Next Article in Journal
Eigenvalue Problem for Discrete Jacobi–Sobolev Orthogonal Polynomials
Previous Article in Journal
Local Convergence for Multi-Step High Order Solvers under Weak Conditions
Previous Article in Special Issue
Nonlinear Multigrid Implementation for the Two-Dimensional Cahn–Hilliard Equation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strong Solutions of the Incompressible Navier–Stokes–Voigt Model

by
Evgenii S. Baranovskii
Department of Applied Mathematics, Informatics and Mechanics, Voronezh State University, 394018 Voronezh, Russia
Mathematics 2020, 8(2), 181; https://doi.org/10.3390/math8020181
Submission received: 30 December 2019 / Revised: 23 January 2020 / Accepted: 28 January 2020 / Published: 3 February 2020
(This article belongs to the Special Issue The Application of Mathematics to Physics and Nonlinear Science)

Abstract

:
This paper deals with an initial-boundary value problem for the Navier–Stokes–Voigt equations describing unsteady flows of an incompressible non-Newtonian fluid. We give the strong formulation of this problem as a nonlinear evolutionary equation in Sobolev spaces. Using the Faedo–Galerkin method with a special basis of eigenfunctions of the Stokes operator, we construct a global-in-time strong solution, which is unique in both two-dimensional and three-dimensional domains. We also study the long-time asymptotic behavior of the velocity field under the assumption that the external forces field is conservative.

1. Introduction

In this work, we study an initial-boundary value problem for the Navier–Stokes–Voigt (NSV) equations that model the unsteady flow of an incompressible viscoelastic fluid:
u t + ( u · ) u - ν Δ u - α 2 Δ u t + p = f in Ω × ( 0 , + ) , · u = 0 in Ω × ( 0 , + ) , u = 0 on Ω × ( 0 , + ) , u ( · , 0 ) = u 0 in Ω ,
where Ω denotes the bounded domain of flow in R n , n = 2 , 3 , with boundary Ω ; the vector function u represents the velocity field; p denotes the pressure; ν > 0 is the viscosity coefficient; α is a length scale parameter such that α 2 / ν is the relaxation time of the viscoelastic fluid; f is the external forces field; and u 0 is the initial velocity.
Note that when α = 0 the NSV system becomes the incompressible Navier–Stokes equations that describe Newtonian fluid flows. If α = 0 and ν = 0 , then we arrive at the incompressible Euler equations governing inviscid flows.
In the literature, the NSV equations are often called the Kelvin–Voigt equations or Oskolkov’s equations. The NSV model and related models of viscoelastic fluid flows have been studied extensively by different mathematicians over the past several decades starting from the pioneering papers by Oskolkov [1,2]. It should be mentioned at this point that Oskolkov later admitted [3] that these works contain some errors and not all obtained results hold. In this regard, Ladyzhenskaya remarked in her note [4] that the method of introduction of auxiliary viscosity used in [1,2] is incorrect under the no-slip boundary condition and explained the reasons for this. However, it is certain that the series of Oskolkov’s works played a major role in the study of the NSV equations and stimulated further research in this direction.
Let us shortly review available literature on mathematical analysis of NSV-type models. Sviridyuk [5] established the solvability of the weakly compressible NSV equations. In [6], the local-in-time unique solvability of problem (1) is proved. Korpusov and Sveshnikov [7] investigated the blowup of solutions to the NSV equations with a cubic source. Various slip problems are studied in the papers [8,9,10]. Kaya and Celebi [11] proved the existence and uniqueness of weak solutions of the so-called g-Kelvin–Voigt equations that describe viscoelastic fluid flows in thin domains. The solvability of the inhomogeneous Dirichlet problem for the equations governing a polymer fluid flow is proved in [12]. Berselli and Spirito [13] showed that weak solutions to the Navier–Stokes equations obtained as limits α 0 + of solutions to the NSV model are “suitable weak solutions” [14] and satisfy the local energy inequality. Fedorov and Ivanova [15] dealt with an inverse problem for the NSV equations. An algorithm for finding of numerical solution of an optimal control problem for the two-dimensional Kelvin–Voigt fluid flow was proposed by Plekhanova et al. [16]. Antontsev and Khompysh [17] established the existence and uniqueness of the global and local weak solutions to the NSV equations with p-Laplacian and a damping term. Artemov and Baranovskii [18] proved the existence of weak solutions to the coupled system of nonlinear equations describing the heat transfer in steady-state flows of a polymeric fluid. Mohan [19] investigated the global solvability, the asymptotic behavior, and some control problems for the NSV model with “fading memory” and “memory of length τ ”.
Most of the papers mentioned above deal with the study of weak (generalized) solutions to the NSV equations in the framework of the Hilbert space techniques. Therefore, it is a relevant question to prove the existence and uniqueness of strong solutions of system (1) in a Banach space under natural conditions on the data. Another important objective is to develop convenient algorithms for finding strong solutions or their approximations. Motivated by this, in the present work, we propose the strong formulation of problem (1) as a nonlinear evolutionary equation in suitable Banach spaces with the initial condition u ( 0 ) = u 0 . Using the Faedo–Galerkin procedure with a special basis of eigenfunctions of the Stokes operator and deriving various a priori estimates of approximate solutions in Sobolev’s spaces H 1 ( Ω ) and H 2 ( Ω ) , we construct a global-in-time strong solution of (1), which is unique in both two-dimensional and three-dimensional domains. We also derive the energy equality that holds for strong solutions. Moreover, it is shown that, if the external forces field f is conservative, then the H 1 -norm of the velocity field u decays exponentially as t + .

