Ergodicity of a nonlinear stochastic reaction-diffusion equation with memory
Abstract.
We consider a class of semi-linear differential Volterra equations with memory terms, polynomial nonlinearities and random perturbation. For a broad class of nonlinearities, we show that the system in concern admits a unique weak solution. Also, any statistically steady state must possess regularity compatible with that of the solution. Moreover, if sufficiently many directions are stochastically forced, we employ the generalized coupling approach to prove that there exists a unique invariant probability measure and that the system is exponentially attractive. This extends ergodicity results previously established in [Bonaccorsi et al., SIAM J. Math. Anal., 44 (2012)].
1. Introduction
1.1. Overview
Let be a bounded open domain with smooth boundary. We consider the following system in the unknown variable
(1.1) | ||||
modeling the temperature of a heat flow by conduction in viscoelastic materials [17, 18, 28, 41]. In (1.1), is the heat flux memory kernel satisfying , is a nonlinear term with polynomial growth, is a cylindrical Wiener process, and is a linear bounded map on certain Hilbert spaces. For simplicity, we set all physical constants to 1.
In the absence of memory effects, that is when , we note that (1.1) is reduced to the classical stochastic reaction-diffusion equation
(1.2) | ||||
Under different assumptions on as well as the noise term, statistically steady states of (1.2) are well studied. Using a Krylov-Bogoliubov argument together with tightness, (1.2) always admits at least one invariant probability measure [22]. For both Lipschitz and dissipative assumptions on , it is a classical result that such an invariant measure is unique for (1.2) [10, 21, 22]. Under more general conditions on , provided that noise is sufficiently forced in many directions of the phase space, (1.2) satisfies the so-called asymptotic strong Feller property and hence exponentially attractive toward equilibrium [31].
On the other hand, in the presence of memory kernels without random perturbation, i.e., , the well-posedness of the following equation
(1.3) | ||||
was studied as early as in the work of [40]. Moreover, positivity of the solutions for (1.3) as well as large-time asymptotic results were established in [15, 16]. The topic of global attractors was explored for a variety of deterministic systems related to (1.3) [12, 13, 19, 20, 43]. Equation (1.3) was also the motivation for many works of differential Volterra equations in the context of memory [2, 3, 4, 5].
Although there is a rich literature on (1.2) and (1.3), much less is known about asymptotic behaviors of the stochastic equation (1.1). To the best of the author’s knowledge, first result in this direction seems to be established in the work of [6]. For a wide class of nonlinearities, mild solutions of (1.1) was constructed [6, Section 4.2] via the classical Yosida approximation. Moreover, under the assumption of either or dissipative nonlinearity, that is , it was shown that (1.1) admits a unique invariant probability measure and that the system is mixing [6, Theorem 5.1]. Similarly method was also employed in [8, 9] to prove the existence of random attractors. The goal of this note is to make further progress toward statistically steady states of (1.1) under more general conditions on and the stochastic forcing.
In general, due to the memory effect, as in (1.1) is not a Markov process. It is thus convenient to transform (1.1) to a Cauchy problem on suitable spaces accounting for the whole past information. To see this, following the framework in [6, 40], we introduce the auxiliary memory variable given by
and observe that formally satisfies the following transport equation
(1.4) |
So that the above equation together with (1.1) forms a Markovian dynamics in the pair (see (2.14) below). It is thus necessary to construct memory spaces in which the dynamic evolves over time (see Section 2 below). The method of augmenting (1.1) by memory variables and spaces was employed as early as in the work of [40] and later popularized in [6]. One of the main difficulties dealing with these memory spaces, however, is the lack of compact embeddings that are typically found in classical Sobolev spaces. Therefore, when it comes to studying large-time asymptotics, it is a challenging problem to prove the existence of invariant measures via the Krylov-Bogoliubov procedure. To circumvent the issue, in [6], based on the assumption that , it was shown that the invariant measure uniquely exists. Because of such dissipative condition, however, the method employed therein does not cover more general potential instances, e.g., the Allen-Cahn function . On the other hand, in the absence of memory, it is well-known that one can employ the so-called generalized coupling argument [7, 31, 37, 36] to study invariant structures of (1.2) with the Allen-Cahn function. One of our goal here therefore is to generalize the ergodic results previously established in [6] making use of the framework from [7, 31, 37, 36] applied to our settings in the context of memory.
