, including estimates of the form $\|\mathbf r_\tau\|_ {C(0,T;L^2(\mathbb{R}^n))}=O(\tau^{\alpha/2})$ for an initial function $\mathbf w_0$ of smoothness $\alpha$, ${0<\alpha\leqslant 2}$, in the sense of Sobolev and Nikolskii, where the case $\alpha=1/2$ covers a wide class of discontinuous functions $\mathbf w_0$, and for relevant smoothness conditions on the free term of the first-order system. For coefficients constant in $x$, $\tau$-uniform estimates for the derivatives $\partial^{\mathbf k}$ with respect to $x$ of solutions of the above systems and the estimates $\|\partial^{\mathbf k}\mathbf r_\tau\|_ {C(0,T;L^2(\mathbb{R}^n))}=O(\tau^{\alpha/2})$ have also been obtained for any $\mathbf k$.

An application of these results to the linearized (at a constant solution) first-order system of gas dynamic equations and its linearized perturbations that are second-order parabolic and hyperbolic quasi-gasdynamic systems of equations has been presented.


Bibliography

1. B. N. Chetverushkin, Kinetic schemes and quasi-gasdynamic system of equations, MAKS Press, Moscow, 2004, 328 pp.  zmath; English transl., CIMNE, Barcelona, 2008, 298 pp.
2. T. G. Elizarova, Quasi-gasdynamic equations and methods for computation of viscous flows, Nauchnyi Mir, Moscow, 2007, 349 pp. (Russian); English transl., Quasi-gas dynamic equations, Springer, Berlin, 2008, xiv+286 pp.  crossref
3. B. N. Chetverushkin, “Hyperbolic quasi-gasdynamic system”, Mat. Model., 30:2 (2018), 81–98  mathnet  mathscinet  zmath; English transl. in Math. Models Comput. Simul., 10:5 (2018), 588–600  crossref
4. L. C. Evans, Partial differential equations, Grad. Stud. Math., 19, Amer. Math. Soc., Providence, RI, 1998, xviii+662 pp.  crossref  mathscinet  zmath
5. J. D. Cole, Perturbation methods in applied mathematics, Blaisdell Publishing Co. [Ginn and Co.], Waltham, MA–Toronto, ON–London, 1968, vi+260 pp.  mathscinet  zmath
6. J. Genet and M. Madaune, “Singular perturbations for a class of nonlinear hyperbolic-hyperbolic problems”, J. Math. Anal. Appl., 64:1 (1978), 1–24  crossref  mathscinet  zmath
7. L. R. Volevich and M. G. Dzhavadov, “Uniform bounds of solutions of hyperbolic equations with a small parameter multiplying the leading derivatives”, Differ. Uravn., 19:12 (1983), 2082–2090  mathnet  mathscinet  zmath; English transl. in Differ. Equ., 19 (1983), 1516–1525
8. A. van Harten and R. R. van Hassel, “A quasi-linear, singular perturbation problem of hyperbolic type”, SIAM J. Math. Anal., 16:6 (1985), 1258–1267  crossref  mathscinet  zmath
9. S. Schochet, “Hyperbolic-hyperbolic singular limits”, Comm. Partial Differential Equations, 12:6 (1987), 589–632  crossref  mathscinet  zmath
10. H. O. Fattorini, “The hyperbolic singular perturbation problem: an operator theoretic approach”, J. Differential Equations, 70:1 (1987), 1–41  crossref  mathscinet  zmath  adsnasa
11. E. M. de Jager and F. Jiang, The theory of singular perturbations, North-Holland Ser. Appl. Math. Mech., 42, North-Holland Publishing Co., Amsterdam, 1996, xii+340 pp.  mathscinet
12. S. Mizohata, Henbidun hôteisiki ron [The theory of partial differential equations], Contemp. Math., 9, Iwanami Shoten, Tokyo, 1965, viii+462 pp.  mathscinet
13. S. K. Godunov, Equations of mathematical physics, 2nd revised and augmented ed., Nauka, Moscow, 1979, 391 pp. (Russian)  mathscinet; French transl. of 1st ed., S. Godounov, Équations de la physique mathématique, Éditions Mir, Moscow, 1973, 452 pp.  mathscinet  zmath
