1. Introduction
The aim of this work is to study the optimal feedback control problem for the alpha-model with the Voigt fractional rheological relation, taking into account the background of a fluid along the trajectory. Note that memory properties in general arise not only in the fluid dynamics field but in many absolute different fields [
1]. So the results of this paper can be useful in many fields. A large number of papers have been devoted to the investigation of control problems [
2,
3,
4]. Although the control problems for linear systems are sufficiently well studied, the situation is not so good for nonlinear systems (even for finite-dimensional cases or local domains). However, due to the complexity of nonlinear systems describing the fluids motion the control of non-Newtonian fluids motion, such as bitumen, polymers, various solutions, emulsions and suspensions, blood, and many others, has not been fully studied. In hydrodynamics the control (optimal control) problems often connected with the fluid control by external forces. Usually in solving such problems, a control is considered from a given (finite) set. In our situation, we consider the external forces control depending on the velocity field. Such types of problems are called feedback control problems [
2,
3,
4,
5]. In this situation the control is chosen more accurately, since in such a way the control belongs to the image of some multi-valued map. This is more naturally due to the fact that control is not chosen from a finite set of available options.
Also in this paper the alpha model case of fractional Voigt model is considered. Alpha-models are some kind of regularized approximate systems that depend on some positive parameter
, and regularization is carried out by some filtering of the velocity vector, which is contained in the argument of the nonlinear term. The
parameter reflects the width of the spatial filtering scale for the modified speed. The Helmholtz operator
is most often used as the filtration kernel. The choice of such an operator is associated with its good mathematical properties. Thus, we ready to proceed to the formulation of the problem. In a bounded domain
(in 2D and 3D cases, that is,
) with a sufficiently smooth boundary
on a time interval
where
, we consider the initial-boundary value problem:
Here
v is a vector-function of the velocity of a medium particle,
u is a vector-function of a modified velocity of a medium particle, defined by equality (
2),
is the trajectory of a medium particle, indicating at time
the location of a medium particle located at time moment
t at point
x,
p is a pressure function,
f is a function of the density of external forces,
is scalar parameter,
,
,
are some constants.
is the strain rate tensor.
is the Euler gamma function [
6] defined through an absolutely convergent integral
This initial-boundary value problem (
1)–(
5) is an alpha model for the mathematical model of viscoelastic Voigt medium with fractional rheological relation. The idea of using this kind of approximation (the alpha-model) first appeared in paper of J. Leray [
7] (in this work, J. Leray used the general form of the filtration kernel) to prove the existence of a weak solution for the Navier-Stokes system of equations. Later, various alpha-models for the Euler equations [
8,
9], the Navier-Stokes system [
10] and others were built on this idea. In general, each alpha model is characterized by its first-order vector differential operator
, in which components
are linear combinations of all kinds of operators of form
:
where
are some real coefficients. Note that in representation (
6) monomials of the form
are not used, since they do not contain the components of the <<smoothed>> vector field
u.
Interest in the study of alpha-models is primarily associated with their application to the study of turbulence effects for fluid flows. It is also associated with obtaining better numerical results for alpha-models in comparison to the original models. However, most of the works on the solvability of alpha-models are devoted to models of the motion of an ideal or Newtonian fluid [
11,
12,
13,
14]. Only in the last few years, works began to appear on the study of alpha-models of non-Newtonian fluid [
15,
16,
17,
18]. This work continues the study of alpha-models for non-Newtonian fluids, namely, for the fractional Voigt model of the viscoelastic medium [
19]. This mathematical model describes a viscoelastic fluid flow with a rheological relation
, considered along the trajectories of fluid motion. Here
is the left-side fractional Riemann-Liouville derivative and
is the Riemann-Liouville fractional integral. This model is a fractional analog of the Voigt model, which describes the motion of a linearly elastic-retarded fluid. In order to study a large class of polymers with creep and relaxation effects one must to consider models with fractional derivatives. It turns out that the models with fractional derivatives are most suitable for this [
20,
21]. Note that the advantage of this model is that, together with the definition of the vector-velocity
v of the particle’s motion, the trajectory of the particles of this medium motion
z is also determined. Also, note that the consideration of fractional derivatives in fluid dynamics has many physical applications [
22,
23,
24]. One of the possible continuations of this model studies is laid out in References [
25] and [
26].