Unlike the mild solution approach of [6, 8, 9], we will study the well-posedness of (1.1) in a weak formulation. More specifically, under a general condition on the nonlinearities, we will construct the solution via a Galerkin approximation. As a byproduct of doing so, we are able to obtain useful moment bounds on the solutions. In turn, this will establish our first main result about the regularity of statistically steady states that is compatible to that of the weak solutions, cf. Theorem 3.8. In the second main result concerning unique ergodicity, under a slightly stronger assumption on the random perturbation, namely, noise is sufficiently forced in many Fourier directions, we employ the approach of generalized coupling developed in [7, 31] to prove the existence and uniqueness of the invariant probability measure to which the system (1.1) is exponentially attracted in suitable Wasserstein distances, cf. Theorem 3.11.
1.2. Physical motivation
Equation (1.1) is derived from the heat conduction in a viscoelastic material such as polymer [18, 41]. More specifically, by Fick’s law of balance of heat, it holds that [6]
where represents the internal energy, is an external heat supply and is the heat flux of the form
In Fourier conductors, , where is the instantaneous conductivity constant. So that, is reduced to
which yields the classical nonlinear heat equation (1.2) (by setting ). This amounts to the assumption that there is no time correlation between the heat transfer and the surrounding medium. However, as pointed out elsewhere [18, 17, 28], the above type of does not account for memory effects in viscoelastic materials, particularly at low temperature [41]. It is thus more appropriate to consider given by
where is the memory kernel that is decreasing on . With regard to the sign of the memory integral, there have been many works considering , as representing the thermal drag in addition to instantaneous conductivity. See [8, 9, 18, 19, 17, 20, 28, 41] and the references therein. On the other hand, [6, 15, 16] as well as this note study , which yields (1.1) by setting . It is worth to mention that for physical reasons, one should require [15]
Later in Sections 4-6, we will see that the above condition is actually critical in the analysis of (1.1).
Finally, we remark that in literature, stochastic equations with memory was studied as early as in the seminal work of [34] for finite-dimensional settings. The theory of stationary solutions was then developed further in [1] and was employed to study a variety of finite-dimensional and infinite-dimensional settings, e.g., Navier-Stokes equation [24], Ginzburg-Landau equation [23] and Langevin equation [33]. In particular, heat equation with memory was studied in a series of work in [2, 3, 4, 5]
The rest of the paper is organized as follows: in Section 2, we introduce all the functional settings needed for the analysis. In particular, we will see that (1.1) induces a Markovian dynamics as an abstract Cauchy equation evolving on an appropriate product space. In Section 3, we introduce the main assumptions on the non-linearities and noise structure. We also state our main results in this section, including Theorem 3.8 on the regularity of invariant measures and Theorem 3.11 on geometric ergodicity. In Section 4, we collect a priori moment bounds on the solutions that will be employed to prove the main results. In Section 5, we establish regularity of an invariant measure. We then discuss the coupling approach and prove geometric ergodicity in Section 6. In Appendix A, we employ Galerkin approximation to construct the solutions of (1.1).
2. Functional setting
Letting be a smooth bounded domain in , we denote by the Hilbert space endowed with the inner product and the induced norm .
Let be the realization of in endowed with the Dirichlet boundary condition and the domain . It is well-known that there exists an orthonormal basis in that diagonalizes , i.e.,
(2.1) |
for a sequence of positive numbers diverging to infinity.
For each , we denote
(2.2) |
endowed with the inner product
In view of (2.1), the inner product in may be rewritten as [11, 19, 20]
The induced norm in then is given by
It is well-known that the embedding is compact for . For , we denote by the projection onto the first wavenumbers , i.e.,
(2.3) |
The above projection will be useful in Sections 5 and 6 when we study the asymptotic behavior of (1.1).
In order to treat the memory term of (1.1) on an extended phase space, following the framework in [6], we introduce the function given by
(2.4) |
and the weighted space
(2.5) |
endowed with the inner product
(2.6) |
As mentioned in Section 1.1, unlike the compact embedding , , the embedding is only continuous [19, 20].
Next, we define to be the product space given by
(2.7) |
endowed with the norm
To simplify notation, we shall use and instead of and , respectively. For , the projections of onto the marginal spaces are given by
(2.8) |
On the Hilbert space , we consider the operator defined for
(2.9) |
with the domain [6, Section 2.3]
It can be shown that generates a strong continuous semigroup of contractions in [6, Theorem 2.5]. Concerning the operator , by the choice of as in (2.4), observe that
(2.10) |
We will make use of the above identity later in the analysis of (1.1). Furthermore, given , the following transport equation
(2.11) |
admits a unique solution given by (see [6, Expression (2.17)] and [44, Proposition 1.2])
(2.12) |
Also, for such that , by integration by parts, we note that [6]
(2.13) |
3. Main results
3.1. Well-posedness
We begin this section by stating the following condition on the kernel :
Assumption 3.1.
The kernel satisfies
(3.1) |
and
(3.2) |
for some constant . Furthermore,
(3.3) |
Remark 3.2.