14. S. Benzoni-Gavage and D. Serre, Multidimensional hyperbolic partial differential equations. First-order systems and applications, Oxford Math. Monogr., The Clarendon Press, Oxford Univ. Press, Oxford, 2007, xxvi+508 pp.  mathscinet  zmath
15. S. M. Nikol'skiĭ, Approximation of functions of several variables and imbedding theorems, Nauka, Moscow, 1969, 480 pp.  mathscinet  zmath; English transl., Grundlehren Math. Wiss., 205, Springer-Verlag, New York–Heidelberg, 1975, viii+418 pp.  crossref  mathscinet  zmath
16. A. A. Zlotnik and B. N. Chetverushkin, “Parabolicity of the quasi-gasdynamic system of equations, its hyperbolic second-order modification, and the stability of small perturbations for them”, Zh. Vychisl. Mat. Mat. Fiz., 48:3 (2008), 445–472  mathnet  mathscinet  zmath; English transl. in Comput. Math. Math. Phys., 48:3 (2008), 420–446  crossref
17. A. A. Zlotnik and B. N. Chetverushkin, “Stability of implicit difference schemes for a linearized hyperbolic quasi-gasdynamic system of equations”, Differ. Uravn., 56:7 (2020), 936–947  crossref  mathscinet  zmath; English transl. in Differ. Equ., 56:7 (2020), 910–922  crossref
18. A. A. Zlotnik and B. N. Chetverushkin, “On second-order parabolic and hyperbolic perturbations of a first-order hyperbolic system”, Dokl. Ross. Akad. Nauk. Mat. Inform. Protsessy Upr., 506 (2022), 9–15  mathnet  crossref  zmath; English transl. in Dokl. Math., 106:2 (2022), 308–314  crossref
19. H. O. Fattorini, “Singular perturbation and boundary layer for an abstract Cauchy problem”, J. Math. Anal. Appl., 97:2 (1983), 529–571  crossref  mathscinet  zmath
20. A. Z. Ishmukhametov, “Controllability of hyperbolic systems under singular perturbations”, Differ. Uravn., 36:2 (2000), 241–250  mathnet  mathscinet  zmath; English transl. in Differ. Equ., 36:2 (2000), 272–284  crossref
21. T. E. Moiseev, E. E. Myshetskaya and V. F. Tishkin, “On the closeness of solutions of unperturbed and hyperbolized heat equations with discontinuous initial data”, Dokl. Ross. Akad. Nauk, 481:6 (2018), 605–609  crossref  zmath; English transl. in Dokl. Math., 98:1 (2018), 391–395  crossref  mathscinet
22. B. N. Chetverushkin and A. A. Zlotnik, “On a hyperbolic perturbation of a parabolic initial-boundary value problem”, Appl. Math. Lett., 83 (2018), 116–122  crossref  mathscinet  zmath
23. O. A. Ladyženskaja, V. A. Solonnikov and N. N. Ural'ceva (Ural'tseva), Linear and quasi-linear equations of parabolic type, Nauka, Moscow, 1967, 736 pp.  mathscinet  zmath; English transl., Transl. Math. Monogr., 23, Amer. Math. Soc., Providence, RI, 1968, xi+648 pp.  crossref  mathscinet  zmath
24. O. A. Ladyzhenskaya, The boundary value problems of mathematical physics, Nauka, Moscow, 1973, 407 pp.  mathscinet  zmath; English transl., Appl. Math. Sci., 49, Springer-Verlag, New York, 1985, xxx+322 pp.  crossref  mathscinet  zmath
25. L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, Textb. Math., 2nd rev. ed., CRC Press, Boca Raton, FL, 2015, xiv+299 pp.  mathscinet  zmath
26. H. Gajewski, K. Gröger and K. Zacharias, Nichtlineare Operatorgleichungen und Operatordifferentialgleichungen, Math. Lehrbucher und Monogr., 38, Akademie-Verlag, Berlin, 1974, ix+281 pp.  mathscinet  zmath