2. Preliminary Information and Statement of the Main Results
We introduce the main notation and auxiliary statements.
By , we denote the set of measurable vector functions , summable with p degree. By , , , we denote Sobolev spaces. We consider the space of infinitely differentiable vector functions from to with compact support in . Denote by the set . Also by and we denote the closure of with respect to the norm of and , respectively, and by we denote the space .
We introduce from Reference [
27] the scale of spaces
,
. For this we consider the Leray projector
and the operator
defined on
. From this operator we can get a self-adjoint positive operator with compact inverse in
. Let
be the eigenvalues of the operator
A. We can get an orthonormal basis in
by the eigenfunctions
of the operator
A due to the Hilbert theorem on the spectral decomposition of compact operators. Denote by
the set of finite linear combinations of
. Thus, we get the space
,
as the completion of
with respect to the norm
In Reference [
27] it is shown that on the space
,
, norm (
7) is equivalent to the ordinary norm
of the space
. In addition, according to Reference [
28], the norms in the spaces
,
and
can be defined as follows:
Here the symbol denotes the component-wise matrix product, that is, for , , , we put .
Further, through the , , we denote the space dual to .
Note that is the Banach space of continuous on functions, is the Banach space of weakly continuous on functions, is the Banach spaces of summable on with p degree functions with values in a Banach space F, respectively.
The set consists of one-to-one mappings coinciding with the identity mapping on and having continuous first-order partial derivatives on such that at every point of the domain . For this set the norm of continuous functions is used. Further, we will consider the following set Note that therefore, in what follows, is considered a metric space with a metric defined by the norm of the space .
We introduce the space in which the solvability of the considered problem will be proved:
with the norm
.
Denote by
,
the operator
, where
, and
I is the identity operator. By virtue of Reference [
28], the operator
is invertible. If we apply the Leray projection
to the equality
for
and express from the last equality
u:
. Then, since
, we get that
Note that for the correct formulation of the considered initial-boundary value problem the trajectories
z must be uniquely determined by the velocity field
In other words, it is necessary that Equation (
3) has a unique solution for the velocity field
v. However, the existence of solutions to Equation (
3) for a fixed
v is known in Reference [
29] only in case
and this solution is unique for
such that
. Therefore, the trajectories of motion are not uniquely determined even for strong solutions whose partial derivatives that appear in Equation (
3) are contained in
One possible way out of this situation is to regularize the velocity field at each time instant
t by averaging over the variable
x and determine the trajectories
for the regularized velocity field [
30]. However, relatively recently [
31,
32], the solvability of Cauchy integral problem (
3) was investigated in the case when the velocity
v belongs to the Sobolev space. Also the existence and uniqueness of regular Lagrangian flows, which are a generalization of the concept of a classical solution, are established.
Definition 1. Regular Lagrangian flow associated to v is the function , satisfying conditions:
- 1.
the function is absolutely continuous and satisfies Equation (3) for almost all and ; - 2.
the equality holds for any and an arbitrary Lebesgue measurable set with Lebesgue measure ;
- 3.
for all , , and almost all
We give the necessary results from a regular Lagrangian flow.
Theorem 1 [
31]
. Let , with conditions , , and . Then there exists a unique regular Lagrangian flow associated to v (where is the Banach space of continuous functions on with values in the metric space of vector functions L measurable on Ω). Moreover, up to a set of measure zero and Theorem 2. Let v, , for some Let , , , . Also, let the inequalitiesare valid. Here and are the Jacobi matrices of the vector functions v and . Let converges to v in as Let and are regular Lagrangian flows associated to and v, respectively. Then the sequence converges (up to a subsequence) to z with respect to the Lebesgue measure on the set uniformly on . This result was proved in Reference [
33] in the general case.
Thus, by virtue of Theorem 1 for each
and for almost all
, the Equation (
3) has a unique solution
, where
, in the class of regular Lagrangian flows.
As a control function, we consider the multi-valued map . Assume that satisfies the following conditions:
- (Ψ1)
is defined on the space and has nonempty, compact, and convex values;
- (Ψ2)
is compact and upper semicontinuous (that is, for any function and any open set such that , there exists a neighborhood such that );
- (Ψ3)
is globally bounded, that is, there exists a constant
such that
- (Ψ4)
is weakly closed, that is: if and in then .