Concerning the noise term, we assume that is a cylindrical Wiener process of the form
where are as in (2.1) and is a sequence of mutually independent Brownian motions, all defined on the same stochastic basis . With regard to the operator , we impose the following condition: [11, 26]
Assumption 3.3.
The operator is diagonalized by , i.e., there exists a sequence such that
(3.5) |
Furthermore,
(3.6) |
Assumption 3.4.
The function is satisfying . Moreover, the followings hold:
1. There exist such that
(3.7) |
3. There exists such that
(3.9) |
A concrete example of is the class of odd–degree polynomials with negative leading coefficients, i.e.,
for some constants , and are constants.
In light of relation (2.13) together with Assumptions 3.1, 3.3 and 3.4, we are now in a position to define weak solutions for (2.14).
Definition 3.5.
We now state the following proposition giving the existence of a solution for (2.14).
Proposition 3.6.
We remark that in [6], under a stronger assumption on the nonlinearities, the authors studied the notion of mild solutions for (2.14). The existence of such solutions was established using a classical Yosida approximation for SPDEs that can be found in literature [6, Section 4.2] (see also [10, 21, 22]). Similar method was also employed in the work of [8, 9]. On the other hand, in this note, we will construct the weak solutions for (2.14) via a Galerkin approximation, following the framework in [27, 45]. The explicit construction will be presented later in Appendix A.
3.2. Geometric ergodicity
We now turn to the main topic of the paper concerning statistically steady states of (2.14).
Given the well-posedness result in the previous subsection, we can thus introduce the Markov transition probabilities of the solution by
which are well-defined for , initial states and Borel sets . Letting denote the set of bounded Borel measurable functions , the associated Markov semigroup is defined and denoted by
(3.13) |
Remark 3.7.
Recall that a probability measure is said to be invariant for the semigroup if for every
In literature, the existence of invariant probability measures is typically established via the Krylov-Bogoliubov argument combined with the tightness of a sequence of auxiliary probability measures [1, 24, 31, 32, 34]. However, as mentioned in Section 2, since the embedding of , , is only continuous [19, 20], it is not clear whether under the same hypothesis of Proposition 3.6, an invariant probability measure exists. Nevertheless, we are able to assert the following moment bounds of any such .
Theorem 3.8.
Following the framework of [7, 11, 29, 30, 32], we recall that a function is called distance-like if it is symmetric, lower semi-continuous, and [32, Definition 4.3]. Let be the Wasserstein metric in associated with , defined by
(3.15) |
where
By the dual Kantorovich Theorem, it is well-known that
(3.16) |
where the infimum is taken over all pairs such that and . In our settings, we will particularly pay attention to the following two distances in : the former is the discrete metric, i.e., . The corresponding is the usual total variation distance, denoted by . The latter is the distance , , given by [7, 32, 36, 37]
(3.17) |
which we will employ to estimate the convergent rate of (2.14) toward equilibrium. The relation between and will be become clearer in Section 6, cf. Lemma 6.7.
As mentioned in Section 1.1, given the generality of the potential , we will make use of the noise term that is sufficiently forced in many directions of the phase space. So that can be dominated by the noise together with the Laplacian. More precisely, we make the following additional assumption on and : [11, 26]
Assumption 3.9.
Remark 3.10.
We note that Assumption 3.9 is actually a condition about the noise structure. Since , cf. (3.9), , cf. (3.1), and the sequence of eigenvalues as in (2.9) is diverging to infinity, there always exists an index such that condition (3.18) holds. We then require that noise be forced in at least directions, , hence the condition (3.19).
We are now in a position to state the main result of the paper ensuring geometric ergodicity of (2.14).
Theorem 3.11.
4. A priori bounds of (2.14)
Throughout the rest of the paper, and denote generic positive constants that may change from line to line. The main parameters that they depend on will appear between parenthesis, e.g., is a function of and .
In this section, we collect several useful a priori moment bounds on the solutions of (2.14). These results will be employed to prove Theorems 3.8 and 3.11 in later sections.
We start off by setting
(4.1) |
Recalling condition (3.1), we observe that and are both positive. In Lemma 4.1 below, we assert two moment bounds in .
Lemma 4.1.
Under the same hypothesis as in Proposition 3.6, the followings hold:
1.
(4.2) |
for some positive constants and independent of initial condition and time .
2. For all
(4.3) |
there exist positive constants and independent of and such that
(4.4) |
Proof.
To simplify notations, throughout the proof, we will omit the subscript in and .