2. Preliminaries

To suggest the concept of a strong solution to problem (1), we introduce some notations, function spaces, and auxiliary results.
For vectors x , y R n and matrices X , Y R n × n by x · y and X : Y , we denote the scalar products, respectively:
x · y = def i = 1 n x i y i , X : Y = def i , j = 1 n X i j Y i j .
Let Ω R n be a bounded domain with sufficiently smooth boundary Ω . By D ( Ω ) denote the set of C functions with support contained in Ω . We use the standard notation for the Lebesgue spaces L s ( Ω ) , s 1 , as well as the Sobolev spaces H k ( Ω ) = def W k , 2 ( Ω ) , k N . When it comes to classes of R n -valued functions, we employ boldface letters, for instance,
D ( Ω ) = def D ( Ω ) n , L s ( Ω ) = def L s ( Ω ) n , H k ( Ω ) = def H k ( Ω ) n .
It is well known that the space Sobolev H 1 ( Ω ) is compactly embedded in L 4 ( Ω ) .
Let us introduce the following spaces:
V ( Ω ) = def { v D ( Ω ) : · v = 0 } , V 0 ( Ω ) = def the closure of the set V ( Ω ) in the space L 2 ( Ω ) , V 1 ( Ω ) = def the closure of the set V ( Ω ) in the space H 1 ( Ω ) , V 2 ( Ω ) = def H 2 ( Ω ) V 1 ( Ω ) .
It is obvious that V 0 ( Ω ) , V 1 ( Ω ) , and V 2 ( Ω ) are Hilbert spaces with the scalar products induced by L 2 ( Ω ) , H 1 ( Ω ) , and H 2 ( Ω ) , respectively. However, when studying problem (1), in the spaces V 1 ( Ω ) and V 2 ( Ω ) , it is more convenient to use the scalar products and the norms defined as follows:
( v , w ) V 1 ( Ω ) = def ( v , w ) L 2 ( Ω ) + α 2 ( v , w ) L 2 ( Ω ) , v V 1 ( Ω ) = def ( v , v ) V 1 ( Ω ) 1 / 2 ,
( v , w ) V 2 ( Ω ) = def ( P Δ v , P Δ w ) L 2 ( Ω ) , v V 2 ( Ω ) = def ( v , v ) V 2 ( Ω ) 1 / 2 .
Here, P : L 2 ( Ω ) V 0 ( Ω ) is the Leray projection, which corresponds the well-known Leray (or Hodge–Helmholtz) decomposition for the vector fields in L 2 ( Ω ) into a divergence-free part and a gradient part (see, e.g., [20], Chapter IV):
L 2 ( Ω ) = V 0 ( Ω ) G ( Ω ) ,
where the symbol ⊕ denotes the orthogonal sum and the subspace G ( Ω ) is defined as follows
G ( Ω ) = def { h : h H 1 ( Ω ) } .
Note that the norm · V i ( Ω ) is equivalent to the norm · H i ( Ω ) , i = 1 , 2 .
We introduce the equivalence relation on the space H 1 ( Ω ) by stating that φ ψ if φ - ψ = const . As usual, H 1 ( Ω ) / R denotes the quotient of H 1 ( Ω ) by R .
For a function ξ H 1 ( Ω ) , we set
ξ ¯ = def { ω H 1 ( Ω ) : ω ξ } H 1 ( Ω ) / R .
Let us define the gradient and the norm of ξ ¯ as follows
ξ ¯ = def ξ , ξ ¯ H 1 ( Ω ) / R = def ξ ¯ L 2 ( Ω ) .
Using Proposition 1.2 from ([21], Chapter I, § 1), it is easy to verify that the norm · H 1 ( Ω ) / R is well defined.
The following lemmas are needed for the sequel.
Lemma 1.
Suppose E is a Banach space and T is a positive number. A set K of the space C ( [ 0 , T ] ; E ) is relatively compact if and only if:
  • for any number t ( 0 , T ) , the set K ( t ) = def { w ( t ) : w K } is relatively compact in E ;
  • for any number ε > 0 , there exists a number η > 0 such that the inequality
    w ( t 1 ) - w ( t 2 ) E ε
    holds for any function w K and any numbers t 1 , t 2 [ 0 , T ] such that | t 1 - t 2 | η .
The proof of this lemma is given in [22].
Lemma 2.
The embedding C 1 ( [ 0 , T ] ; V 2 ( Ω ) ) C ( [ 0 , T ] ; V 1 ( Ω ) ) is completely continuous.
Proof. 
Let S be a bounded set of C 1 ( [ 0 , T ] ; V 2 ( Ω ) ) . Then
max ( w , t ) S × [ 0 , T ] w ( t ) V 2 ( Ω ) + max ( w , t ) S × [ 0 , T ] w ( t ) V 2 ( Ω ) r
with some constant r. Clearly, this implies that the set
S ( t ) = def { w ( t ) : w S }
is bounded in V 2 ( Ω ) for any t [ 0 , T ] .
From the Rellich–Kondrachov theorem (see, e.g., [23], Chapter 1, Theorem 1.12.1), it follows that the space V 2 ( Ω ) is compactly embedded into V 1 ( Ω ) . Therefore, the set S ( t ) is relatively compact in the space V 1 ( Ω ) .
By I denote the embedding operator from V 2 ( Ω ) into V 1 ( Ω ) . Taking into account inequality (5), we get the estimate
w ( t 1 ) - w ( t 2 ) V 1 ( Ω ) I L ( V 2 ( Ω ) , V 1 ( Ω ) ) w ( t 1 ) - w ( t 2 ) V 2 ( Ω ) I L ( V 2 ( Ω ) , V 1 ( Ω ) ) max τ [ t 1 , t 2 ] w ( τ ) V 2 ( Ω ) | t 1 - t 2 | r I L ( V 2 ( Ω ) , V 1 ( Ω ) ) | t 1 - t 2 | ε
for any function w S and for any numbers t 1 , t 2 [ 0 , T ] such that
| t 1 - t 2 | ε r I L ( V 2 ( Ω ) , V 1 ( Ω ) ) ,
where I L ( V 2 ( Ω ) , V 1 ( Ω ) ) is the operator norm of I .
Applying Lemma 1 with E = V 1 ( Ω ) , we conclude that the set S is relatively compact in the space C ( [ 0 , T ] ; V 1 ( Ω ) ) . Lemma 2 is proved. □
Lemma 3.
Let
L d = def { x R n : | x n | < d / 2 } , d Ω = def inf { d > 0 : Ω L d } .
Then, we have
4 d Ω 2 + 4 α 2 v L 2 ( Ω ) 2 + α 2 v L 2 ( Ω ) 2 v L 2 ( Ω ) 2 ,
for any v V 1 ( Ω ) .
Proof. 
The estimate (6) is a direct consequence of the Poincaré inequality (see, e.g., [24], Chapter II, Theorem II.5.1). □

