Spatially Periodic Solutions for Evolution Anisotropic Variable-Coefficient Navier–Stokes Equations: II. Serrin-Type Solutions
Funding: The research has been funded by the general research grant from UK Research and Innovation (UKRI) to Brunel University of London.
ABSTRACT
We consider evolution (nonstationary) space-periodic solutions to the -dimensional nonlinear Navier–Stokes equations of anisotropic fluids with the viscosity coefficient tensor variable in space and time and satisfying the relaxed ellipticity condition. Employing the Galerkin algorithm, we prove the existence of Serrin-type solutions, that is, the weak solutions with velocity in the periodic space , . The solution uniqueness and regularity results are also discussed.
1 Introduction
Analysis of Stokes and Navier–Stokes equations is an established and active field of research in applied mathematical analysis; see, for example, [1-11] and many other publications. These works were mainly devoted to the flows of isotropic fluids with constant-viscosity coefficient, and some of the employed methods were heavily based on these properties.
On the other hand, in many cases, the fluid viscosity can vary in time and spatial coordinates, for example, due to variable ambient temperature. Moreover, from the point of view of rational mechanics of continuum, fluids can be anisotropic, and this feature is indeed observed in liquid crystals, electromagnetic fluids, and so forth; see, for example, [12] and references therein. In [13-18], the classical Navier–Stokes equations analysis has been extended to the transmission and boundary-value problems for stationaryStokes and Navier–Stokes equations of anisotropic fluids, particularly with relaxed ellipticity condition on the viscosity tensor.
In Part I, [19], we considered evolution (nonstationary) spatially periodic solutions in , , to the Navier–Stokes equations of an anisotropic fluid with the viscosity coefficient tensor variable in spatial coordinates and time and satisfying the relaxed ellipticity condition. We implemented the Galerkin algorithm but unlike the traditional approach, for example, in [10, 11], where the Galerkin basis consisted of the eigenfunctions of the corresponding isotropic constant-coefficient Stokes operator, we employed the basis constituted by the eigenfunctions of the periodic Bessel-potential operator having an advantage that it is universal, that is, independent of the analyzed anisotropic variable-coefficient Navier–Stokes operator. To analyze the solution in higher dimensions, the definition of the weak solution was generalized to some extent. Then, the periodic weak solution existence was considered in the spaces of Banach-valued functions mapping a finite-time interval to periodic Sobolev (Bessel-potential) spaces on -dimensional flat torus, . The periodic setting is interesting on its own, modeling fluid flow in periodic composite structures, and is also a common element of homogenization theories for inhomogeneous fluids and in the large eddy simulation.
In this paper, Part II, we prove the existence, uniqueness, and regularity of the weak solutions that belongs to the space (we call them Serrin-type solutions). It is well known that the regularity results available at the moment for evolution Navier–Stokes equations are rather different for dimensions and , even for isotropic constant-viscosity fluids. The weak solution global regularity under arbitrarily large smooth input data for is proved and can be found, for example, in [1-4, 6-11]. However, for , it is still an open question and constitutes one of the Clay Institute famous Millennium problems. Our motivation for considering arbitrary is particularly to place the cases and in a more general set and to see which of them is an exception.
The paper material is presented as follows. In Section 1.1, we provide essentials on anisotropic Stokes and Navier–Stokes equations. Section 1.2 gives an introduction to the periodic Sobolev (Bessel-potential) functions spaces in spatial coordinates on -dimensional flat torus and to the corresponding Banach-valued functions mapping a finite-time interval to these periodic Sobolev spaces. In Section 2, we describe the existence results for evolution spatially periodic anisotropic Navier–Stokes problem available from Part I, [19]. Sections 3–5 contain the main results of the paper. In Section 3, we define the Serrin-type solutions and prove the energy equality for them and also their uniqueness, for the -dimensional periodic setting, . We also remark on their relations with the strong solutions and show that for , any weak solution is a Serrin-type solution. In Section 4, we analyze the Serrin-type solution existence and regularity for constant anisotropic viscosity coefficients, while in Section 5, we generalize these results to variable anisotropic viscosity coefficients. In Section 6, we collect some technical results used in the main text of the paper, several of which might be new.
1.1 Anisotropic Stokes and Navier–Stokes PDE Systems
1.2 Periodic Function Spaces
Let us introduce some function spaces on torus, that is, periodic function spaces (see, e.g., [22, p. 26], [23, 24], [25, Chapter 3], [7, Section 1.7.1], [10, Chapter 2] for more details).
Let be an integer and be the -dimensional flat torus that can be parametrized as the semiopen cube ; compare [26, p. 312]. In what follows, denotes the (test) space of infinitely smooth real or complex functions on the torus. As usual, denotes the set of natural numbers, the set of natural numbers augmented by 0, and the set of integers.