Denote by the function given by
(4.5) |
A routine calculation gives
Recalling condition (3.6), we have
Concerning the convolution involving , by Young’s inquality,
where is as in (4.1). Recalling the identity (2.10) and the fact that , it holds that
Also, since satisfies (3.2) and , it holds that
(4.6) |
Recalling as in (4.1), we then derive the bound
(4.7) | |||
(4.8) |
To bound the non-linear term, we invoke (3.9) to see that
In the above, denotes the volume of in . Collecting everything now yields the estimate
(4.9) |
which proves (4.2) by virtue of Gronwall’s inequality.
With regard to (4.4), for to be chosen later, the partial derivatives of along the direction of is given by
and
Then, by Ito’s formula,
where is as in Assumption 3.3. We note that
Together with the estimates as in the proof of (4.2), we arrive at the bound
Since , we infer
for some positive constants and independent of . We now employ the elementary inequality
to arrive at the bound
whence
In the above, and are positive constants independent of initial condition and time . This establishes (4.4), thereby concluding the proof. ∎
Next, we state and prove Lemma 4.2 giving moment bounds in of the solutions.
Lemma 4.2.
For all , , and , it holds that
(4.10) |
and
(4.11) |
where is as in (4.3), and are positive constants independent of and .
Proof.
We proceed to prove (4.10) by induction on .
We first start with the base case and set
(4.12) |
A routine calculation yields
(4.13) | ||||
In the above,
by virtue of condition (3.6). Similarly to (2.10), we have
where in the last implication above, we employed again the fact that . Also,
where is given by (4.1). Using an argument similarly to (4.8), we have the bound
where , are respectively as in (3.2), (4.1) and is the first eigenvalue of . With regard to the nonlinear term in (4.13), we invoke Assumption 3.4 to see that
Combining the above estimates with (4.13), we arrive at the moment bound
(4.14) |
To estimate the term involving on the above right-hand side, we invoke (4.2) to see that
We then infer the estimate
and
Now assume (4.10)-(4.11) hold for up to . Let us consider the case . We first compute partial derivatives of in :
and
By Ito’s formula, the following holds
Similarly to the estimates in the base case , we readily have
Using -Young’s inequality, we note that
Likewise,
By taking small enough, we arrive at the following useful estimate in expectation
(4.15) | ||||
In view of Lemma 4.1,
(4.16) |
Combining (4.16) and (4.15) and integrating with respect to time, we obtain
Recalling as in (4.12), estimate (4.10) immediately follows from the above inequality. Also, by variation formula,
which proves (4.11). The proof is thus finished. ∎
5. Moment bounds of invariant probability measures
In this section, we provide the proof of Theorem 3.8 giving regularity of any invariant measure . Following the framework in [25], we introduce the auxiliary system
(5.1) | ||||
In the above, is as in (4.1), is the eigenvalue of associated with as in (2.1), and is the projection on to as in (2.3). Also, we chose sufficiently large such that
(5.2) |
Remark 5.1.
We note that the choice of as in (5.2) is always valid owing to the fact that , cf. (3.9), , cf. (3.1), and is diverging to infinity. Although (5.2) seems similarly to (3.18) as in Assumption 3.9 for geometric ergodicity, cf. Remark 3.10, for higher regularity properties of invariant probability measures, we do not require noise be forced in enough many directions. Only the same conditions for the well-posedness as in Proposition 3.6 are sufficient to prove Theorem 3.8.
It is worth to point out that system (5.1) only differs from (2.14) by the appearance of the linear term . More importantly, since (5.1) starts from the origin, it enjoys better regularity compared with (2.14). This is precisely summarized in the following lemma whose proof will be deferred to the end of this section.
Lemma 5.2.
Let be the process as in (5.1). For all and , the followings hold:
(5.3) |
and
(5.4) |
for some positive constants and independent of and .
Assuming the above result, we are now ready to conclude Theorem 3.8.
Proof of Theorem 3.8.
We first show that
(5.5) |
for all , cf. (4.3). To see this, for , consider such that
where
Given , we set . By invariance of , since is bounded,
Also, by the choice of , we have
Considering , we invoke (4.4) to see that
whence
It follows that for all we have the bound
We may take small and then take t sufficiently large to arrive at the following uniform bound in :
The Dominated Convergence Theorem then implies (5.5).
Next, we aim to show that must concentrate in . To do so, for , we set
We invoke invariance of again to see that
(5.6) |
Recalling the pair solving the auxiliary system (5.1), we estimate as follows:
In light of (5.3) and Sobolev embedding, we have
Also, we invoke (5.4) to see that
It follows that
where and are independent of and . Combining with (5.6), we infer that
By virtue of (5.5), we readily have , implying
Since the above inequality holds for arbitrarily , we may take sufficiently large so that . As a consequence, we obtain the following uniform bound in
We invoke the Monotone Convergence Theorem again to obtain
which proves that .