3. Strong Formulation of Problem (1) and Main Results

Let us suppose that
f C ( [ 0 , + ) ; L 2 ( Ω ) ) , u 0 V 2 ( Ω ) .
Definition 1.
We say that a pair ( u , p ¯ ) is a strong solution to problem (1) if
u C 1 ( [ 0 , + ) ; V 2 ( Ω ) ) , p ¯ C ( [ 0 , + ) ; H 1 ( Ω ) / R )
and the following equalities are valid:
u ( t ) + ( u ( t ) · ) u ( t ) - ν Δ u ( t ) - α 2 Δ u ( t ) + p ¯ ( t ) = f ( t ) , t > 0 , u ( 0 ) = u 0 .
Remark 1.
Equation (8) with the initial condition u ( 0 ) = u 0 is a natural interpretation of the initial-boundary value problem (1) as an evolutionary equation in suitable function spaces. Note that, if a pair ( u * , p * ) is a classical solution to problem (1), then ( u * , p ¯ * ) satisfies Equation (8), i.e., this pair is a strong solution. On the other hand, if ( u , p ¯ ) is a strong solution and the functions u and p are sufficiently smooth in the usual sense, then ( u , p ) is a classical solution to (1).
We are now in a position to state our main results.
Theorem 1.
Assume that the boundary of the domain Ω belongs to the class C 2 and condition (7) holds. Then problem (1) has a unique strong solution ( u , p ¯ ) . This strong solution satisfies the energy equality
u ( t ) L 2 ( Ω ) 2 + 2 ν 0 t u ( τ ) L 2 ( Ω ) 2 d τ + α 2 u ( t ) L 2 ( Ω ) 2 = u 0 L 2 ( Ω ) 2 + α 2 u 0 L 2 ( Ω ) 2 + 2 0 t Ω f ( τ ) · u ( τ ) dx d τ , t 0 .
If there exists a function q C ( [ 0 , T ] ; H 1 ( Ω ) ) such that q = f , then
u ( t ) L 2 ( Ω ) 2 + α 2 u ( t ) L 2 ( Ω ) 2 exp - 8 ν t d Ω 2 + 4 α 2 u 0 L 2 ( Ω ) 2 + α 2 u 0 L 2 ( Ω ) 2 , t 0 ,
where the positive constant d Ω is defined in Lemma 3.