27. A. A. Zlotnik and B. N. Chetverushkin, “Spectral stability conditions for an explicit three-level finite-difference scheme for a multidimensional transport equation with perturbations”, Differ. Uravn., 57:7 (2021), 922–931  crossref  mathscinet  zmath; English transl. in Differ. Equ., 57:7 (2021), 891–900  crossref
28. J. Bergh and J. Löfström, Interpolation spaces. An introduction, Grundlehren Math. Wiss., 223, Springer-Verlag, Berlin–New York, 1976, x+207 pp.  crossref  mathscinet  zmath
29. B. N. Chetverushkin and A. A. Zlotnik, “On some properties of multidimensional hyperbolic quasi-gasdynamic systems of equations”, Russ. J. Math. Phys., 24:3 (2017), 299–309  crossref  mathscinet  zmath  adsnasa
30. A. A. Zlotnik, Projection-difference methods for nonstationary problems with nonsmooth data, Kandidat Dissertation, Moscow State University, Moscow, 1979 (Russian)
31. L. Tartar, An introduction to Sobolev spaces and interpolation spaces, Lect. Notes Unione Mat. Ital., 3, Springer, Berlin; UMI, Bologna, 2007, xxvi+218 pp.  crossref  mathscinet  zmath
32. M. S. Agranovich, Sobolev spaces, their generalizations and elliptic problems in smooth and Lipschitz domains, Moscow Center of Continuous Mathematical Education, Moscow, 2013, 378 pp.; English transl., Springer Monogr. Math., Springer, Cham, 2015, xiii+331 pp.  crossref  mathscinet  zmath
33. A. Zlotnik and T. Lomonosov, “$L^2$-dissipativity of the linearized explicit finite-difference scheme with a kinetic regularization for 2D and 3D gas dynamics system of equations”, Appl. Math. Lett., 103 (2020), 106198, 7 pp.  crossref  mathscinet  zmath
34. A. A. Zlotnik and A. S. Fedchenko, “Properties of an aggregated quasi-gasdynamic system of equations for a homogeneous gas mixture”, Dokl. Ross. Akad. Nauk Mat. Inform. Protsessy Upr., 501 (2021), 31–37  mathnet  crossref  zmath; English transl. in Dokl. Math., 104:3 (2021), 340–346  crossref  mathscinet

Citation: A. A. Zlotnik, B. N. Chetverushkin, “Properties and errors of second-order parabolic and hyperbolic perturbations of a first-order symmetric hyperbolic system”, Mat. Sb., 214:4 (2023), 3–37; Sb. Math., 214:4 (2023), 444–478
Citation in format AMSBIB
\Bibitem{ZloChe23}
\by A.~A.~Zlotnik, B.~N.~Chetverushkin
\paper Properties and errors of second-order parabolic and hyperbolic perturbations of a~first-order symmetric hyperbolic system
\jour Mat. Sb.
\yr 2023
\vol 214
\issue 4
\pages 3--37
\mathnet{http://mi.mathnet.ru/sm9800}
\crossref{https://doi.org/10.4213/sm9800}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4653191}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023SbMat.214..444Z}
\transl
\jour Sb. Math.
\yr 2023
\vol 214
\issue 4
\pages 444--478
\crossref{https://doi.org/10.4213/sm9800e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001086876100001}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85176577744}
Linking options:
  • https://www.mathnet.ru/eng/sm9800
  • https://doi.org/10.4213/sm9800e
  • https://www.mathnet.ru/eng/sm/v214/i4/p3
  • This publication is cited in the following articles:
    1. Tuowei Chen, Yun Pu, “On a relaxation approximation of the 2-D incompressible magnetohydrodynamic equations”, Applicable Analysis, 2024, 1  crossref
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Математический сборник Sbornik: Mathematics
    Statistics & downloads:
    Abstract page:266
    Russian version PDF:12
    English version PDF:35
    Russian version HTML:125
    English version HTML:86
    References:20
    First page:11
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024