We now give the proof of Lemma 5.2.
Proof of Lemma 5.2.
We start with (5.3) and set , . Observe that satisfies the following deterministic equation with random coefficients
It follows that (recalling as in (4.5))
Similarly to (4.7), we readily have
In light of condition (3.9),
As a consequence, we obtain the almost sure bound
which together with the choice of as in (5.2) and clearly implies (5.3).
With regard to (5.4), we employ an argument similarly to the proof of Lemma 4.2 (in the base case ). In particular, following (4.14), we have the bound
Concerning the cross term on the above right-hand side, we employ Cauchy-Schwarz inequality to see that
It follows that
Gronwall’s inequality then implies
This establishes (5.4), thereby finishing the proof. ∎
6. Geometric Ergodicity of (2.14)
As mentioned in Section 1.1, the proof of Theorem 3.11 makes use of the generalized coupling in [7, 26, 32, 36, 37, 38]. The method is based on two important concepts: a suitable distance in that is contracting for the semigroup and a -small set to which the Markov process returns often enough. Given these ingredients, one is able to conclude the existence and uniqueness of an invariant probability measure . Moreover, the convergent rate toward can be quantified by the recurrence rate, i.e., how quickly the system returns to the -small set [7, 32, 36]. In turn, this can be estimated via suitable Lyapunov functions.
For the reader convenience, we recall the notions of contracting distances, d-small sets and Lyapunov functions [7, 32].
Definition 6.1.
A distance-like function bounded by 1 is called contracting for if there exists such that for any with , it holds that
(6.1) |
Definition 6.2.
A set is called -small for if for some ,
(6.2) |
Definition 6.3.
A function is called a Lyapunov function for if
1. as .
2. There exist positive constants independent of and such that
In Lemma 6.4 below, we assert the existence of a contracting distance and the corresponding -small set. The proof of Lemma 6.4 will be deferred to the end of this section.
Lemma 6.4.
Proof of Theorem 3.11.
We first recall as in (4.5). By integrating both sides of (4.9) with respect to time, we obtain
So that plays the role of the required Lyapunov function in the sense of Definition 6.3. In light of [7, Proposition 2.1], we combine the above estimate with Lemma 6.4 to conclude the unique existence of as well as the convergent rate (3.20). ∎
We now turn to the proof of Lemma 6.4. Before diving into the details, it is illuminating to review the generalized coupling argument in [7, 32]. In order to establish required bounds on the Wassertstein distance between and , it is sufficient to compare the two solutions and of (2.14) starting from and , respectively. However, we will not do so directly. To help with the analysis, we will consider instead an auxiliary system, denoted by , obtained from (2.14) by shifting the Wiener process in a suitably chosen direction. We note that the change of measures is valid thanks to the assumption that noise is sufficiently forced in enough many directions in , cf. Assumption 3.9. As it turns out, the choice of allows us to deduce two crucial estimates: firstly, and can be arbitrarily close to one another as tends to infinity. Secondly, and can be efffectively controlled. Altogether, we are able to conclude Lemma 6.4.
Recall from Assumption 3.9 that
where is as in (3.18). We set
(6.3) |
Observe that and are both positive by virtue of (3.18). Moreover,
(6.4) |
Next, we introduce the following “shifted” system
(6.5) | ||||
where is as in (6.3), is the projection on to as in (2.3), and is the -component in the solution of the original equation (2.14). Similarly to (5.1), we note that (6.5) only differs from (2.14) by the appearance of the term . For notational convenience, we denote
the solution of (6.5) with initial condition .
Three of the main ingredients in the generalized coupling argument are given in the following results to be proved at the end of this section.
Lemma 6.5.
Lemma 6.6.
Lemma 6.7.
For all probability measures , in , and ,
(6.9) |
where is the Wasserstein distance associated with as in (3.17).
Assuming the above results, we are ready to conclude Lemma 6.4 whose argument is based on [7, Proof of Theorem 2.4]. See also [32].
Proof of Lemma 6.4.
1. Suppose such that . Recalling as in (3.17), this implies that
By triangle inequality, for all ,
(6.10) | ||||
In view of (6.7) and (6.9), we readily have
where is given by (6.20) below. Concerning , by the dual formula (3.16), it holds that
(6.11) |
In the last estimate above, we made use of (6.6) with as in (6.16) below. Altogether, we deduce the bound for all
By choosing large enough such that
(6.12) |
we obtain
which establishes that is contracting for as claimed.
We now turn to the auxiliary results in Lemmas 6.5–6.6–6.7. To prove Lemma 6.5, we will mainly invoke the choice of the index as in (3.18) to derive the exponential estimate (6.6). In turn, we will combine (6.6) with the fact that is invertible in span, cf. (3.19), to conclude Lemma 6.6. Finally, the proof of Lemma 6.7 is quite standard relying on the fact that the discrete metric dominates .