4. Proof of Theorem 1

To prove the existence of a strong solution to problem (1), we use the Faedo–Galerkin method with a special basis of eigenfunctions of the Stokes operator
A : V 2 ( Ω ) V 0 ( Ω ) , A w = def - P Δ w .
This linear operator is invertible and A - 1 is self-adjoint and compact as a map from V 0 ( Ω ) into V 0 ( Ω ) . From the spectral theorem for self-adjoint compact operators (see, e.g., [25], Chapter 10, Theorem 10.12), it follows that there exist sequences { w j } j = 1 V 2 ( Ω ) and { λ j } j = 1 ( 0 , + ) such that
A w j = λ j w j , j { 1 , 2 , } ,
and { w j } j = 1 is an orthonormal basis of the space V 0 ( Ω ) .
Let
w ˜ j = def λ j - 1 w j , j { 1 , 2 , } .
It is easily shown that { w ˜ j } j = 1 is an orthonormal basis in the space V 2 ( Ω ) .
Let us fix an arbitrary number T > 0 . For each fixed integer m 1 , we would like to define the approximate solution as follows:
v m ( t ) = def i = 1 m g m i ( t ) w i , t [ 0 , T ] ,
where g m 1 , , g m m are unknown functions such that
Ω v m ( t ) · w j dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · w j dx - ν Ω Δ v m ( t ) · w j dx - α 2 Ω Δ v m ( t ) · w j dx = Ω f ( t ) · w j dx , t ( 0 , T ) , j = 1 , , m , v m ( 0 ) = i = 1 m ( u 0 , w ˜ i ) V 2 ( Ω ) w ˜ i .
Let us define the matrix Q m R m × m and the vector a m R m by the rules:
Q m i j = def Ω w i · w j dx - α 2 Ω Δ w i · w j dx , i , j = 1 , , m , a m i = def λ i - 2 ( u 0 , w i ) V 2 ( Ω ) , i = 1 , , m .
Then, system (12) can be rewritten in the form
Q m g m ( t ) = F m ( t , g m ( t ) ) , t ( 0 , T ) , g m ( 0 ) = a m ,
where F m : [ 0 , T ] × R m R m is a known nonlinear vector function and g m = def ( g m 1 , , g m m ) .
Using integration by parts, we obtain
Q m i j = ( w i , w j ) V 1 ( Ω ) , i , j = 1 , , m .
Therefore, the matrix Q m is symmetric and invertible.
Applying Q m - 1 to the first equation of problem (13), we obviously get
g m ( t ) = Q m - 1 F m ( t , g m ( t ) ) , t ( 0 , T ) , g m ( 0 ) = a m .
The local existence of g m on an interval [ 0 , T m ] is insured by the Cauchy–Peano theorem. Thus, we have a local solution v m of problem (12) on [ 0 , T m ] . Below, we obtain a priori estimates (independent of m) for vector function v m , which entail that T m = T .
Let us assume that v m satisfies system (12). We multiply the jth equation of (12) by g m j ( t ) and sum with respect to j from 1 to m. Since
i = 1 n Ω v m i ( t ) v m ( t ) x i · v m ( t ) dx = 1 2 i = 1 n Ω v m i ( t ) x i | v m ( t ) | 2 dx = - 1 2 i = 1 n Ω v m i ( t ) x i | v m ( t ) | 2 dx = - 1 2 Ω · v m ( t ) = 0 | v m ( t ) | 2 dx = 0 , t ( 0 , T ) ,
we get
Ω v m ( t ) · v m ( t ) dx - ν Ω Δ v m ( t ) · v m ( t ) dx - α 2 Ω Δ v m ( t ) · v m ( t ) dx = Ω f ( t ) · v m ( t ) dx .
Integrating by parts the second and third terms on the left-hand side of equality (14), we arrive at the following relation
Ω v m ( t ) · v m ( t ) dx + ν Ω | v m ( t ) | 2 dx + α 2 Ω v m ( t ) : v m ( t ) dx = Ω f ( t ) · v m ( t ) dx ,
which, in turn, gives
1 2 d d τ v m ( τ ) L 2 ( Ω ) 2 + ν v m ( τ ) L 2 ( Ω ) 2 + α 2 2 d d τ v m ( τ ) L 2 ( Ω ) 2 = Ω f ( τ ) · v m ( τ ) dx ,
for any τ [ 0 , T ] . Further, we multiply the last equality by 2 and integrate from 0 to t with respect to τ ; this yields
v m ( t ) L 2 ( Ω ) 2 + 2 ν 0 t v m ( τ ) L 2 ( Ω ) 2 d τ + α 2 v m ( t ) L 2 ( Ω ) 2 = v m ( 0 ) L 2 ( Ω ) 2 + α 2 v m ( 0 ) L 2 ( Ω ) 2 + 2 0 t Ω f ( τ ) · v m ( τ ) dx d τ .
Taking into account (3) and (4), we easily derive from equality (15) that
v m ( t ) V 1 ( Ω ) 2 v m ( 0 ) V 1 ( Ω ) 2 + 2 0 t Ω f ( τ ) · v m ( τ ) dx d τ v m ( 0 ) V 1 ( Ω ) 2 + 0 t Ω | f ( τ ) | 2 dx d τ + 0 t Ω | v m ( τ ) | 2 dx d τ C 1 u 0 V 2 ( Ω ) 2 + 0 T f ( τ ) L 2 ( Ω ) 2 d τ + C 2 0 t v m ( τ ) V 1 ( Ω ) 2 d τ .
Here and in the succeeding discussion, the symbols C i , i = 1 , 2 , , designate positive constants that are independent of m. Using Grönwall’s inequality, we get
v m ( t ) V 1 ( Ω ) 2 C 1 u 0 V 2 ( Ω ) 2 + 0 T f ( τ ) L 2 ( Ω ) 2 d τ exp ( C 2 t ) , t ( 0 , T ) .
Hence,
v m C ( [ 0 , T ] ; V 1 ( Ω ) ) = max t [ 0 , T ] v m ( t ) V 1 ( Ω ) C 1 u 0 V 2 ( Ω ) 2 + 0 T f ( τ ) L 2 ( Ω ) 2 d τ 1 / 2 exp ( C 2 T ) 1 / 2 .
Next, by multiplying the jth equation of (12) with g m j and summing over j = 1 , , m , we obtain
Ω | v m ( t ) | 2 dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · v m ( t ) dx - ν Ω Δ v m ( t ) · v m ( t ) dx - α 2 Ω Δ v m ( t ) · v m ( t ) dx = Ω f ( t ) · v m ( t ) dx , t ( 0 , T ) .
Integrating by parts the third and fourth terms on the left-hand side of the last equality, we arrive at
Ω | v m ( t ) | 2 dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · v m ( t ) dx + ν Ω v m ( t ) : v m ( t ) dx + α 2 Ω | v m ( t ) | 2 dx = Ω f ( t ) · v m ( t ) dx , t ( 0 , T ) .
From here, using (3) and Hölder’s inequality, one can obtain