In this paper, a weak statement of the feedback control problem for initial-boundary value problem (
1)–(
5) is considered. By feedback, we mean the condition
We formulate the definition of a weak solution to feedback control problem (
1)–(
5), (
8):
Definition 2. A pair of functions is called a weak solution of feedback control problem (
1)–(
5), (
8),
if it for all and almost all satisfies the equalitythe initial condition and feedback condition (
8).
Here is a regular Lagrangian flow associated to v. Remark 1. It is known that [
34].
Therefore, initial condition (
5)
has sense. The following theorem is the first result of the paper:
Theorem 3. Let a multi-valued mapping Ψ satisfy conditions . Then there is at least one weak solution of feedback control problem (
1)–(
5), (
8).
We denote by
the set of all weak solutions of problem (
1)–(
5), (
8). Consider an arbitrary cost functional
satisfying the following conditions:
- (Φ1)
For all a number exists such that
- (Φ2)
If in and in then
As an example of this functional, we can take
Here is some specified velocity field. This functional characterizes the deviation of velocity from the required, and its minimum yields the minimal deviation of velocity from the one specified by the minimal control. One of the possible applications of the proposed approach is an optimal feedback control problem and the results are in the consideration, analysis and calculation of different such problems with special (necessary in industry) cost functionals .
The following theorem is the second result of this paper.
Theorem 4. If the mapping Ψ
satisfies conditions – and the functional Φ
satisfies conditions , then optimal feedback control problem (
1)–(
5), (
8)
has at least one weak solution such that The proof of Theorems 3 and 4 is based on the approximation-topological method for investigating fluid dynamics problems [
35]. To do this, first, we pass to the operator interpretation of the problem under consideration (operator inclusion) in suitable function spaces. Further, since the operators in the obtained operator inclusion do not have the necessary properties, we consider a problem that approximates the original one (in this case, it is also an operator inclusion, but with a better operator that has the required properties and in better functional spaces). Then, based on a priori estimates of solutions and the theory of the topological degree of multi-valued vector fields, the existence of a solution to the approximation problem is proved. Finally, it is shown that from the sequence of solutions of the approximation problem, one can extract a subsequence that converges in a weak sense to the solution of the original operator inclusion. After proving the solvability of the control problem, it is shown that in the set of solutions there is at least one solution that gives a minimum to a given cost functional (this is why this type of problem is called the optimal feedback control problem for fluid motion).
The work is organized as follows—in
Section 3 we consider the family of auxiliary problems and prove the necessary properties of an introduced operators. Also on the basis of the topological degree theory for multivalued maps we prove the solvability of the auxiliary problem and establish necessary estimates for solutions to the auxiliary problem.
Section 4 is devoted to the passage, the limit and the proof of Theorem 3.
Section 5 is devoted to the proof of Theorem 4. The final
Section 6 contains conclusions.
3. The Family of Auxiliary Problems
Throughout this section we will assume that .
Consider the following auxiliary family of systems of equations (
) with a small parameter
:
For this family we consider another functional space:
with the norm
.
Equation (
10) includes the integral calculated along the trajectories of motion of the fluid particles. As was noted in the previous section, it is necessary that the trajectories are uniquely determined by the velocity field
. In other words, Equation (
11) must have a unique solution for the velocity field
. Note that for the family of auxiliary problems (
10)–(
14), the velocity
v from the space
has sufficient smoothness (due to the embedding of the space
in
for
). Thus, it follows from Reference [
29] that the Cauchy problem (
11) is non-locally uniquely solvable.
Analogously with the definition of a weak solution for feedback control problem (
1)–(
5), (
8), we formulate the definition of a weak solution to auxiliary problem (
10)–(
14), (
8) for fixed
.
Definition 3. A pair of functions is called a weak solution to auxiliary problem (
10)–(
14), (
8)
if it satisfies for any and almost all the equality feedback condition (
8)
and initial condition (
14).
Here z is the trajectory associated to the velocity v. To prove the existence of a weak solution to auxiliary problem (
10)–(
14),(
8) for
, we rewrite the auxiliary family in operator form. Using the terms in equality (
15), we introduce the operators using the following equalities:
Since the function
is arbitrary in (
15), for almost all
this equality is equivalent to the following operator equation in
:
Thus, a weak solution to auxiliary problem (
10)–(
14), (
8) for a fixed
is a solution
of the following operator inclusion
satisfying initial condition (
14).