Proof of Lemma 6.5.
To simplify notation, we will omit the subscripts in the proof.
For a slightly abuse of notation, we set and , from (6.5) and (2.14), and observe that
(6.14) | ||||
Following the same argument as in the proofs of Lemma 4.1 and (5.3), equation (6.14) implies (recalling as in (4.5))
Recalling ,
where is given by (6.3). Also, recalling (2.10),
We then deduce that
whence
To control the nonlinear term, we invoke condition (3.18) to infer
Altogether, we obtain
(6.15) |
where and are as in (6.3). Thanks to (3.18), the choice of satisfies (6.4), namely,
We now set
(6.16) |
which is positive. Estimate (6.6) now follows from (6.15)-(6.16).
∎
Next, we present the proof of Lemma 6.6.
Proof of Lemma 6.6.
In order to prove (6.7)-(6.8), we first consider the following cylindrical Wiener process
(6.17) |
where
(6.18) |
and and are the first components of and , respectively. Since is invertible on by virtue of condition (3.19), we note that
In the last estimate above, we employed (6.6) with given by (6.16). As a consequence,
(6.19) |
where
(6.20) |
In light of [7, Inequality (A.1) and Theorem A.2] together with (6.19), it holds that
(6.21) |
On the other hand, we invoke [7, Inequality (A.2)] to see that for all
(6.22) |
Now, observe that (6.21)-(6.22) imply (6.7)-(6.8) if we can show that
(6.23) |
To see this, consider any coupling for and denote by , respectively the solutions of (6.5) and (2.14) associated with and . It is clear that is a coupling for . We note that by the uniqueness of weak solution,
It follows that if
then
In particular,
By the dual formula (3.16), we establish (6.23), thereby concluding the proof. ∎
Acknowledgment
The author thanks Nathan Glatt-Holtz and Vincent Martinez for fruitful discussions on the topic of this paper. The author also would like to thank the anonymous reviewer for their valuable comments and suggestions.
Appendix A Well-posedness of (2.14)
In this section, we discuss Proposition 3.6 whose proof relies on the construction of the weak solutions for (2.14). We start with the Galerkin finite-dimensional approximation.
A.1. Finite-dimensional approximation
Recalling the projection onto the first wavenumbers as in (2.3), we set
We then consider the pair solving the following finite-dimensional system
(A.1) | ||||
By Ito’s formula, we have
Similarly to the a priori bounds in Section 4, e.g., the proof of Lemma 4.1, we proceed to estimate the above right-hand side as follows:
Concerning the non-linear term, we invoke (3.8) to see that
where is the exponent constant from Assumption 3.4.
Using Burkholder’s inequality, the Martingale term can be bounded in expectation by
Altogether, we arrive at the following estimate
(A.2) | ||||
In particular, this implies the existence and uniqueness of the strong solution for (A.1) [35, 39, 42]. Moreover, combining Burkholder’s and Gronwall’s inequalities, we deduce the following bound in sup norm and
for some positive constant independent of and initial condition . Also, setting , (3.7) combined with (A.2) implies that
(A.3) |
Furthermore, since solves the transport equation in
admits the following representation [6]
(A.4) |
A.2. Passage to the limit
As a consequence of the preceding subsection, we deduce the following limits (up to a subsequence)
Furthermore, (see [45, pg. 224])
Next, we proceed to prove that a.s. satisfies (3.12), i.e.,
(A.5) |
To see this, consider any arbitrary . We first note that
which converges to
as tends to infinity, since in . Also, for each , Cauchy-Schwarz inequality and the fact that is decreasing on yield the bound
which is a.s. integrable on . The Dominated Convergence Theorem then implies a.s.
(A.6) |
as . On the other hand, since converges to in , for all , it holds that
Using the Dominated Convergence Theorem again, we obtain a.s.
(A.7) |
We now combine (A.6) and (A.7) to deduce
in . This proves (3.12) by the uniqueness of weak limit.
Next, we turn to establish (3.11). Considering such that , we multiply both sides of -equation in (A.1) with and perform integration by parts to obtain
In the last implication above, we employed the fact that . Integrating the above equation with respect to time yields
By sending to infinity, we deduce the identity (3.11) provided that
(A.8) |
To see this, recall from (3.4) that
Thus, , whence
We are left to establish (3.10). Considering any , it holds that
(A.9) | ||||
We first claim that
To this end, recalling (3.4) again, we see that for all , , implying
In other words, . It follows that as ,
As a consequence, sending in (A.9) yields
It therefore remains to establish that . The argument follows along the lines of [45, Theorem 8.4] tailored to our setting (see also [9, 27]).