v m ( t ) V 1 ( Ω ) 2 = - i = 1 n Ω v m i ( t ) v m ( t ) x i · v m ( t ) dx - ν Ω v m ( t ) : v m ( t ) dx + Ω f ( t ) · v m ( t ) dx i , j = 1 n v m i ( t ) L 4 ( Ω ) v m j ( t ) x i L 2 ( Ω ) v m j ( t ) L 4 ( Ω ) + ν v m ( t ) L 2 ( Ω ) v m ( t ) L 2 ( Ω ) + f ( t ) L 2 ( Ω ) v m ( t ) L 2 ( Ω ) C 3 v m ( t ) V 1 ( Ω ) 2 + f ( t ) L 2 ( Ω ) v m ( t ) V 1 ( Ω ) ,
whence
v m ( t ) V 1 ( Ω ) C 3 v m ( t ) V 1 ( Ω ) 2 + f ( t ) L 2 ( Ω ) , t ( 0 , T ) .
With the help of inequality (16), we get
v m ( t ) V 1 ( Ω ) C 3 C 1 u 0 V 2 ( Ω ) 2 + 0 T f ( τ ) L 2 ( Ω ) 2 d τ exp ( C 2 t ) + C 3 f ( t ) L 2 ( Ω ) ,
for all t ( 0 , T ) . Therefore, we have
v m C ( [ 0 , T ] ; V 1 ( Ω ) ) = max t [ 0 , T ] v m ( t ) V 1 ( Ω ) C 3 C 1 u 0 V 2 ( Ω ) 2 + 0 T f ( τ ) L 2 ( Ω ) 2 d τ exp ( C 2 T ) + C 3 max t [ 0 , T ] f ( t ) L 2 ( Ω ) .
We now multiply the jth equation of (12) by - λ j g m j ( t ) and sum with respect to j from 1 to m. Taking into account equality (11), we get
Ω v m ( t ) · P Δ v m ( t ) dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · P Δ v m ( t ) dx - ν Ω Δ v m ( t ) · P Δ v m ( t ) dx - α 2 Ω Δ v m ( t ) · P Δ v m ( t ) dx = Ω f ( t ) · P Δ v m ( t ) dx , t ( 0 , T ) ,
which leads to
Ω v m ( t ) · P Δ v m ( t ) dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · P Δ v m ( t ) dx - ν Ω | P Δ v m ( t ) | 2 dx - α 2 Ω P Δ v m ( t ) · P Δ v m ( t ) dx = Ω f ( t ) · P Δ v m ( t ) dx , t ( 0 , T ) .
From this equality, with the help of Hölder’s and Young’s inequalities, we derive
ν P Δ v m ( t ) L 2 ( Ω ) 2 + α 2 2 d d t P Δ v m ( t ) L 2 ( Ω ) 2 = Ω v m ( t ) · P Δ v m ( t ) dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · P Δ v m ( t ) dx - Ω f ( t ) · P Δ v m ( t ) dx v m ( t ) L 2 ( Ω ) + i = 1 n v m i ( t ) L 4 ( Ω ) v m ( t ) x i L 4 ( Ω ) + f ( t ) L 2 ( Ω ) P Δ v m ( t ) L 2 ( Ω ) 1 2 ν v m ( t ) L 2 ( Ω ) + i = 1 n v m i ( t ) L 4 ( Ω ) v m ( t ) x i L 4 ( Ω ) + f ( t ) L 2 ( Ω ) 2 + ν 2 P Δ v m ( t ) L 2 ( Ω ) 2 , t ( 0 , T ) .
Therefore, the following inequality holds
ν P Δ v m ( t ) L 2 ( Ω ) 2 + α 2 d d t P Δ v m ( t ) L 2 ( Ω ) 2 1 ν v m ( t ) L 2 ( Ω ) + i = 1 n v m i ( t ) L 4 ( Ω ) v m ( t ) x i L 4 ( Ω ) + f ( t ) L 2 ( Ω ) 2 , t ( 0 , T ) ,
and, using estimates (17) and (18), we deduce that
ν v m ( τ ) V 2 ( Ω ) 2 + α 2 d d τ v m ( τ ) V 2 ( Ω ) 2 C 4 + C 5 v m ( τ ) V 2 ( Ω ) 2 , τ ( 0 , T ) .
Integrating both sides of this differential inequality with respect to τ from 0 to t, we deduce
ν 0 t v m ( τ ) V 2 ( Ω ) 2 d τ + α 2 v m ( t ) V 2 ( Ω ) 2 α 2 v m ( 0 ) V 2 ( Ω ) 2 + C 4 t + C 5 0 t v m ( τ ) V 2 ( Ω ) 2 d τ α 2 u 0 V 2 ( Ω ) 2 + C 4 T + C 5 0 t v m ( τ ) V 2 ( Ω ) 2 d τ .
It follows easily that
v m ( t ) V 2 ( Ω ) 2 u 0 V 2 ( Ω ) 2 + C 4 α - 2 T + C 5 α - 2 0 t v m ( τ ) V 2 ( Ω ) 2 d τ , t ( 0 , T ) .
Applying Grönwall’s inequality, we obtain
v m ( t ) V 2 ( Ω ) 2 u 0 V 2 ( Ω ) 2 + C 4 α - 2 T exp ( C 5 α - 2 t ) , t ( 0 , T ) .
This implies that
v m C ( [ 0 , T ] ; V 2 ( Ω ) ) = max t [ 0 , T ] v m ( t ) V 2 ( Ω ) u 0 V 2 ( Ω ) 2 + C 4 α - 2 T exp ( C 5 α - 2 T ) 1 / 2 .
Finally, we multiply the jth equation of (12) by - λ j g m j ( t ) and sum with respect to j from 1 to m. Bearing in mind equality (11), we obtain
Ω v m ( t ) · P Δ v m ( t ) dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · P Δ v m ( t ) dx - ν Ω Δ v m ( t ) · P Δ v m ( t ) dx - α 2 Ω Δ v m ( t ) · P Δ v m ( t ) dx = Ω f ( t ) · P Δ v m ( t ) dx , t ( 0 , T ) .
Using Hölder’s inequality, from the last equality one can derive
α 2 P Δ v m ( t ) L 2 ( Ω ) 2 = Ω v m ( t ) · P Δ v m ( t ) dx + i = 1 n Ω v m i ( t ) v m ( t ) x i · P Δ v m ( t ) dx - ν Ω P Δ v m ( t ) · P Δ v m ( t ) dx - Ω f ( t ) · P Δ v m ( t ) dx ( v m ( t ) L 2 ( Ω ) + i = 1 n v m i ( t ) L 4 ( Ω ) v m ( t ) x i L 4 ( Ω ) + ν P Δ v m ( t ) L 2 ( Ω ) + f ( t ) L 2 ( Ω ) ) P Δ v m ( t ) L 2 ( Ω ) ( v m ( t ) L 2 ( Ω ) + C 6 v m ( t ) V 1 ( Ω ) v m ( t ) V 2 ( Ω ) + ν v m ( t ) V 2 ( Ω ) + f ( t ) L 2 ( Ω ) ) P Δ v m ( t ) L 2 ( Ω ) , t ( 0 , T ) .
Clearly, this yields the estimate
P Δ v m ( t ) L 2 ( Ω ) α - 2 ( v m ( t ) L 2 ( Ω ) + C 6 v m ( t ) V 1 ( Ω ) v m ( t ) V 2 ( Ω ) + ν v m ( t ) V 2 ( Ω ) + f ( t ) L 2 ( Ω ) ) , t ( 0 , T ) .
Taking into account (17)–(19), from the last inequality, we easily obtain that
v m ( t ) V 2 ( Ω ) = P Δ v m ( t ) L 2 ( Ω ) C 7 , t ( 0 , T ) ,
and, hence,
v m C ( [ 0 , T ] ; V 2 ( Ω ) ) = max t [ 0 , T ] v m ( t ) V 2 ( Ω ) C 7 .
From estimates (19) and (20) and Lemma 2, it follows that there exist a subsequence { m k } k = 1 and a function u such that v m k converges strongly to u in the space C ( [ 0 , T ] ; V 1 ( Ω ) ) as k . Without loss of generality, we can assume that
v m u strongly in C ( [ 0 , T ] ; V 1 ( Ω ) ) as m ,
v m u weakly in L 2 ( 0 , T ; V 2 ( Ω ) ) as m .
Moreover, we have
v m ( 0 ) u 0 strongly in V 2 ( Ω ) as m .
On the other hand, from (21) it follows that
v m ( 0 ) u ( 0 ) strongly in V 1 ( Ω ) as m .
Comparing the convergence results (23) and (24), we obtain
u ( 0 ) = u 0 .
Integrating the jth equation of (12) from 0 to s, we obtain