The space dual to , that is, the space of linear bounded functionals on , called the space of torus distributions, is denoted by and can be identified with the space of periodic distributions acting on .
For any , let be the Euclidean norm in and let us denote . Evidently, .
For , , we can write , where the Fourier series converges in the sense of norm (1.5). Moreover, because is an arbitrary distribution from , this also implies that the space is dense in for any (cf. [25, Exercise 3.2.9]).
There holds the compact embedding if , embeddings if , , and moreover, (cf. [25, Exercises 3.2.10 and 3.2.10, Corollary 3.2.11]). Note that the periodic norms on are equivalent to the corresponding standard (nonperiodic) Bessel-potential norms on as an -cubic domain; see, for example, [23, Section 13.8.1].
Denoting , then . The corresponding spaces of -component vector functions/distributions are denoted as , , and so forth.
The following assertion is produced in [19, Theorem 1].
Theorem 1.1.Let and .
- a.
The space has the Helmholtz–Weyl decomposition, , that is, any can be uniquely represented as , where and .
- b.
The spaces and are orthogonal subspaces of in the sense of inner product, that is, for any and .
- c.
The spaces and are orthogonal in the sense of dual product, that is, for any and .
- d.
There exist the bounded orthogonal projector operators and (the Leray projector), while for any .
Unless stated otherwise, we will assume in this paper that all vector and scalar variables are real valued (however, with complex-valued Fourier coefficients).
2 Existence Results Available for Evolution Spatially Periodic Anisotropic Navier–Stokes Problem
We use the following definition of weak solution given in [19, Definition 1].
Definition 2.1.Let , , , and . A function is called a weak solution of the evolution space-periodic anisotropic Navier–Stokes initial value problem (2.1-2.3) if it solves the initial-variational problem
The following assertion is proved in [19, Lemma 1].
Lemma 2.2.Let , , , and . Let solve Equation (2.11).
- i.
Then,
(2.13)whileIn addition, for a.e. and also in the distribution sense on .
- ii.
Moreover, is almost everywhere on equal to a function , and is also -weakly continuous in time on , that is, .
- iii.
There exists the associated pressure that for the given is the unique solution of Equation (2.1) in this space.
Remark 2.3.The initial condition (2.12) should be understood for the function redefined as the function that was introduced in Lemma 2.2(ii) and is -weakly continuous in time.
The following existence theorem was proved in [19, Theorem 2].
Theorem 2.4. (Existence)Let and . Let and the relaxed ellipticity condition (1.2) hold. Let , .
- i.
Then, there exists a weak solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) in the sense of Definition 2.1.Particularly, .There exists also the unique pressure associated with the obtained , that is the solution of Equation (2.1) in .
- ii.
Moreover, satisfies the following (strong) energy inequality,
(2.14)
3 Serrin-Type Solutions and Their Properties
In this paper, we limit ourself to the -based Sobolev spaces with respect to the spatial variables and hence introduce a corresponding particular counterpart of the class of solutions satisfying the conditions close to (3.1) and (3.2) and leading to the Serrin-type results.
3.1 Serrin-Type Solutions and Their Properties for
Definition 3.1.Let , , , and . If a solution of the initial-variational problem (2.11)–(2.12) belongs to , we will call it a Serrin-type solution.
Definition 3.2.Let , , , and . If a solution of the initial-variational problem (2.11) and (2.12) belongs to , we will call it a a strong solution.
The above definition of the strong solution is a bit weaker than, for example, in [7, Definition 6.1] or [6, Chapter 1, Section 6.7], because it does not explicitly require the additional inclusion or .
Remark 3.3.Definitions 3.1 and 3.2 imply that the strong solutions are also Serrin-type solutions if .
The Serrin-type solutions are also strong solutions if . Some sufficient conditions for the Serrin-type solution existence are provided in Section 5.2 further on in the paper.
If , then for a Serrin-type solution to be also a strong solution, the Serrin-type solution should have an additional regularity and the sufficient conditions for this are provided by the regularity theorems and corollaries in Sections 4 and 5, with the parameter there.
Lemma 3.4.Let , , , , and . Let be a Serrin-type solution. Then, , , and hence, and Moreover,
Proof.By relation (1.10), multiplication Theorem 6.1(b), and the Sobolev interpolation inequality (6.16),
To prove the lemma claim about the associated pressure , we remark that it satisfies (2.10), where due to the lemma conditions and the inclusion . By Lemma 6.5 for gradient, with , Equation (2.10) has a unique solution in .
Let us prove, in the variable-coefficient anisotropic setting, the energy equality and solution uniqueness for the Serrin-type solutions.
Theorem 3.5. (Energy equality for Serrin-type solutions)Let , , , and . If is a Serrin-type solution of the initial-variational problem (2.11) and (2.12), then the following energy equality holds for any ,
Proof.By Lemma 3.4, the function satisfies Equation (3.4), where for a.e. , we can employ as to obtain
By Lemma 3.4, , while the integrals in (3.7) are continuous in as well. Then, we conclude that the energy equality (3.7) holds for any , implying also (3.8).
Theorem 3.6. (Uniqueness of Serrin-type solutions)Let , , , and . Let be a Serrin-type solution of the initial-variational problem (2.11)–(2.12) on the interval and be any solution of the initial-variational problem (2.11)–(2.12) satisfying the energy inequality (2.14) on the interval . Then, on .
Proof.We will here generalize the proof of Theorem 6.10 in [7].
By Lemma 3.4, the function satisfies Equation (3.4), where for a.e. , we can employ as to obtain
By Lemma 2.2(i), , and hence, by Lemma 6.8, the traces are well defined. On the other hand, by Lemma 3.4, , and hence, by Lemma 6.8, the traces are well defined. Then, due to Lemma 6.9(ii) with and ,
Let us denote . Because and , we obtain
Taking into account inequality (2.6) for the quadratic form , (3.18-3.20) imply
3.2 Serrin-Type Property of the Two-Dimensional Weak Solution
By Definitions 2.1 and 3.1, any weak solution of the evolution space-periodic anisotropic Navier–Stokes initial value problem (2.1-2.3) is a Serrin-type solution for . Then, Lemmas 2.2 and 3.4 along with Theorems 3.5 and 3.6 lead to the following results for any and arbitrarily large data (unlike the higher dimensions discussed further on).
Theorem 3.7.Let , , , and the relaxed ellipticity condition (1.2) hold. Let , .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type, and hence, , , is almost everywhere on equal to a function belonging to and
4 Serrin-Type Solution Existence and Regularity for Constant Anisotropic Viscosity Coefficients
In this section, we analyze the existence and regularity of Serrin-type solutions for any in the anisotropic constant-coefficient case. This gives a generalization of Theorem 10.1 in [7], where similar results were obtained for , for the smoothness index , and for the isotropic constant-viscosity coefficients.
4.1 Vector Heat Equation
4.2 Preliminary Results for Constant Anisotropic Viscosity Coefficients
For some , , and , let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let also and .
4.3 Serrin-Type Solution Existence for Constant Anisotropic Viscosity Coefficients
Employing the results from Section 4.2 for , we are now in the position to prove the existence of Serrin-type solutions.
Theorem 4.1.Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
- i.
Then, there exist constants , , and that are independent of and but may depend on , and , such that if , and satisfy the inequality
(4.28)where is the operator defined in (4.3), then there exists a solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) in , which is thus a Serrin-type solution. - ii.
In addition, , , , and .
- iii.
Moreover, satisfies the following energy equality for any ,
(4.29)It particularly implies the standard energy equality,(4.30) - iv.
The solution is unique in the class of solutions from satisfying the energy inequality (2.14) on the interval .
Proof.
- i.
Let . Estimate (4.27) implies
(4.31)where . Then, by (4.31), we obtain from (4.25),(4.32)Let us now apply to (4.32) Lemma 6.12 with(4.33)then(4.34)(4.35)Estimates (4.13) and (4.14) were taken into account in (4.34) and (4.35).Taking into account inequalities (4.13) and (4.14)again, we obtain that condition (4.33) is satisfied if is such that
(4.36)Note that condition (4.36) gives condition (4.28) with
This implies that is a Serrin-type solution on the interval , and we thus proved item (i) of the theorem.
- ii.
Repeating for the reasoning related to inequality (4.27), we obtain
(4.37)that is, . By (1.1) and (1.3), we haveThen, (2.9) implies that , and hence, by Theorem 6.8, we obtain that , which also means that as . To prove the theorem claim about the associated pressure , we remark that it satisfies (2.10), where due to the theorem conditions and the inclusion . By Lemma 6.5 for gradient, with , Equation (2.10) has a unique solution in .
- iii.
The energy equalities (4.29) and (4.30) immediately follow from Theorem 3.5.
- iv.
The solution uniqueness follows from Theorem 3.6.
Remark 4.2.Because is integrable on by the theorem condition and is integrable on by inequality (4.6), we conclude that due to the absolute continuity of the Lebesgue integrals, for arbitrarily large data and , there exists such that condition (4.28) holds.
Estimating the integrand in the second integral in (4.28) according to (4.5), we arrive at the following assertion allowing an explicit estimate of for arbitrarily large data if and .
Corollary 4.3. (Serrin-type solution for arbitrarily large data but small time or vice versa)Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
Then, there exist constants that are independent of and but may depend on , , and , such that if satisfies the inequality
Estimating the second integral in (4.28) according to (4.8), we arrive at the following assertion.
Corollary 4.4. (Existence of Serrin-type solution for arbitrary time but small data)Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
Then, there exist constants that are independent of and but may depend on , , and , such that if and satisfy the inequality
4.4 Spatial Regularity of Serrin-Type Solutions for Constant Anisotropic Viscosity Coefficients
Theorem 4.5. (Spatial regularity of Serrin-type solution for arbitrarily large data)Let , , and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (4.28) from Theorem 4.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , , and .
Proof.The existence of the Serrin-type solution is proved in Theorem 4.1(i), and we will prove that it has a higher smoothness. We will employ the same Galerkin approximation used in Section 4.2 and in the proof of Theorem 4.1(i).
-
- Step a.
-
- Let us estimate the last term in (4.26) for the case . By (4.27), we obtain from (4.26),
(4.40)implying(4.41)By Gronwall's inequality (6.22), we obtain from (4.41) that(4.42)We have , and by (4.35), the sequence is bounded. Then, (4.42) implies that the sequence is bounded as well. Integrating (4.40), we conclude that(4.43)Inequalities (4.42) and (4.43) mean that the sequences(4.44) - Let us estimate the last term in (4.26) for the case . By (4.27), we obtain from (4.26),
-
- Step b.
-
Let now . Then, by multiplication Theorem 6.1(a) and relation (1.10),
(4.45)where .Then, by (4.45), we obtain from (4.26),
(4.46)implying(4.47)By Gronwall's inequality (6.22), we obtain from (4.47) that(4.48)We have , and by (4.44), the sequence is bounded as well. Then, (4.48) implies that the sequence is also bounded. Integrating (4.46), we conclude that(4.49)Inequalities (4.48) and (4.49) mean that the sequences(4.50) -
- Step c.
-
Let now , By multiplication Theorem 6.1(a) and relation (1.10),
(4.51)where .Then, by (4.51), we obtain from (4.26),
(4.52)implying(4.53)By Gronwall's inequality (6.22), we obtain from (4.53) that(4.54)where .If , then the sequence in (4.54) is bounded due to (4.44) and (4.50). Then, (4.54) implies that the sequence is bounded as well. Integrating (4.52), we also conclude that for ,
(4.55)Inequalities (4.54) and (4.55) mean that for , the sequences(4.56)If we assume that properties (4.56) hold for some integer , then by the similar argument, properties (4.56) hold with replaced by , and thus, by induction, they hold for any integer . Hence, collecting properties (4.44), (4.50), and (4.56), we conclude that the sequences
(4.57)Properties (4.57) imply that there exists a subsequence of converging weakly in and weakly star in to a function . Then, the subsequence converges to also weakly in and weakly star in . Because is the subsequence of the sequence that converges weakly in and weakly star in to the weak solution, , of problem (2.1-2.3) on , we conclude that , for any , and we thus finished proving that
(4.58) -
- Step d.
-
Repeating for the reasoning related to inequalities (4.27), (4.45), and (4.51), corresponding to the considered , we obtain
(4.59)Due to (4.58), then . By (1.1) and (1.3), we haveTo prove the theorem claim about the associated pressure , we remark that satisfies (2.10). By Lemma 6.5 for gradient, with , Equation (2.10) has a unique solution in .
As in Corollaries 4.3 and 4.4, condition (4.28) in Theorem 4.5 can be replaced by simpler conditions for particular cases, which leads to the following two assertions.
Corollary 4.6. (Serrin-type solution for arbitrarily large data but small time or vice versa)Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and , . Let satisfies inequality (4.38) in Corollary 4.3.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , , and .
Corollary 4.7. (Serrin-type solution for arbitrary time but small data)Let , , and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let the data and satisfy inequality (4.39) in Corollary 4.4.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , and .
Theorem 4.5 leads also to the following infinite regularity assertion.
Corollary 4.8.Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (4.28) from Theorem 4.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , , and .
Proof.Taking into account that , Theorem 4.5 implies that , , , . Hence, , , and .
4.5 Spatial-Temporal Regularity of Serrin-Type Solutions for Constant Anisotropic Viscosity Coefficients
Theorem 4.9.Let and . Let if , while if . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (4.28) from Theorem 4.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , while .
Proof.By Theorems 4.1 and 4.5, we have the inclusions , , and . Then, we only need to prove the inclusions and .
Let, first, (and also if ). By relation (1.10) and multiplication Theorem 6.1(b), we have
Let now . Again, by relation (1.10) and by multiplication Theorem 6.1(a), we have
Then, (2.9) implies that , while (2.10) and Lemma 6.5 for gradient, with , imply that .
Theorem 4.10.Let and . Let . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let be an integer. Let , , and , while , , and satisfy inequality (4.28) from Theorem 4.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , , while , .
Proof.Some parts of the following proof are inspired by [30, Theorem 3.1] and [11, Chapter 3, Section 3.6], see also [7, Section 7.2].
We will employ the mathematical induction argument in the proof. We first remark that by Theorems 4.1, 4.5, and 4.9, if , then , , while . This means that the theorem holds true for .
Let us assume that the theorem holds true for some , that is,
Let us denote . Then,
- Step 1.
Convection term.
(4.64)where are the binomial coefficients.- Case A.
Let
. Then,
,
.
- Subcase A1.
Let
. Then,
. By the theorem conditions,
and
, and hence, there exists
, and thus,
- Subcase A2.
Let .
Then, . Hence, by relation (1.10) and multiplication Theorem 6.1(a), we have
Thus, combining Cases (A1) and (A2), we obtain that for any and for ,
(4.65)(4.66)where is or , respectively.
- Subcase A1.
Let
. Then,
. By the theorem conditions,
and
, and hence, there exists
, and thus,
- Case B.
Let . Then, taking into account that
Thus, for any and any integer , we obtain the estimates
(4.67)
- Case A.
Let
. Then,
,
.
- Step 2.
Linear terms and right-hand side.
Due to (4.61),
(4.68)We also have . Then, (4.63), (4.67), and (4.68) imply that(4.69)Thus, by (4.62), . - Step 3.
Pressure.
The associated pressure satisfies (2.10). Differentiating it in time, we obtain
(4.70)By the same reasoning as in the proof of (4.69), the similar inclusions for junior derivatives also hold:
Corollary 4.11.Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (4.28) from Theorem 4.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , .
Proof.Taking into account that , Theorem 4.10 implies that for any integer , , , for any . Hence, , .
4.6 Regularity of Two-Dimensional Weak Solution for Constant-Viscosity Coefficients
The regularity results of Sections 4.4 and 4.5 hold for , but as for the isotropic constant-coefficient case (cf., e.g., [11, Chapter 3, Sections 3.3 and 3.5.1], [7, Section 6.5]), these results can be essentially improved for also in the anisotropic setting with constant coefficients.
Let us give a counterpart of Theorem 4.5 that for , it is valid on any time interval (and not only on its special subinterval ).
Theorem 4.12. (Spatial regularity of solution for arbitrarily large data)Let , , and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and belongs to . In addition, , , , and .
Proof.The proof coincides word for word with the proof of Theorem 4.5 if we take there while replacing by and the reference to (4.35) for the boundedness of the sequence for by the reference to the corresponding inequality
The following assertion can be proved similarly to Corollary 4.8.
Corollary 4.13.Let and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , , and .
The next three assertions on spatial-temporal regularity for are the corresponding counterparts of Theorems 4.9 and 4.10 and Corollary 4.11 and are proved in a similar way after replacing there by .
Theorem 4.14.Let , , and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , .
Theorem 4.15.Let , , and . Let the coefficients be constant and the relaxed ellipticity condition (1.2) hold. Let be an integer. Let , , and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , ; , .
5 Serrin-Type Solution Existence and Regularity for Variable Anisotropic Viscosity Coefficients
In this section, we generalize to the anisotropic variable viscosity coefficients the analysis of the existence and regularity of Serrin-type solutions for any given in Section 4 for the anisotropic constant-viscosity coefficients.
5.1 Preliminary Results for Variable Anisotropic Viscosity Coefficients
For some , , and , let , , and the relaxed ellipticity condition (1.2) hold. Let also and .
Let us now estimate the terms in the right-hand side of (5.2). For the first two terms and for the last one, estimates (4.21), (4.22), and (4.24) still hold.
5.2 Serrin-Type Solution Existence for Variable Anisotropic Viscosity Coefficients
Employing the results from Section 5.1 for , we are now in the position to prove the existence of Serrin-type solutions.
Theorem 5.1.Let and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and .
- i.
Then, there exist constants , , and that are independent of and but may depend on , , , and , such that if , , and satisfy the inequality
(5.8)where is the operator defined in (4.3), then there exists a solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) in , which is thus a Serrin-type solution. - ii.
In addition, , , , and .
- iii.
Moreover, satisfies the following energy equality for any ,
(5.9)It particularly implies the standard energy equality,(5.10) - iv.
The solution is unique in the class of solutions from satisfying the energy inequality (2.14) on the interval .
Proof.
- i.
Let . The estimate (4.27) still holds. Let us fix any small such that . Then, by (4.31), we obtain from (5.6),
(5.11)where , .Let us now apply to (5.11) Lemma 6.13 with
(5.12)where(5.13)(5.14)Estimates (4.13) and (4.14) were taken into account in (5.13) and (5.14).Taking into account inequality (4.13), we obtain that condition (5.12) is satisfied if is such that
(5.15)Note that condition (5.15) gives condition (5.8) with
Inequalities (5.13) and (5.14) imply that there exists a subsequence of converging weakly in and weakly star in to a function . Then, the subsequence converges to also weakly in and weakly star in . Because is the subsequence of the sequence that converges weakly in and weakly star in to the weak solution, , of problem (2.1-2.3) on , we conclude that .
This implies that is a Serrin-type solution on the interval , and we thus proved item (i) of the theorem.
- ii.
As in step (ii) of the proof of Theorem 5.1, estimate (4.37) implies that . By (1.1) and (1.3), we have
Then, (2.9) implies that , and hence, by Theorem 6.8, we obtain that , which also means that as .
To prove the theorem claim about the associated pressure , we remark that it satisfies (2.10), where due to the theorem conditions and the inclusion . By Lemma 6.5 for gradient, with , Equation (2.10) has a unique solution in .
- iii.
The energy equalities (5.9) and (5.10) immediately follow from Theorem 3.5.
- iv.
The solution uniqueness follows from Theorem 3.6.
Remark 5.2.Note that by the Sobolev embedding theorem, the condition , , in Theorem 5.1 and further on implies .
Remark 5.3.Because is integrable on by the theorem condition and is integrable on by the inequality (4.5), we conclude that due to the absolute continuity of the Lebesgue integrals, for arbitrarily large data and , there exists such that condition (5.8) holds.
Estimating the integrand in the second integral in (5.8) according to (4.5), we arrive at the following assertion allowing an explicit estimate of for arbitrarily large data if .
Corollary 5.4. (Serrin-type solution for arbitrarily large data but small time or vice versa)Let and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and .
Then, there exist constants that are independent of and but may depend on , , , and , such that if satisfies the inequality
Estimating the second integral in (5.8) according to (4.8), we arrive at the following assertion.
Corollary 5.5. (Existence of Serrin-type solution for arbitrary time but small data)Let and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and .
Then, there exist constants that are independent of and but may depend on , , , and , such that if and satisfy the inequality
5.3 Spatial Regularity of Serrin-Type Solutions for Variable Anisotropic Viscosity Coefficients
Theorem 5.6. (Spatial regularity of Serrin-type solution for arbitrarily large data)Let , , and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (5.8) from Theorem 5.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , , and .
Proof.The existence of the Serrin-type solution is proved in Theorem 5.1(i), and we will prove that it has a higher smoothness. We will employ the same Galerkin approximation used in Section 5.1 and in the proof of Theorem 5.1(i).
- Step a.
Let us estimate the last term in (5.7) for the case . By (4.27), we obtain from (5.7),
(5.18)implying(5.19)By Gronwall's inequality (6.22), we obtain from (5.19) that(5.20)We have and by (5.14), the sequence is bounded. Then, (5.20) implies that the sequence is bounded as well. Integrating (5.18), we conclude that(5.21)Inequalities (5.20) and (5.21) mean that the sequences(5.22) - Step b.
Let now . By (4.45), we obtain from (5.7),
(5.23)implying(5.24)By Gronwall's inequality (6.22), we obtain from (5.24) that(5.25)We have , and by (5.22), the sequence is bounded as well. Then, (5.25) implies that the sequence is also bounded. Integrating (5.23), we conclude that(5.26)Inequalities (5.25) and (5.26) mean that the sequences(5.27) - Step c.
Let now , By (4.51), we obtain from (5.7),
(5.28)implying(5.29)By Gronwall's inequality (6.22), we obtain from (5.29) that(5.30)where .If , then the sequence in (5.30) is bounded due to (5.22) and (5.27). Then, (5.30) implies that the sequence is bounded as well. Integrating (5.28), we also conclude that for ,
(5.31)Inequalities (5.30) and (5.31) mean that for , the sequences(5.32)If we assume that properties (5.32) hold for some integer , then by the similar argument, properties (5.32) hold with replaced by , and thus, by induction, they hold for any integer . Hence, collecting properties (5.22), (5.27), and (5.32), we conclude that the sequences
(5.33)Properties (5.33) imply that there exists a subsequence of converging weakly in and weakly star in to a function . Then, the subsequence converges to also weakly in and weakly star in . Because is the subsequence of the sequence that converges weakly in and weakly star in to the weak solution, , of problem (2.1-2.3) on , we conclude that , for any , and we thus finished proving that
(5.34) - Step d.
Estimate (4.59) and inclusion (5.34) imply that . By (1.1) and (1.3), we have
To prove the theorem claim about the associated pressure , we remark that satisfies (2.10). By Lemma 6.5 for gradient, with , Equation (2.10) has a unique solution in .
As in Corollaries 5.4 and 5.5, condition (5.8) in Theorem 5.6 can be replaced by simpler conditions for particular cases, which leads to the following two assertions.
Corollary 5.7. (Serrin-type solution for arbitrarily large data but small time or vice versa)Let and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and , . Let satisfy inequality (5.16) in Corollary 5.4.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , , and .
Corollary 5.8. (Serrin-type solution for arbitrary time but small data)Let , , and . Let , , and the relaxed ellipticity condition (1.2) hold. Let the data and satisfy inequality (5.17) in Corollary 5.5.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) belongs to . In addition, , , , and .
Theorem 5.6 leads also to the following infinite regularity assertion.
Corollary 5.9.Let and . Let and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (5.8) from Theorem 5.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , , and .
Proof.Taking into account that , Theorem 5.6 implies that , , , . Hence, , and .
5.4 Spatial-Temporal Regularity of Serrin-Type Solutions for Variable Anisotropic Viscosity Coefficients
Theorem 5.10.Let and . Let if , while if . Let , , and the relaxed ellipticity condition (1.2) hold. Let and , while , , and satisfy inequality (5.8) from Theorem 5.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , while .
Proof.By Theorems 5.1 and 5.6, we have the inclusions , , and . Then, we only need to prove the inclusions and .
As in the proof of Theorem 5.10, we arrive at estimate (4.60).
By (1.1), (1.3), and multiplication Theorem 6.1(a), we have
Then, (2.9) implies that , while (2.10) and Lemma 6.5 for gradient, with , imply that .
To simplify the following two assertions we assume there that the viscosity coefficients are infinitely smooth it time and in the space coordinates. This smoothness condition can be relaxed if we instead assume that all the norms of these coefficients encountered in the proof are bounded.
Theorem 5.11.Let and . Let . Let and the relaxed ellipticity condition (1.2) hold. Let be an integer. Let , , and , while , , and satisfy inequality (5.8) from Theorem 5.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , , while , .
Proof.The proof coincide with the proof of the corresponding constant-coefficient Theorem (4.10), except the proof of inclusion (4.68) in Step 2, that for the variable coefficients is replaced by the following argument.
On the other hand, by (1.1), (1.3), and Theorem 6.1(a), we have
Corollary 5.12.Let and . Let and the relaxed ellipticity condition (1.2) hold. Let , and , while , , and satisfy inequality (5.8) from Theorem 5.1.
Then, the Serrin-type solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) is such that , .
Proof.Taking into account that , Theorem 5.11 implies that for any integer , , , for any . Hence, , .
5.5 Regularity of Two-Dimensional Weak Solution for Variable Anisotropic Viscosity Coefficients
Here, we provide a counterpart of Section 4.6 generalized to variable viscosity coefficients.
Theorem 5.13. (Spatial regularity of solution for arbitrarily large data)Let , , and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and belongs to . In addition, , , and .
Proof.The proof coincides word for word with the proof of Theorem 5.6 if we take there while replacing by and the reference to (5.14) for the boundedness of the sequence for by the reference to the corresponding inequality
The following assertion is proved similar to Corollary 5.9.
Corollary 5.14.Let and . Let and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , , and .
The next three assertions on spatial-temporal regularity for are the corresponding counterparts of Theorems 5.10 and 5.11 and Corollary 5.12 and are proved in a similar way after replacing there by .
Theorem 5.15.Let , , and . Let , , and the relaxed ellipticity condition (1.2) hold. Let and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , .
Theorem 5.16.Let , , and . Let and the relaxed ellipticity condition (1.2) hold. Let be an integer. Let , , and .
Then, the solution of the anisotropic Navier–Stokes initial value problem (2.1–2.3) obtained in Theorem 2.4 is of Serrin type and is such that , ; , .
6 Auxiliary Results
6.1 Advection Term Properties
Equations (6.3) and (6.4) evidently hold also for , , and from the more general spaces, for which the dual products in (6.3) and (6.4) are bounded and in which and , respectively, are densely embedded.
6.2 Some Point-Wise Multiplication Results
Let us accommodate to the periodic function spaces in , , a particular case of a much more general Theorem 1 in Section 4.6.1 of [31] about point-wise products of functions/distributions.
Theorem 6.1.Assume , , and . Then, there exists a constant such that for any and ,
- a.
;
- b.
.
Proof.Items (a) and (b) follow, respectively, from items (i) and (iii) of [31, Theorem 1 in Section 4.6.1] when we take into account the norm equivalence in the standard and periodic Sobolev spaces.