We first consider the pair and note that they obey the finite-dimensional system
(A.10) | ||||
Setting and subtracting (A.1) from (A.10), observe that
(A.11) | ||||
It follows that
Similarly to (4.7), we readily have the bound
where we recall and as in (4.1). Concerning the non-linear term involving , we write
Concerning , since both and converge weakly to in , we obtain the limit
Also, using Holder’s inequality,
for some positive constant . To bound , we note that since , condition (3.7) implies . Similarly to , we have
and
With regard to , since is uniformly bounded in , cf. (A.3), it holds that
In particular, we have
and
Concerning , we invoke condition (3.9) again to infer
Setting
by Gronwall’s inequality, we infer the bound a.e. in
for some positive constant . We observe that converges to zero and that
By the Dominated Convergence Theorem, it holds that
whence for a.e.
We invoke the Dominated Convergence Theorem again to deduce
In particular, this implies that (up to a subsequence)
and thus, (up to a further subsequence) converges to a.e. . It follows that converges to a.e. since is continuous. In view of [45, Lemma 8.3], we obtain
whence a.s., a.e. . This concludes the proof of Proposition 3.6.
References
- [1] Y. Bakhtin and J. C. Mattingly. Stationary solutions of stochastic differential equations with memory and stochastic partial differential equations. Communications in Contemporary Mathematics, 7(05):553–582, 2005.
- [2] V. Barbu. Nonlinear Volterra equations in a Hilbert space. SIAM Journal on Mathematical Analysis, 6(4):728–741, 1975.
- [3] V. Barbu. Nonlinear semigroups and differential equations in Banach spaces. Noordhoff, Leyden, The Netherlands, 1976.
- [4] V. Barbu. Existence for nonlinear Volterra equations in Hilbert spaces. SIAM Journal on Mathematical Analysis, 10(3):552–569, 1979.
- [5] V. Barbu. Nonlinear differential equations of monotone types in Banach spaces. Springer Science & Business Media, 2010.
- [6] S. Bonaccorsi, G. Da Prato, and L. Tubaro. Asymptotic behavior of a class of nonlinear stochastic heat equations with memory effects. SIAM Journal on Mathematical Analysis, 44(3):1562–1587, 2012.
- [7] O. Butkovsky, A. Kulik, and M. Scheutzow. Generalized couplings and ergodic rates for SPDEs and other Markov models. The Annals of Applied Probability, 30(1):1–39, 2020.
- [8] T. Caraballo, I. Chueshov, P. Marín-Rubio, and J. Real. Existence and asymptotic behaviour for stochastic heat equations with multiplicative noise in materials with memory. Discrete & Continuous Dynamical Systems-A, 18(2&3):253, 2007.
- [9] T. Caraballo, J. Real, and I. Chueshov. Pullback attractors for stochastic heat equations in materials with memory. Discrete & Continuous Dynamical Systems-B, 9(3&4, May):525, 2008.
- [10] S. Cerrai. Second order PDE’s in finite and infinite dimension: a probabilistic approach, volume 1762. Springer Science & Business Media, 2001.
- [11] S. Cerrai and N. Glatt-Holtz. On the convergence of stationary solutions in the Smoluchowski-Kramers approximation of infinite dimensional systems. Journal of Functional Analysis, 278(8):108421, 2020.
- [12] M. D. Chekroun, F. Di Plinio, N. E. Glatt-Holtz, and V. Pata. Asymptotics of the Coleman-Gurtin model. arXiv preprint arXiv:1006.2579, 2010.
- [13] M. D. Chekroun and N. E. Glatt-Holtz. Invariant measures for dissipative dynamical systems: Abstract results and applications. Communications in Mathematical Physics, 316(3):723–761, 2012.
- [14] P. Clément and G. Da Prato. White noise perturbation of the heat equation in materials with memory. Dynam. Systems Appl., 6:441–460, 1997.
- [15] P. Clément, R. MacCamy, and J. A. Nohel. Asymptotic properties of solutions of nonlinear abstract Volterra equations. The Journal of Integral Equations, pages 185–216, 1981.
- [16] P. Clément and J. A. Nohel. Asymptotic behavior of solutions of nonlinear Volterra equations with completely positive kernels. SIAM Journal on Mathematical Analysis, 12(4):514–535, 1981.
- [17] B. D. Coleman and M. E. Gurtin. Equipresence and constitutive equations for rigid heat conductors. Zeitschrift für angewandte Mathematik und Physik ZAMP, 18(2):199–208, 1967.