Ω v m ( s ) · w j dx + i = 1 n 0 s Ω v m i ( t ) v m ( t ) x i · w j dx d t - ν 0 s Ω Δ v m ( t ) · w j dx d t - α 2 Ω Δ v m ( s ) · w j dx = Ω v m ( 0 ) · w j dx - α 2 Ω Δ v m ( 0 ) · w j dx + 0 s Ω f ( t ) · w j dx d t , j { 1 , 2 , } , s [ 0 , T ] .
Integrating by parts the third and fourth terms on the left-hand side of this equality, we arrive at
Ω v m ( s ) · w j dx + i = 1 n 0 s Ω v m i ( t ) v m ( t ) x i · w j dx d t + ν 0 s Ω v m ( t ) : w j dx d t + α 2 Ω v m ( s ) : w j dx = Ω v m ( 0 ) · w j dx - α 2 Ω Δ v m ( 0 ) · w j dx + 0 s Ω f ( t ) · w j dx d t .
Using the convergence results (21)–(23), we can pass to the limit m in the last equality and obtain
Ω u ( s ) · w j dx + i = 1 n 0 s Ω u i ( t ) u ( t ) x i · w j dx d t + ν 0 s Ω u ( t ) : w j dx d t + α 2 Ω u ( s ) : w j dx = Ω u 0 · w j dx - α 2 Ω Δ u 0 · w j dx + 0 s Ω f ( t ) · w j dx d t , j { 1 , 2 , } , s [ 0 , T ] .
Applying integration by parts again, we get
Ω u ( s ) · w j dx + i = 1 n 0 s Ω u i ( t ) u ( t ) x i · w j dx d t - ν 0 s Ω Δ u ( t ) · w j dx d t - α 2 Ω Δ u ( s ) · w j dx = Ω u 0 · w j dx - α 2 Ω Δ u 0 · w j dx + 0 s Ω f ( t ) · w j dx d t , j { 1 , 2 , } , s [ 0 , T ] .
Because { w j } j = 1 is a basis of V 0 ( Ω ) , equality (26) remains valid if we replace w j with an arbitrary vector function w from the space V 0 ( Ω ) , that is
Ω u ( s ) · w dx + i = 1 n 0 s Ω u i ( t ) u ( t ) x i · w dx d t - ν 0 s Ω Δ u ( t ) · w dx d t - α 2 Ω Δ u ( s ) · w dx = Ω u 0 · w dx - α 2 Ω Δ u 0 · w dx + 0 s Ω f ( t ) · w dx d t , s [ 0 , T ] .
From the last equality it follows that
u ( s ) + i = 1 n P 0 s u i ( t ) u ( t ) x i d t - ν P 0 s Δ u ( t ) d t - α 2 P Δ u ( s ) = u 0 - α 2 P Δ u 0 + P 0 s f ( t ) d t .
Using the Stokes operator A , we can rewrite this equality as follows
( I + α 2 A ) u ( s ) = - i = 1 n P 0 s u i ( t ) u ( t ) x i d t + ν P 0 s Δ u ( t ) d t + ( I + α 2 A ) u 0 + P 0 s f ( t ) d t , s [ 0 , T ] ,
where I : V 2 ( Ω ) V 0 ( Ω ) is the embedding operator.
Applying the operator ( I + α 2 A ) - 1 : V 0 ( Ω ) V 2 ( Ω ) to both sides of equality (28), we get
u ( s ) = - i = 1 n ( I + α 2 A ) - 1 P 0 s u i ( t ) u ( t ) x i d t + ν ( I + α 2 A ) - 1 P 0 s Δ u ( t ) d t + u 0 + ( I + α 2 A ) - 1 P 0 s f ( t ) d t , s [ 0 , T ] .
Since
u C ( [ 0 , T ] ; V 1 ( Ω ) ) L 2 ( 0 , T ; V 2 ( Ω ) ) ,
we conclude from (29) that
u C ( [ 0 , T ] ; V 2 ( Ω ) ) .
Next, differentiating both sides of (29) with respect to s, we get
u ( s ) = - i = 1 n ( I + α 2 A ) - 1 P u i ( s ) u ( s ) x i + ν ( I + α 2 A ) - 1 P Δ u ( s ) + ( I + α 2 A ) - 1 P f ( s ) , s [ 0 , T ] .
Taking into account (30), from the last equality we deduce that u C ( [ 0 , T ] ; V 2 ( Ω ) ) . Hence,
u C 1 ( [ 0 , T ] ; V 2 ( Ω ) ) .
Next, from equality (27) it follows that there exists an element π ¯ ( t ) H 1 ( Ω ) / R such that
u ( t ) + i = 1 n 0 t u i ( τ ) u ( τ ) x i d τ - ν 0 t Δ u ( τ ) d τ - α 2 Δ u ( t ) - u 0 - α 2 Δ u 0 - 0 t f ( τ ) d τ = π ¯ ( t ) .
It is readily seen that π ¯ C 1 ( [ 0 , T ] ; g ( Ω ) ) and, consequently, we have
π ¯ C 1 ( [ 0 , T ] ; H 1 ( Ω ) / R ) .
Letting p ¯ ( t ) = def - π ¯ ( t ) , from (33) we get
p ¯ C ( [ 0 , T ] ; H 1 ( Ω ) / R ) .
Finally, differentiating both sides of (32) with respect to t, we arrive at
u ( t ) + i = 1 n u i ( t ) u ( t ) x i - ν Δ u ( t ) - α 2 Δ u ( t ) + p ¯ ( t ) = f ( t ) , t ( 0 , T ) .
Bearing in mind (25), (31), (34), and (35), we conclude that the pair ( u , p ¯ ) is a strong solution to problem (1) on the interval [ 0 , T ] . The uniqueness of a strong solution can be proved by using arguments similar to those that are presented in [9], thus we choose to omit the details of the corresponding proof. Since T is arbitrary, we see that ( u , p ¯ ) is a solution of (1) in the sense of Definition 1.
Next, we take the L 2 -scalar product of (8) with the vector function u . Using integration by parts, one can easily arrive at the energy equality (9).
The rest of the proof consists in proving inequality (10). If there exists a function q from the space C ( [ 0 , T ] ; H 1 ( Ω ) ) such that q = f , then we have
Ω f ( τ ) · u ( τ ) dx = Ω q ( τ ) · u ( τ ) dx = - Ω q ( τ ) · u ( τ ) = 0 dx = 0 , τ 0 ,
i. e., the total work done by external forces f is zero.
In view of (36), the energy equality (9) reduces to
u ( t ) L 2 ( Ω ) 2 + 2 ν 0 t u ( τ ) L 2 ( Ω ) 2 d τ + α 2 u ( t ) L 2 ( Ω ) 2 = u 0 L 2 ( Ω ) 2 + α 2 u 0 L 2 ( Ω ) 2 , t 0 .
Differentiating the last equality with respect to t, we get
d d t u ( t ) L 2 ( Ω ) 2 + α 2 u ( t ) L 2 ( Ω ) 2 + 2 ν u ( t ) L 2 ( Ω ) 2 = 0 , t 0 .
Using inequality (6), we obtain
d d t u ( t ) L 2 ( Ω ) 2 + α 2 u ( t ) L 2 ( Ω ) 2 + 8 ν d Ω 2 + 4 α 2 u ( t ) L 2 ( Ω ) 2 + α 2 u ( t ) L 2 ( Ω ) 2 0 , t 0
and, hence,
d d t exp 8 ν t d Ω 2 + 4 α 2 u ( t ) L 2 ( Ω ) 2 + α 2 u ( t ) L 2 ( Ω ) 2 0 , t 0 .
Then, by integrating (37) with respect to t, we derive inequality (10). Thus, the proof of Theorem 1 is complete.