Lemma 6.2.Let , . Then, the convolution of sequences and belongs to and
Proof.By the triangle inequality, we obtain
Theorem 6.3.Assume . Let , , , . Then, and
Proof.By (1.6), we have
Here,
By the Schwarz inequality for any , we have
Similarly, for any , we have
Then, for , we obtain
6.3 Spectrum of the Periodic Bessel-Potential Operator
The following assertion is given in [19, Theorem 4, Remark 2].
Theorem 6.4.Let , .
- i.
Then, the operator in possesses a (nonstrictly) monotone sequence of real eigenvalues and a real orthonormal sequence of associated eigenfunctions such that
(6.9)(6.10)(6.11) - ii.
Moreover, the sequence is an orthonormal basis in , that is,
(6.12)where the series converges in for any . - iii.
In addition, the sequence is also an orthogonal basis in with
(6.13)converges to in as . The operator defined by (6.13) is for any the orthogonal projector operator from to .
6.4 Isomorphism of Divergence and Gradient Operators in Periodic Spaces
The following assertion proved in [19, Lemma 2] provides for arbitrary and dimension the periodic version of Bogovskii/deRham-type results well known for nonperiodic domains and particular values of ; see, for example, [33, 34] and references therein.
Lemma 6.5.Let and . The following operators are isomorphisms,
6.5 Some Functional Analysis Results
Let us provide the Sobolev embedding theorem that can be considered, for example, as a particular case of [31, Section 2.2.4, Corollary 2] adapted to the periodic spaces.
Theorem 6.6.Let be the dimension, and . The periodic Bessel-potential space is continuously embedded in if and only if .
The following version of the Sobolev interpolation inequality without a multiplicative constant, generalized also to any real (including negative) smoothness indices, on periodic Bessel-potential spaces was given in [19, Theorem 5].
Theorem 6.7.Let , , , , be real numbers such that , and . Then, for any ,
Theorem 3.1 and Remark 3.2 in Chapter 1 of [27] imply the following assertion.
Theorem 6.8.Let and be separable Hilbert spaces and with continuous injection. Let . Then, almost everywhere on equals to a function , where is the intermediate space. Moreover, the trace is well defined as the corresponding value of at .
The following assertion was proved in [19, Lemma 4].
Lemma 6.9.Let , , and be real valued.
- i.
Then,
(6.17)for a.e. and also in the distribution sense on . - ii.
Moreover, for any real-valued and ,
(6.18)
Let us now prove the first Korn inequality for the periodic Sobolev spaces by adapting the proof available for the standard Sobolev spaces, for example, in [35, Theorem 10.1]; compare also [20, Theorem 2.8].
Theorem 6.10.Let , . Then,
Proof.By the norm definition (1.5), we obtain
6.6 Gronwall's Inequalities
Gronwall's inequality is well known and can be found, for example, in [36, Appendix B.2.j], [7, Lemma A.24]. Here, we provide its slightly more general version valid also for arbitrary-sign coefficients.
Lemma 6.11.Let be an absolutely continuous function that satisfies the differential inequality
- a.
Then,
(6.21) - b.
Moreover, for nonnegative and , (6.21) implies
(6.22) - c.
In particular, if is nonnegative, while on and , then on .
Proof.Multiplying (6.20) by
Let us slightly generalize and give an alternative proof of [7, Lemma 10.3].
Lemma 6.12.Let be an absolutely continuous function that satisfies the differential inequality
If
Proof.By Lemma 6.11(a), inequality (6.23) and condition (6.24) lead to
Let us consider the function on the interval . One can elementary obtain that is reached at and equals to . But due to (6.27), this maximum for is not reached for , and hence, , giving the second inequality in (6.25).
Further, (6.26) implies that
Let us give a generalization of [7, Lemma 10.3] and of Lemma 6.12.
Lemma 6.13.Let be an absolutely continuous function that satisfies the differential inequality
If
Proof.By Lemma 6.11(a), inequality (6.28) and condition (6.29) lead to
Let us consider the function on the interval . One can elementary obtain that is reached at and equals to . But due to (6.32), this maximum of is not reached for , and hence, , giving the second inequality in (6.30).
Further, (6.31) implies that
Let us give a version of integral Gronwall's inequality implied, for example, by Theorem 1.3 and Remark 1.5 in [37].
Lemma 6.14.Let , , and be measurable functions in , such that . Suppose that is nonnegative a.e. on . Suppose
Author Contributions
Sergey E. Mikhailov: conceptualization, investigation, writing – original draft, methodology, validation, writing – review and editing, formal analysis.
Acknowledgments
The research has been funded by the general research grant from UK Research and Innovation (UKRI) to Brunel University of London.
Conflict of Interest
The author declares no conflicts of interest.
Open Research
Data Availability Statement
This paper has no associated data.