- [18] B. D. Coleman and W. Noll. Material symmetry and thermostatic inequalities in finite elastic deformations. Archive for Rational Mechanics and Analysis, 15(2):87–111, 1964.
- [19] M. Conti, V. Pata, and M. Squassina. Singular limit of dissipative hyperbolic equations with memory. DYNAMICAL SYSTEMS, 2005:200–208.
- [20] M. Conti, V. Pata, and M. Squassina. Singular limit of differential systems with memory. Indiana University mathematics journal, pages 169–215, 2006.
- [21] G. Da Prato and J. Zabczyk. Ergodicity for infinite dimensional systems, volume 229. Cambridge University Press, 1996.
- [22] G. Da Prato and J. Zabczyk. Stochastic equations in infinite dimensions. Cambridge university press, 2014.
- [23] W. E and D. Liu. Gibbsian dynamics and invariant measures for stochastic dissipative PDEs. Journal of Statistical Physics, 108(5-6):1125–1156, 2002.
- [24] W. E, J. C. Mattingly, and Y. Sinai. Gibbsian Dynamics and Ergodicity for the Stochastically Forced Navier–Stokes Equation. Communications in Mathematical Physics, 224(1):83–106, 2001.
- [25] N. Glatt-Holtz, V. R. Martinez, and G. H. Richards. On the long-time statistical behavior of smooth solutions of the weakly damped, stochastically-driven KdV equation. arXiv preprint arXiv:2103.12942, 2021.
- [26] N. Glatt-Holtz, J. C. Mattingly, and G. Richards. On unique ergodicity in nonlinear stochastic partial differential equations. Journal of Statistical Physics, 166(3-4):618–649, 2017.
- [27] N. Glatt-Holtz and M. Ziane. The stochastic primitive equations in two space dimensions with multiplicative noise. Discrete & Continuous Dynamical Systems-B, 10(4):801, 2008.
- [28] M. E. Gurtin and A. C. Pipkin. A general theory of heat conduction with finite wave speeds. Archive for Rational Mechanics and Analysis, 31(2):113–126, 1968.
- [29] M. Hairer and J. C. Mattingly. Ergodicity of the 2D Navier-Stokes equations with degenerate stochastic forcing. Annals of Mathematics, pages 993–1032, 2006.
- [30] M. Hairer and J. C. Mattingly. Spectral gaps in Wasserstein distances and the 2D stochastic Navier–Stokes equations. The Annals of Probability, 36(6):2050–2091, 2008.
- [31] M. Hairer and J. C. Mattingly. A theory of hypoellipticity and unique ergodicity for semilinear stochastic PDEs. Electronic Journal of Probability, 16:658–738, 2011.
- [32] M. Hairer, J. C. Mattingly, and M. Scheutzow. Asymptotic coupling and a general form of Harris’ theorem with applications to stochastic delay equations. Probability Theory and Related Fields, 149(1):223–259, 2011.
- [33] D. P. Herzog, J. C. Mattingly, and H. D. Nguyen. Gibbsian dynamics and the generalized Langevin equation. arXiv preprint arXiv:2111.04187, 2021.
- [34] K. Itô and M. Nisio. On stationary solutions of a stochastic differential equation. J. Math. Kyoto Univ., 4(3):1–75, 1964.
- [35] R. Khasminskii. Stochastic stability of differential equations, volume 66. Springer Science & Business Media, 2011.
- [36] A. Kulik. Ergodic Behavior of Markov Processes. de Gruyter, 2017.
- [37] A. Kulik and M. Scheutzow. Generalized couplings and convergence of transition probabilities. Probability Theory and Related Fields, pages 1–44, 2015.
- [38] J. C. Mattingly. Exponential convergence for the stochastically forced Navier-Stokes equations and other partially dissipative dynamics. Communications in Mathematical Physics, 230(3):421–462, 2002.
- [39] S. P. Meyn and R. L. Tweedie. Markov chains and stochastic stability. Springer Science & Business Media, 2012.
- [40] R. Miller. Linear Volterra integrodifferential equations as semigroups. Funkcial. Ekvac, 17:39–55, 1974.
- [41] J. W. Nunziato. On heat conduction in materials with memory. Quart. Appl. Math., 29(2):187–204, 1971.
- [42] B. Øksendal. Stochastic Differential Equations: An Introduction with Applications. Springer Science & Business Media, 2003.
- [43] V. Pata and A. Zucchi. Attractors for a damped hyperbolic equation with linear memory. 2001.
- [44] J. Prüss. Evolutionary Integral Equations and Applications, volume 87. Birkhäuser, 2013.
- [45] J. C. Robinson. Infinite-dimensional dynamical systems: an introduction to dissipative parabolic PDEs and the theory of global attractors, volume 28. Cambridge University Press, 2001.