5. Concluding Remarks

In this paper, we prove the existence and uniqueness of a strong solution to the incompressible Navier–Stokes–Voigt model. The construction of a strong solution proceeds via the Faedo–Galerkin procedure with a special basis of eigenfunctions of the Stokes operator. Note that this approach allows easily obtaining approximations of strong solutions, which frequently reduce to approximate analytic or semi-analytic solutions when the flow domain has a simple symmetric shape. Such solutions favor a better understanding of the qualitative features of unsteady flows of viscoelastic fluids and can be used to test the relevant numerical, asymptotic, and approximate methods.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

Symbols and NotationsMeaning
Ω flow domain
Ω boundary of Ω
x 1 , , x n space variables
ttime
u velocity field
u 0 initial velocity field
ppressure
ν viscosity coefficient
α relaxation coefficient
f external forces field
qscalar potential for f
Tfixed point in time
x 1 , , x n
Δ i = 1 n 2 x i 2
differentiation with respect to t
orthogonal sum of subspaces
weak convergence
strong convergence
embedding
A × B Cartesian product of two sets A and B
x · y scalar product of vectors x , y R n
X : Y scalar product of matrices X , Y R n × n
( v , w ) H scalar product of functions v and w from a Hilbert space  H
v E norm of function v from a Banach space  E
L ( E 1 , E 2 ) space of all bounded linear mappings from E 1 to E 2
D ( Ω ) space of C functions with support contained in Ω
V ( Ω ) space of C divergence-free vector functions with support contained in Ω
L s ( Ω ) Lebesgue space
H k ( Ω ) Sobolev space
V i ( Ω ) special Hilbert space defined by (2) for i { 0 , 1 , 2 }
G ( Ω ) { h : h H 1 ( Ω ) }
H 1 ( Ω ) / R quotient of H 1 ( Ω ) by R
equivalence relation on H 1 ( Ω )
L d layer with thickness d
d Ω inf { d : Ω L d }
ξ ¯ { ω H 1 ( Ω ) : ω ξ }
I embedding operator
P Leray projection
A Stokes operator
λ j eigenvalue of Stokes operator
w j eigenfunction of Stokes operator
v m Galerkin solution
C i positive constant independent of m