We also define the operators using the following equalities:
Thus, from our problem of finding a solution to operator inclusion (
16) for a fixed
satisfying initial condition (
14) we get the problem of finding a solution for a fixed
to the following operator inclusion
We need the following properties of the operators from inclusions (
16) and (
17). In order to not to pile up the notation, we will use the same letter to denote the same operators acting in different function spaces.
Lemma 1. - 1.
For any function it holds that the function and the operator is continuous and the following estimates hold: - 2.
The operator is linear, continuous, invertible and the following estimate holds: In addition, the operator is also continuous.
- 3.
For any function , the function belongs to and the operator is continuous and invertible. In addition, the following estimate holds: Moreover, the inverse to it operator is continuous and for any we have the estimate - 4.
The operator is invertible and the operator is a continuous operator.
Proof. The proof is carried out in the same way as in Reference [
36]. □
Lemma 2. - 1.
The map is continuous and the following estimate holds: - 2.
For any the function and the map is continuous.
- 3.
For any function the function and the map is compact.
Proof. - 1.
For any
,
using Holder’s inequality, we obtain
This implies inequality (
23). Note that here we used the following well-known estimate [
37,
38]:
We show the continuity of the map
. For arbitrary
,
we have
Assuming that in , we obtain that the map is continuous.
- 2.
To prove this item, it is necessary to use the last estimate and repeat the proof of Lemma 2.5.4 (item 2) from Reference [
28].
- 3.
To prove this item, we use the Aubin-Simon theorem:
Theorem 5. [
28,
39,
40]
Let are Banach spaces, the embedding is compact and the embedding is continuous. Also let , . We assume that for any its generalized derivative belongs to , . Now let:F is bounded in ;
is bounded in .
Then for the set F is relatively compact in , and for and the set F is relatively compact in .
Consider the set . Since the embedding is compact, the embedding is also compact.
From continuity of embeddings
the continuous embedding
follows. In addition, also we have that the operator
is continuous (from the second item of this lemma). Thus, we have the superposition of embeddings:
where the first embedding is continuous, the second is compact, and the map
B is continuous. Therefore, for any function
we obtain that the function
, and the map
is compact. The proof is complete.
☐
We proceed to investigate the properties of the map C. We introduce the following norm equal to the norm where . Then the following lemma holds.
Lemma 3. For any , we have that and the map is continuous and bounded. In addition, for any fixed function and arbitrary u, the following estimate holds: Proof. The first part of this lemma is proved similarly to the Lemma 2.2 [
30]. We prove necessary estimate (
25). Let
. By definition, for almost all
we have
and obtain
Then, using the Holder inequality, we obtain
The last inequality holds by virtue of the following estimate [
41] (Theorem 2.6)
Estimate the remaining integral:
Thus, we obtain the estimate
From where necessary estimate (
25) follows. □
We formulate one more necessary property of the operator C.
But first we define several concepts concerning the measure of noncompactness and
L-condensing operators [
30,
42].
Definition 4. A non-negative real function ψ defined on a subset of a Banach space F is called a measure of non-compactness if for any subset of this space the following properties are satisfied:
Here, by we denote the convex closure of the set . As an example of a measure of non-compactness, we give the Kuratowski measure of non-compactness: the exact lower bound for which the set can be divided into a finite number of subsets whose diameters are less than d. Kuratowski’s non-compactness measure has several important properties:
, if is a relatively compact subset;
if K is a relatively compact set.
Definition 5. Let X be bounded subset of a Banach space, and is a map from X into a Banach space F. A map is called L-condensing if for any set such that .
Let be the Kuratowski measure of noncompactness in the space with the norm . Then the following lemma holds.
Lemma 4. The map is L-condensing with respect to the Kuratowski measure of noncompactness .
Proof. Let
be an arbitrary bounded set. By virtue of Theorem 2, the set
is the set of trajectories
z that are uniquely determined by the velocities
and this set is relatively compact. Then for any fixed
the set
is relatively compact. In addition, for any
, the map
satisfies the Lipschitz condition with constant
in the norms
and
. Then, by Theorem 1.5.7 [
42], the map
and, therefore, the map
G are
-bounded with respect to the Hausdorff measure
It is known, see Theorem 1.1.7 [
42], that the non-compactness measures of Hausdorff and Kuratowski satisfy the following inequalities
. Therefore, the estimate
hold. Choosing
k so that
, we obtain the statement of the lemma. □
Using the above estimates and the properties of the operators, we prove the following a priori estimates for auxiliary family (
10)–(
14), (
8).
Lemma 5. Let . Then for any solution of operator inclusion (
16)
the following estimates hold:where the constants , , do not depend on ε and ξ. Proof. Let
be a solution of operator inclusion (
16). Then for any
and almost all
equality (
15) holds. Since it is valid for all
, we assume that
, where
. Then
Let us replace
and separately transform the terms in the left side of the last equality as follows. Consider the first term:
Now we turn to the consideration of the following term:
Finally, we transform the last term:
Thus, equality (
29) can be rewritten as follows:
We estimate modulo the right-hand side of the resulting equality. Using the Cauchy inequality
for
, we obtain:
Multiplying both sides of equality (
31) on
, for almost all
we have
We integrate the last inequality with respect to
t from 0 to
, where
. Then
We use estimate (
25) for
. In this way,
We assume that the number
k is sufficiently large such that
. The nonnegativity of the quantities
,
and
yields the following estimates:
Since the right-hand side in all the above inequalities does not depend on
, we pass to the maximum in
in the left-hand sides of these inequalities. Then
From this and feedback condition (
8) the required estimates (
26)-(
28) directly follow. The proof is complete. □
Lemma 6. Let . For any solution for operator inclusion (
16)
we have the following estimateswhere the constants , , do not depend on ε, v and ξ. Proof. Let
be a solution of (
16). Then it satisfies the following operator equality
We estimate the right-hand side of the last equality. By estimates (
18) and (
25) for
, we get:
We separately estimate the
. Using (
23), and the continuity of the embedding
, we have:
We rewrite inequality (
37) as follows
From the a priori estimates (
26) and (
28) it immediately follows that
To prove estimate (
32), it remains to use the left (
21) for
:
Hence, inequality (
32) is established.
We pass to the proof of estimate (
33). Represent the function
as follows:
Since the right-hand side of the resulting inequality does not depend on
t, we pass to the maximum in
in the left-hand side. Then, taking into account estimate (
32), we obtain
Thus, we received estimate (
33).
Now we prove inequality (
34). As before,
is a solution of operator Equation (
36). Then
We separately consider the terms on the right-hand side of the last inequality. First, we estimate
. Given from Reference [
34] the well-known inequality for
and estimate (
23), we obtain (for the case
the proof is similar):
Consider the following term. We use the Holder inequality and estimate (
18). Then
Similarly, using the Holder inequality and estimate (
25) for
, we obtain an estimate for the next term:
Finally, we consider the last term. Using inequality (
20), we get:
Let us estimate the right-hand side of the last inequality. We use the left side of estimate (
22) for
. Thus, to obtain an estimate of
, it is necessary to obtain an estimate of
. To do this, we again use operator Equation (
36). From its appearance, it follows that
From (
38), estimates (
39)–(
41) and a priori estimates (
26) and (
27), we get
This completes the proof of inequality (
34), where
.
Finally, applying again estimates (
39) and (
40), for the right-hand side of (
41), as well as a priori estimates (
26) and (
27), we obtain
Thus, inequality (), where is established. The proof is complete. □
Lemma 7. Let . Then for any solution of operator Equation (
16)
we have the following estimate:where is a constant that depends on ε. Theorem 6. Let . Then there is at least one solution of auxiliary problem (
10)–(
14), (
8)
for . Proof. To prove this theorem, we use the topological degree theory for multi-valued vector fields [
2,
43]. Consider operator inclusion (
17). From Corollary 7 it follows that all solutions of inclusion (
17) are in the ball
of radius
centered at zero. By item 4 of Lemma 1 the operator
is invertible. Then there is no solution of the family of following inclusions
on the boundary of the same ball
.
By virtue of item 4 of Lemma 1 the operator is continuous. By the Lemmas 2 and 4 the map is L - condensing with respect to the Kuratowski non-compactness measure. Therefore, the operator is condensing with respect to the Kuratowski non-compactness measure.
Thus, the vector field
is non-degenerate on the boundary of the ball
, which means that the topological degree
is defined for this vector field. By the properties of homotopy invariance and normalization of degree we obtain that
The non-zero degree of the mapping ensures the existence of at least one solution
of inclusion (
17) for
, and therefore of auxiliary problem (
8), (
10)–(
14) for
. The theorem is proved. □