References

  1. Oskolkov, A.P. Solvability in the large of the first boundary value problem for a certain quasilinear third order system that is encountered in the study of the motion of a viscous fluid. Zap. Nauchn. Semin. LOMI 1972, 27, 145–160. [Google Scholar]
  2. Oskolkov, A.P. On the uniqueness and solvability in the large of the boundary-value problems for the equations of motion of aqueous solutions of polymers. Zap. Nauchn. Semin. LOMI 1973, 38, 98–136. [Google Scholar]
  3. Oskolkov, A.P. Some quasilinear systems that arise in the study of the motion of viscous fluids. Zap. Nauchn. Semin. LOMI 1975, 52, 128–157. [Google Scholar]
  4. Ladyzhenskaya, O.A. On some gaps in two of my papers on the Navier–Stokes equations and the way of closing them. J. Math. Sci. 2003, 115, 2789–2791. [Google Scholar] [CrossRef]
  5. Sviridyuk, G.A. On a model of the dynamics of a weakly compressible viscoelastic fluid. Russian Math. (Iz. VUZ) 1994, 38, 59–68. [Google Scholar]
  6. Sviridyuk, G.A.; Sukacheva, T.G. On the solvability of a nonstationary problem describing the dynamics of an incompressible viscoelastic fluid. Math. Notes 1998, 63, 388–395. [Google Scholar] [CrossRef]
  7. Korpusov, M.O.; Sveshnikov, A.G. Blow-up of Oskolkov’s system of equations. Sb. Math. 2009, 200, 549–572. [Google Scholar] [CrossRef]
  8. Ladyzhenskaya, O.A. On the global unique solvability of some two-dimensional problems for the water solutions of polymers. J. Math. Sci. 2000, 99, 888–897. [Google Scholar] [CrossRef]
  9. Baranovskii, E.S. Mixed initial–boundary value problem for equations of motion of Kelvin–Voigt fluids. Comput. Math. Math. Phys. 2016, 56, 1363–1371. [Google Scholar] [CrossRef]
  10. Baranovskii, E.S. Global solutions for a model of polymeric flows with wall slip. Math. Meth. Appl. Sci. 2017, 40, 5035–5043. [Google Scholar] [CrossRef]
  11. Kaya, M.; Celebi, A.O. Existence of weak solutions of the g-Kelvin–Voight equation. Math. Comput. Model. 2009, 49, 497–504. [Google Scholar] [CrossRef]
  12. Baranovskii, E.S. Flows of a polymer fluid in domain with impermeable boundaries. Comput. Math. Math. Phys. 2014, 54, 1589–1596. [Google Scholar] [CrossRef]
  13. Berselli, L.C.; Spirito, S. Suitable weak solutions to the 3D Navier–Stokes equations are constructed with the Voigt approximation. J. Differ. Equ. 2017, 262, 3285–3316. [Google Scholar] [CrossRef] [Green Version]
  14. Caffarelli, L.; Kohn, R.; Nirenberg, L. Partial regularity of suitable weak solutions of the Navier–Stokes equations. Comm. Pure Appl. Math. 1982, 35, 771–831. [Google Scholar] [CrossRef]
  15. Fedorov, V.E.; Ivanova, N.D. Inverse problem for Oskolkov’s system of equations. Math. Meth. Appl. Sci. 2017, 40, 6123–6126. [Google Scholar] [CrossRef]
  16. Plekhanova, M.V.; Baybulatova, G.D.; Davydov, P.N. Numerical solution of an optimal control problem for Oskolkov’s system. Math. Meth. Appl. Sci. 2018, 41, 9071–9080. [Google Scholar] [CrossRef]
  17. Antontsev, S.N.; Khompysh, K. Kelvin–Voight equation with p-Laplacian and damping term: Existence, uniqueness and blow-up. J. Math. Anal. Appl. 2017, 446, 1255–1273. [Google Scholar] [CrossRef]
  18. Artemov, M.A.; Baranovskii, E.S. Solvability of the Boussinesq approximation for water polymer solutions. Mathematics 2019, 7, 611. [Google Scholar] [CrossRef] [Green Version]
  19. Mohan, M.T. On the three dimensional Kelvin–Voigt fluids: Global solvability, exponential stability and exact controllability of Galerkin approximations. Evol. Equ. Control Theory 2019. [Google Scholar] [CrossRef] [Green Version]
  20. Boyer, F.; Fabrie, P. Mathematical Tools for the Study of the Incompressible Navier–Stokes Equations and Related Models; Springer: New York, NY, USA, 2013. [Google Scholar] [CrossRef]
  21. Temam, R. Navier–Stokes Equations—Theory and Numerical Analysis; North-Holland Publishing Co.: Amsterdam, The Netherlands, 1977. [Google Scholar]
  22. Simon, J. Compact sets in the space Lp(0,T;B). Ann. Mat. Pura Appl. 1986, 146, 65–96. [Google Scholar] [CrossRef]
  23. Agranovich, M.S. Sobolev Spaces, Their Generalizations, and Elliptic Problems in Smooth and Lipschitz Domains; Springer: Cham, Switzerland, 2015. [Google Scholar] [CrossRef]
  24. Galdi, G.P. An Introduction to the Mathematical Theory of the Navier–Stokes Equations—Steady-State Problems; Springer: New York, NY, USA, 2011. [Google Scholar] [CrossRef]
  25. Siddiqi, A.H. Functional Analysis and Applications; Springer: Singapore, 2018. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Baranovskii, E.S. Strong Solutions of the Incompressible Navier–Stokes–Voigt Model. Mathematics 2020, 8, 181. https://doi.org/10.3390/math8020181

AMA Style

Baranovskii ES. Strong Solutions of the Incompressible Navier–Stokes–Voigt Model. Mathematics. 2020; 8(2):181. https://doi.org/10.3390/math8020181

Chicago/Turabian Style

Baranovskii, Evgenii S. 2020. "Strong Solutions of the Incompressible Navier–Stokes–Voigt Model" Mathematics 8, no. 2: 181. https://doi.org/10.3390/math8020181

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop